Ferrofluids. Magnetically Controllable Fluids and Their Applications S. Odenbach
Ferrofluids. Magnetically Controllable Fluids and Their Applications S. Odenbach
Ferrofluids. Magnetically Controllable Fluids and Their Applications S. Odenbach
Editorial Board
R. Beig, Wien, Austria
B.-G. Englert, Ismaning, Germany
U. Frisch, Nice, France
P. Hänggi, Augsburg, Germany
K. Hepp, Zürich, Switzerland
W. Hillebrandt, Garching, Germany
D. Imboden, Zürich, Switzerland
R. L. Jaffe, Cambridge, MA, USA
R. Lipowsky, Golm, Germany
H. v. Löhneysen, Karlsruhe, Germany
I. Ojima, Kyoto, Japan
D. Sornette, Nice, France, and Los Angeles, CA, USA
S. Theisen, Golm, Germany
W. Weise, Trento, Italy, and Garching, Germany
J. Wess, München, Germany
J. Zittartz, Köln, Germany
3
Berlin
Heidelberg
New York
Barcelona
Hong Kong
London
Milan
Paris
Tokyo
Editorial Policy
The series Lecture Notes in Physics (LNP), founded in 1969, reports new developments in
physics research and teaching -- quickly, informally but with a high quality. Manuscripts
to be considered for publication are topical volumes consisting of a limited number of
contributions, carefully edited and closely related to each other. Each contribution should
contain at least partly original and previously unpublished material, be written in a clear,
pedagogical style and aimed at a broader readership, especially graduate students and
nonspecialist researchers wishing to familiarize themselves with the topic concerned. For
this reason, traditional proceedings cannot be considered for this series, though volumes
to appear in this series are often based on material presented at conferences, workshops
and schools (in exceptional cases the original papers and/or those not included in the
printed book may be added on an accompanying CD-ROM, together with the abstracts
of posters and other material suitable for publication, e.g. large tables, colour pictures,
program codes, etc.).
Acceptance
A project can only be accepted tentatively for publication, by both the editorial board and the
publisher, following thorough examination of the material submitted. The book proposal
sent to the publisher should consist of at least a preliminary table of contents outlining the
structure of the book together with abstracts of all contributions to be included.
Final acceptance is issued by the series editor in charge, in consultation with the publisher,
only after receiving the complete manuscript. Final acceptance, possibly requiring minor
corrections, usually follows the tentative acceptance unless the final manuscript differs
significantly from expectations (project outline). In particular, the series editors are entitled
to reject individual contributions if they do not meet the high quality standards of this
series. The final manuscript must be camera-ready, and should include both an informative
introduction and a sufficiently detailed subject index.
Contractual Aspects
Publication in LNP is free of charge. There is no formal contract, no royalties are paid,
and no bulk orders are required, although special discounts are offered in this case. The
volume editors receive jointly 30 free copies for their personal use and are entitled, as are the
contributing authors, to purchase Springer books at a reduced rate. The publisher secures
the copyright for each volume. As a rule, no reprints of individual contributions can be
supplied.
Manuscript Submission
The manuscript in its final and approved version must be submitted in camera-ready form.
The corresponding electronic source files are also required for the production process, in
particular the online version. Technical assistance in compiling the final manuscript can be
provided by the publisher’s production editor(s), especially with regard to the publisher’s
own LaTeX macro package which has been specially designed for this series.
Ferrofluids
Magnetically Controllable Fluids
and Their Applications
13
Editor
Stefan Odenbach
ZARM
Universität Bremen
Am Fallturm
28359 Bremen, Germany
This work is subject to copyright. All rights are reserved, whether the whole or part of the
material is concerned, specifically the rights of translation, reprinting, reuse of illustra-
tions, recitation, broadcasting, reproduction on microfilm or in any other way, and
storage in data banks. Duplication of this publication or parts thereof is permitted only
under the provisions of the German Copyright Law of September 9, 1965, in its current
version, and permission for use must always be obtained from Springer-Verlag. Violations
are liable for prosecution under the German Copyright Law.
Springer-Verlag Berlin Heidelberg New York
a member of BertelsmannSpringer Science+Business Media GmbH
http://www.springer.de
c Springer-Verlag Berlin Heidelberg 2002
Printed in Germany
The use of general descriptive names, registered names, trademarks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
Typesetting: Camera-ready by the authors/editor
Camera-data conversion by Steingraeber Satztechnik GmbH Heidelberg
Cover design: design & production, Heidelberg
Printed on acid-free paper
SPIN: 10886791 54/3141/du - 5 4 3 2 1 0
Preface
Magnetic control of the properties and the flow of liquids can be a challenging
field for basic research as well as for applications. An important condition for a
technically interesting magnetic control of fluids is that the requested magnetic
fields should be as small as possible.
An excellent material fulfilling this condition are suspensions of magnetic
nanoparticles - commonly called ferrofluids. These fluids exhibit coupled liquid
and superparamagnetic properties, leading to the possibility to control their
flows and properties with magnetic fields having a strength of about 10mT. After
the first synthesis of stable ferrofluids in 1964, a rapid development of ideas for
applications as well as of basic fluid mechanics investigations took place. Till now
the research field experiences a strong ongoing growth, recently triggered by the
establishment of a national research program on magnetic fluids in Germany and
documented by numerous presentations of the 9th International Conference on
Magnetic Fluids (ICMF9), held in Bremen in 2001.
One of the most challenging aspects of ferrofluid research is the interdisci-
plinarity of the field. Besides chemistry for the preparation of magnetic fluids,
basic theoretical physics for the description of their properties and behavior, fluid
physics and rheology for the investigation of flows and rheological properties un-
der influence of magnetic fields are needed to cover the basic research interests.
In addition engineering and medical applications contribute to the importance
of ferrofluid research for everyday life.
Most of these aspects are covered in this issue of Lecture Notes in Physics,
which is mainly based on a series of plenary talks given during ICMF9. Starting
with a review of preparation techniques for ferrofluids the chemical aspect of pro-
duction of stable magnetic colloids is highlighted, followed by two contributions
showing different approaches for a quantitative characterization of ferrofluids.
Part two of the issue is devoted to basic questions concerning the behavior
of ferrofluids in the presence of magnetic fields. Three contributions highlight
different ways to describe the fluids theoretically. Macro- and microscopic ap-
proaches are shown together to enable a comparison of their advantages and
problems. The section ends with a specific example of magnetic control - the
influence of fields on heat and mass transfer in ferrofluids.
In the third part, the question of magnetic control of the fluids properties is
addressed using the example of magnetoviscosity - i.e. the change of the viscous
properties of magnetic fluids in a magnetic field. The combination of theoret-
VI Preface
Ferrofluid Dynamics
H.W. Müller and M. Liu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Part IV Applications
Konstantin I. Morozov
A. Schmidt
Institute of Continuous Media
Department of Otorhinolaryngology
Mechanics
Head and Neck Surgery,
UB of Russian Academy of Sciences
Klinikum rechts der Isar
614013 Perm
Technical University of Munich
Russia 81675 Munich
[email protected]
Mark I. Shliomis
Hanns Walter Müller Department of Mechanical Engineer-
Max-Planck Institut für Polymer- ing
forschung Ben-Gurion University of the Negev
Ackermannweg 10 P.O.B. 653
D-55128 Mainz Beer-Sheva 84105
Germany Israel
[email protected] [email protected]
Stuart W. Charles
1 Ferrofluids
Ferrofluids are stable colloidal dispersions of nano-sized particles of ferro- or fer-
rimagnetic particles in a carrier liquid. A wide range of carrier liquids have been
employed, and many ferrofluids are commercially available to satisfy particular
applications, e.g. in rotary vacuum lead-throughs, it is essential that the carrier
liquid has a very low vapor pressure. In other applications, temperature, either
high, low or both, may be a critical consideration. Theoretically it should be
possible to produce dispersions in any liquid thereby being able to tailor the
requirements of viscosity, surface tension, temperature and oxidative stability,
vapor pressure, stability in hostile environments, etc., to the particular applica-
tion in mind, whether it be technological or biomedical.
When a ferrofluid is subjected to a magnetic field, magnetic field gradient
and/or gravitational field, in order that the colloidal suspension remains stable
the magnetic particles generally have to be of approximately 10 nm in diam-
eter. Particles of this size, whether they be ferrite or metal, possess a single
magnetic domain only, i.e., the individual particles are in a permanent state
of saturation magnetization. Thus a strong long-range magnetostatic attraction
exists between individual particles, the result of which would lead to agglomer-
ation of the particles and subsequent sedimentation unless a means of achieving
1
For further information and availability of Ferrofluids and Magnetorheological Flu-
ids the reader is recommended to view www.liquidsresearch.com or contact Liquids
Research Limited by e.mail [email protected]
substituted ferrites. Numerous papers and patents have been published on this
subject. See the references given in bibliographies quoted earlier. The method
of precipitation of the particles is essentially the same as that used to prepare
magnetite. Thus it is not important that details of individual preparations be
given here, except to say that in some cases the precipitate needs to be hy-
drothermally aged to facilitate the conversion of the precipitated hydroxides to
the ferrite [15]. The reason for the interest in substituted ferrite particles stems
from the wide-differences in the magnetic properties of these materials [14], [16].
Thus it is possible to tailor the properties of these particles and dispersions of
such particles to satisfy various applications. For example, where particles are re-
quired with high magnetocrystalline anisotropy, cobalt ferrites can be employed
[20]. Other ferrites, which have relatively low Neel temperatures ( 100o C) and
thus high thermomagnetic coefficients at these temperatures, have been used in
studies of thermomagnetic convection and heat transfer [22].
tion still showed the reduced saturation magnetization. However evidence from
X-ray diffraction indicated that the particles were poorly crystallized or amor-
phous with similar results being obtained for Fe particles produced by a similar
process [30]. X-ray diffraction and Mössbauer studies indicated the presence of
carbon in the particles which is thought responsible for the poorly crystallized
state. The particles, when annealed showed an X-ray pattern consistent with α-
Fe and χ-Fe5 C2 . Some researchers, using transmission electron diffraction, have
reported that the particles produced by thermolysis have a crystalline structure.
Whether this is a true indication of the structure of the particles or that an-
nealing has taken place in the electron beam is not known. Measurements of the
magnetic anisotropy of Co particles have been undertaken by several groups of
workers [31,29].
Iron particles of colloidal dimensions can also be produced by thermolysis
[30]; in this case of iron pentacarbonyl in decahydronaphthalene. Particle size
is again controlled by the presence of surfactants. The actual structure of the
particles is open to question. Some groups have reported a crystalline struc-
ture [32] whilst others [30] reported an amorphous structure which undergoes
crystallization on annealing.
Hoon et al. [33] have reported the preparation of colloidal nickel particles
by the irradiation of nickel carbonyl with ultraviolet light and by the reduc-
tion of nickelocene. The particles were reported to have an fcc structure from
transmission electron diffraction.
Nakatani et al. [34,35] have prepared iron nitride particles by two methods.
Firstly by a plasma CVD reaction of iron pentacarbonyl and nitrogen gas and
secondly by heating a solution of iron pentacarbonyl in kerosene containing a
surfactant through which a constant stream of ammonia is passed. The crystal
structure observed is -Fe3 N.
It is also possible to prepare alloy particles of the transition metals by the
thermolysis of mixed-metal organometallic compounds in the presence of surfac-
tants. Lambrick et al. [36,37] prepared organo-metallic compounds from which
particles of Fe-Co and Ni-Fe alloys with median diameter 8 nm were obtained.
The saturation magnetizations of the particles are >84% of that of the bulk
values.
appearance, see Fig. 1(b). Further the metals invariably have a narrow particle-
size distribution with a standard deviation of <0.2.
3 Preparation of Ferrofluids
For most applications it is absolutely essential that the ferrofluids be very stable
with regard to temperature and in the presence of magnetic fields. The presence
of agglomeration of particles must be avoided at all costs. The literature con-
tains numerous papers where phenomena can be attributed to the presence of
agglomerates and yet their presence has been omitted in any discussion of the
phenomena.
For most applications a low viscosity, low vapor pressure and chemical in-
ertness are desirable for the carrier liquid and surfactant. Scholten [44] gives
a review of the chemical and physical problems associated with these parame-
ters and lists the advantages and disadvantages associated with different carrier
liquids.
The Preparation of Magnetic Fluids 11
Magnetite can be coated with oleic acid by adding the acid to the precipitated
phase in alkaline solution at pH 9.5. The mixture is usually left stirring for
approximately 1 hour after which time it is heated to 95o C to facilitate the
conversion of hydroxides formed to ferrite. After cooling, the product is acidified
to pH 5 using nitric acid. The oleate ion produced chemisorbs strongly on to
the magnetite particles, and the precipitate coagulates and falls out of solution.
The coagulation may also be accompanied by the appearance of a thin film of
oleic acid on the surface of the supernatant liquid which is unadsorbed oleate
ions reconverted to oleic acid in acid solution. Sufficient oleic must be added
to form a monolayer coverage [20]. The supernatant liquid can be decanted
12 S.W. Charles
4 Magnetorheological Fluids
References
1. Zahn M, Shenton K E 1980 Magnetic fluids bibliography. IEEE Trans. Mag. MAG-
16(2): 387-415.
2. Charles, S W, Rosensweig R E 1983 Magnetic fluids bibliography. J. Magn. Magn.
Mat. 39: 190-220
3. Kamiyama S, Rosensweig R E 1987 Magnetic fluids bibliography. J. Magn. Magn.
Mat. 65: 401-39.
4. Blums E, Ozols R, Rosensweig R E 1990 Magnetic fluids bibliography. J. Magn.
Magn. Mat. 85: 303-82
5. Cabuil V, Neveu S, Rosensweig R E 1993 Magnetic fluids bibliography. J. Magn.
Magn. Mat. 122: 437-82
6. Bhatnagar S P, Rosensweig R E 1995 Magnetic fluids bibliography. J. Magn. Magn.
Mat. 149: 198-232
7. Vekas L, Sofonea V, Balau O 1999 Magnetic fluids bibliography. J. Magn. Magn.
Mat. 39: 454-89
8. Rosensweig R E 1979 Fluid dynamics and science of magnetic liquids. In: Advances
in Electronics and Electron Physics. Vol. 48: Martin M (ed.) Academic Press, New
York, pp. 103-99
9. Rosensweig R E 1985 Ferrohydrodynamics. Cambridge University Press, Cam-
bridge, UK
10. Rosensweig R E 1988 Special Issue on Ferrofluids. Chem. Eng. Comm. 67: 1-340
11. Charles S W, Popplewell J 1980 Ferromagnetic liquids. In: Wohlfarth E P (ed.)
Ferro-magnetic Materials Vol. 2: North-Holland, Amst. pp. 509-59
12. Martinet A 1983 The case of ferrofluids. In: Aggregation Processes in Solution,
Elsevier, New York, Ch. 18, pp. 1-41
13. Scholten P C 1978 In: Berkovsky B M (ed.) Thermomechanics of magnetic fluids.
Hemisphere, Washington D C, pp. 1-26
14. Smit J, Wijn H P J 1959 Ferrites. Wiley, New York
15. Schuele W J, Deetscreek V D 1963 Ultrafine particles. Kuhn W E (ed.) Wiley, New
York
16. Dormann J-L, Nogues M 1990 Magnetic structure of substituted ferrites. J. Phys:
Condens. Mat. 2: 1223-37
17. Papell S S 1965 Low viscosity magnetic fluid obtained by the colloidal suspension
of magnetic particles. U S Patent 3,215,572
18. Khalafalla S E, Reimers G W 1973 Magnetofluids and their manufacture. U S
Patent 3,764,540
19. Massart R 1981 Preparation of aqueous magnetic liquids in alkaline and acidic
media. IEEE Trans. Mag. MAG-17: 1247-8
The Preparation of Magnetic Fluids 17
20. Davies K J, Wells, S, Charles S W 1993 The effect of temperature and oleate
adsorption on the growth of maghemite particles. J. Magn. Magn. Mat. 122: 24-8
21. Bee A, Massart R, Neveu S 1995 Synthesis of very fine maghemite particles. J.
Magn. Magn. Mat. 149: 6-9
22. Nakatsuka K, Hama Y, Takahashi J 1990 Heat transfer in temperature-sensitive
magnetic fluids. J. Magn. Magn. Mat. 85: 207-9
23. Pileni M P 1989 Structure and Reactivity of Reverse Micelles, Elsevier, Amst.
24. Gobe M, Kon-No K, Kitahara A 1984 Preparation of magnetite superfine sol in
w/o microemulsion. J. Coll. Int. Sci 93: 293-5
25. Thomas J R 1966 Preparation and magnetic properties of colloidal cobalt particles.
J. Appl. Phys. 37(7): 2914-15
26. Hess P H, Parker P H 1966 Polymers for stabilization of colloidal cobalt particles.
J. Appl. Polymer Sci 10: 1915-17
27. Smith T W 1981 Surfactant-catalyzed decomposition of dicobalt octacarbonyl. U S
Patent 4,252,673
28. Papirer E, Horny P, Balard H, Anthore R, Petipas R, Martinet A 1983 The prepa-
ration of a ferrofluid by the decomposition of dicobalt octacarbonyl. J. Coll. Int.
Sci. 94: 207-20
29. Charles S W, Wells S 1990 Magnetic properties of colloidal suspensions of cobalt.
Magnetohydrodynamics 26: 288-92
30. Wonterghem J van, Mørup S, Charles S W, Wells S, Villadsen J 1986 Formation of
a metallic glass by thermal decomposition of Fe(CO)5 . Phys. Rev. Lett. 55: 410-13
31. Chantrell, R W, Popplewell J, Charles S W 1980 The coercivity of a system of
single domain particles with randomly oriented easy axes. J. Magn. Magn. Mat.
15-18: 1123-4
32. Kilner M, Hoon S R, Lambrick D R, Potton J A, Tanner B K 1984 Preparation
and properties of metallic iron ferrofluids. IEEE Trans. Mag. MAG-20: 1735-7
33. Hoon S R, Kilner M, Russell G J, Tanner B K 1983 Preparation and properties of
nickel ferrofluids. J. Magn. Magn. Mat. 39: 107-10
34. Nakatani I, Furubayashi T 1990 Iron-nitride magnetic fluids prepared by plasma
CVD technique. J. Magn. Magn. Mat. 85: 11-3
35. Nakatani I, Hijikata M, Ozawa K 1993 Iron-nitride fluids prepared by vapor-liquid
reaction. J. Magn. Magn. Mat. 122: 10-4
36. Lambrick D B, Mason N, Harris N J, Russell G J, Hoon S R, Kilner M 1985 An
iron-cobalt alloy magnetic fluid. IEEE Trans. Mag. MAG-21: 1891-3
37. Lambrick D B, Mason N, Hoon S R, Kilner M 1987 Preparation and properties of
Ni-Fe magnetic fluids. J. Magn. Magn. Mat. 65: 257-60
38. Lopez-Quintela M A, Rivas J 1992 Covering of magnetic particles to produce stable
magnetic fluids. Spanish Patent 9,201,984 (see also 1994, Structural and magnetic
characterisation of Co particles coated with Ag. J. Appl. Phys. 76: 6564-6)
39. Akashi G, Fujyama M 1969 Process for the production of magnetic substances. US
Patent 3,607,218
40. Harada S, Yamanashi T, Ugaji M 1972 Preparation and magnetic properties of
cobalt alloy particles. IEEE Trans. Mag. MAG-8: 468-72
41. Charles S W, Issari B 1986 The preparation and properties of small acicular par-
ticles of cobalt. J. Magn. Magn. Mat. 54-57:
743-4
42. Chantrell R W, Popplewell J, Charles S W 1978 Measurements of particle-size
distribution parameters in ferrofluids. IEEE Trans. Mag. MAG-14: 975-7
43. O’Grady K, Bradbury A 1983 Particle-size analysis in ferrofluids. J. Magn. Magn.
Mat. 39: 91-4
18 S.W. Charles
44. Scholten P C 1988 Some material problems in magnetic fluids. Chem. Eng. Comm.
67: 331-40
45. O’Grady K, Stewardson H R, Chantrell R W, Fletcher D, Unwin D, Parker M R
1986 Magnetic filtration of ferrofluids. IEEE Trans. Mag. MAG-22: 1134-6
46. Wyman J E 1984 Ferrofluid composition and method of making the same. U S
Patent 4,430,239
47. Bottenberg W R, Chagnon M S 1982 Low vapor-pressure ferrofluids using
surfactant-containing polyphenyl ether. U S Patent 4,315,827
48. Chagnon M S 1982 Stable ferrofluid compositions. U S Patent 4,356,098
49. Shimoiizaka, J, Nakatsuka K, Chubachi R 1978 In: Thermomechanics of magnetic
fluids. Berkovsky B M (ed.) Hemisphere, Washington DC, pp. 67-76
50. Scholten P C 1983 How magnetic can a fluid be? J. Magn. Magn. Mat. 39: 99-106
51. Bacri J C, Perzynski R, Salin D, Cabuil V, Massart R 1990 Ionic ferrofluids: a
crossing of chemistry and physics. J. Magn. Magn. Mat. 85: 27-32
52. Nakatani I, Furubayashi T, Takahashi T, Hanaoka H 1987 Preparation and mag-
netic properties of colloidal ferromagnetic metals. J. Magn. Magn. Mat. 65: 261-4
53. Shepherd P G, Popplewell J, Charles S W 1970 A method of producing a ferrofluid
with Gd ferromagnetic particles. J. Appl. Phys. D 3: 1985-7
54. Shepherd P G, Popplewell J, Charles S W 1972 Ferrofluids containing Ni-Fe alloy
particles in mercury. J. Phys D 5: 2273-82
55. Keeling L, Charles S W, Popplewell J 1984 The prevention of diffusional growth
of cobalt particles in mercury. J. Phys. F 14: 3093-3100
56. Berkowitz A E, Walter J L 1983 Ferrofluids produced by spark erosion. J. Magn.
Magn. Mat. 39: 75-8
57. Charles S W 1988 Aggregation in magnetic fluids and magnetic fluid composites.
Chem. Eng. Comm. 67: 145-80
58. Fannin P C, Scaife B K P, Charles S W 1987 The study of the complex susceptibility
of ferrofluids and rotational Brownian motion. J. Magn. Magn. Mat. 65: 279-81
59. Beresford J, Smith F M 1973 In: Parfitt G D (ed.) Dispersion of Powders in Liquids.
Publ. Appl. Science.
60. Cebula D J, Charles S W, Popplewell J 1983 Aggregation in ferrofluids studied by
neutron small angle scattering. J. Physique 44: 207-13
61. Bissell P R, Chantrell R W, Spratt G W D, Bates P A, O’Grady K 1984 Long term
stability measurements on magnetic fluids. IEEE Trans. Mag. MAG-20: 1738-40
62. Shulman Z P, Kordonsky V I, Zaltsgendler E A, Prokhorov I V, Khusid B M and
Demchuk S A 1986 Structure, physical properties and dynamics of magnetorheo-
logical suspensions. Int. J. Multiphase Flow 12: 935-55
63. Bibik E E 1981 Rheology of Dispersed Systems. Leningrad State Univ. Press. 171
64. Carlson J D, Weiss KD 1995 Magnetorheological materials based on alloy particles.
U S Patent 5,382,373
65. Shulman Z P 1991 Magnetorheological systems and their application. European
Advanced Short Course of UNESCO. Minsk, USSR
66. Weiss K D, Nixon D A, Carlson J D, Margida A J 1994 Thixotropic magnetorhe-
ological materials. W O Patent 9,410,693
Magnetic Spectroscopy
as an Aide in Understanding Magnetic Fluids
Paul C. Fannin
1 Relaxation Mechanisms
The major advances of recent times in the investigation of the relaxation mech-
anisms of ferrofluids are in no small way due to the application of dielectric
formalism in the representation of ferrofluid data [1-5]. There are two distinct
mechanisms by which the magnetisation of ferrofluids may relax after an applied
field has been removed: either by rotational Brownian motion of the particle
within the carrier liquid, with its magnetic moment, m, locked in an axis of easy
magnetisation, or by rotation of the magnetic moment within the particle.
m = Ms v (1)
where v is the volume of the particle and Ms its saturation magnetisation. The
time associated with the rotational diffusion is the Brownian relaxation time τB
[6] where
τB = 3V η/kT (2)
V is the hydrodynamic volume of the particle and η is the dynamic viscosity of
the carrier liquid. In the case of the second relaxation mechanism, the magnetic
moment may reverse direction within the particle by overcoming an energy bar-
rier, which for uniaxial anisotropy, is given by Kv, where K is the anisotropy
constant of the particle. The probability of such a transition is exp(σ) where σ
is the ratio of anisotropy energy to thermal energy (Kv/kT ). This reversal time
τN = τ0 exp(σ) (3)
where
χ0 = nm2 /3kT μ0 (8)
is the static susceptibility and
Fig. 1 shows a plot of the Debye equation (7) which clearly indicates how
χ (ω) falls monotonically and that the maximum possible value of the absorption
component, χ (ω), is equal to half that of the χ (ω) component. Here τ = τef f =
τ .
Fig. 1. Debye plot of χ (ω) and χ (ω) against ωτ
Fig. 2. Normalised plot of χ (ω) and χ (ω) against f (Hz) for samples 1, 2 and 3
For the Frohlich [3,17] distribution function for which it is assumed that the
distribution of relaxation times is confined to a certain range from τ1 to τ2 ;
f (τ )dτ = G(ln τ )d(ln τ ) where
and as τ1 and τ2 are proportional to the cube of the maximum and minimum
hydrodynamic radii respectively, then (rmax /rmin ) = (τ1 /τ2 )1/3
As an example of obtaining an estimate of the aggregate size distribution in
the case of sample 2, Fig 3 shows the result of fitting the measured susceptibility
plots to profiles generated by the Debye equations modified by Frohlich and Cole-
Cole distribution functions. The fits obtained from the two fitting functions are
found to be similar with the Frohlich distribution function giving the value for
the particle radii of from 37 nm to 174 nm whilst the Cole-Cole fit was realised
with αc = 0.31.
Magnetic Spectroscopy 23
Fig. 3. Normalised plot of χ (ω) and χ” (ω) against f (Hz) for samples 2. Curve 1
represents the measured data whilst curves 2 and 3 represent Frohlich and Cole-Cole
fits
and from equation (18), assuming that X(f ) is an even function it follows that
∞
x̂(t) = X(f ) sin(2πf )df (20)
0
Fig. 4. Plot of χ (ω), χ (ω) and its Hilbert transform χ̂ (ω), against f (Hz), for the
Debye case with χ∞ = 0 and N= 8192 points
In reference [22] the technique is applied to a dynamic situation, and this ex-
ample illustrates the usefulness of the Hilbert transform in determining complex
susceptibility components once one component is known. Of course, as already
mentioned, the technique can also be used to generate data on χ (ω) from a
knowledge of χ (ω), simply by taking the inverse Hilbert transform of χ (ω) .
This is due to the fact that the Hilbert transform of a Hilbert transform returns
the original signal with a change in sign [23].
3 Resonance
In the GHz frequency range the character of the dispersion changes from relax-
ation to one of resonance and it is convenient to describe χ(ω) in terms of its
Magnetic Spectroscopy 25
Fig. 5. Plot of χ (ω), χ (ω) and its Hilbert transform χ̂ (ω) , against f(Hz), for the
Debye case with χ∞ = 0.2 and N= 8192 points
Fig. 6. Plot of χ (ω) and χ over the frequency range 50 MHz-18 GHz for 17 values
of polarising field over the range 0-100 kA/m, for sample 4
Fig. 8. Plot of χ (ω) and χ (ω), and corresponding fits, for unpolarised plot of Fig 6
1 χ (ω) iωt
b(t) = 2 e dω
2π −∞ ω
−1 χ (ω)
= 2 F (30)
ω
A plot of the normalised after-effect functions, b(t)/b(0), for the data in Fig 6,
is shown in Fig 9. Here it is seen that the effect of the polarising field is to change
b(t)/b(0) from an exponential type decay for H = 0, with a time-constant, or
1/e value, of 2.5·10−10 s to a damped oscillatory type waveform having a time-
constant of 6.9·10−11 s at H =100 kA/m. Thus there is an approximate fourfold
increase in the rate of decay over the polarising field range.
As previously mentioned χ can be described in terms of parallel, (χ ), and
perpendicular, (χ⊥ ), components (Eq. (21)) with the corresponding relaxation
times τ and τ⊥ef f . However, as the effect of the biasing field is to cause χ(ω)
to be dominated by the χ⊥ (ω) component, the decay time at H =100 kA/m
corresponds approximately to that of τ⊥ef f . This decrease in the time-constant
with increasing polarising field arises because a ferrofluid has particles of different
30 P.C. Fannin
5 Magnetic Losses
In the context of device application, losses in magnetic fluids are of particular in-
terest. The permeability of a magnetic fluid, μ(ω) = μ (ω)−iμ (ω), is a complex
quantity which expresses the loss of energy which occurs as the magnetisation
alternates. The loss mechanisms cause the flux density, B, to lag behind the ap-
plied alternating field, H, by a phase angle δ, and in this context two important
relations are the loss tangent, tan δ, which is also known as the dissipation factor
[19,28] and sin δ, which is referred to as the power factor [19]. The dissipated
energy per cm3 of the ferrofluid sample is directly proportional to both tan δ
and sin δ where in direct analogy with the dielectric case, tan δ = μ (ω)/μ (ω)
and sin δ = μ (ω)/|μ(ω)| [19].
As μ (ω) = χ (ω), and μ (ω) = χ (ω) + 1, the corresponding equations for
the loss factor and power factor are,
χ (ω)
tan δ = (31)
(1 + χ (ω))
and
χ (ω)
sin δ =
1/2 (32)
2
(1 + χ (ω)) + χ (ω)2
Thus measurement of χ(ω) enables one to investigate the frequency depen-
dence of the loss tangent (tan δ) and power factor (sin δ) as a function of particle
packing fraction and hence fluid saturation magnetization where, Fig 10 shows
the high frequency dependence of tan δ and sin δ for a 200 G fluid (sample 5).
Here the profile of tan δ is identical to that of sin δ with both differing from the
χ (ω) profile.
In [28] such measurements are reported for four magnetic fluids with cor-
responding saturation magnetizations of 720 G, 360 G, 175 G and 75 G. The
results show that, with reducing fluid magnetisation, the profiles of both tan δ
and sin δ gradually become closer to that of the χ (ω) profile, until, in the case
of a 75 G Fluid, the frequency dependent profiles become almost identical. Thus,
apart from the results giving an indication of the measure of fluid magnetisation
(and hence packing fraction) required for the above condition to occur it is of
significance to note that the results also enable one to conclude that, in the case
of the most dilute fluid, the high frequency dependent loss tangent and power
factor can be modelled by the imaginary component of equation (23), with,
ω 2 τ22 τ⊥ + ωτ⊥ (1 + Δ)
tan δ = sin δ = χ⊥ (ω) = χ⊥ (0) 2 2. (33)
(1 − ω 2 τ2 τ⊥ + Δ) + ω 2 (τ⊥ + τ2 )
Magnetic Spectroscopy 31
Fig. 10. Normalised plot of χ (ω), χ (ω), tan δ and sin δ, against f in Hz, for sample 5
Finally, with the advent of magnetoelectronics and the possible use of nano-
particles as devices, such as switches, one is interested in the signal-to-noise(SNR)
[29] ratio of such systems. Here again the usefulness of the availability of complex
susceptibility data come to the fore since the SNR for a field driven magneti-
sation of a nanoparticle is a function of the parameters τ0 , α, γ, η and K; all of
which can be determined from susceptibility measurements as demonstrated in
this Chapter.
Acknowledgements
References
1. P. Debye. Polar Molecules, The Chemical Catalog Company, New York (1929).
2. K.S. Cole, R.H. Cole. J. Chem. Phys., vol. 9, (1941), pp. 341.
3. H. Frohlich. Theory of Dielectrics, 2 nd edn (Oxford: Clarandon), (1958).
4. D.W. Davidson, R.H. Cole. J. Chem. Phys., vol. 9, (1951), pp. 341.
5. S. Havriliak, S. Negami. Polymer , vol. 8 (1967), pp. 161.
6. W.F. Brown. 5.Appl.Phys., vol. 34 (1963), pp. 1319.
7. L. Néel. Ann. Géophys., (1949), pp. 99.
8. D.P.E. Dickson, N.M.K. Reid, C.A. Hunt, H.D. Williams, M. El-Hilo,
K. O’Grady. J. Magn. Magn. Mater., (1993), pp. 345.
9. V. Schunemann, H. Winkler, H.M. Ziethen, A. Schiller, A.X. Trautwein.
Magnetic Properties of Fine Particles Elsevier Science Publishers, Netherlands,
(1992), pp. 371.
10. E. Kneller, E.P. Wohlfahrt. J.Appl.Phys., vol. 37, (1966), pp. 4816.
11. W.F. Brown. Phys. Rev., vol. 130, (1963), pp. 1677.
32 P.C. Fannin
12. L. Bessais, L.B. Jaffel, J.L. Dormann. Phys, Rev B., vol. 45, (1992), no. 14,
pp. 7805.
13. A. Aharoni. Phys. Rev. B , vol. 46, (1992), no. 9, pp. 5434.
14. M.I. Shliomis. Sov. Phys.-Usp., vol. 17, (1974), pp. 53.
15. P.C. Fannin, B.K.P. Scaife, S.W. Charles. J. Phys. E.Sci. Instrum, (1986),
pp. 238.
16. M.M. Mairov. Magnetohydrodynamics (Trans. of Magnitnia Hidrodinamika),
vol. 2, (1979), no. 21, pp. 53.
17. C.J.F. Bottcher, P. Bordewijk. Theory of Electric Polarisation Amsterdam,
Elsevier Scientific Publishing Company), vol. 2, (1978).
18. V.V. Daniel. Dielectric Relaxation Academic press, London, (1967).
19. B.K.P. Scaife. Principles of Dielectrics rev.edn (London, Oxford Science Publi-
cations),(1998).
20. L. Lundgren, P. Svedlindh, O. Beckman. J. Magn. Magn. Mater , vol. 25
(1986), pp. 33.
21. C.C. Paulsen, S.J. Williamson, H. Maletta. J. Magn. Magn. Mater., vol. 54-
57, (1986), pp. 209.
22. P.C. Fannin, A. Molina, S.W. Charles. J. Magn. Magn. Mater., vol. 26, (1993),
pp. 194.
23. M.S. Roden. Analog and Digital Communication Systems Prentice Hall Interna-
tional, London, (1991).
24. Y.L. Raikher, M.I. Shliomis. Sov.Phys.JETP , vol. 40, (1975), pp. 526.
25. W.T. Coffey, J.P. Kalmykov, E.S. Massawe. Modern Nonlinear Optics
Adv.Chem.Phys., vol. 85 part 2, (1993), pp. 667, Wiley Interscience, New York.
26. S. Roberts, A.R. von Hippel. J. Appl. Phys., vol. 17, (1946), pp. 610.
27. P.C. Fannin, T. Relihan, S.W. Charles. J. Phys. D; Appl. Phys., vol. 28,
(1995), pp. 2003.
28. E.C. Snelling, A.D. Giles. Ferrites for Inductors and Transmission Lines Wiley,
New York ,(1983).
29. P.C. Fannin, S.W. Charles. Investigation of the frequency dependent loss tan-
gent and power factor of colloidal suspension of nano-particles J. Magn. Magn.
Mater. , vol. 226, (2001), pp. 1887.
Magnetic and Crystalline Nanostructures
in Ferrofluids as Probed by Small Angle
Neutron Scattering
Albrecht Wiedenmann
1 Introduction
Small angle neutron scattering (SANS) is known to be well suited for studying
density- and concentration fluctuations on a length scale between 0.5 nm and 300
nm which corresponds to typical sizes of microstructural features in nanoscaled
(n-) materials [1,2]. This technique allowed different types of inhomogeneities to
be identified in crystalline, amorphous and liquid materials. In addition, magneti-
sation fluctuations in domain like structures resulting from strong intergranular
correlations have been monitored by SANS in compacted n-Fe, n-Co, or n-Ni
alloys [3,4]. In some diluted systems such as n-Fe3 O4 embedded in a glass ce-
ramic [5] and n-Fe3 Si in an ferromagnetic amorphous Fe-Si-B-Nb-Cu matrix [6,7]
magnetic single domain behaviour of the nanocrystalline grains and the nature
of interfaces have been evaluated.
Here we’ll focus on the determination of the crystalline and magnetic mi-
crostructures of magnetic colloids. Magnetic colloids are stable dispersions of
ferromagnetic materials. In such “Ferrofluids” (denoted as FF) nanoscaled mag-
netic particles are stabilized against coagulation either by electrostatic repulsion
or by coating the core with organic chain molecules acting as surfactants [8]. Cur-
rently great effort is undertaken to prepare new bio-compatible Ferrofluids for
potential biomedical applications [9]. Such applications are based on the super-
paramagnetic behaviour of nanosized particles, which disappears when aggrega-
tion takes place as the consequence of an inefficient screening [10,11]. Therefore
a precise knowledge of the microstructural parameters is a pre-requisite for the
interpretation of macroscopic phenomena and for a tailored fabrication of FF.
Only few methods give access to these parameters. From macroscopic tech-
niques such as magnetisation M(H,T) measurements [12] or magneto-viscous
effects [13] average parameters of concentration, size and arrangements of mag-
netic particles are derived. However, the actual values obtained depend strongly
on the basic assumptions for the underlying processes [14]. Wide angle and small
angle X-ray scattering as well as transmission electron microscopy [15] are sensi-
tive to the particle core only, since the light elements of the organic shell give no
sufficient contrast. The advantages of neutrons in such liquids are twice: First,
the strong scattering power of hydrogen contained in the organic surfactants
gives access to the shell and second, the interaction of the neutron spin with
magnetic moments allows to visualise the magnetism of the core. Some SANS
studies have been performed mainly on concentrated systems FF, which allowed
the stability phase diagram to be established [16,17]. However, complications
arise in poly-disperse systems when different constituents are present. Then we
face the problem that weak magnetic scattering signals have to be analysed be-
side strong nuclear contributions from other sources or vice versa which can lead
to considerable inconsistencies in the interpretation of results.
In our SANS investigations we intended to study the mesoscopic contituents
of such poly-disperse FF using polarised neutrons as a labelling technique. Size,
distribution and composition of the particles as well as magnetic nanostructures
are determined systematically depending on the magnetic materials, surfactants
and carrier liquids and on the particle concentration.
The paper is organised as follows: In section 2 we recall the basic concept
of SANS and present the newly developed technique of nuclear and magnetic
contrast variation by using polarised neutrons. We will show how this combi-
nation allows constituents of magnetic liquids to be identified. Results of our
recent investigations on Co-FF, Fe3 04 -FF and Ba-hexaferrite-FF are presented
in the third section. It will be shown that this technique provides additional
information about the nature of the nanocrystalline magnetic core and magnetic
correlation as well as of the nonmagnetic shell which are not available by other
techniques. Some common features of the different materials investigated are
summarized in section 4.
Scattering of neutrons result from both nuclear and magnetic interactions be-
tween neutron and the nucleus and with the magnetic moment, respectively. The
formalism is described in detail in various monographs e.g. [18]. In this section
we recall some basic terms which are important for the understanding of the
present subject.
The concept and experimental set-up for SANS is very simple and schemati-
cally shown in Fig. 1. From the “white” spectrum of the neutron source (reactor
or spallation source) a small band of wavelengths Δλ is filtered out, collimated
Magnetic and Crystalline Nanostructures in Ferrofluids 35
and directed to the sample. The transmitted beam is absorbed in a beam stop
whereas neutrons scattered around the primary beam are registered in a de-
tector. The intensity distribution I(Q) has to be analysed as a function of the
scattering vector Q(reciprocal space) which by Fourier transform gives access to
correlation functions in real space i.e. to size, composition and magnetization of
inhomogeneities present in the material. Small angle neutron scattering (SANS)
is a special regime of low values of Q between typically 10−2 nm−1 and 5 nm−1
which allows to probe fluctuations on a length scale D∝2π/Q ranging from 1
nm to 500 nm.
Fig. 1. Sketch of a typical SANS instrument (V4 at HMI Berlin) with the option for
polarised neutrons (SANSPOL): The velocity selector picks out a small wavelength
band Δλ from the reactor spectrum. The beam is focussed in the collimator and di-
rected to the sample where part of the neutrons are scattered elastically at nanosized
inhomogeneities in an angle 2θ. The scattering intensity around the primary beam is
counted in a position sensitive area detector an analysed a function of the scattering
vector Q. For SANSPOL polarised neutrons are produced in the transmission polariser.
Two scattering intensities are measured subsequently with neutron spin (n+) and (n-)
after reversal in a spin flipper.
where b is the nuclear scattering length, c the concentration and Ω the atomic
volume of the species i. η is easily calculated from the tabulated values of each
element e.g. for water and heavy water η (H2 O,D2 O) = ρ NA (2bH,D + bO )/MW
which, with the mass density ρ, the Avogadro number NA and the molecular
weight MW yields values of -0.56 1010 cm−2 and 6.33 1010 cm−2 , respectively.
For the interaction between the neutron spin S=1/2 with an assembly of
magnetic moments a magnetic scattering amplitude is defined similarly by
ηM = e2 γ/2mc2 Σci M ⊥ i /Ωi (3)
where only the projection of the magnetic moment Mi ⊥ onto a plane perpen-
dicular to the scattering vector Qcontributes to the interaction.Mi ⊥ is given in
units of Bohr magnetons μB and (e2 γ/2mc2 ) = 0.27 · 10−12 cm.
Throughout this paper we restrict our formulations to two phase-systems
where particles are embedded in a homogeneous matrix. The total scattering
amplitude of the particle is called “form factor” and defined by
F (QR) = dr3 Δηexp(iQr j ) = ΔηVp f(QR) (4)
For such simple mono-disperse systems the scattered intensity follows general
approximations which are independent of the shape of the particle: At low Q
(QR <1) the scattered intensity I(Q) is described by the Guinier law
I(Q → 0) = Δη 2 Np Vp2 exp(−Q2 R2g /3) for (QR < 1), (7)
from which a radius of gyration Rg can be derived. When the shape is known
the particle dimensions are obtained from Rg , e.g. for spheres of radius R,
Rg =(5/3)0.5 R. At large Q, scattering arises from the total surface of the particle
S according to the Porod approximation
Combining Eq. (7 - 9) one obtains some indications of size, shape and con-
centration of the particle. However, in most of materials the two extreme regions
are in practice not well defined due to additional scattering contributions, such
as grain boundaries, surface defects , dislocations or strong background.
In this low-Q regime of SANS there is no access to distances between atoms
or molecules; e.g. in simple liquids scattering in the SANS regime is independent
of Q and given by
I(Q) = I(0) = N < b >2 kB T χ(T ) + Iinc . (10)
The first term results from thermodynamic fluctuations, where χ(T) is the
isothermal compressibility and <b> is the average atomic scattering length ,
the second term results from the incoherent cross section of each atom which in
some cases might be very high such as for protons where Iinc is 80 barn/atom.
The scattered intensity therefore depends on the form factors F (Qr) and
partial structure factors Sij (Q) of all particles of type i and j. This general
case is difficult to treat for which different approximations have been proposed
[19-21]. Analytical solutions are available only for mono-disperse systems; for
spherical particles of identical size it simplifies to the product
The structure factor S(Q) will be unity when particles are fully uncorrelated
which is the case only for a perfect gas. In disordered condensed matter the
scattering at large values of Q is determined mainly by the particle form factor
alone, where S(Q)=1. For a repulsive potential such as excluded volume effects
or electrostatic repulsion S(Q) decreases at low Q and oscillates with Q when the
particles form a pseudo-periodic arrangement. Attractive interaction potentials
will give rise to an increase of the intensity at low Q which in practice might
also be assigned to polydispersity or aggregates.
B ± (Q) = FM
2
− 2P (1 − 2f ± )FN FM . (17)
The cross term is linear in the magnetic amplitude and hence makes it pos-
sible to determine the absolute value of the magnetic contrast with respect to
the nuclear contrast, i.e. magnetic moment and compositions of particles and
matrix.
40 A. Wiedenmann
The ratios FR and R2 can be measured very precisely which allows the magni-
tude of γ to be determined much more accurately than for nonpolarised neutrons
using Eq. (6), for which the modulus of γ is given by |γ| = [B(Q)/A(Q)]1/2 .
Note that the average of the intensities for both neutron polarisations is given
by
which corresponds to the intensity of non-polarised neutrons (Eq. 14) since the
second term in brackets vanishes for f → 1. For the paramagnetic case where
all moment directions are equally probable the scattering is fully isotropic and
independent of the polarisation and given by
Now we consider the case of non perfect alignment of the magnetic mo-
ments of a ferromagnetic single domain particle of saturation magnetization Ms p
embedded in a nonmagnetic matrix [25]. For such particles superparamagnetic
behaviour is expected, where the orientation distribution of the magnetic mo-
ments as a function of an effective magnetic field Hef f and temperature follows
the Langevin statistics
L(x) = coth(x) − 1/x (23)
where x= M(R’)H ef f / k B T.
The total moment, M(R’ ), depends on the radius R’ of the particles accord-
ing to
M (R ) = 4πR3 m0 /3Ωat (24)
where m0 is the saturation value of the atomic magnetic moment. Using the
formalism presented in [6] SANSPOL cross sections for the case of a dilute
system of non-interacting particles are again described by Eq. (16). If we include
steric interaction effects via phenomenological structure factors S(Q//H) and
S(Q⊥H) the intensities parallel and perpendicular to the magnetic field are then
given by
(25)
Magnetic and Crystalline Nanostructures in Ferrofluids 41
The isotropic term as obtained from the intensity parallel to His independent
of the polarization state. Beside the nuclear term F N 2 S(QH) it contains an
2
additional term of magnetic origin, FM 2L(x)/x , which vanishes for complete
alignment of all moments along H, when L(x)/x = 0. The anisotropic part
B ± (Q) is different for the two polarization states. Since the cross-term B int (Q)
as obtained from the intensity difference
is linear in the magnetic amplitude the field variation of the magnetic moment
according to the Langevin statistics can be precisely determined from the cross-
term. Note that in Eq. (25) inter-particle correlations parallel to the field affect
only the nuclear part but correlations perpendicular to H influences nuclear,
magnetic and cross terms. We will see later that these features of Eq. (25) will
help to distinguish between inter-particle correlations and particular form-factors
which both may give rise to characteristic peaks in the SANS curves at low Q.
Fig. 2. Illustration of the labelling technique for SANSPOL combined with contrast
matching as a grey scale representation of the scattering length densities η in a com-
(±)
posite particle: η1 of the central magnetic depends on the neutron polarisation. The
contrast between the nonmagnetic shell and matrix is adjusted using different isotope
mixtures of the solvent.
I+ (Q)-I− (Q) is proportional to the product η 1 mag ·η 1 nuc of one and the same
particle whereas nuclear and magnetic contrasts from different particles do not
contribute to the cross-term. This allows additional nonmagnetic particles to be
distinguished from magnetic particles even when they are of the same size.
In addition, the “colour” of the surrounding matrix can be adjusted using
the above mentioned contrast matching with different mixtures of isotopes in
the solvent.
Both scattering curves I+ (Q) and I− (Q), the difference I+ (Q)-I− (Q) and
FN (Q) und FM 2 (Q) as obtained from the averaged signal of Eq. 14 in all iso-
2
tope mixtures must be adjusted using the same structural model. This combined
contrast variation thus reduces considerably the number of possible structural
models and ensures the consistence and accuracy of the constrained model pa-
rameters.
2.6 Experimental
neutron flux of more than 30% of non-polarized neutrons, a high degree of polar-
ization (P = 95%) and by the high efficiency of the spin flipper (f(-)≈ 98%) at the
wavelength λ=0.6 nm used in this experiments. The reliability of this option has
been demonstrated by comparing the results of a SANSPOL study on magnetic
glass ceramics to those of a conventional SANS study [28]. By choosing three
distances between the sample and detector a Q range of 0.08 nm−1 to 3 nm−1
could be accessed which allows particles to be detected with sizes ranging from
0.5 to 80 nm.
We investigate ferrofluids where the nanoscaled magnetic particles are stabi-
lized against coagulation either by electrostatic repulsion or by coating the core
with organic chain molecules acting as surfactants. Magnetic moments of the
core materials (Co, Magnetite Fe3 O4 , and Ba-Hexaferrite), shell forming surfac-
tants (mono- and bi-layers) and carrier liquids (Water, organic solvents) have
been systematically varied in order to study the influence on the microstructures.
Co- and Magnetite ferrofluids have been prepared by Berlin Heart AG [41] and
Ba-Hexaferrite by R. Müller of IPHT, Jena Germany [31].
3 Results
3.1 Cobalt Ferrofluides
Diluted Co-FF. Colloidal solutions of ferromagnetic Cobalt can be prepared by
coating the nanosized particles with organic chain molecules C21 -H39 -N-O3 . The
hydrophilic head groups are considered to be in contact with the metal surface
and the hydrophobic tail pointing to the non-polar toluene as carrier liquid. We
performed a SANS study on dilute solutions with a nominal concentration of 0.5
vol. % of Cobalt as derived from saturation magnetisation measurements. Four
different mixtures of protonated and deuterated toluene, denoted as AF1-AF4
(Table 1) have been studied in a magnetic field of 1.1 T in which full alignment
of the magnetic moments is expected.
Table 1. Scattering length densities as obtained from a fit to a shell model using the
bulk value of η1 (nuclear) = 2.531010 cm−2 for Co.
Sample Content η (solvent) η 1 (mag) η (mag) η2 (shell)
C 7 D8 1010 cm−2 1010 cm−2 1010 cm−2 1010 cm−2
SANSPOL SANS
AF2 100% 5.65 4.29 4.38 0.56
AF4 43% 3.0 4.5 5.3 0.3
AF3 14% 1.6 1.52 1.62 0.9
AF1 0% 0.97 0.94 1.24 0.8
A first conventional SANS study showed that the scattering on of the sample
AF4 is dominated in the high Q range by a high level of the incoherent back-
ground (mainly from hydrogen contained in the solvent and surfactant). Nu-
clear and magnetic scattering contributions have been extracted and analyzed
44 A. Wiedenmann
Fig. 3. SANSPOL patterns in Co-ferrofluid AF4 for neutron spins antiparallel (I− )
and parallel (I+ ) to the horizontal field of strength H=1.1 T. The arithmetic mean [(I− )
+ (I+ )]/2 corresponds to the 2D pattern of non-polarized neutrons. The difference (I− )-
(I+ ) yields the interference term [equation (18)]which presents negative values in the
center.
The scattering contrasts with respect to the matrix are different for the
magnetic core and non-magnetic shell and given by
(±)
Δη1 = η1nuc ± η1mag − ηmatrix andΔη2 = η2nuc − ηmatrix (28)
respectively. In the present case, only Δη 1 (±) depends on the polarization. The
intensities were calculated according to I (Q⊥H)= N p F shell 2 (Q,R) N (R’ )
dR, where for the diluted case inter-particle correlations have been neglected i.e
S(Q)=1 in Eq. (25) . The number distribution N (R’ ) has been parameterized
using a log-normal number distribution of the core radius R’ according to
√
N (R )1/( 2πRσ)exp[− ln2 (R /R0 )/2σ 2 ] (29)
where R0 denotes the median and s the width of the distribution.
The thickness of the shell D is assumed to be constant, i.e. R=R’+D. The
parameters η 1 nuc , η 1 mag , η 2 nuc have been adjusted in a non-linear least square
fitting routine using the constraints of Eq. (28) for the contrasts Δη 1 (±) and
Δη 2 . The scattering length density of the solvent η matrix was known for the
H/D ratio and confirmed by the incoherent background level. The parameters
N p , R’, D and σ have been constrained to be identical for both polarization
states. The solid lines in Fig. 4 represent the calculated intensities I ± (Q⊥H). It
turned out that for all solvents this model function lead to one set of parameters
which are consistent with the intensities FN (Q)2 and FM (Q)2 as calculated for
the nuclear and magnetic contributions from the averaged signal according to
Eq. (14).
The distribution N (R’ ) was found to be rather sharp (Fig. 4d) corresponding
to a volume weighted average of the core radius of < R > = 3.7 nm and a
constant thickness of the shell of D= 2.47 nm. We emphasize that from non-
polarized neutrons alone it would not have been possible to derive the shell
model due to the very low nuclear contrast of the shell beside the high incoherent
background. The first maximum of the size distribution of spheres as obtained
from classical SANS is therefore identified as artificial and reflects in fact half of
the thickness of the shell.
The scattering length densities resulting from fits of both polarization states
using the same model function are included in Table 1. The values of η 2 for the
shell of the same order of magnitude as calculated for densely packed surfactant
molecules (η 2 = 0.33 1010 cm−2 ) and do not depend significantly on the solvent
composition. This supports strongly the conclusion that the shell of thickness
of about 2.4 nm formed by organic surfactants is nearly impenetrable for the
solvent in i.e. the Co-core is not in direct contact with the solvent. The values
of η 1 (mag) in AF2 and AF4 experimentally derived at H = 1.1 T correspond
closely to η 1 (mag) = 4.14 1010 cm−2 expected from Eq. (3) for bulk Co (m0 =
1.715 μB / atom , Ωat (Co) = 0.01099 nm3 /at.) Discrepancies appeared in the
samples AF1 and AF3 where much lower values for η 1 (mag) were found. The
reason for such anomalous low values is not yet clear; it might result from a lower
particle density, to some oxidation , or to non-magnetic amorphous surface layers
as reported for magnetic glass ceramics [32].
Magnetic and Crystalline Nanostructures in Ferrofluids 47
Variation of the magnetic field. For the solvent composition AF4 the cross-
term B int (Q,H ) has been evaluated from the difference I − (Q) − I + (Q) for dif-
ferent values of the external magnetic field and plotted in Fig. 5. All additional
non-magnetic contributions are eliminated in this difference [25].
B int (Q,H ) could be well fitted according to Eq. (26) by the product
FN FM L(x) alone showing that for this diluted sample no significant inter-particle
correlations are present or induced by the magnetic field, i.e. S(Q⊥H )=1 [38].The
ratio B int (Q,H ) / B int (Q,H=1.1T ) was found to be constant for all values of
Q. The H -dependence of this ratio is plotted in the insert of Fig. 5. The ex-
perimental points follow closely a Langevin function corresponding to spherical
Co-particles with average core radius < R > = 3.7 nm. This is in good agree-
ment with the prediction of Eq. (24) and with < R > as derived from the direct
model fit [33].
Fig. 6. Radial averaged SANS intensity of 1-vol% (triangles) and 5-vol% (circles)
Co-Ferrofluids in C7 D8 at H = 0 (full symbols) and H = 1.1 T (open symbols).
Fig. 7. SANSPOL intensities I + (Q⊥H) (solid triangles ) and I − (Q⊥H) (open tri-
angles) of 5vol% Co-Ferrofluids in C7 D8 and magnetic (squares) and nuclear (circles)
contributions from non-polarised neutrons. The solid line represents the common struc-
ture factor S(Q) used in the model fits.
Fig. 8. Charge stabilized Magnetite FF (ELEC) in H0.2 D1.8 O,: a) SANSPOL inten-
sities I + (Q⊥H) and I − (Q⊥H) where solid lines corresponds to a fit with using a
bimodal distribution of spheres with b) scattering length density profiles of small and
c) large spherical particles.
Fig. 9. Magnetite FF with DEX coating in H0.2 D1.8 O,: a) SANSPOL intensi-
tiesI + (Q⊥H) and I − (Q⊥H) where solid lines corresponds to a fit with a distribution
of small core-shell particles and larger spherical aggregates with the corresponding
scattering length density profiles shown in b) and c). The low value of η 1 (nuc) in the
aggregates reflects inclusion of free surfactant molecules.
It turned out that for all samples this model function lead to consistent
parameters. The volume weighted size distributions are compared in Fig. 11.
For the small particles a rather sharp distribution N (R’ ) corresponding to a
volume weighted average of the core radius of < R >=6.73nm (LM), 4.8 nm
(DEX) and 4.0 (ELEC) a constant thickness of the shell of D = 2.61 nm and 2.4
nm (DEX). For the second fraction the volume averaged radius is by a factor of
2.5-3.5 times larger than that of the core.
The scattering length densities resulting from simultaneous model fits of both
polarization states as well as Inuc and Imag using the same model function are
presented in Fig. 8-10 (b,c).
From SANSPOL it turned out that for the small particles the scattering
length density of the core (η 1) is clearly higher than η solvent and higher than
52 A. Wiedenmann
that of the shell (η 2 ) for the samples DEX and LM. The magnetic moment for
Fe3 O4 as derived from the fit is very close to the bulk value. In addition, the
resulting scattering length density (Fig. 10) for the shell in LM was found to be
independent on the H/D composition of the solvent, which indicates that the
organic surfactant is nearly impenetrable for the solvent.
Fig. 10. a) SANSPOL intensities I − (Q ⊥ H)(open symbols) and I + (QH) (full sym-
bols) of Magnetite FF with LM-coating in different isotope mixtures of D2 O/H2 O and
b) corresponding scattering length density profiles of small bi-layer coated composites
particles and c)of spherical aggregates.
For the large particles η 1 is found to be lower than η solvent in LM and DEX
but higher in ELEC. Since the scattering intensity for the large particles de-
pends on the polarisation, they must also contain magnetic materials. However,
the observed ratio η (−) /η (+) was much smaller than that expected for pure mag-
netite. It corresponds to a magnetite content of only 22%. These result suggests
that the large particles consist of a loose and mixed arrangement of magnetite
together with surfactant molecules forming a larger aggregate. Size and volume
fraction of these aggregates seem to depend on the preparation conditions (see
Table 2). In ELEC where surfactants are absent the magnetic contrasts of both
fractions are very similar which indicates that also the aggregates presents the
ferrimagnetic ordering of magnetite.
Magnetic and Crystalline Nanostructures in Ferrofluids 53
Fig. 11. Volume weighted size distribution of magnetic particles in charge stabilized
ELEC (without shell), monolayer coated DEX and bilayer coated LM . The size of the
central core are very similar while that of the aggregates depends on the preparation
conditions.
encapsulated with surfactants. The remaining 1/3 of ferrite particles are in the
aggregates which have volume fractions of about 4.5% to 7%.
4 Summary
The use of polarized neutrons in Small Angle Neutron Scattering is a power-
ful labeling technique for investigations of magnetic materials. In the present
study of magnetic colloids we combined this technique with the conventional
contrast variation using different isotope mixtures of the carrier liquid. This
combination allows magnetic and nonmagnetic particles to be distinguished and
density-, composition-, and magnetization profiles to be precisely determined.
The microstructure parameters have been evaluated in polydisperse multiphase
systems where magnetic materials (Co, Magnetit, Ba-Hexaferrite), shell forming
surfactants (mono- and bi-layers) and carrier liquids (Water, organic solvents)
have been systematically varied. The results are compared in Table 2 and sum-
marized as follows:
As a common feature three different components were identified in the mag-
netic liquids, magnetic composite particles, free organic shell molecules and mag-
netic aggregates. In diluted Ferrofluids which are stabilized by surfactants com-
posite particles are well described by a magnetic core and an organic shell of
constant thickness. The size of the core depends on material and preparation
conditions while the thickness of the shell is characteristic for the surfactant
materials. The shell was found to be of homogeneous density and almost impen-
etrable for the solvent. No shell structure was found in charge stabilized samples
where solvent molecules are in touch with the magnetic core. The magnetic or-
dering of diluted systems corresponds to non-interacting ferromagnetic single
domain particles. Small amounts of aggregates with lower densities and mag-
netization and a typical size of 2-4 times the core radius were identified. This
indicates that screening is incomplete in some systems leading to aggregates in
which surfactant molecules are included. Depending on preparation conditions
nonmagnetic contributions have been found which are ascribed to free organic
surfactants or micelles.
The set-up of strong field induced inter-particle correlations was detected in
Co- Ferrofluids with higher concentrations. The observed structure factor S(Q)
corresponds to an interaction of hard spheres with a radius which is by a factor
of about 1.5 larger than the total radius of the composite particles. Surprisingly,
no evidence for an anisotropic structure factor was found which implies a dense
packing of the particles rather than formation of chains as predicted for dipole
interactions.
The application of the combined contrast variation technique will enable a
more detailed insight in the nature of more complicated core-shell structures,
which are generally not accessible from other microscopic methods. This knowl-
edge is a necessary pre-requisite for potential biomedical applications of fer-
rofluids, where immunoassays or therapeutic agents should be connected to the
surfactant shell.
Table 2. Structural parameters resulting from SANSPOL investigations in Ferrofluids with different core and surfactant materials. 56
Average radii < R >, shell thickness D and volume fractions f of core in composites, aggregates (agg) and micelles (mic). fmagn
denotes the total volume fraction of magnetic particles as obtained from magnetization measurements.
Sample Shell fmagn < Rcore > Dshell fcore < RAgg > fAgg < RM ic > fM ic
vol % [nm] [nm] vol % [nm] vol % [nm] vol %
Acknowledgements
The author acknowledges the fruitful cooperation with his colleagues Dr. A.
Heinemann, Dr. A. Hoell, M. Kammel, and Dr. U. Keiderling of the SANS team
at HMI Berlin. Many thanks are given for sample preparation to Dr. N. Buske
and Dr. M. Gansau from Berlin Heart AG, and to Dr. R. Müller from IPHT
Jena, Germany.
This work is supported by the German DFG priority program “Colloidal
magnetic fluids”, Grant No WI-1151/2.
References
1. J. Loeffler, W. Wagner, H. Van Swygenhoven, & A. Wiedenmann, Nanostruct.
Mat., 9 (1997), 331-334.
2. A. Wiedenmann , J. Appl. Cryst., 30 (1997), 580-585.
3. J. F. Loeffler, H. B. Braun, W. Wagner, G. Kostorz, A. Wiedenmann
Materials Science and Engineering A, 304-306, 31 (2001), 1050-1054.
4. A. Michels,, J. Weissmüller, A. Wiedenmann, J.S. Pedersen, J.G. Barker
Philosoph. Magazine Letters 80 (2000), 785-792.
5. A. Wiedenmann, U. Lembke, A Hoell,., R. Müller, & W. Schüppel Nanostruct.
Mat., 12 (1999), 601-604.
6. J. Kohlbrecher, A. Wiedenmann, H. Wollenberger, Z. Physik, 104(1997),1-4.
7. A. Heinemann, H. Hermann, A. Wiedenmann, N. Mattern, K. Wetzig J. Applied
Cryst. 33(2000),1386-1392.
8. S.W. Charles, JMMM (2002).
9. P.P.Gravina et al., JMMM (2002).
10. E. Dubois et.al, Langmuir 13(2000),5617.
11. G. A.van Ewijk, PhD Thesis Utrecht (2001).
12. B. Huke et al., Phys. Rev. E 62(2000),6875.
13. S. Odenbach, K.Raj, Magnetohydrodynamics 36(2000),379-386.
14. B. U. Felderhof Phys. Rev. E 62(2000),3848.
15. J.P.Embs et al. Phys. Rev. E(2001), preprint
16. E. Dubois, V. Cabuil, F. Boué, R. Perzyski, J.Phys. Chem. 111(1999),7147-7160.
17. F. Cousin, V. Cabuil, Prog. Colloid Polym. Sci. 2000),77-83.
18. G. Kostorz in Treatease on Materials Science and Technology, Acad. Press (1979)
ed. G. Kostorz, 227-290.
19. J.K. Percus, G.J. Yevick Phys. Rev 110(1959),1-13.
20. J. Brunner-Popela, O. Glatter J. Appl. Cryst 30(1997),431-442.
21. J.S. Pedersen J. Appl. Cryst. 27(1994),595-608.
22. R.M. Moon, T. Riste, & W.C. Koehler, Phys. Rev.(1969) 181, 920-931.
23. R. Pynn, &J. B.Hayter. Phys. Rev. Letters , 51, 710-713.
24. A. Wiedenmann, Mat. Science Forum, 312-314(1999) 315. J. Metastable and
Nanocrystalline Materials, 2-6 (1999), 315-324.
25. A. Wiedenmann, Physica B 297 (2001), 226-233.
26. R. P. May and K. Ibel and J. Haas, J. Appl. Cryst., 15 (1982), 15-19.
27. R. Stegmann, E. Manakova, , M. Rössle, H. Heumann, S. Axmann, A. Plückthun,
T. Hermann, R. May, A. Wiedenmann. J. Structural Biology 121(1998),30-30.
28. A. Wiedenmann, J. Appl. Cryst. 33 (2000), 428-432.
29. U. Keiderling & A. Wiedenmann, A. Physica B, 213&214(1995). 895-897.
58 A. Wiedenmann
Ronald E. Rosensweig
Abstract. Several authors have attempted with varying success to derive a complete
set of basic equations for magnetic fluids having internal rotations. In this work a com-
plete set of governing equations is derived on the basis of dynamic balance relationships
with the dissipation function derived from thermodynamic consideration. The magneti-
zation relaxation equation is thereby determined from requirement of positive entropy
production along with a complete set of well-accepted constitutive laws. The analysis
employs the Minkowski expression of electromagnetic momentum and assumes that the
product of Maxwell stress and velocity for polarized media contributes to the energy
balance on the same footing as contact stresses of pressure and viscous origin.
1 Introduction
Magnetic fluids, also known as ferrofluids, are stable suspensions of subdomain
ferrimagnetic or ferromagnetic fine particles in a liquid carrier. Early theory of
magnetic fluids assumed that the magnetization is colinear with the magnetic
field at all times [1]. More generally, the magnetic particles lag the changing
field and the description of the flow becomes more complex. The work of Dahler,
Scriven, and Condiff [2] exposed the role of asymmetric stress in this circum-
stance, although the relaxation equation was postulated rather than derived.
Independently the work of Shliomis [3] has been influential in the development
of the science, albeit the relaxation law is derived separately [4] from the main
theory [5].
Irreversible thermodynamics provides a framework for the consistent descrip-
tion of dissipative processes. In their well-known monograph deGroot and Mazur
[6] arrive at a plausible expression for entropy production of a polarized medium
only after introducing an artificial definition of electromagnetic energy density.
Also, their work did not treat the case when the direction of the magnetization
differs from the direction of the field. Felderhof and Kroh [7] building on the De-
Groot and Mazur treatment present a theory that yields a plausible form of the
relaxation relationship when magnetization and field are not colinear. Yet the
derivation is not transparent to us for the reason cited, and is incomplete; e.g.,
the asymmetric hydrodynamic stress tensor is postulated rather than deduced.
A main objective of the present work is to circumvent the infelicities cited
above while employing the irreversible thermodynamic framework. We utilize
with modifications the formulation of Shizawa and Tanahashi [8] with its method-
ology based on integral formulation of equations governing the balance of linear
d
where dt ∂
= ∂t + v · ∇ is the material derivative and is the mass density. The
Reynolds transport theorem for any function Ψ of position and time (scalar,
Basic Equations for Magnetic Fluids 63
where n is the unit outward facing normal to the surface. In words, the total
time rate of change equals the time rate of change within the volume plus the
net convection out of the volume through the surrounding surface. Applying the
Gauss divergence theorem to the surface integral term transforms it to a volume
integral yielding the transport theorem in alternate form as
d ∂Ψ
Ψ dV = + ∇ · (vΨ ) dV (7)
dt ∂t
Vm Vm
Next, rewriting the first term of the right side of Eq. (11) using Cauchy’s
fundamental theorem: tn = n · T̃ where T̃ is the total stress vector (sum of
viscous, pressure, and electromagnetic components), transforming the surface
integral into a volume integral by applying Gauss’ theorem, collecting terms and
noting that the volume of integration is arbitrary, the equation of motion in
Cauchy (unconstituted) form is obtained.
dv ∂g
ρ + = ∇ · T + ρb (14)
dt ∂t
where
T ≡ T̃ − vg (15)
where the first equality results from the Reynolds corollary of Eq. (10), and the
last equality from the circumstance that dr/dt = v.
Multiplying and dividing the integrand of the next term in Eq. (16) by ,
using the Reynolds transport theorem of Eq. (10) and then the equation of
continuity of Eq. (9), and noting that dr/dt = v, the term transforms as follows.
d d g
(r × g)dV = ρ r× dV
dt dt ρ
Vm Vm
dg
= (v × g) + r × + g∇ · v dV
dt
Vm
∂g
= − : vg + r × + ∇ · (vg) (18)
∂t
Vm
Using the Reynolds transport theorem of Eq. (10) the spin term on the left
side of Eq. (16) becomes
d ds
ρsdV = ρ dV (19)
dt dt
Vm Vm
Substituting the definition of T from Eq. (15), noting that the bracketed
term in the cross product with the radius vector disappears by virtue of the
equation of motion of Eq. (14), and invoking the arbitrariness of the volume
element yields the balance equation of internal angular momentum.
ds
ρ = ρl + ∇ · Λ − : T (24)
dt
For a monodispersion of spherical particles spin can be expressed as s = IΩ/
so that spin rate is given by
ds dΩ
ρ =I (25)
dt dt
where I is moment of inertia per unit mass and Ω is average angular velocity of
particles about their own center.
work is given by E · dD and H · dB [10]. The combined first and second law
expression (Gibbs equation) written for the contents of a fixed unit volume of
space and incorporating these work terms is absent a pressure term and, in its
place contains a term proportional to the chemical potential of the medium,
considered to be of constant composition. Later in the present study the Gibbs
equation is required in a form in which internal energy and entropy are expressed
on the unit mass basis. Conversion from the unit volume basis to unit mass basis
yields the following expression [11]
du(e) = T ds(e) + p(e) dρ/ρ2 + E (e) · d D (e) /ρ + H (e) · d B (e) /ρ (28)
where superscript (e) denotes equilibrium value. u(e) is internal energy per unit
mass and s(e) entropy per unit mass. The same expression is derived by Chu
[12] based on a roundabout treatment, and adopted by Tarapov [13] and Blums
et al. [14]; our development in [11] shows that the result follows from classical
thermodynamics alone. The p(e) denotes the value of pressure in the presence of
the fields and it differs from the pressure p0 (, T ) that attains in the absence of
fields.
The analysis of [11] relates the two pressures as follows:
p(e) = p0 (ρ, T ) + 0 E (e)2 + μ0 H (e)2 /2
E H
(e) (e)
+ (∂υP/∂υ)T,H,E dE (e) + (∂υM/∂υ)T,H,E dH (e)
0 0
≡ p + 0 E (e)2
+ μ0 H (e)2
/2 (29)
E H
(e) (e)
p = p0 (ρ, T ) + (∂υP/∂υ)T,H,E dE (e) + (∂υM/∂υ)T,H,E dH (e) (30)
0 0
Substituting for p(e) from Eq. (29) into Eq. (28) and expanding the elec-
tromagnetic work terms yields the following form of the Gibbs equation after
rearrangement and substitution from Eqs. (26,27).
ρdu(e) = ρT ds(e)
+ p − 0 E (e)2 /2 − μ0 H (e)2 /2 − P (e) · E (e) − M (e) · H (e) dρ/ρ
This form of the Gibbs equation with each term expressed on the per-unit-volume
basis is employed subsequently below.
68 R.E. Rosensweig
The u is internal energy per unit mass, q is the heat flux vector and R is
volumetric heat release rate, e.g., due to radioactive decay. The integrand of
the first term on the right side treats all components (viscous, pressure and
electromagnetic) of the stress vector on equal basis in expressing the work rate
as the scalar product with the velocity vector 1 .
tn = n · T̃ (33)
T = T̃ − vg (34)
The integrand of the sixth term on the right side is the Poynting vector
expressing electromagnetic energy flux density. Since the integration is over a
closed material surface, the Poynting vector is evaluated in the moving medium
as indicated by the asterisks suffixed to the vector fields. The electric field vector
E ∗ and magnetic field vector H ∗ are related to the components of vectors E
and H observed in a stationary coordinate system as follows [10].
E ∗ = E // + γ(E ⊥ + v × B) (35)
H ∗ = H // + γ(H ⊥ − v × D) (36)
where −1/2
γ 1 − (v/c)2 (37)
In Eqs. (35,36), v × B and v × D are respectively the apparent electric field
and magnetic field produced when an observer crosses B and D with velocity
v. Since the speed v of the suspensions can be assumed much smaller than light
1
Inclusion of the electromagnetic stress in the energy balance in this manner has
been questioned recently as unphysical [9]. Our thesis is, because Maxwellian stress
applies in linear and angular momentum balances, for consistency it should apply in
the energy balance an well, on the same footing as contact stresses of pressure and
viscosity.
Basic Equations for Magnetic Fluids 69
By substituting Eqs. (44) and (41) into Eq. (40) and substituting the result into
Eq. (39), taking account of the linear momentum balance of Eq. (14) and the
internal angular momentum balance of Eq. (24), and the definition of T̃ from
Eq. (15)the balance of energy relationship is expressed as follows:
du
ρ = TT : ∇v + ΛT : ∇Ω − T : · Ω − ∇ · (E × H)
dt
− [E · {∇ × (v × D)} + H · {∇ × (v × B)}]
+j ∗ · (v × B) − ∇ · q + ρr (45)
Unknown parameters at this stage are stress tensor T, couple stress tensor Λ,
polarizations M and P , electric current density vector j ∗ and heat flux density
vector q.
Eq. (45) for internal energy is not yet in the form that is needed as the Poynt-
ing term in the equation requires further transformation. From the mathematical
identity for a vector triple product we have
∇ × (v × B) = (B · ∇) v − B (∇ · v) − (v · ∇) B (46)
and
∇ × (v × D) = (D · ∇) v − D (∇ · v) − (v · ∇) D + ρe v (47)
where use is made of Eqs. (3) and (4). Next the left sides of Eqs. (1) and (2) are
rewritten in the material time derivative form by adding (v · ∇)B and (v · ∇)D
to both sides of the equations, respectively. Then using Eqs. (46) and (47) the
following equations are obtained.
dB
= −∇ × E + (B · ∇) v − B (∇ · v) − ∇ × (v × B) (48)
dt
dD
= −j + ∇ × H + ρe v + (D · ∇) v − D (∇ · v) − ∇ × (v × D) (49)
dt
Scalar multiplication of both sides of (48) and (49) by H and E respectively,
then adding the equations and rearranging, yields the desired expression for the
divergence of the Poynting vector.
dD dB
∇ · (E × H) = −j ∗ · E − E · −H ·
dt dt
+ {ED + HB − (D · E + B · H) I} : ∇v
− [E · {∇ × (v × D)} + H · {∇ × (v × B)}] (50)
Substituting this expression into Eq. (45) yields the energy balance in the
following form.
du
ρ = TT : ∇v + ΛT : ∇Ω − T : · Ω
dt
− [ED + HB − (D · E + B · H) I] : ∇v
dD dB
+E · +H · + j ∗ · E ∗ − ∇ · q + ρr (51)
dt dt
Basic Equations for Magnetic Fluids 71
where, again
E∗ = E + v × B (52)
Substituting the definitions of D and B from Eqs. (26,27) into the electro-
magnetic work terms of Eq. (51) after minor algebraic transformation yields the
energy balance in final form.
dum
ρ = TT : ∇v + ΛT : ∇Ω − T : · Ω
dt
− [ED + HB − (D · E + B · H) I] : ∇v
dP dM
+E · +H · + j∗ · E∗ − ∇ · q
dt
dt
μ0 H 2 0 E 2
+ρR − + (∇ · v) (53)
2 2
where
0 E 2 μ0 H 2
um ≡ u − − (54)
2ρ 2ρ
The um of Eq.(54) represents internal energy of the medium exclusive of field
energy associated with the space the medium occupies.
ds
ρT = Ttv : ∇v + Λt : ∇Ω − ∇ · q + ρR − T : · Ω
dt
+ [(D · E + B · H)] I : ∇v − P · E (e) + M · H (e) I : ∇v
dM
dP
H − H (e) · + E − E (e) · + j∗ · E∗
dt dt
− 0 E 2 + μ0 H 2 ∇ · v (60)
The fifth term on the right side of Eq. (60) transforms as follows:
T : · Ω = Tv : · Ω + (P × E + M × H) · Ω
= Tv : · Ω − E · (P × Ω) − H · (M × Ω)
= Tv : · Ω − E − E (e) · (P × Ω)
− H − H (e) · (M × Ω) (63)
Also,
1 t
1
t
∇v = ∇v + (∇v) + ∇v − (∇v)
2 2
= D +·ω
(v)
(64)
t
where D (v) = 12 ∇v + (∇v) is the strain rate tensor and ω = 12 ∇ × v is the
fluid rate of rotation, i.e., one-half the fluid vorticity. Thus, the first term on the
right side of Eq. (60) can be written as:
Ttv : ∇v = Ttv : D (v) + · ω (65)
The sixth term on the right hand side of Eq. (60) transforms as follows:
[(D · E + B · H)] I : ∇v − P · E (e) + M H (e) I : ∇v
= 0 E 2 + μ0 H 2 + P · E − E (e) + M · H − H (e) ∇ · v (66)
9 Clausius-Duhem Inequality
The second law of thermodynamics states that the entropy increase dS in a
system is larger than the entropy dS(s) supplied with the difference between the
two compensated by positive internal entropy production dS(i) . This relationship
is expressed as follows:
dS(i) = dS − dS(s) ≥ 0 (68)
The inequality applied to a continuum can be written in the following form.
⎡ ⎤
d d q ρR ⎦
ρs(i) dV = ρsdV − ⎣− n · dS + dV ≥ 0 (69)
dt dt T T
Vm Vm Sm Vm
where q/T is local entropy flux. Applying the Reynolds transport theorem to the
time derivative terms and Gauss’ theorem to the surface integral term, noting
the arbitrariness of the volume, and multiplying by T on both sides yields
ds(i) ds q
ρT = ρT + T ∇ · − ρR ≥ 0 (70)
dt dt T
74 R.E. Rosensweig
Substituting this expression into the inequality of (70) together with the
expression for entropy production rate from Eq. (67) yields
ds(i) 1
ρT = q · ∇ ln + j∗ · E∗ + Φ ≥ 0 (72)
dt T
which is the Clausius-Duhem inequality for this system, where Φ is the dissipa-
tion function:
Φ = Ttv : D (v) + · (ω − Ω) + Λt : ∇Ω
dP
+ E − E (e) · + P (∇ · v) − Ω × P
dt
dM
+ H − H (e) · + M (∇ · v) − Ω × M ≥ 0 (73)
dt
and where it is assumed that heat flow and current flow produce dissipation
in the presence of mechanical dissipation and each other, and that no cross
couplings exist. Accordingly, the following conditions must be met.
1
q · ∇ ln ≥0 (74)
T
and
j∗ · E∗ ≥ 0 (75)
q = −k∇T (76)
j ∗ = σE ∗ (77)
Basic Equations for Magnetic Fluids 75
where (76) is Fourier’s first law, and (77) is Ohm’s law. Transforming the quan-
tities in (77) to the stationary coordinate system using Eqs. (42) and (52) leads
to
j = σ (E + v × B) + ρe v (78)
This implicity recovers the Lorentz force law and gives the relationship between
electric field and electric current density in a conductor containing free charge
of density e moving with velocity v when a magnetic flux density B exists.
where T is any dyadic, trT is the trace of T, T(a) the antisymmetric part of T,
and T − (1/3)ItrT. The following identity applies for the indicated product of
two dyadics:
1 (s) (s)
T:V= (trT) (trV) + T(a) : V(a) + T̆ : V̆ (80)
9
The first term on the right is a product of scalars. The second term on the
right can alternatively be written as the scalar product of two axial vectors.
Applying the identity to the first term on the right side of Eq. (73) gives:
1
Ttv D(v) + · (ω − Ω) = tr Ttv trD(v)
9
(a)
t (s) (v)
+ TTv : · (ω − Ω) + T̆v : D̆ (81)
Thus, non-negative values of entropy production are assured with the follow-
ing relationships.
t (s) (v)
T̆v = 2η D̆ , (82)
(a)
Ttv = −2ζ · (ω − Ω) , (83)
tr Ttv = γ trD(v) (84)
Substituting values from (82,83,84) now yields the following constituted form
of Tv .
(v) γ
Tv = 2η D̆ + 2ζ · (ω − Ω) + ItrD(v)
3
t
= η ∇v + (∇v) + 2ζ · (ω − Ω) + λ (∇ · v) I (86)
where η and λ are the shear and bulk coefficients of spin viscosity, respectively.
P ∼
= P 0 = χe 0 E (89)
∂ρ
ρ + v · ∇v = −∇p + (η + λ) ∇ (∇ · v) + η∇2 v
∂t
+∇ × [2ζ (ω − Ω)] + ρe E + j × B + (P · ∇) E
+ (M · ∇) H + P × (∇ × E) + M × (∇ × H) + ρb(91)
dΩ
I = (λ + η ) ∇ (∇ · Ω) + η ∇2 Ω − 4ζ (ω − Ω)
dt
+M × H + P × E + ρI (92)
ds η t
2
2 2
ρT = ∇v + (∇v) + 4ζ (ω − Ω) + λ (∇v)
dt 2
+η ∇Ω + ∇Ω t : ∇Ω + λ (trΩ)
2
j ∗2 1 2
+k∇2 T + + ρR + (M 0 − M ) (93)
σ χm τm
dM 1
+ M (∇ · v) = Ω × M + (M 0 − M ) (94)
dt τm
P = χe 0 E (95)
∂B
= −∇ × E (96)
∂t
∂D
+j =∇×H (97)
∂t
∇·B =0 (98)
∇ · D = ρe (99)
j = σ (E + v × B) + ρe v (100)
∗
j = j − ρe v (101)
D = 0 E + P B = μ0 H + M (102)
Note in Eq.(91) that terms may be combined as follows.
P · ∇E + P × (∇ × E) = (∇E) · P (103)
78 R.E. Rosensweig
M · ∇H + M × (∇ × H) = (∇H) · M (104)
The term M × (∇ × H) disappears for non-conductive fluid unless the E
field is time-varying and the corresponding electrical term disappears unless B
is time-varying, with both forces small, even in high frequency fields.
The last term in Eq. (93) is useful in computing radio frequency heating of
ferrofluids [16].
For negligible angular acceleration, spin diffusion, electric polarization, and
distant source of body couple, Eq. (92) simplifies to
1
Ω= M ×H +ω (105)
4ζ
Substituting for Ω in Eq. (94) yields the relaxation equation in the commonly
useful form for incompressible fluid as
dM 1 1
=ω×M − (M − M 0 ) − M × (M × H) (106)
dt τm 4ζ
12 Conclusion
This work derived a general set of basic equations for flow and response of con-
ductive magnetic fluids with internal rotation based on integral balance equa-
tions and thermodynamics. The work is free of arbitrary definitions of electro-
magnetic energy density and stress. Employing a strict formulation of entropy
production in which applied electromagnetic field is distinguished from its equi-
librium counterpart, the usual constitutive relationships emerge from the analy-
sis without empirical assumption, and include the magnetization relaxation rela-
tionship. The work gives support to the Minkowski form of the electromagnetic
momentum density D × B.
Basic Equations for Magnetic Fluids 79
This study has treated the Maxwell type stress of a magnetic fluid on the same
basis as ordinary contact stresses of pressure and viscous stress and produces
the accepted form of the linear and angular momentum balances. In addition,
the work shows that the product of Maxwell type stress and velocity is useful in
formulating the integral energy balance.
As a recommendation for further work it is desirable that a broader treatment
be developed such that the stress tensor in Eq. (62) is expressed in rest frame
variables. Such a theory would be exact to first order in velocity throughout. For
additional remarks, see Appendix C.
Acknowledgments
References
1. J. L. Neuringer and R. E. Rosensweig, Phys. Fluids 7 (12), 1927 (1964).
2. J. S. Dahler and L. E. Scriven, Nature 192, 36 (1961); D. W. Condiff and J. S.
Dahler, Phys. Fluids 7 (6), 842 (1964).
3. M. I. Shliomis, Soviet Phys. JETP (Engl. transl.) 34 (6), 1291 (1972).
4. M. I. Shliomis, Soviet Phys. Uspekhi (Engl. transl.) 17 (2), 153 (1974).
5. M. I. Shliomis, Phys. Rev. E, Comment on “Magnetoviscosity and relaxation in
ferrofluids” Phys.Rev.E.64 (2001).
6. S. R. deGroot and P. Mazur,Non-equilibrium thermodynamics, Dover, Mine-
ola,NY(1984).
7. B. U. Felderhof and H. J. Kroh, Jl. Chem. Phys. 110 (15), 7403 (1999).
8. K. Shizawa and T. Tanahashi, Bulletin of JSME, 29 (250), 1171 (1986); 29 (255),
2878 (1986).
9. C. Rinaldi and H. Brenner, “Body vs surface forces in continuum mechanics: Is the
Maxwell stress tensor physically objective?”, Phys. Rev E 65 (3); 036615 (2002).
10. J. A. Stratton, Electromagnetic theory, (McGraw-Hill, New York, 1941).
11. R. E. Rosensweig, Chapter 13 in G. Astarita, Thermodynamics:An advanced text-
book for Chemical Engineers, Plenum Press, New York (1989).
12. B. T. Chu, Phys. Fluids 2 (5), 473 (1959).
13. I. E. Tarapov, Magnit. Gidrodin. 1, 3 (1973).
14. E. Blums, Yu. A. Mikhailov, and R. Ozols, Heat and mass transfer in MHD flows,
World Scientific, Singapore (1987).
15. R. E. Rosensweig, Ferrohydrodynamics, Cambridge University Press, New York
(1985); reprinted with slight corrections by Dover, Mineola, New York (1997).
16. R. E. Rosensweig, Magnetohydrodynamics, 36 (4), 300 (2000).
80 R.E. Rosensweig
Symbols
Latin
b Body force per unit mass (N kg−1 )
B Magnetic induction, tesla (kg s−2 A−1 )
D Displacement vector (m−2 s A)
(v)
D Strain rate tensor (s−1 )
E Electric field vector in stationary coordinates, (m kg s A−1 ) −3
volts/meter
E ∗ Electric field vector in the moving medium, (m kg s−3 A−1 )
volt/meter
g Electromagnetic momentum vector (N s m−3 )
H Magnetic field vector in stationary coordinates (m−1 A)
∗
H Magnetic field vector in the moving medium (m−1 A)
I Unit tensor (dimensionless)
I Moment of inertia per unit volume (kg m−1 )
j Current density in stationary coordinates (A m−2 )
∗
j Current density in the moving medium (A m−2 )
l Body couple vector per unit mass transmitted from (N m kg−1 )
afar
M Magnetization, tesla (kg s−2 A−1 )
M 0 Equilibrium magnetization in field H (kg s−2 A−1 )
n Unit outward facing normal (dimensionless)
P Electric polarization vector (m−2 s A)
P 0 Equilibrium electric polarization vector in field E (m−2 s A)
p Pressure-like variable defined in Eq. (30) (N m−2 )
p0 Pressure in absence of electromagnetic fields (N m−2 )
p (e)
Pressure in presence of electromagnetic fields (N m−2 )
q Heat flux vector (J m−2 s−1 )
r Position vector (m)
R Internal heat release rate per unit mass (J kg−1 s−1 )
s Entropy per unit mass (J K−1 kg−1 )
s Spin vector, i.e., angular momentum per unit mass (m2 s−1 )
S Total entropy of a system (J K−1 )
S m Surface area enclosing volume Vm (m2 )
t Time (s)
tn Stress vector (N m−2 )
T Temperature (K)
T Stress tensor less electromagnetic momentum flux (N m−2 )
T̃ Stress tensor sum of pressure, viscous and electro- (N m−2 )
magnetic contributions
u Internal energy per unit mass including field energy (J kg−1 )
um Internal energy per unit mass excluding field energy (J kg−1 )
u (e)
Value of u in equilibrated system (J kg−1 )
(J kg−1 )
(e)
um Value of um in equilibrated system
Vm Volume containing mass (m3 )
Basic Equations for Magnetic Fluids 81
Greek
γ Lorentz transformation coefficient (dimensionless)
0 Permittivity of free space, 8.854 x 10−12 farad/meter (m3 kg−1 s4 A2 )
Polyadic alternator (dimensionless)
ζ Vortex vorticity (N s m−2 )
η Shear coefficient of viscosity (N s m−2 )
η Shear coefficient of spin viscosity (N s)
λ Bulk coefficient of viscosity (N s m−2 )
λ Bulk coefficient of spin viscosity (N s)
λn Couple stress vector (N m−1 )
Λ Couple stress tensor (N m−1 )
−7
μ0 Permeability of free space, 4π x 10 henry/meter (m kg s−2 A−2 )
Mass density (kg m−3 )
e Charge density, coulomb/meter 3
(A s m−3 )
σ Electrical conductivity, siemen (m kg−1 s3 A2 )
−3
Appendix A: Notation
Vectors are set in bold Roman such as v or B, or lower case bold Greek such as
λ.
Dyadics are indicated in bold sans serif such as T or bold upper case Greek
such as Λ.
T corresponds to êi êj Tij in indicial notation employing the Einstein summa-
tion convention; êi and êj are a unit vectors. Tij is defined as the j th component
of force acting on the surface having normal orientated in the i th direction.
TT is the transpose of T and corresponds to êi êj Tji .
The operator ∇ corresponds to êi ∂/∂xi .
The nesting convention is followed as shown by these examples:
ab · cd = (b · c) ad
ab : cd = (b · c) (a · d)
∇·T corresponds by the nesting convention to êi ∂Tij /∂xi and not êi ∂Tij /∂xj .
I is the unit dyadic corresponding to êi êj δij where δij is the Kronecker delta
function.
0 (i = j)
δij =
1 (i = j)
82 R.E. Rosensweig
⎧
⎨ 1 (ijk = 123, 231, or312)
ijk = 0 (i = j, i = k, orj = k)
⎩
−1 (ijk = 132, 213, or321)
where E ∗ is the electric field in the frame (rest frame) of an observer moving
with the local velocity, dl is the differential of path length, and Φ is the number
of flux linkages defined by
Φf = B · ndS (110)
Applying Stokes’s theorem to the left side of (109) converts the line integral
to a surface integral.
E ∗ · dl = − (∇ × E ∗ ) · ndS (112)
5
R. Aris, Vectors, tensors and the basic equations of fluid mechanics. Dover, Mineola,
NY (1989). An alternate derivation is given in W. K. H. Panofsky and M Phillips,
Classical electricity and Magnetism, Second Edition, Addison-Wesley, Reading, Mas-
sachusetts (1962, pages 160-300).
Basic Equations for Magnetic Fluids 83
E∗ = E + v × B (115)
This relationship is given in Eq.(38) of the text. The observer in the moving
frame senses an additional electric field due to motion through the induction
field.
Applying Stokes’s theorem to the left side of (116) converts the line integral
to a surface integral.
H ∗ · dl = − (∇ × H ∗ ) · ndS (118)
This may be compared with the form of Ampere’s extended equation written in
laboratory coordinates which is
∂D
∇×H =j+ (120)
∂t
Comparison of the two foregoing expressions yields the transformation rela-
tionships for magnetic field and electrical current density. These are:
H∗ = H − v × D (121)
j ∗ = j − ρe v (122)
Motion relative to charges alters the perception of electric current density as
shown by (122). Motion relative to the displacement field from (121) produces
a component of magnetic field in the observer’s frame.
Appendix C:
Comment on Electromagnetic Body Couple Density
From the exact expression for the rest frame stress tensor T∗m known from anal-
ysis of motionless fluid [15] and having the form given by Eq.(62), the associated
body couple density is given by vecT∗m = M ∗ × H ∗ + P ∗ × E ∗ . Expansion of
the semi-relativistic expressions [10] relating field vectors between frames gives
to first order in velocity,
M ∗ = M + μ0 v × P (123)
H∗ = H − v × D (124)
∗
P = P − 0 v × M (125)
∗
E = E+v×B (126)
Thus,
M ∗ × H ∗ + P ∗ × E ∗ = M × H + P × E − B × (v × D) + D × (v × B)
H E
+ 2 × (v × E) − 2 × (v × H)
c c
∼
= M ×H +P ×E+v×g (127)
where the second equality results as a vector identity following neglect of the
terms reciprocal to c2 , and g = D × B is the electromagnetic momentum den-
sity. In comparison to (127), the body couple corresponding to Eq. (62) in the
text is absent the v × g term. This limits the text result to quasi-static flows of
low velocity. Newton’s equations are galilei-invariant but electromagnetism au-
tomatically brings in Lorentz-invariance, hence it should be expected that only
a completely relativistic formulation could be totally satisfactory.
Ferrohydrodynamics: Retrospective and Issues
Mark I. Shliomis
Abstract. Two basic sets of hydrodynamic equations for magnetic colloids (so-called
ferrofluids) are reviewed. Starting from the quasistationary ferrohydrodynamics, we
then give a particular attention to an expanded model founded on the concept of
internal rotation. A specific relation between magnetic and rotational degrees of free-
dom of suspended grains provides a coupling of the fluid magnetization with the fluid
dynamics. Hence a complete set of constitutive equations consists of the equation of
ferrofluid motion, the Maxwell equations, and the magnetization equation. There are
three kinds of the latter. Two of them were derived phenomenologically as a gener-
alization of the Debye relaxation equation in case of spinning magnetic grains, while
one of them was derived microscopically from the Fokker-Planck equation. Testing the
magnetization equations, we compare their predictions about the dependence of the
rotational viscosity on the magnetic field and the shear rate.
dv d ∂
ρ = −∇p + η∇2 v + M ∇H , = + (v · ∇) , (1)
dt dt ∂t
the magnetic state equation M = M (H, T ), for which it is natural to employ
[3,4] the Langevin formula
H ∂M
F = (M · ∇)H − ∇ · M −ρ , (5)
2 ∂ρ H ,T
where the first is called sometimes the Kelvin force, and the second represents
the magnetostrictive force. Being the potential force, the latter is always in-
cluded into the pressure gradient in (1), i.e., it is equilibrated automatically by
the hydrostatic or hydrodynamic pressure. Further, since the equilibrium mag-
netization is collinear with the local magnetic field, M = (M/H)H, the Kelvin
force takes the form (M · ∇)H = (M/H)(H · ∇)H. Hence, using an identity
(H · ∇)H = 12 ∇H 2 − H × rot H
and taking into account the condition (3) of the absence of free electrical currents,
rot H = 0, we arrive at the pointed above expression F = M ∇H.
With allowance for (2) and the equality
d sinh ξ
L(ξ) = ln ,
dξ ξ
the Kelvin force can be written in the case of isothermal fluids, T = const, as
sinh ξ
F = nkB T ∇ ln . (6)
ξ
Ferrohydrodynamics: Retrospective and Issues 87
If magnetic grains are distributed uniformly all over the ferrofluid sample, n =
const, the last expression takes the form
sinh ξ
F = ∇ nkB T ln , (7)
ξ
i.e., the magnetic force proves to be potential. It makes an evidence of Bernoulli
theorem for non-vortical ferrofluid flow [2]
sinh ξ
p + ρgz + 12 ρv 2 − nkB T ln = const (8)
ξ
(g is the gravity acceleration and z is the vertical coordinate). This relationship
is to high extent useful in qualitative investigations of equilibriums and flows of
ferrofluids under nonuniform magnetic fields. Note, however, that a nonunifor-
mity of the field induces an inhomogeneity of the particle distribution which in
its turn leads to an alteration in magnetic force. Indeed, as seen from (6), each
a particle is acted upon by the force
sinh ξ
F /n = −∇U, U = −kB T ln , (9)
ξ
where U is the potential. The equilibrium (Boltzmann’s or “barometric”) particle
distribution is given by the formula n = n0 exp(−U/kB T ). Thus we find, using
(9),
sinh ξ sinh ξ(r) 3
n = n0 , n0 d r = n̄V. (10)
ξ V ξ(r)
The constant n0 is determined from the indicated normalization condition ex-
pressing conservation of the total particle number in the ferrofluid volume V; n̄
is the mean density of the particle number. Substituting n(ξ) from (10) into (6)
or (9) gives the resulting self-consistent magnetic force density
sinh ξ sinh ξ sinh ξ
F = n0 kB T ∇ ln ≡ n0 kB T ∇ = ∇(nkB T ) . (11)
ξ ξ ξ
When the fluid is at rest, this force is equilibrated by the pressure gradient,
F = ∇p [see (1)], which emerges automatically. Thus, one can say of magnetic
pressure pm = nkB T created by the thermal agitation of the magnetic grains in
an inert liquid. As it should be, pm is similar to the pressure of an ideal gas of
molecules.
One should not, however, be in hurry with the replacement
sinh ξ sinh ξ
n̄kB T ln =⇒ n0 kB T (12)
ξ ξ
in the Bernoulli theorem (8), because an equilibrium particle concentration n(ξ)
sets in for a very long time. The equilibrium is settled owing to diffusion and mag-
netophoresis which proceed extremely slow due to smallness of the particle diffu-
sion coefficient D. Actually, according to the Einstein formula, D = kB T /3πηd, it
88 M.I. Shliomis
larger the particle size the better it holds the condition of “freezing up” of the
particle magnetic moment into the particle body.
The quasistationary ferrohydrodynamics is valid just for colloids of Néel par-
ticles. Indeed, the coupling between magnetic and mechanical degrees of freedom
of such particles breaks down (hence M does not depends on Ω), and the magne-
tization relaxes to the direction of H almost immediately – for d (0.3 − 0.5)dS
there is τ τN ∼ 10−9 s (!) – as the model assumes.
The opposite case of suspensions of rigid magnetic dipoles (i.e., the grains
with “freezing in” magnetic moments, τN τB ) is considered below. In this
model with the infinitely strong coupling and the finite relaxation time τ τB ∼
10−4 − 10−5 s, the vector M is not obliged to be collinear with H. Therefore,
besides the magnetic force density (M · ∇)H, ferrofluid can undergo a magnetic
torque of the density M × H. The model of rigid dipoles is widely used in
the theory of magnetic fluids. Despite its simplicity, it allows one to explain a
wide complicated tangle of magnetic and hydrodynamic phenomena emerging
in ferrofluids under the field. Below we derive and discussed the basic equations
for the model.
dA d A
= ω p ×A + . (20)
dt dt
Substituting here A = M , ω p from (2), and d M /dt from (7), we obtain the
equation sought [3]:
dM 1 1
= Ω ×M − (M − M 0 ) − M ×(M × H). (21)
dt τB 6ηφ
whence it follows that the last (relaxation) term in (21) describes a process of
approach of the vector M to its equilibrium orientation without change of its
length. As the result, the relaxation rates of the longitudinal and transverse (to
the field) components of magnetization appears to be different. Let us calculate
them. In the linear approximation in M − M 0 one can split the nonequilibrium
part of magnetization into components parallel and perpendicular to the field:
H [H · (M − M 0 )] H ×(M ×H)
M − M0 = + . (22)
H2 H2
92 M.I. Shliomis
Substituting this expression into (21) gives the linear magnetization equation
dM H [H · (M − M 0 )] H ×(M ×H)
= Ω ×M − − , (23)
dt H 2 τ H 2 τ⊥
where τ = τB and
1 1 nmL(ξ)H 1 1
= + = 1 + ξL(ξ) ,
τ⊥ τB 6ηφ τB 2
so that
2τB
τ⊥ = . (24)
2 + ξL(ξ)
Let us revert to the Debye–like equation (18) and demonstrate that this postu-
lated by myself [3] relaxation equation can be easily derived [12] from irreversible
thermodynamics (IT). With this purpose it is convenient to take advantage of
the method of IT proposed by Landau and first applied by him just to the de-
scription of relaxation of the order parameter in a nonequilibrium system [16].
An equilibrium value of the parameter (M 0 in our case) corresponds to the
minimum of an appropriate thermodynamic potential Φ (usually the Gibbs or
Helmholtz free energy) depending on the magnetization M and other thermo-
dynamic variables. Thus, at the equilibrium ∂Φ/∂M = 0. Out of equilibrium
this condition is not satisfied, so the relaxation process occurs: M changes in
time approaching M 0 . For small deviations from equilibrium, the derivative
∂Φ/∂M and the relaxation rate dM /dt are small. The relation between the two
derivatives in the Landau theory is reduced to simple proportionality:
dM ∂Φ
= −γ (25)
dt ∂M
with a constant coefficient γ > 0. Hence we have
2
dΦ ∂Φ dM ∂Φ
= · = −γ <0 (26)
dt ∂M dt ∂M
as it should be: when a system moves to equilibrium, its free energy decreases.
In the case of a weakly nonequilibrium state of the system, one can substitute
in (25) and (26) the expansion
2
∂Φ ∂Φ ∂ Φ
= + (M − M 0 ) + ... ,
∂M ∂M 0 ∂M 2 0
where subscript 0 marks the point of equilibrium. As the first derivative in this
point is equal to zero and the second one is positive, equation (25) turns into
(18) with τB−1 = γ(∂ 2 Φ/∂M 2 )0 and (26) takes the form
dΦ (M − M 0 )2
=− . (27)
dt γ τB2
Ferrohydrodynamics: Retrospective and Issues 93
Thus, (18) and hence (21) are well corroborated by the method of IT. Notice that
some different thermodynamic method based on a strict formulation of entropy
production leads [17] to the same phenomenological magnetization equation: Eq.
(21) coincides with Eq. (106) from [17]. Equations (3), (17) and (21) constitute
the complete set of conventional ferrohydrodynamic equations.
equations. Ideally, however, one would like to have only one equation since only
the first moment – magnetization – has a clear physical meaning. An original
scheme of closure of the first-moment equation (29), titled the effective field
method (EFM), has been proposed in [11]. Let us explain the fruitful physical
idea.
In equilibrium (Ω = 0) under a constant magnetic field the stationary solu-
tion of (28) is the Gibbs distribution
ξ
W0 (e) = exp(ξ · e) . (30)
4π sinh ξ
An averaging of the vector e with function (30) gives expression (19) for the
equilibrium magnetization. Note that only in true equilibrium the magnetiza-
tion is one or another function of the field. In a nonequilibrium state there is no
connection between M and H : any arbitrary magnetization may be created –
in principle – even in the absence of the field. Nevertheless, one may consider
any value of M as an equilibrium magnetization in a certain – specially pre-
pared – magnetic field. This effective field H e is related to the nonequilibrium
magnetization by the equilibrium relation:
During the equilibrium settling process, the magnetization (31) relaxes to its
equilibrium value (19) as the effective field H e (or ζ) approaches the true field
H (or ξ). Comparing (19) and (31) we see that the latter is obtained by averaging
of e with the distribution function
ζ
We (e) = exp(ζ · e) , (32)
4π sinh ζ
which differs from the Gibbs distribution (30) by replacement of the true dimen-
sionless field ξ by the effective field ζ. Carrying out the averaging in (29) with
the function (32), we find the sought equation [11]
dM (ξ · ζ) M 1 1 1 M × (M × H)
= Ω ×M − 1 − − − . (33)
dt ζ2 τB L(ζ) L(ζ) ζ 6ηφ
are parallel and perpendicular to the true field, respectively. Employing the rela-
tion (34), one can reduce (33) to the linear magnetization equation (23), where
relaxation times of the longitudinal and transverse components of magnetization
are
d ln L(ξ) 2L(ξ)
τ = τB , τ⊥ = τB . (35)
d ln ξ ξ − L(ξ)
It is well-established that equation (33) derived by EFM describes magne-
tization processes in real magnetic colloids very well. Predictions of the theory
are well corroborated by experiments on positive [27] and negative [28,29] rota-
tional viscosities, and on magneto-vortical birefringence [24] in ferrofluids. So-
lutions of (33) agree perfectly with the results of numerical integration of the
non-stationary Fokker-Planck equation [13], and what is more, with the numer-
ical simulation of Brownian dynamics performed by Cebers [21] on the basis of
Langevin equation for rotational motion of magnetic grains in the presence of
magnetic field. Phenomenological equation (21) is certainly far simpler for anal-
ysis than (33). However, the latter guarantees the correct description of magneti-
zation even if its deviation from equilibrium value is large (e.g., at Ωτ 1), that
is when (21) leads to erroneous results. Interestingly, a new phenomenological
magnetization equation derived recently [12] from irreversible thermodynamics
(see below) is free partially from the above mentioned shortcoming of (21).
dH e ∂ Φ̃
= −γ̃ , γ̃ > 0.
dt ∂H e
d H e 1
= − (H e − H) , (36)
dt τB
where we set γ̃ −1 = (∂ 2 Φ̃/∂H 2e )τB . Under this choice, (36) turns into (18) in the
low-field limit, when the true magnetization and its equilibrium value take the
form M = χH e and M 0 = χH, respectively; here χ = nm2 /3kB T stands for
the initial magnetic susceptibility.
Equation (36) is written out in a rotating coordinate system Σ . Reverting
to the immobile system Σ by the general formula (20) and eliminating ω p with
the aid of torque-balance equation (15), we obtain [12]
dH e 1 1
= Ω ×H e − (H e − H) − H e ×(M ×H). (37)
dt τB 6ηφ
96 M.I. Shliomis
This equation (marked below as Sh’01) coincides with the previous phenomeno-
logical equation (21) (marked as Sh’72) in the limit of low magnetic field. How-
ever, due to nonlinearity of the Langevin magnetization law, equations (21) and
(37) predict very different magnetization values for large enough magnitudes of
ξ. Note that as shown above, both these equations are in full agreement with
fundamental principles and hence both they are correct from phenomenological
point of view: irreversible thermodynamics gives preference neither to (21) nor
to (37).
In the case of small deviations from equilibrium, equation (37) can be lin-
earized with respect to H e − H and M − M 0 . Substituting ν = ζ − ξ =
(m/kB T )(H e − H) into (34) gives
dL(ξ) L(ξ)
(M − M 0 ) = 3χ (H e − H) , M ⊥ = 3χ (H e )⊥ .
dξ ξ
Employing these relationships and the equality
dM dM dH e L(ξ) d(H e ) L(ξ) d(H e )⊥
= · = 3χ +
dt dH e dt ξ dt ξ dt
For the Poiseuille or Couette flow, v = (0, v(x), 0), in a transversal magnetic
field, H = (H, 0, 0), the magnetization has two components: M = (Mx , My , 0),
so (38) may be written in the form fτ = 2(η + ηr )Ω, where 2Ω = ∂v/∂x and
rotational viscosity is defined as
ηr = My H/4Ω . (39)
M ⊥ = τ⊥ Ω × M 0 , i.e., My = τ⊥ M0 Ω . (40)
Substituting here M0 = nmL(ξ) and τ⊥ from (24), we arrive at the formula [3]
3 ξL(ξ) 3 ξ − tanh ξ
ηr (ξ) = ηφ = ηφ . (42)
2 2 + ξL(ξ) 2 ξ + tanh ξ
So, both phenomenological equations, (21) and (37), predict the same depen-
dence of rotational viscosity on magnetic field strength. In the absence of the
field an individual particle “rolls” freely along corresponding shear surface with
angular velocity ω p equal to Ω , so that ηr (0) = 0 . Conversely, ηr (ξ) attains
its limiting value ηr (∞) = 32 ηφ (the saturation) when rolling of the particle is
replaced by slipping: the field of sufficiently large intensity guarantees constancy
of the particle’s orientation, not allowing it to twist with the fluid. Note that
the saturation value of ηr does not depend on a concrete form of the magne-
tization equation but follows directly from the equation of fluid motion (17).
Actually, in the limit under consideration ω p = 0 , so that (15) takes the form
M × H = −6ηφΩ . Substituting this torque in (17) and grouping there the
second and fourth terms,
η∇2 v + 12 rot (M × H) = η + 32 ηφ ∇2 v ,
3 ξL2 (ξ)
ηr (ξ) = ηφ . (43)
2 ξ − L(ξ)
Figure 1 shows that though at first sight functions (42) and (43) do not
appear alike, they agree closely in the entire range of their argument. Both of
them approach the saturation value ηr (∞) = 32 ηφ at ξ 1. In the figure we
98 M.I. Shliomis
1.0
Ωτ=0
0.8
ηr(ξ)/ηr(∞)
0.6
0.4
MRSh
0.2 Sh’72,’01
0.0
0 5 10 15 20 25 30
ξ
Fig. 1. Reduced rotational viscosity in the Newtonian limit ΩτB → 0 versus the di-
mensionless field strength, as calculated from (42) (Sh’72 [3] and Sh’01 [12]) and (43)
(MRSh [11]).
plot the reduced rotational viscosity ηr (ξ)/ηr (∞) as a function of ξ. The upper
curve calculated by (43) of the EFM [11] represents a very good approximation.
Actually, as shown in [9,22,23], it hardly differs from the direct solution of the
Fokker-Plank equation (28) in linear approximation in ΩτB . The phenomenolog-
ical equations, (21) and (37), also provide a quite satisfactory description of the
rotational viscosity. Both they result in the lower curve in Fig. 1 that is described
by the Shliomis’ formula (42). This function agrees with (43) in the low- and
high-field limits and deviates from it, at most, on 15 % in the entire range of the
argument ξ. As seen from (39), ηr does not depend on the flow vorticity till My
is proportional to Ω, what takes place only if ΩτB 1. For finite values of ΩτB
the viscosity (39) does depend on Ω. As a result, the function στ n (Ω) deviates
from the linear one, i.e., a ferrofluid acquires rheological properties.
Proceeding to the calculation of the non-Newtonian viscosity on the basis of
(37), it is convenient to pass from the fields H and H e to their nondimensional
values ξ and ζ :
dζ 1 L(ζ)
=Ω×ζ− (ζ − ξ) − ζ × (ζ × ξ) . (44)
dt τB 2τB ζ
At the stated above arrangement of the applied magnetic field with respect to
the fluid flow, the last equation admits a steady solution in which the effective
field ζ tracks the true field ξ with lag angle α, i.e., ζ = (ζ cos α, ζ sin α, 0). The
Ferrohydrodynamics: Retrospective and Issues 99
3 ζL(ζ)
ηr = ηφ . (Sh 01) (46)
2 2 + ζL(ζ)
2ΩτB ζL(ζ) ζ
ξ2 − ζ 2 = , cos α = , (47)
ζ − L(ζ) ξ
that results in
3 ζL2 (ζ)
ηr = ηφ . (MRSh) (48)
2 ζ − L(ζ)
The solution of equation (21) can be presented in a similar form. Let us in-
troduce a new variable ζ instead of M by the relation M = M0 (ζ/ξ) where
M0 = nmL(ξ). It is worth noting that ζ is no more an effective field unlike
ζ in preceding relationships (45)–(48). By substituting the components Mx =
M0 (ζ/ξ) cos α and My = M0 (ζ/ξ) sin α in (21), we get cos α = M/M0 = ζ/ξ and
2ΩτB ξζ
ξ2 − ζ 2 = . (49)
2ξ + ζ 2 L(ξ)
3 ζ 2 L(ξ)
ηr = ηφ . (Sh 72) (50)
2 2ξ + ζ 2 L(ξ)
In the limit ΩτB 1, one can neglect the value ΩτB in (45), (47) and (49), after
what all three of these relationships are reduced to ζ = ξ. Eliminating now ζ from
(13) and (17), we see that, as it should be, expressions (50) and (46) obtained
from the old and the new phenomenological magnetization equations turn into
(42), while the EFT formula (48) is transformed into (43). When, however, the
ferrofluid is subjected to a sufficiently large shear rate, ΩτB ≥ 1, the flow induces
– along with the Brownian motion – a quotient demagnetization since under the
viscous shear the magnetic grains tend to be rotated out of alignment with the
magnetic field. Formally, this effect originates from decreasing the parameter ζ
determined by (45), (47) and (49). According to these equations, ζ = ξ when
ΩτB = 0 but the more there is of Ωτ at constant ξ, the less there is of ζ. The
reduction of the magnetization leads in turn to some decrease in the rotational
viscosity. This decrease, imperceptible in practice up to ΩτB 1, then becomes
very significant. Figures 2–4 illustrates the dependence of the viscosity increase
on the magnetic-field strength for three values of the product ΩτB . Interestingly,
under the finite shear rate the viscosities given by (46) and (50) do not coincide
100 M.I. Shliomis
1.0
Ωτ=2
0.8
ηr(ξ)/ηr(∞)
0.6
0.4
MRSh
0.2 Sh’01
Sh’72
0.0
0 5 10 15 20 25 30
ξ
Fig. 2. Dependence of the rotational viscosity on the field for the dimensionless shear
rate ΩτB = 2, as calculated from (47)–(48) MRSh, (45)–(46) Sh’01, and (49)–(50)
Sh’72.
1.0
Ωτ=4
0.8
ηr(ξ)/ηr(∞)
0.6
0.4
MRSh
0.2 Sh’01
Sh’72
0.0
0 5 10 15 20 25 30
ξ
Fig. 3. Same as Fig. 2, but for ΩτB = 4.
Ferrohydrodynamics: Retrospective and Issues 101
1.0
Ωτ=6
0.8
ηr(ξ)/ηr(∞)
0.6
0.4
MRSh
0.2 Sh’01
Sh’72
0.0
0 5 10 15 20 25 30
ξ
Fig. 4. Same as Figs. 2 and 3, but for ΩτB = 6.
with each other any more. As seen from the plot, the higher the shear the
more discrepancy between viscosity values predicted by the new and the old
phenomenological equations. At high shear in a high field, the old equation
[(21), Sh’72] predicts a hysteresis of viscosity (see Fig. 4), which however is
corroborated neither by direct calculations of [13,20] nor by the solution (47)–
(48) of the EFM equation (33). The new equation [(37), Sh’01] also does not
predict such a hysteresis, but it provides us with a quite satisfactory viscosity
description in a wide region of parameters ξ and ΩτB . Indeed, in this entire
region the solutions of (33) and (37) agree closely, as shown in Figs. 2–4. Thus,
in the case of a stationary magnetic field, equation (37) can be recommended
for an employment on the same level with (33). It is worth nothing that all the
above calculations, carried out for a shear flow, apply equally to a rigid rotation
of a ferrofluid with an angular velocity Ω in a constant transversal magnetic
field, H ⊥ Ω, and to a quiescent ferrofluid subjected to a uniform rotating field
H = (H cos Ωt, H sin Ωt, 0) as well.
the effective field is evidently parallel to the true field, while its magnitude is
determined by the equation
−1
dζ d ln L(ζ)
=− (ζ − ξ0 cos ωt). (56)
dt d ln ζ
where h = H/H is the unit vector along the field and function Ψ (ζ) is satisfied
to the linear equation
dΨ ξ0 1 1
=1− − Ψ cos ωt. (57)
dt 2 L(ζ) ζ
The magnetic torque M×H can be averaged over the period of field variation
2π/ω because we are interested in ω ∼ τB−1 while τB is always much less than
the characteristic hydrodynamic time ∼ ρl2 /η, where l is the reference spatial
scale of the fluid motion. For the mean magnetic torque one gets
Averaging now the equation of fluid motion (17) over the time and substituting
into (17) the mean torque (58), one may group the second and fourth terms in
this equation:
η∇2 v − 3ηφg rotΩ = η (1 + 32 φg)∇2 v.
Thus the rotational viscosity is ηr = 32 ηφg. In accordance with (52) and (58),
the angular velocity of the particles may be written as ωp = (1 − g)Ω, so
that ηr can be presented in the form (53). For arbitrary ξ0 and ωτB the prob-
lem (56)–(57) was solved numerically [28]. Results of the computations are dis-
played in the map of viscosity, Fig. 5, where a set of isolines of reduced viscosity
g(ξ0 , ωτB ) = ηr (ξ0 , ωτB )/ηr (∞, 0) is presented in the plane (ξ0 , ωτB ). The iso-
line g = 0 parts the plane into two regions: g > 0 for ωτB ≤ 1 and g < 0 for
ωτB ≥ 1. A similar map of viscosity, but calculated from the phenomenologi-
cal magnetization equation (21), Sh’72, is shown in Fig. 6. A strong similarity
between a plot of experimentally obtained isolines and that of Fig. 5 has been
observed in [28]. As for the map in Fig. 6, the experimental data [28] agree
with it on the whole, but there are some discrepancies in details. Namely, in
the experiment [28] as well as in Fig. 5 the neutral isoline, g = 0, bends to
the right, while in Fig. 6 it bends slightly to the left. Still more there is the
difference between the isolines of Fig. 5 and those as calculated from the new
phenomenological equation (37), Sh’01. Generally, this equation is very appro-
priate for description of ferrofluid magnetization in a stationary magnetic field
(see Sect. 6.1), whereas in an alternating field it is working satisfactory only in
the low-frequency limit ωτB 1. [The only exception to the rule represents itself
the case of rotating magnetic field of the kind of H = (H0 cos ωt, H0 sin ωt, 0).
104 M.I. Shliomis
20
−0.7
−0.5
−0.3
−0.1
15
0
+0.1
+0.3
+0.5
10
ξ0
+0.7
0
0 1 2 3 4 5 6
ωτB
Fig. 5. Isolines of reduced viscosity ηr (ξ0 , ωτB )/ηr (∞, 0) = g in the plane (ξ0 , ωτB )
for −0.7 ≤ g ≤ +0.7 , as calculated from EFM equation (33).
20
−0.7
−0.5
−0.3
−0.1
15
0
+0.1
+0.3
+0.5
10
ξ0
+0.7
0
0 1 2 3 4 5 6
ωτB
i.e., it does not depend on ξ . Analogously, (59 c) has the solution He (t) =
H exp(−t/τB ), so that we find
The last decay predicted by the new magnetization equation (37), Sh’01, is
exponential only in the limit ξ 1 , while the old phenomenological equation
(21) and – what is much more important – the EFM equation (33) predicts
exponential decay of the magnetization for any values of ξ . This difference in
relaxation behavior side by side with the difference in the ferrofluid viscosity can
be of relevance for testing the magnetization equations and the interpretation of
corresponding experiments. At the description of nonstationary situations (like
that as presented in Fig. 7) one should, however, a priori give preference to (21)
and (33) before (37) since all previous predictions of (33) always were realized.
H = (H0 cos ωt, H0 sin ωt, 0), M = [M cos(ωt − α), M sin(ωt − α), 0]. (62)
106 M.I. Shliomis
1.0
ξ=0.05
0.8 ξ=5
ξ=10
ξ=20
0.6
M/M0
0.4
0.2
0.0
0 1 2 3 4 5 6
t/τB
Fig. 7. Time dependence of the reduced magnetization M (t)/M0 after the field ξ is
switched off, as described by (61). The lowest curve also represents the solution (60)
for any field.
while the fluid remains quiescent [31]. The point is, each a grain forces to spin
the near-by mass of the viscous liquid thereby becoming a center of microscopic
Ferrohydrodynamics: Retrospective and Issues 107
vortex, the size of which does not exceed the mean distance between the grains
∼ φ−1/3 d. But such a motion is not hydrodynamic yet, because averaging of the
microwhirls over physically small fluid volume does not result in macroscopic
vorticity Ω = 12 rot v. Indeed, microwhirls created by the neighboring grains
compensate each other like elementary electric currents of neighboring molecules
in Ampere’s model of ferromagnetism do that.
The uniformity of both the internal moment of momentum S = Iω p and
the magnetization M breaks off at the ferrofluid border. A jump-like increase
in magnetization on the boundary surface (outside the fluid M = 0) results
in magnetic tangential surface stress. The latter just provide the coupling of
rotating field and the fluid motion.
Let us consider ferrofluid confined within the layer by horizontal planes x =
0, l. Under the field (62) rotating in the vertical plane (x, y), the fluid can come
into the motion along axis y, v = [0, v(x), 0], so that Ω = (0, 0, Ω) where
Ω = 12 (dv/dx). The tangential force density (38) acting the upper surface is
dv
fy = [σyx ] = η − 1 (Mx Hy − My Hx ) . (65)
dx 2
Substituting from (62) and (63) into (65) yields
dv (ω − Ω)τ⊥
fy = η − 12 M H0 sin α = 2ηΩ − 12 M0 H0 2 . (66)
dx 1 + (ω − Ω)2 τ⊥
As seen from (66), if the surface x = l is free, fy = 0, the rotating field generates
the flow with the constant vorticity
M0 H0 ωτ⊥
Ω≈ · 2 , (67)
4η 1 + ω 2 τ⊥
so the plane Couette flow – with the linear velocity profile v(x) = 2Ωx – arises.
[Note that we have neglected in the right-hand side of (67) the value Ω in
comparison with ω since Ω/ω ≤ 14 φξ 2 at ξ 1 and Ω/ω ≤ 32 φ at ξ 1]. If,
conversely, the upper surface is fixed, Ω = 0, to hold it one needs to apply the
force in the opposite direction:
ωτ⊥
fy = − 12 M0 H0 2 . (68)
1 + ω 2 τ⊥
Let the field (62) rotates in horizontal plane around a long (l R) vertical
cylinder of the radius R completely filled with ferrofluid. Then the lateral walls
of the cylinder are acted upon by the tangential force density
dv v
fϕ = [σϕr ] = η − + 12 M H0 sin α , (69)
dr r
cf. (65), where v(r) is the azimuthal component of the fluid velocity and H0
stands for the field amplitude inside the fluid. Restricting ourselves by the case
of low enough field strength, ξ < 1, we obtain from (63) and (69)
dv v (ω − Ω)τB dv v
fϕ = η − 1
+ 2 χH02
, Ω=2 1
+ . (70)
dr r 1 + (ω − Ω)2 τB2 dr r
108 M.I. Shliomis
This relationship allows us to calculate the magnetic torque T = 2πR2 lfϕ acting
on the motionless cylinder (v = dv/dr = 0):
ωτB
T = χH02 V , (71 a)
1 + ω 2 τB2
χH02 V ωτB
T = . (71 b)
(1 + ω 2 τB2 )[(1 + 2πχ)2 + ω 2 τB2 ]
This expression agrees well with experimental data [33,34]. To interpret (71 b), it
should be noted that the stationary fields H and H are coupled by the relation
H
H=
1 + 2πχ
where 2π is the demagnetization factor of cylinder in the direction transverse
to its axis of symmetry (axis z). Therefore, at the low frequency of the field
rotation, ωτB 1, the magnetic torque (71 b) may be presented in its usual
form [5]
T = M × H, (72)
where we have introduced the total magnetic moment of the cylinder
χH
M = M V, M=
1 + 2πχ
and have taken into account that the small angle between M and H is equal to
ωτB .
Revert to Eq. (70). If the cylinder is not fixed, the equation fϕ = 0 determines
the velocity of free fluid rotation v(r) = ωr, i.e., Ω = ω: the fluid rotates as a
whole about the vertical axis. Finally, if the cylinder is fixed, but not completely
filled with ferrofluid, the latter comes into motion due to magnetic tangential
stresses on the free fluid surface. The moving surface involves an adjacent fluid
layer and forms the circular hydrodynamic flow first observed in [34]. This phe-
nomenon known as rotational effect has long held fascination for investigators.
Many theories of the spin-up motion were devoted to the effect, but only re-
cently it was clarified by experiments and theory of Rosensweig et al. [35] and
Pshenichnikov et al. [32,33,36].
7 Conclusion
Thus, there exist two basic models of the ferrofluid dynamics. One of the two
represents a quasistationary theory (Sect. 1), which provides an irreproachable
description of any ferrohydrostatic problems. It is also widely employed at the
Ferrohydrodynamics: Retrospective and Issues 109
Acknowledgements
I thank Ronald Rosensweig for his comments on the manuscript and Alexei
Krekhov for providing Figs. 5 and 6. This work was supported by the Alexander
von Humboldt Foundation by dint of the Meitner–Humboldt research award,
and the Israel Science Foundation under Grant No. 336/00.
References
1. J.L. Neuringer, R.E. Rosensweig: Phys. Fluids 7, 1927 (1964)
2. R.E. Rosensweig: Ferrohydrodynamics (Cambridge University Press, Cambridge
1985)
3. M.I. Shliomis: Sov. Phys. JETP 34, 1291 (1972)
4. M.I. Shliomis: Sov. Phys. Usp. 17, 153 (1974)
5. L.D. Landau, E.M. Lifshitz: Electrodynamics of Continuous Media, second ed.
(Pergamon Press, New York 1984)
6. M.I. Shliomis: ‘Convective Instability of Magnetized Ferrofluids: Influence of Mag-
netophoresis and Soret Effect’. In: Thermal nonequilibrium phenomena in fluid
mixtures, ed. by W. Köhler, S. Wiegand (LNP, Springer, 2001)
7. B.A. Finlayson: J. Fluid Mech. 40, 753 (1970)
8. M.I. Shliomis, B.L. Smorodin: J. Magn. Magn. Mater. (2002) (in press).
9. Yu.L. Raikher, M.I. Shliomis: Adv. Chem. Phys. 87, 595 (1994)
10. M.I. Shliomis: Sov. Phys. JETP 24, 173 (1967)
110 M.I. Shliomis
11. M.A. Martsenyuk, Yu.L. Raikher, M.I. Shliomis: Sov. Phys. JETP 38, 413 (1974)
12. M.I. Shliomis, Phys. Rev. E 64, 063501 (2001)
13. M.I. Shliomis, T.P. Lyubimova, D.V. Lyubimov: Chem. Eng. Comm. 67, 275
(1988)
14. M.I. Shliomis, Phys. Rev. E 64, 060501(R) (2001)
15. P. Debye: Polar Molecules (Dover, New York 1929)
16. L.D. Landau, I.M. Khalatnikov: Dokl. Akad. Nauk SSSR 96, 469 (1954)
17. R.E. Rosensweig: this issue, p. 63.
18. J.P. McTague: J. Chem. Phys. 51, 133 (1969)
19. E.N. Mozgovoi, E.Ya. Blum, A.O. Tsebers: Magnetohydrodinamics 9, 52 (1973)
20. O. Ambacher, S. Odenbach, K. Stierstadt: Z. Phys. B–Condensed Matter 86, 29
(1992)
21. A.O. Cebers: Magnetohydrodynamics 20, 343 (1984); 21, 357 (1985); E. Blums,
A. Cebers, M. Maiorov: Magnetic Fluids (W. de Gruyter, Berlin 1997)
22. A.C. Levi, R.F. Hobson, F.R. McCourt: Canad. J. Phys. 51, 180 (1973)
23. B. U. Felderhof: Magnetohydrodinamics 36, 396 (2000)
24. B.M. Heegaard, J.-C. Bacri, R. Perzynski, M.I. Shliomis: Europhys. Lett. 34, 299
(1996)
25. F. Gazeau, B.M. Heegaard, J.-C. Bacri, A. Cebers, R. Perzynski: Europhys. Lett.
35, 609 (1996)
26. F. Gazeau, C. Baravian, J.-C. Bacri, R. Perzynski, M.I. Shliomis: Phys. Rev. E
56, 614 (1997)
27. J.P. Embs, H.W. Müller, C. Wagner, K. Knorr, M. Lücke: Phys. Rev. E 61, R2196
(2000)
28. J.-C. Bacri, R. Perzynski, M.I. Shliomis, G.I. Burde: Phys. Rev. Lett. 75, 2128
(1995)
29. A. Zeuner, R. Richter, I. Rehberg: Phys. Rev. E 58, 6287 (1998)
30. M.I. Shliomis, K.I. Morozov: Phys. Fluids 6, 2855 (1994)
31. V.M. Zaĭtzev, M.I. Shliomis: J. Appl. Mech. Tech. Phys. 10, 696 (1969)
32. A.F. Pshenichnikov, A.V. Lebedev: Magnetohydrodinamics 36, 317 (2000)
33. A.V. Lebedev, A.F. Pshenichnikov: J. Magn. Magn. Mater. 122, 227 (1993)
34. R. Moskowitz, R.E. Rosensweig: Appl. Phys. Lett. 11 (1967)
35. R.E. Rosensweig, J. Popplewell, R.J. Johnston: J. Magn. Magn. Mater. 85, 171
(1990)
36. A.F. Pshenichnikov, A.V. Lebedev, M.I. Shliomis: Magnetohydrodinamics 36, 339
(2000)
Ferrohydrodynamics: Retrospective and Issues 111
Supplementary Glossary
M magnetization
H magnetic field
H e effective magnetic field
B magnetic induction: B = H + 4πM
F magnetic force (volume density)
v ferrofluid velocity
Ω flow vorticity: Ω = 12 rot v
S internal angular momentum (volume density): S = Iω p
ω p angular velocity of magnetic particles
I moment of inertia of particles in a unit volume
p pressure
σik stress tensor
T temperature
kB Boltzmann’s constant
ρ mass density
φ concentration of magnetic grains (volume fraction)
n number density of magnetic grains
m magnetic moment of a single particle: m = Md V
V volume of a single particle
d particle diameter
D particle diffusion coefficient
Md domain magnetization of the particle material
W orientational distribution function of particle magnetic moments
e unit vector along particle magnetic moment: e = m/m
h unit vector along magnetic field: h = H/H
ξ dimensionless magnetic field: ξ = mH/kB T
ζ dimensionless effective magnetic field: ζ = mH e /kB T
L(ξ) Langevin function: L(ξ) = coth ξ − ξ −1
η shear viscosity
ηr rotational viscosity
χ initial magnetic susceptibility
Φ thermodynamic potential
t time
τB Brownian diffusion time
τN Néel diffusion time
τ⊥ transverse magnetization relaxation time
τ longitudinal magnetization relaxation time
ω magnetic field frequency
Ferrofluid Dynamics
an enhanced effective shear viscosity [6,3]. Even more spectacular is the acceler-
ation of the flow in response to a high-frequency AC-field [7,8,9] (also denoted as
”negative viscosity”, though only a negative viscosity increment was observed ).
That way, the oscillating magnetic field pumps energy into the rotating motion
of the ferromagnetic grains resulting in an acceleration of the flow.
The occurrence of the magneto-viscous effects is intimately related to the
fact that the local magnetization M(r, t) deviates from its equilibrium value
Meq [H(r, t)], where H(r, t) is the local magnetic field. Significant increments
δM = (M−Meq ) are expected to appear whenever the magnetic relaxation time
τ compares to the other relevant hydrodynamic time scales. Phenomena related
to the finiteness of τ are commonly denoted as magneto-dissipative effects.
As outlined in Rosensweig’s textbook, the microscopic mechanism responsible
for the magneto-relaxation is either due to particle rotation against the viscosity
of the liquid carrier (Brownian rotational diffusion) or by re-orientation of the
magnetic moments relative to the crystallographic orientation of the ferromag-
netic grains (Néel relaxation). Which of these mechanisms predominates depends
on the specific anisotropy energy of the employed ferromagnetic material, the
size of the suspended grains, and the viscosity of the carrier liquid. Since real
ferrofluid suspensions usually exhibit a broader particle size distribution, it is in
general a combination of both processes which determines the effective magnetic
relaxation time for a given ferrofluid species.
Assuming that the Brownian mechanism is the principal source of dissipa-
tion, the intuitive picture of particles rotating against the viscous carrier lead
Shliomis [3] to his theory for magneto-dissipative ferrohydrodynamics. To that
end he included both the magnetization M and the mechanical angular mo-
mentum density S of the grains as additional thermodynamic variables. After
eliminating the latter, an extra momentum flux remains, which enters the stress
tensor in the form
ΔΠij = 12 εijk (H × M)k , (1)
This term exactly compensates the antisymmetric part of Maxwell’s stress Hi Bj ,
if H and M are non-parallel. Clearly, being treated as a separate independent
variable, the magnetization requires an extra evolution equation. According to
Shliomis there are two versions of this equations, of which the first is a phe-
nomenological relaxation equation with a Debye-like relaxation term in the form
δM/τ . In the ferrofluid literature this approach is frequently referred to as the
Debye theory. In combination with Eq (1), many magneto-dissipative phenom-
ena, especially the elevated shear viscosity, were successfully explained. The sec-
ond variant of his relaxation equation for M is more elaborate as it is derived
from a microscopic, statistical investigation of the rotary diffusion of magnetic
particles. Since the problem was solved with the assistance of the effective field
method, this second variant is commonly denoted as the effective-field theory,
or EFT. The latter is rather more complicated and unwieldy than Debye as
it provides an evolution equation for a quantity called the “effective magnetic
field” from which the magnetization is to be determined in a subsequent step.
114 H.W. Müller and M. Liu
The EFT was found to explain the “negative viscosity” experiment much more
convincingly than the Debye-like approach [8].
Comparing both theories, and emphasizing that the EFT is the more rigor-
ous and accurate one, Shliomis concluded that it is valid for all experimentally
relevant situations [8]. The Debye theory, on the other hand, he considers to be
adequate only in the limit of small deviations from the magnetization equilib-
rium [8], δM M eq , implying the hydrodynamic low frequency limit ωτ 1,
where 1/ω is the characteristic time scale of the experiment.
We do think that these assessments, referring to both (i) the general validity
of the EFT and (ii) the limited validity of the phenomeonological Debye approach
are in need of a clarification:
First, owing to their microscopic input, EFT is in it essence a microscopic
theory, with necessarily rather specific inputs. In the present case, ferrofluids
are considered as suspensions of noninteracting, spherical, equal sized Brownian
rigid dipoles. In the framework of these restricting simplifications, EFT is rigor-
ously valid. But one always has to be aware of the limitations and deficiencies
purchased with the above idealizing assumptions. Regarding to the fact that real
ferrofluids are suspension of interacting, non-spherical, poly-dispersed particles,
whose magnetic relaxation usually involves both Brownian and Néel processes,
EFT cannot be expected to be a sufficient approach under all circumstances.
Second, the potential of a proper macroscopic theory is much larger than for
any microscopic approach, because it is constructed on the sole base of general
principles, without any specific microscopic information. For ferrofluids such a
macroscopic theory is very similar to Shliomis’ Debye theory but it will not suffer
from the above constraint to the low frequency regime. In the following we shall
denote such a modified approach as the rectified Debye theory. Demonstrating
how such a macroscopic theory – we shall call it ferrofluid dynamics (FFD) – can
be derived from the concepts of non-equilibrium thermodynamics is the purpose
of this article.
Generally speaking, any macroscopic theory consists of two separate ingre-
dients. First the structure of the equations, which is solely based on conserva-
tion laws and symmetries, and second the material-specific parameters such as
susceptibilities and transport coefficients. The aim of the present lecture is to
show how the general structure of a hydrodynamics for ferrofluids is derived. No
attempt will be undertaken to provide values for the material-dependent coeffi-
cients. Following the standard approach in macroscopic physics, those quantities
(such as viscosities or susceptibilities) are usually measured. So we leave them
to be determined by a series of suitable experiments. An alternative way is to
calculate the coefficients from an appropriate microscopic model. Such a model
is for example EFT but one must not forget that it is valid in its specified range
of validity. In particular it does not necessarily yield – and indeed it does not,
as we shall see – the most general structure of equations, which are compatible
with symmetries and conservation laws.
Ferrofluid Dynamics 115
Eq. (2) is to be understood as the definition for the conjugate variables such as
temperature T , chemical and relative chemical potentials μ and μc , velocity field
v etc. . In particular, the quantity h is associated to the magnetization. With
M ≡ B − H, or ∂Hi /∂Mj = −δij for given B, together with the thermodynamic
Maxwell relation, ∂Hi /∂Mj = ∂hj /∂Bi , we obtain
Eq.(3) results from the requirement that u has to be minimal with respect to
M at M = Meq (B), or equivalently h ≡ ∂u/∂M = 0. So Beq (M) is the inverse
function of the equilibrium magnetization curve Meq (B). Subtracting M from
both Beq and B, we may also write h = Heq − H, where again Heq (M) is the
inverse function of Meq (H). Note that the function Heq (M) is frequently referred
to in the ferrofluid literature as the ”effective field”.
The conserved variables satisfy continuity equations,
(See [10,12] for the cases where an external electric field is applied.) As a result,
we may use the Maxwell equations in the static approximation
∇ · B = 0, ∇ × H = 0. (8)
The fluxes in Eqs (4-7) still need to be derived – although for some of them
their convective contributions such as ρc v or (v · ∇)M +M × Ω have already
been made explicit. To derive the unknown flux contributions we employ the
so-called standard procedure of hydrodynamics: Take the temporal derivative of
Eq (2), substitute u̇, T ṡ, μρ̇ . . . using the above equations of motion, and most
importantly, require that the resultant equation to hold identically (cf [10,13]
and references therein). This yields the energy flux Q, the momentum flux Πij ,
and the entropy production R as
rectified relaxation term. Although δM and h are in general not linearly related
they do in two special cases: (i) For small deviations from local equilibrium,
where δM and Heq − H are proportional to each other and (ii) if the applied
field is sufficiently weak for the linear constitutive relation Meq = χH to apply.
We now turn to compare our stress tensor Πij with the traditional formula-
tion. We start with the observation that the last term in (9) is equivalent to the
magneto-dissipative element as given in Eq.(1). This term accounts for magneto-
dissipation when the vectors H and M are twisted relative to each other. This
happens for instance in the McTague experiment [6], where the vorticity of the
flow deflects the magnetization vector out of the equilibrium direction.
We emphasize that the condition for the appearance of magneto-dissipative
effects, M = Meq , does not necessarily require the two vectors H and M to
point in different directions. For instance, magneto-dissipation is also expected
to occur if M and H oscillate parallel to each other but with a temporal phase
lag. Note that in this situation Eq.(1) is inoperative, so the Shliomis theory does
not yield a magneto-dissipative stress here. However, as will become clear in
a moment, FFD does. To facilitate further comparison with the conventional
notation of the stress tensor we rewrite the diagonal contribution to [Eq. (9)]
in terms of the zero-field pressure p0 (ρ, T ). To that end we perform a Legendre
transformation to the independent variables T , ρ, v, H, and M (for simplicity
we neglect the concentration dynamics by assuming ρc = const). The associated
thermodynamic free energy f = u − sT − v · g − H · B has the total differential
The last term h · M is missing from previous works. As outlined above it accounts
for magneto-dissipation in a situation where the off-equilibrium increment h is
parallel to the magnetization. For an illustration of what are the physical signif-
icances of the two different magneto-dissipative stresses, (1/2)εijk (H × M)k ≡
(1/2)εijk (M × h)k and δij (h · M) let us consider the force exerted by a ferrofluid
on a rigid container wall. Multiplying the antisymmetric element with the sur-
face normal vector leads to a tangential traction proportional to H × M. On the
Ferrofluid Dynamics 119
Although these expressions appear complicated, one must realize that the uniax-
ial, and not the isotropic, case is the generic one: If we take M as small to arrive
at the isotropic case, we must for consistency also neglect all term ∼ M 2 in the
Maxwell stress, which is considered too crude an approximation to be employed
frequently.
The appearance of the parameters ζ, ζ and ζ× in XiD implies different relax-
ation times for the respective magnetization components parallel to h, parallel
to M and perpendicular to both of these directions. The latter one is analogous
to the Righi-Leduc effect in heat conduction or the Hall effect of an electrical
conductor.
In the Debye theory, the EFT, or the isotropic case considered in the previous
subsection, the only velocity gradient entering Ṁ is the flow vorticity Ω. This is
in contrast to Eq. (18), which suggests that a compressional flow vkk , and even
0
more importantly, an elongational flow vij will do this too. The coefficients λi
are material dependent and need to be measured for each ferrofluid. They are
reactive transport coefficients (as opposed to dissipative ones), because they do
not enter the expression for the entropy production (11). Nevertheless, as these
coefficients appear as a product either with velocity gradients [Eq. (18)] or with
the magneto-dissipative increment h [Eq. (19)], they can only be evaluated by
an appropriate off-equilibrium experiment. It is noteworthy that a term similar
0
to vij Mj is known to appear in the dynamics of nematic liquid crystals [15],
where it gives rise to the “flow alignment” of the nematic director in response
to a shear flow.
For a quantitative experimental test of the whether λ2 for a given ferrofluid
material is of significant size one has to study the magnetization dynamics in flow
geometries which allow to control the relative strength between elongational and
rotational flow contributions. This requirement is easy to fulfill by the flow be-
tween two rotating cylinders (Couette-Taylor setup). Allowing the two cylinders
to rotate independently with distinct angular frequencies Ω1 and Ω2 , enables
to pass over from a rigid rotation (vij = 0), where Ω1 = Ω2 , to a flow with a
finite shear rate, where vij ∝ (Ω1 − Ω2 ) = 0. Recording the magnetization vector
within the sample (or equivalently the magnetic field outside of the sample) as
a function of Ω1 − Ω2 allows to evaluate the transport coefficient λ2 .
λ1 = λ2 = λ3 = λ4 = ζ× = 0, (22)
1 M eq (M eq )2
ζ= + , (23)
τB H 6η1 ϕ
1 ∂M eq M eq (M eq )2
ζ (M eq )2 = − − . (24)
τB ∂H H 6η1 ϕ
where Le = L(ξe ). This equation has to be solved for ξe for given ξ. Then in a
second step, the magnetization is to be deduced from ξe via (25).
We point out that this somewhat unwieldy two-step procedure can be circum-
vented. Without approximation Eq. (26) can be recast in the following, rather
more explicit form
d
2τB M−Ω×M = (27)
dt
M M M·h
− 3χ − eq h − 3 eq
−χ M.
H H M2
λ1 = λ2 = λ3 = λ4 = ζ× = 0, (28)
1 M
ζ= 3χ − eq , (29)
2τB H
3 1 1 M
ζ = − χ . (30)
2 τB M 2 H eq
Clearly, owing to its microscopic input the EFT provides a dependence for the
transport coefficients ζ and ζ on the strength of the magnetization, a feature
which cannot be accomplished by the macroscopic FFD. That is why EFT is a
complementary approach rather than being competitive. On the other hand, the
disappearance of many of the transport coefficients [see Eq. (28)], an immedi-
ate consequence of the idealizing assumptions, reflects the limitations of EFT.
Real ferrofluids generally do not meet these approximations. So one either has
to develop a more elaborate theory, or – following the standard approach in
macroscopic physics – to determine the set of transport coefficients by a series
of appropriate experiments.
Ferrofluid Dynamics 123
5 Conclusion
The general structure of the hydrodynamic equations for ferrofluids is derived
here. The gain in rigour is paid by a loss of specific information on the transport
parameters. We did not provide the numerical values of the transport coefficients
here, nor their dependence on the thermodynamic variables. This task is left
open for a series of experiments. An alternative way – albeit less reliable and
less quantitative – is to derive this information from a microscopic theory such
as EFT. Besides giving numerical estimates for the transport coefficients their
primary advantage is to provide physical insight into their dependencies on the
thermodynamic variables.
In spite of the complete lack of microscopic specifics in the present derivation,
the resultant theory does have some restrictions that we need to keep in mind.
They arise due to the assumption that a unique characteristic time τ is sufficient
to characterize the magnetic relaxation process. As a result, any microscopic
features (such as poly-dispersity) that influence this time are to be handled with
some care. For instance, a ferrofluid consisting of two populations, each with
a distinct relaxation time, will not be well accounted for at higher frequencies,
outside the hydrodynamic regime.
References
*. [email protected], [email protected]
1. R.E. Rosensweig, Ferrohydrodynamics, (Cambridge University Press, Cambridge,
1985).
2. E. Blums, A. Cebers, M.M. Maiorov, Magnetic Fluids, (Walter de Gruyter, Berlin
1997).
3. M. I. Shliomis, Sov. Phys. JETP, 34, 6 1291 (1972); Usp. Fiz. Nauk 112, 427
(1974) [Sov. Phys. Usp. 17, 153 (1974)]; M.I. Shliomis, J. Mag. Mag. Mat. 159 236
(1996).
4. U. Felderhof, B. Kroh, J. Chem. Phys. 110, 7403 (1999).
5. H. W. Müller, M. Liu, Phys. Rev. E 64, 061405 (2001).
6. J. P. McTague, J. Chem. Phys. 51, 133 (1969).
7. M.I. Shliomis and K. I. Morozov, Phys. Fluids 6,2855 (1994).
8. J.-C. Bacri, R. Perzynski, M.I. Shliomis, G.I. Burde, Phys. Rev. Lett. 75, 2128
(1995).
9. A. Zeuner, R. Richter, I. rehberg, Phys. Rev. E 58, 6287 (1998).
10. Mario Liu, Phys. Rev. Lett. 70, 3580 (1993); 74, 4535 (1995); 80, 2937, (1998).
11. S. Lissek, H.W. Müller, M. Liu, Hydrodynamic Maxwell Theory for Ferrofluids,
submitted.
12. Y.M. Jiang and M. Liu, Phys. Rev. Lett. 77, 1043 (1996); Phys. Rev. E, 6685
(1998).
13. K. Henjes and M. Liu, Ann. Phys. 223, 243 (1992);
M. Liu and K. Stierstadt, cond-mat/0010261.
14. L.D. Landau, E.M. Lifshitz, Fluid Mechanics (Pergamon, Oxford, 1987).
15. P. G. de Gennes, J. Prost, The Theory of liquid crystals Clarendon, Oxford (1983).
16. M.A. Martsenyuk, Y.L. Raikher, and M.I. Shliomis, Sov. Phys. JETP 38, 413
(1974); Y. L. Raikher, M. I. Shliomis, in Relaxation Phenomena in Condensed
Matter, ed. by W. Coffey, Advances in Chemical Physics Series 87, Wiley (1994).
Heat and Mass Transfer Phenomena
Elmars Blums
Abstract. This section deals with main problems of the heat and mass transfer in
magnetic colloids. The analysis is mainly based on the general model given in the
Chapter written by R. E. Rosensweig. Hydrodynamic and thermal problems are sim-
plified considering incompressible liquids and neglecting the effects of polarization and
electric conductivity as well as ignoring some other secondary effects that usually can
be neglected in ferrofluid experiments. Contrarily, the analysis of mass transfer ac-
counts for new sedimentation phenomena and cross effects of interrelated heat and
mass transfer. Since the description given by Rosensweig is of general theoretical na-
ture, while the present work mainly focusses on experimental problems, the various
equations needed will not be cited from Rosensweig’s article but from the original lit-
erature since the transfer from the general model to the experimentally needed relations
is often cumbersome and would exceed the frame of this presentation.
The following problems shortly are discussed: the equation of energy conservation
accounting for new adiabatic effects and dissipation due to internal rotation, the ther-
momagnetic and magnetosolutal instabilities, the specifics of magnetic convection in
fluids of non-uniform temperature and concentration, the magnetophoresis and ther-
modiffusion of colloidal particles under the effect of a magnetic field and some phe-
nomena of combined thermal and Soret-driven convection.
1 Introduction
In main applications the magnetic fluids mostly are used to position the colloid at
a certain part of devices by means of magnetic forces. Therefore, for a long time
the most important research problems are the increase of fluid magnetization,
the specific properties of free ferrofluid surfaces in the presence of an external
field and the magnetoviscous and magnetorheological effects. Presently the re-
search topics undergo serious changes. Particularly, there start to be popular
non-potential bulk forces in fluids of a nonuniform magnetization. Possibility to
induce an intensive convection by means of a magnetic field opens the promising
design of new challenging applications (magnetically controlled thermosyphons
for technological purposes, augment of heat transfer for cooling of high power
electric transformers etc.). Besides, the temperature or the concentration gra-
dients cause new thermo-and magnetophoretic transport processes, which can
influence the long-time stability of ferrocolloids in many technical devices. In
the present paper main problems of heat and mass transfer in the presence of a
magnetic field shortly are discussed.
2 Energy Conversion
Many authors have analyzed the energy conservation equation in magnetizable
media. From the second law of thermodynamics for the quasi-equilibrium model
of ferrofluids it becomes [1]
dT 1 ∂ V̄ dp ∂M dH d ∂
ρcp =T −T μ0 + Q − divjq = +v·∇ (1)
dt V̄ ∂T dt ∂T dt dt ∂t
The left part of equation (1) characterizes the accumulated thermal energy
of the fluid (ρ is the fluid density and cp is the specific heat capacity). The first
two terms of the right side reflect the adiabatic effects of fluid compression and
magnetization (V̄ and M are the specific volume and the magnetization), the
Q represents the internal heat source density, and the last term describes the
conductive heat transfer. According to Fik’s law the heat flux jq is proportional
to the temperature gradient:
jq = −λ∇T (2)
(λ is the fluid thermal conductivity). The internal gradients of fluid temperature
are created by the conductive heat input or output through the volume borders
as well as by the internal heat source Q.
Due to a low compressibility of liquids the adiabatic rise of their temperature
under the effect of an increasing pressure always is very small [1]. The thermal
effect of fluid magnetization at room temperatures also is not strong. Only close
to the Curie temperature, the adiabatic magnetization can cause a considerable
change in the fluid temperature. This phenomenon at early stage of the mag-
netic fluid research was proposed to use for a thermomagnetic energy conversion
[2]. The term Q in (1) represents the absorption of external radiation and the
heat generation due to a viscous dissipation of the fluid kinetic energy. Magnetic
colloids are not transparent to the radiation of optical and infrared frequencies;
therefore, the internal effects of light absorption should be taken into account
only in the event of thin fluid layers. If macroscopic volumes of liquids are con-
sidered, only the viscous dissipation of flow energy in Q usually should be taken
into account. Apart from the conventional viscous term, the dissipation func-
tion for magnetic fluids contains an additional term, which reflects the effect of
internal rotations. For Langevin-type magnetization M = ϕMs L(ξ) of colloids
(ξ = μ0 mH/kT , ϕ is the volume fraction of magnetic phase of a bulk saturation
magnetization Ms ), if they contain magnetically hard particles of a magnetic
moment m, the heat source Q becomes [1]
2 2
η ∂vi ∂vk ξL(ξ) H
Q= + + 6ϕη Ω −
2
·Ω (3)
2 ∂xk ∂xi 2 + ξL(ξ) H
(η is the fluid dynamic viscosity). The thermal effect of internal rotations can be
observed in the presence of a constant magnetic field only if it is oriented non-
parallel to the flow vorticity Ω = (1/2)rotv. At low flow velocities the heating
intensity usually causes a relatively small effect upon the fluid temperature.
126 E. Blums
Only for high shear rates typical for some ferrofluid applications (for example,
in high-speed rotary seals and in loudspeakers) the viscous heating of the fluid
can cause a significant rise of its temperature.
Fig. 1. Specific heat induced in magnetic fluids by a rotating field [3]. Left side: mag-
netite particles in kerosene, frequency f = 1.75 kHz, particle volume concentration
ϕ = 0.096 (1) and ϕ = 0.081 (2). Right side: cobalt ferrite in kerosene, ϕ = 0.086,
frequency f = 1.75 kHz (1), f = 0.6 kHz (2), f = 0.2 kHz (3)
Heating of the colloid due to internal rotations in the most pure form man-
ifests in a rotating magnetic field. Despite the presence of anti-symmetric tan-
gential stress, a liquid may remain quiescent, and only the second term in (3)
is responsible for a rise of the fluid temperature. Figure 1 represents the heat-
ing intensity, which is measured employing magnetite and cobalt ferrite based
ferrofluids [3]. From the presented results we see that in colloids of magnetically
“hard” particles, the field-induced internal rotation causes a significant thermal
effect. For example, in a cobalt ferrite based ferrofluid the rotating field 40 kA/m
of frequency ω = 1.75 kHz generates a heating density Q up to 600 W/kg [3]. In
principle, the microconvection under the effect of field rotation induces also some
changes in the fluid thermal conductivity. Microvortices alter the magnitude of
heat flux along the temperature gradient and cause a transversal heat flux in the
plane of rotation axis. Still, the numerical estimates show [4] that this magnetic
analogue of Rigi-Leduc effect in colloids usually is negligible.
A detailed thermodynamic analysis of the energy conservation equation re-
quires considering all thermodynamic parameters as functions of a magnetic
field. However, from a practical viewpoint the corresponding corrections, except
the situation when temperature is close to the Curie one, are negligibly small. A
more important problem is the dependence of physical properties and transfer
coefficients on the particle concentration and on the structure of colloids. The
Heat and Mass Transfer Phenomena 127
additivity law applies for the fluid density and the specific heat capacity; there
is no observable effect of a magnetic field upon these. The transfer coefficients
are affected by a magnetic field due to several reasons. It is well known that
the magnetic braking of internal rotations causes a fluid magnetoviscosity. In-
ternal rotations exert the Hall -type influence upon the heat and mass transfer
coefficients as well. Still, the microconvective analog of the Rigi-Leduc effect of
thermal conductivity is small [1]. More significant seem to be the magnetic field
induced effects which are related to the aggregate forming and orientation.
Here λ0 and λp are the thermal conductivity’s of the carrier liquid and of parti-
cles.
The dependence of conductive heat transfer on a magnetic field in experi-
ments employing stable ferrocolloids usually is not observed, but it is informative
to note that most of the known experimental data suggest a more pronounced
rise in thermal conductivity predicted by the relation (4). Obviously, this is due
to the peculiarities of surfactant layers on particles. The presence in a colloid
of various additional ingredients plays a more sufficient rule when the electrical
conductivity of the colloid is considered.
3 Thermomagnetic Convection
To analyze the convective processes, the energy conservation equation should
be considered together with the equation of a fluid motion. Let us consider the
problem in a Bussines approximation, i.e. assuming the physical properties and
the transfer coefficients (with exception of the density in the gravitation force
ρg and of the magnetization) constant. The equation of motion contains a new
term of magnetostatic force:
dv
ρ = −∇p + ηΔv + ρg + μ0 (M∇)H (5)
dt
If the magnetic susceptibility χ is dependent solely upon the field and ρ =
const, then under an approximation of equilibrium magnetization M = χH
both the gravitation and the magnetic forces are potential and there cannot
arise internal convective flows in the fluid. The non-potentiality of bulk forces
appears only if a fluid possesses the spatial non-uniformity of ρ and M due their
dependence on temperature or on particle concentration ϕ. From (5) it follows
that the free convection develops spontaneously if
∇ρ × g + μ0 ∇χ × ∇H 2 = 0 (6)
We can see that there exists a certain analogy between the thermogravita-
tional and the thermomagnetic convection. The difference is only the spatial
non-uniformity of magnetic acceleration ∇H 2 while the gravitation acceleration
g is constant. There are many works published in which the magnetic convection
under various field distributions are calculated and measured (see the review in
Ref. [1]).
If (6) equals to zero, a problem of thermoconvective instability appears. The
stability criterion is the Rayleigh number Ra. The equations (1) and (5), if they
are linearized with respect to small perturbations of velocity, temperature and
pressure, have the same form as those describing the conventional convection
problems. The only difference lies in appearance of a different form of Ra (we
assume the coordinate x being directed toward the vector g) [7]:
4
dH dT T dH cp l
Ra = βT ρg + μ0 αT M − βT gρ + μ0 αT M (7)
dz dz ρcp dz νλ
Heat and Mass Transfer Phenomena 129
Here βT and αT are the expansion and the pyromagnetic coefficients of the fluid
(according to the definition they are positive), l is the characteristic length.
From (7) it follows that, without applying external temperature gradients,
the fluid with respect to the adiabatic compression and magnetization is always
stable (note, if dT /dz = 0, the Rayleigh number Ra is negative ). The tem-
perature gradients necessary to attain the critical value of Ra are so large that
the adiabatic terms (the second part in the square brackets in (7)) usually can
be neglected. In such approximation the Rayleigh number contains two addi-
tive terms, they reflect the thermogravitational (RaT ) and the thermomagnetic
(RmT ) buoyancy forces:
dH dT cp l4
Ra = RaT + RmT = βT ρg + μ0 αT M (8)
dz dz νλ
Ra∗ Rm∗
= (10)
R0 Rm0
Here R0 = 1708 (flat layer of rigid boundaries), whereas Rm0 is the critical
Rm value when the gravitation is absent. The later depends on the magnetic
properties of channel walls and is slightly influenced also by the nonlinearity of
the fluid magnetization curve.
The general conclusions of thermoconvective stability theories (the pioneer-
ing one, obviously, is that published in Ref. [9]) are confirmed experimentally.
Figure 3 represents the results [10] of measurements of the heat transfer through
a horizontal layer of ferrofluid in the presence of a transversal magnetic field
Bz = const. If B = 0, the convection appears (the Nusselt number N u starts
to exceed the value 1) when the layer is heated from below and the Rayleigh
number reaches its critical value R0 . With a magnetic field applied the critical
temperature gradient drops. Besides this, in accordance with (9) the convective
130 E. Blums
Fig. 3. Effect of a transversal magnetic field upon the heat transfer in infinite flat layer
of magnetic fluid: 1 – heated from below, H = 0.2 – heated from above, H = 40.8 kA/m,
3 – heated from below, H = 40.8 kA/m [10]
instability sets on also when the layer is heated from above. Figure 4 represents
the experimental values of Rac measured in the presence of a magnetic field.
These results are in a good agreement with the theoretical dependence (10) if
accounting for the critical values of both the R0 and Rm0 . The Rm0 is higher
than R0 , (compare the points and the corresponding theoretical dependence in
Fig. 4 with the curve Rac + Rmc = 1708), it depends on magnetic properties
of the channel walls [8,9]. The reason of that is a difference in mechanisms of
the gravitational and the magnetic convective instabilities: the thermomagnetic
Heat and Mass Transfer Phenomena 131
instability is affected not only by perturbations of the fluid temperature but also
by those of the magnetic field inside the fluid.
If the external magnetic field has a gradient non-parallel to the ∇T , inside the
fluid sets on a non-threshold convection. Its intensity significantly exceeds the
intensity of the thermogravitational convection and a significant enhancement
of the heat transfer can be achieved. It means that magnetic control of the heat
transfer in ferrofluids is an interesting problem of applications.
4 Mass Transfer
Considering stable colloids without chemical reactions, the mass conservation
equation for two component systems (particles of a mass concentration ρi and a
carried liquid) is the following:
dρi
= −divji (11)
dt
Ultra-fine particles in magnetic fluids obey an intensive Brownian motion.
Therefore, the mass transfer in colloids can be considered similar to that in
molecular liquids involving the concept of a gradient diffusion. The mass diffusion
coefficient of nanoparticles is determined by the relation D = kT /fv , where fv is
the coefficient of the hydrodynamic drag force (for spherical particles of radius
r the fv = 6πρνr). The mass flux [1]
mg
ji = −D∇ρi + −(V̄i − V¯0 )ρg + μ0 (M̄i − M̄0 )∇H ρi (1−ni )−ρi (1−ni )DST ∇T
fν
(12)
contains a new barodifusion term. In addition to the conventional gravitational
sedimentation of particles (mg is the mass of the particle,V̄i and V¯0 are the
specific volume of the solid phase and that of the carrier liquid) it is necessary
to take into account also for their magnetic sedimentation ( M̄i = Mi /ρi and
M̄0 = M0 /ρ0 are correspondingly the specific magnetization of the particles and
the carrier). The mass diffusion coefficient of colloidal particles is several orders
of magnitude less than that of molecules in liquids. Therefore, the gravitation
and magnetic sedimentation processes in ferrofluids play a more significant rule
than those in molecular liquids. The thermodiffusion term in (12) is retained for
purpose of generality, ST there is the Soret coefficient of particles, ni = ρi /ρ is
the mass fraction of solid phase.
The conditions of the fluid convective instability in isothermal liquids also
are determined by the relation (6). The corresponding solute Rayleigh number
now is the following:
dH dρi l4
Ra = Rac + Rmc = − βc ρg + μ0 αc M (13)
dz dz ηD
(According to the definition, the solute expansion (βc ) and magnetic (αc ) co-
efficients are positive). It is important to note that there is a principal differ-
ence between the thermal and the solutal convective stability. For stable colloids
132 E. Blums
That means that internal field gradients always cause a lowering of the convective
stability of fluid layers.
From Maxwell equations it follows that the external magnetic field gradient
cannot be constant, the magnetic force always is spatially non-homogeneous. As
a result, the magnetic sedimentation forms a non-uniform distribution of parti-
cles not only near walls but also outside the boundary layers. As an example,
in Fig. 5 are shown profiles of concentration in the vicinity of a cylinder when
the particles are transferred by magnetophoresis under the effect of a radial
gradient of the azimuthal magnetic field generated by axial electric current I
Heat and Mass Transfer Phenomena 133
direction of particle transfer, starting some critical values of Debay length the
Soret coefficient can be negative. The parameters of ionic ferrofluid used in the
experiment [16] correspond to such requirements.
From (12) it follows that in colloids of a non-uniform magnetization specific
mass transfer effects can be observed even if the external magnetic field is con-
stant. Let us consider a homogeneous magnetic field B = const directed along
the temperature gradient. From equation divH = −divM we obtain (we assume
M0 = 0)
∂Mi ∂Mi ∂Mi
∇H = −∇Mi = − ∇ρi − ∇T − ∇H (16)
∂ρi ∂T ∂H
The mass flux (12) with account for the equation (16) can be rewritten in
the form of (for a simplicity the term of gravitation sedimentation is omitted)
μ0 αc M 2 mg μ0 αT M 2 mg
ji = − D + (1 + ni ) ∇ρi − ST − ρi D(1−ni )∇T
fν (1 + χ) fν Dρi (1 + χ)
(17)
Respectively, the magnetic stratification of ferroparticles under the influence
of internal field gradients can be considered as an increase of the mass diffusion
coefficient and as a reduction (in surfacted colloids) of the Soret coefficient. More
detailed analysis leads to a conclusion, that the change in coefficient D should
be observed also in the case when field is oriented normally to the temperature
gradient [25,26]. The only difference is that now (at B ⊥ ∇ρi ) the field causes a
reduction of the mass diffusion coefficient. Both effects, the increase in D under
the effect of a longitudinal field B∇ρi and its reduction B ⊥ ∇ρi , recently are
confirmed experimentally [26,27], see Fig. 8.
The Soret effect of magnetic particles in the presence of a uniform magnetic
field is considered in Ref. [28]. From the analysis of hydrodynamic Stokes prob-
lem for a spherical magnetic particle it follows, that if B∇T , the nonuniformity
of fluid magnetization causes a motion of particles in the direction of tempera-
ture gradient in accordance with the dependence (17). Contrary, in the presence
of a transversal field B ⊥ ∇T , the particles are transferred toward decreasing
temperatures. This so-called “magnetic Soret effect” and its anisotropy have
been investigated experimentally. The separation measurements in thermodif-
fusion column [29,30] confirm the general predictions of hydrodynamic theory:
in colloids with surfacted particles of positive Soret coefficients the transversal
magnetic field B ⊥ ∇T causes an increase in ST , whereas at B∇T a strong
reduction of ST takes place (see Fig. 9).
Similar results are obtained also in the experiments on optical grating in
thin ferrofluid layers [27]. But, it is interesting to note that the measured mag-
netic field effects always are significantly stronger than those predicted by the
hydrodynamic theory. There might be several reasons of that. Firstly, the con-
centration profiles in nonisotermic fluids dvelop under the combined action of the
thermodiffusion coefficient DT and the mass diffusion coefficient D. Therefore,
the measurd values of Soret coefficient ST = DT /D reflect not only the direct
influence of B on DT but also the dependence D = D(H). The second reason
can be the polidispersity of colloids. In Ref. [31] are reported some calculations
Heat and Mass Transfer Phenomena 137
6 Conclusions
Magnetic convection in non-isothermic and stratified ferrocolloids can be used
to develop new effective cooling systems (for loudspeakers, electric transform-
ers, magnetically controlled thermosyphons) etc. Magnetic and thermodifusional
stratification of colloids should be taken into account considering the long-time
working ability of magnetofluid devices.
The research on separation in thermodiffusion columns reported in Section
5 is supported by BMBF and by Alexander von Humboldt Foundation.
References
1. E. Blums, A. Cebers, M. M. Maiorov: Magnetic Fluids (Walter de Gruyter, Berlin,
New York, 1997)
2. R. E. Rosensweig: Ferrohydrodynamics (Cambridge University Press, Cambridge,
1985)
3. G. Kronkalns, M. M. Maiorov: Magnetohydrodynamics 3, 28 (1984) (in Russian)
4. V. G. Bashtovoy, A. N. Vislovich, B. E. Kashevsky: Zh. Prikl. Mat. Techn. Fiz. 3,
88 (1978) (in Russian)
5. M. A. Martsenyuk: “Thermal conductivity of a suspension of ellipsoidal particles
in a magnetic field”. In: Proceedings of the 8 ght Riga MHD Conference (Zinatne,
Riga, 1975) 1, pp. 108–109 (in Russian)
6. B. M. Tareev: Coll. Journ. 6, 545 (1940) (in Russian)
7. M. I. Shliomis: Fluid Dynamics 6, 957 (1973) (in Russian)
8. K. Gotoh, M. Yamada: J. Phys. Soc. Jap. 51, 3042 (1982)
9. B. A. Finlayson: J. Fluid Mech. 40, 753 (1970)
10. L. Schwab: Ķonvektion in Ferrofluiden. Ph.D. Thesis, Ludwig-Maximillians Uni-
versität München, München (1989)
Heat and Mass Transfer Phenomena 139
Andrey Zubarev
Department of Mathematical Physics, Ural State University, Lenin Av., 51, 620083
Ekaterinburg, Russia
1 Introduction
One of the fundamental problems of physics of ferrofluids (magnetic liquids) is
the determination of their macroscopical characteristics as a function of their
inner composition, i.e. of shape, size distribution, physical properties and con-
centration of the magnetic particles and of the properties of the carrier liquid.
The classical theories for ferrofluids (see references and discussion in [1], [2])
deal with ideal models of very dilute systems in which any interactions between
particles are negligible. However in real magnetic liquids these interactions often
lead to long-range correlations between positions and orientations of the particles
and to the appearance of bulk drop-like, linear chain-like and other heterogeneous
aggregates, consisting of various number of particles. These inner structures play
a fundamental role for the physical properties of magnetic fluids and, therefore,
the study of the structures and internal states of ferrofluids gives a key for the
understanding of the properties and the macroscopical behavior of these systems.
Since the particles in typical ferrofluids are small (mean diameter is about
10 nm), they are subject to intensive translational and rotational Brownian mo-
tion. Any inner structure in a ferrofluid is a result of a competition between
thermal motion of the particles, their magnetic interaction and hydrodynamical
forces, which are especially significant when the magnetic liquid is exposed to
macroscopical flow. That is why the internal states of magnetic liquids can be
understood and described only on the basis of an appropriate statistical the-
ory. From the point of view of statistical physics the following main specifics of
ferrofluids have to be considered.
tunately, the final results of the theory are cumbersome and it is not an easy
thing to use them for calculations of the physical properties of magnetic fluids.
Similar methods of chemical kinetics for the investigation of chain-like structures
in ferrofluids in the absence of external field were used in [22].
Models of chains as heterogeneous fluctuations of density in magnetic fluids
on the basis of statistical thermodynamics were suggested in [23-25]. Methods,
developed for the theory of very long flexible macromolecules, with a tremendous
number of monomers in the order of 104 -109 ), were used in the last mentioned
model. However it is difficult to expect such long linear clusters in real ferroflu-
ids. The models in [23,24] use a very simple approximation of the chains as
straight rod-like aggregates, i.e. thermal fluctuations of the shape of the chains
are ignored in these models. In spite of the fact that this strong simplification
can be used only when most of the chains are short enough, this allows us to
understand the physical origin of group of phenomena in magnetic fluids and to
estimate them, at least, by the order of magnitude. The main points and results
of the models [23,24] are discussed in parts 2-4 of this chapter. It should be noted
that effects driven by the flexibility of the chains are discussed in [26].
The formation and evolution kinetics of the “drops” as a result of the “gas-
liquid” phase transition in a system of interacting, but single particles, was
studied theoretically in [27,28]. As in [3,4] and [6,7], any linear chains are ignored
in these models and the ferroparticles are treated as separate. These theories
describe correctly the principal points of the phase transitions, but numerical
experiments [14-17] and analytical models [25,29] show that linear chains appear
in systems of dipole-dipole interacting particles long before the bulk condensation
phenomena takes place. Therefore these chains must be necessarily taken into
account in theories of phase transitions and other structural transformations in
ferrofluids.
It is very difficult to study various kinds of inner structures in ferrofluids and
their influence on the macroscopic properties of these systems in the framework
of one model. That is why it is reasonable to start with separate analysis of
each kind of these structures. In this paper some results of the theoretical study
of effects of chain-like aggregates on the macroscopic dynamical properties of
ferrofluids are presented. To avoid too cumbersome mathematics and to reach
physically reasonable estimates, we use the following approximations.
1. We start with an analysis of monodisperse systems. A bidisperse model of
a polydisperse ferrofluid is considered in the last part of the paper.
2. We assume that the considered particles are big enough and that the mag-
netic moment of each of them is frozen into its body. Estimates of the minimal
size of the particle necessary to provide this approximation are given, for ex-
ample, in [1] or [2]. One can show that the magnetic interaction of very small
superparamagnetic particles with magnetic moments rotating freely inside their
bodies, is too weak to provide any heterogeneous structures. Therefore, the as-
sumption of magnetically hard particles is in agreement with our consideration
of chain-like structures consisting of reasonably sized particles.
Statistical Physics of Non-dilute Ferrofluids 147
3. We neglect any interactions between the chains. One can show that, be-
cause of the fact that the chains can not overlap, particles from the different
chains can not be situated in places, where their interaction is strong, since
these positions, for every particle, are occupied by other particles from the same
chain. Therefore interaction between particles of different chains is weaker than
interaction between those from one chain. That is why the interparticle interac-
tion inside one chain plays the main role in the formation of the chain structure.
Thus, the model of non interacting chains seems to be a reasonable first approxi-
mation, at least when concentration of the interacting particles is not high. This
assumption is not suitable for very dense systems, where the chains lose their
individuality and are destroyed due to interaction between them [16].
4. We restrict ourselves to situations where the interaction of the magnetic
particles with the external magnetic field is much smaller than the magnetic
interaction of two neighboring particles in the chain. It means that the chains
react with the field as whole aggregates. This restriction is valid in many real
experimental situations.
mH μ0 m2
e = 2.72..., κ = μ0 , ε=
kT 16πa3 kT
Here kT is the absolute temperature in energetic units, fn is the dimensionless
free energy of the chain due to its inner structure, nc is the maximal number
of particles in the chain, estimated below. For equilibrium states one may put
nc = ∞ (see eq. (15)). The first term in the parentheses in equation (1) represents
the entropy of the ideal gas of n-particle chains determined by their translation
motion, H is the magnetic field inside the sample, μ0 is the vacuum permeability,
148 A. Zubarev
fn = − ln Zn (4)
⎛ ⎞
!n !
n
Zn = exp ⎝ (ν i κi ) − Ud (ν i , ν j , r ij )⎠ dΩ
i i>j
Here ν i is the unit vector aligned along the magnetic moment of the i − th
particle, r ij is the radius-vector, connecting the centers of the i − th and j − th
particle, Ω includes all ν i and r ij and Ud is the dimensionless potential of the
dipole-dipole interaction.
As usual, the statistical integral Zn can not be calculated in a general form.
To reach reasonable estimates we may take into account only the interaction be-
tween nearest particles in the chain. This approximation was used in the models
[22-26]. The interaction between nearest particles and throughout the nearest
neighbors was taken into account in [21] and led to cumbersome final results. In
the “nearest neighbor” approximation the statistical integral (4) is
& n '
! !
n
Zn = exp (ν i κi ) − Ud (ν i , ν i+1 , r i,i+1 ) dΩ (5)
i i
exp (Ud (ν i , ν i+1 , r i,i+1 ) by δ(ν i − ν i+1 )δ(ν i − r i,i+1 )δ(ri,i+1 − 2a) as a first ap-
proximation. In this approximation we ignore thermal fluctuations of the chains
and treat them as straight rod-like aggregates. This is justified when the con-
tour length of the chain is less or about the persistent length l0 of the chain.
Using standard considerations of the theory of linear polymers, one can show
that l0 ∼ 2aε. Therefore the “rod” model of the chains can be considered as
justified when the mean contour length 2a < n > of the chain is not more than
2aε, i.e. < n >≤ ε where < n > is the mean number of particles in the chain.
In the framework of this model the dimensionless free energy fn may be
written down as:
sinh κn
fn = − ε (n − 1) + ln (6)
κn
The first term in the square brackets represents the absolute value of the
energy of magnetic interaction of particles inside the rod-like chain, estimated
in the approximation of nearest neighbors. The second term corresponds to the
Langevin energy of the rod in the dimensionless field κ. It should be noted that
relation (6) is the upper estimate for the absolute magnitude of fn .
Substituting equation (6) into (3) and, then, into (2), taking into account
that one can put nc = ∞ for equilibrium systems (see equation (15)), we get
after some calculations (details in [23])
xn0 sinh κn
gn = exp(−ε), (7)
v κn
where
(
2
2y cosh κ − sinh κ − (2y cosh κ − sinh κ) − 4y 2
x0 = , y = κϕ exp(ε)
2y
Figure 1 illustrates calculations of the volume concentration vngn of the par-
ticles collected in the n-particle chains. This volume concentration, as function
of n, has a maximum increasing with κ and ε. Similar maxima were obtained
also in [29].
The mean number of particles in the chain is
)
ngn ϕ
< n >= )n = ) (8)
n gn v n gn
Fig. 2. The critical volume concentration ϕs of the particles for the model of straight
aggregates vs the dimensionless magnetic field. Solid and dashed lines correspond to
ε = 5 and ε = 3, respectively
Here η0 is the viscosity of the carrier liquid. This inequality is valid for most
part of experimental situations.
It is well known since Einstein’s classical work, that the influence of dispersed
particles on the rheological properties of suspensions is connected with their per-
turbations of the macroscopic flow of the suspension. It is practically impossible
to calculate the perturbations induced by the chains because of their complex
shape. In order to get reasonable estimates, we model the n−particle chain, like
in [23] and [24], as a prolate spheroid with semiminor axis a and semimajor
one an . It is of principal importance that the volume of this spheroid is equal
to the sum of the volumes of the particles in the chain. Therefore, the volume
concentration of the model spheroids is equal to the total volume concentration
of the particles.
Using the well known results of the theory of statistical hydromechanics of
dilute suspensions of rigid spheroids (see, for example, [23,24,31]), we can write
down expressions for the Cartesian components of the average viscous stress
tensor σ̂ as follows
!nc
σik = 2η0 γik + Φn (γlm , ωlm , < ej es >n , < el em ej es >n )gn (10)
n=1
< ... >n = ...φn (e)de, i, k...s = x, y, z
The symbol ˆ here and below indicates tensor magnitudes, e is a unit vec-
tor aligned along axis of the chain, φn (e) is a distribution function over the
orientations of the chain normalized to unity, γij and ωij are the symmetrical
152 A. Zubarev
Fig. 5. (a) Real and (b) imaginary part of complex dimensionless effective viscosity
vs frequency Ω of an oscillating flow (Ω = Ω/D1 ). The external field is aligned along
the gradient of the flow velocity, ε = 3, ϕ = 0.05. Solid and dashed lines correspond to
κ = 1 and κ = 0.5, respectively
Fig. 6. The same curves as in Fig.5 for ferrofluid with single particles
Figure 5 shows calculations of the real θ1 and imaginary θ part of the
dimensionless complex viscosity as a function of the frequency of the flow for a
magnetic field parallel to the gradient of the flow velocity. For comparison, Fig.6
shows the results of calculations for the same magnitudes under the assumption
that all particles are free. Comparing the results in Figs. 5 and 6, one can note a
strong increase of both components of complex viscosity due to the chains and the
fact that the frequency, corresponding to the maximum of θ , is much smaller
in the model with chains than in the model for free particles. It means that
the time of hydrodynamical relaxation in ferrofluids with chains is significantly
longer than the same time predicted by the classical models of separate particles.
Using this approach, one can estimate the normal components of the stress
tensor under shear flow [24]. It is known that the normal stresses can lead to the
so-called Weissenberg effect [33], observed usually in isotropic polymer liquids.
The results of an experimental study of this effect in ferrofluids are discussed
in [34]. It should be noted that this phenomenon can not take place in systems
with separate spherical particles only. The fact that it is observed in ferrofluids
is a result of the presence of anisotropic (chain-like) aggregates and qualitatively
supports the suggested model.
Statistical Physics of Non-dilute Ferrofluids 155
When γ̇ is vanishing, nc tends to infinity. This result was taken into account
in the previous sections. When γ̇ is significant, but the inequality γ̇/D1 << 1 is
fulfilled, we should use the estimate (14) in equations (1),(2) and (10). Thus, in
first approximation, when the inequality γ̇/Dn << 1 is true for all chains, the
influence of shear rate on the chain structure can be taken into account by a use
of the estimate (15) in relations (1,2) and (10), remaining all other considerations
of parts 2 and 3 without changes.
Fig. 7. Experiments [18] (dots) and calculations (lines) of the magnetoviscous param-
eter S. a) Squares corresponds to shear rate γ̇ = 0.1s−1 , line 1 to γ̇ = 0; black circles
and line 2 to γ̇ = 0.5s−1 , white circles and line 3 to γ̇ = 0.9s−1 . b) light squares and
line 1 correspond to γ̇ = 1.05s−1 , black squares and line 2 to γ̇ = 5.23s−1
Fig. 8. Diagram of the size distribution for ferrofluid, used in experiments [18], [19]
One needs to stress that the 10 nm particles are superparamagnetic and the
energy of magnetic interaction between them is much smaller than the thermal
energy for room temperature. Therefore, these particles neither can influence sig-
nificantly the magnetorheological properties of the ferrofluid, nor they can form
any heterogeneous aggregates. It means that only the biggest particles, in spite
of their small concentration, give rise to the observed strong magnetorheological
effects. The analysis, given in Section 3, shows, that the physical origin of this
phenomenon can be chain-like aggregates, formed by the biggest particles.
The problem of a theoretical description of chain-like structure in a polydis-
perse ferrofluid, even under the used strong simplification, is very complicated.
To avoid cumbersome mathematics, we use a model of a bidisperse system con-
sisting of the “small” particles with sizes being near the mean size of particles
in the ferrofluid and “big” particles with sizes and volume concentration to be
determined. For maximal simplification of the calculations we assume that there
are only the “big” particles, which can form the chains. All the “small” particles
are supposed to be free.
158 A. Zubarev
Third. What is the role of the “small” particles in the formation of the inner
structure of ferrofluids? The point is that simple estimates show that the energy
of magnetic interaction between typical magnetite particles with a magnetic
diameter about 10 nm and those with 16-17 nm for room temperature is slightly
more than kT . Therefore, the “small” particles can take part in the formation
of the chains. The previous analysis [37] shows that their influence on the chain
structure is not negligible, however this problem is worth a further detailed
examination.
Forth. What is the role of the chain-chain interaction in ferrofluids? This
question is especially intrigue, since many experiments (for example, [9-13])
demonstrate that the “gas-liquid” phase transition can take place in an ensem-
ble of ferroparticles under suitable conditions. Theories [3,4,6,7,27,28] explain
and describe the main points of these phase transitions, however they ignore
any possibility of the formation of chain-like structures. Computer simulations
[14-17] and theories [25,36] that take into account this possibility, conclude that
the condensation phase transition in ferrofluids is impossible - when tempera-
ture decreases the chains become longer in the spatially homogeneous system.
Thus, there is a qualitative contradiction between experimental and theoretical
(including computer experiment) results on phase transitions in ferrofluids. A
physical nature of this contradiction may lie in the following facts. First, the sur-
face layers on the ferroparticles do not screen van-der-Waals interaction entirely.
The combination of dipole-dipole forces and “tails” of the central van-der-Waals
interaction can lead to a bulk phase transitions even in ferrofluids with chains
[25]. The second factor, which can induce this transition, can be polydispersity
of real ferrofluids. Indeed, the “small” particles, with magnetic diameter about
10 nm and high concentration, can create a strong osmotic pressure that pushes
the “big” particles to regions where concentration of the “small” ones is low. As a
result domains with high concentration of big particles, where the osmotic pres-
sure compensate the pressure in an environment with small particles, can appear
as thermodynamical stable phases. Both of these scenarios of the bulk separation
(first, provided by the van-der-Waals tails and the second one, induced by the
“small” particles) are worth a thorough examination.
Finishing, one can note that in spite of the mentioned unsolved problems,
the simple model of non-interacting rod-like chains can be considered as a robust
basis for a qualitative understanding and study of the dynamical properties of
real non dilute ferrofluids. This model can serve also as a base for a develop-
ment of more advanced theories of internal structures and macroscopic physical
properties of magnetic liquids.
This work has been supported by grants of the Russian Basic Research Foun-
dation, projects NN 02-00-17731, 02-01-6072, 01-01-00058, grant CRDF, REC-
005 and grant of BMBF N RUS 00/196
160 A. Zubarev
References
1. M.I.Shliomis, Sov.Phys.Usp. 17, 153 (1974)
2. E.Blums, A.Cebers, M.Majorov, Magnetic Fluids(de Gruyter, Berlin, New York,
1997)
3. A.O.Tsebers, Magnetohydrodynamics, 18, 345 (1982)
4. K.Sano and M.Doi, J.Phys. Soc.Japan., 52, 2810 (1983)
5. D.Wei, G.N.Patey and A. Perera, Phys.Rev.E, 47, 506 (1993); H.Zhang and
M.Widom, J.Magn.Magn.Mater.122, 119(1993); Phys.Rev.E 49, 3591 (1994);
Phys.Rev.B 51,8951 (1995); B.Groh and S.Dietrich, Phys.Rev.E, 50, 3814 (1994);
Phys.Rev.E, 53, 2509 (1996); Phys.Rev.E, 57, 453518, (1998)
6. K.I.Morozov, Bull. Acad. Sci. Ussr, Phys. Ser. 51, 32 (1987)
7. Yu.A.Buyevich and A.O.Ivanov. Phys. A, 190, 276 (1992)
8. A.Yu.Zubarev and A.V.Yushkov, Zh.Eksp.Theor.Fiz 114, 892 (1998) [JETP 87
484 (1998)]
9. C.F.Hayes, J.Colloid. Interface Sci., 52, 239 (1975)
10. E.A.Peterson and A.A.Krueger, J.Colloi. Interfase Sci., 62, 27 (1977).
11. A.F.Pshenichnikov and I.Yu.Shurubor, Bull. Acad. Sci. USSR, Phys. Ser., 51, 40
(1987)
12. J.C.Bacri, R.Perzynski, D.Salin, V.Cabuil and R.Massart,J. Colloid. Interface Sci.,
132 43 (1989)
13. P.K.Khizhenkov, V.L.Dorman and F.G.Bar’akhtar, Magnetohydrodynamics, 25,
30 (1989)
14. J.M.Caillol, J.Chem. Phys., 98, 9835 (1993)
15. M.E.Van Leeuwen and B.Smit, Phys. Rev.Let. 71, 3991 (1993)
16. D.Levesque and J.J.Weis, Phys. Rev.E 49 5131 (1994)
17. A.Satoh, R.W.Chantrel, S.Iand G.N.Coverdall, Coll., Intersace Sci., 181, 422
(1996)
18. A.Zubarev, S.Odenbach, J. Fleicher J.Magn. Magn.Mat. (submitted as material of
IX Int. Conf. on Magnetic Fluids)
19. S.Odenbach and H.Stork, J.Magn.Magn.Mat. 183, 188 (1998)
20. P.G. de Gennes and P.A.Pincus, Phys. Condens. Mat. 11, 189 (1970)
21. P.G.Jordan, Molecular Physics, 25, 961(1973); ibid. 38, 769 (1979)
22. A.Cebers, Magnitnaja Gidrodinamika, 2 36 (1972)
23. A.Yu.Zubarev, L.Yu. Iskakova, J.Exp. Theoret. Physics, 80, 857 (1995)
24. A.Yu.Zubarev and L.Yu. Iskakova, Phys. Rev. E, 61, 5415 (2000)
25. M.A.Osipov, P.I.C.Teixeira and M.M. Telo da Gama, Phys.Rev.E., 54, 2597 (1998).
M. Tavares, M.M. Telo da Gama and M.A.Osipov, Phys. Rev. E., 56, 6252 (1997),
see, also, Errata, Phys. Rev. E, 57 7367 (1998)
26. K.Morozov and M.Shliomis see p.204 of this volume.
27. A.O.Tsebers, Magnetohydrodynamics, 1 3, (1994)
28. A.Yu.Zubarev and A.O.Ivanov, Phys. Rev. E, 55,7192 (1997); A.O.Ivanov and
A.Yu.Zubarev, Phys. Rev.E, 58,7517 (1998)
29. J.M.Tavares, J.J.Weis and M.M.Telo da Gama, Phys. Review E, 59, 4388 (1999)
30. M.Rasa, D.Bica, A.Philipse and M.Vekas, “Dilution Series Approach for Inves-
tigation of Microstructural properties and Particle Interaction in High Quality
magnetic Fluids” (will be published)
31. V.N. Pokrovskij, Statistical mechanics of Dilute Suspension, (Nauka, Moscow,
1978)
Statistical Physics of Non-dilute Ferrofluids 161
32. M.A.Marstenjuk, Y.A.Raikher and M.I.Shliomis, Zh.Eksp. Teor Fiz 65 834 (1974)
[JETP, 38, 413 (1974)]
33. D.D.Josef, G.S.Beavers and R.L.Fosdick, Arch. Rational Mech. Annals, 49, 381
(1973)
34. K.Melzner and S.Odenbach, J. Magn. Magn. Mat, (submitted as material of IX
Int. Conf. on Magnetic Fluids)
35. H.Muller and M. Liu see page 115 of this volume
36. S.Klapp and F.Forstmann, Phys. Rev. E., 60, 3183 (1999).
37. A.O.Ivanov and S.Kantorovich, J.Magn.Magn.Mat. (submitted as material of IX
Int. Conf. on Magnetic Fluids)
Magnetic Fluid
as an Assembly of Flexible Chains
Abstract. Dipolar chains formed in magnetic fluids out of colloidal magnetic grains
have much in common with polymer molecules. An investigation of spatial and ori-
entational intrachain correlations and elucidation of an important role of the chains
flexibility lead us to natural and fruitful extension of basic concepts of polymer physics
to the case of dipolar chains. Conformations of the chains (statistical coil, globule) in
zero and infinitely strong external magnetic field are studied and the possible coil–
globule phase transition is predicted and discussed.
1 Introduction
The phenomenon of formation of internal structure of magnetic fluids (MF) and
its influence on the macroscopic properties of MF is one of the most exciting
problems of physics of ferrocolloids. It is generally recognized that the origin of
internal structure is due to dipolar interactions between magnetic grains com-
posed MF. The basic features of the effects of dipolar interactions in MF were
outlined long ago in the pioneering work of de Gennes and Pincus [1]. The authors
had brilliantly foreseen the two possible consequences of magnetic interparticle
interactions. First, in the zero external field they predicted the phase separa-
tion of a magnetic fluid into the dilute vapor and condensed liquid phases below
some critical temperature in analogy with the case of simple fluids. A heuristic
explanation of the phenomenon is following. The dipole 1/r3 pair potential (r
the particle separation) being averaged over the dipole orientations becomes an
attractive 1/r6 -potential responsible for the van der Waals condensation of sim-
ple fluids. Second, at high external fields H the association phenomenon should
appear: the dipoles align along the field direction, the head-to-tail configuration
is energetically preferable and the magnetic grains tend to form chains along
the direction of H. Thus, the two types of effects of interparticle interactions
in MF differ by their scales: the condensation manifests itself as macroscopic
phenomenon whereas the grain association as microscopic one. De Gennes and
Pincus discussed also the appearance of chain-like structure in zero external
field. However, the relation between the condensation and the chain formation
had remained unclear in this case. As we will see, just this question became the
subject of the latest intensive investigations.
In the 70-s, the task of condensation of dipolar fluids in the absence of external
field seemed to be well understood. All the theories of the system of interacting
dipoles predicted the appearance of the gas-liquid phase transition under some
critical temperature (see, e.g. [2] and references therein). Among the theories
we note the analytical mean spherical model (MSM) of Wertheim [3] for the
system of hard sphere dipoles. The thermodynamics for MSM was developed
in [4] and the critical value of the coupling parameter λ = m2 /d3 kB T (m is
dipole moment, d is a grain diameter, kB is Boltzmann’s constant and T is the
temperature) was determined. According to [4], this value is λc = 4.445. The
condensation observed in Monte Carlo simulations [5] qualitatively corroborated
the theoretical results.
The condensation of magnetic fluids in zero field as well as in external mag-
netic fields is reliably established experimentally. Starting from the probably first
evidence, given by Hayes [6], it was observed in a numerous works for both types
of MF - ionic [7,8,9] and surfacted ones [10,11,12,13,14]. It has been shown that
an external magnetic field always promotes the phase transition. The droplets
of concentrated “liquid” phase appear among the more dilute “gas” phase with
decreasing of temperature or/and with growth of magnetic field. The real MF
are polydisperse ones. So, there is no simple to determine experimentally the
critical values λc (H) of the beginning of the phase separation. The estimation of
the value in strong magnetic filed is λc (∞) = 2.96 [14]. In zero field the droplets
of concentrated phase are spherical what means the absence of their sponta-
neous magnetization. The droplets have typical sizes of a few microns, i.e., they
are macroscopic formations. Just the macroscopic scale of the droplets allows
to identify them as new concentrated phase originating in the system. So, no
wonder that the elongation of a droplet in external field is well described by the
competition of the surface energy at the gas-liquid boundary and the magnetic
energy of the droplet [7].
To describe the observable condensation of MF under influence of external
magnetic field the statistical mechanics models [15,16,17,18,19] were proposed.
Initially the dipolar interactions were taken into account within the framework
of the classical Weiss model of ferromagnets [15,16] and antiferromagnetic model
[17]. The both models quantitatively correctly describe the phase separation of
MF however they possess the defect common for all the mean field theories. It
is a prediction of spontaneous magnetized state of concentrated liquid phase in
the absence of external field. This property was never observed in experiments
with real MF. In contrast with the mean field theories, the more late models
[18,19] took into account the orientational and spatial correlations of MF grains.
In [18] the generalization of the MSM of Wertheim [3] onto the case of arbitrary
external field was done. The mean spherical model has no the problem with
spontaneous magnetized state of liquid phase [4]. The external magnetic field
lowers the threshold of condensation from zero field value λc = 4.445 to infinite
field value λc (∞) = 3.055 [18] what is in agreement with mentioned estimation
[14]. The analogous results were obtained within the framework of perturbation
theory [19] which is more simple than MSM. The both theories describe perfectly
the many properties of real MF such as the magnetization curve, temperature
and concentration dependence of initial susceptibility [20,21,22].
164 K.I. Morozov and M.I. Shliomis
Here it is necessary to note that the usual magnetic fluids used in practice
are required to be stable against temperature and magnetic field action. In other
words, in many cases the condensation of MF proves to be the undesirable phe-
nomenon. Thus, the ordinary magnetic fluids and commercial MF especially are
characterized by very low values of dipolar interactions, λ ≤ 1. Often due to the
special techniques only, the authors could increase value of λ up to 3 ÷ 4 and ob-
serve the phase transition in MF [7,8,9,13,14]. As we will see, under these small
values of λ the association phenomenon – the second consequence of dipolar
interactions according to de Gennes and Pincus – is negligible. Another aspect
of dealing for a long time with mentioned MF is the problem of synthesizing
of MF with strong magnetic interactions. Only recently such ferrocolloids were
successfully synthesized [23]. Now Nakatani and co-authors achieved the values
of the coupling parameter λ = 3 ÷ 10 [24]. Just in such MF the association of
grains should be essential. In contrast to condensation, to detect the particle
association taking place on microscopic level is a serious experimental problem.
We can judge about MF microstructure using only indirect data on negative
viscosity effect and magneto-vortical resonance in MF under an oscillating mag-
netic field [25]. The data gave sure signals of the presence of a wide spectrum of
magnetization relaxation times τ instead of only one Brownian time for single
magnetic grains. Assuming that this spectrum originates from chains formed out
of the grains, authors of [25] have reached a perfect agreement between the the-
oretical and experimental frequency dependencies of MF viscosity and dynamic
susceptibility.
The intense interest to the systems with strong dipolar interactions arose
however a bit earlier, at the beginning of 90-s, as a corollary of new unexpected
data of numerical experiments. First, a surprising result of Monte Carlo simu-
lations was the absence of vapor-liquid transition in the system of interacting
dipoles. Instead, simulations revealed a phenomenon of formation of chains of
head-to-tail aligned dipoles in the dipolar hard sphere fluid [26,27] as well as
in the Stockmayer [28] and dipolar soft sphere fluid [29,30]. The evidence of
the above transition did was given after all [31]. This transition proved to be
rather unusual, namely, it occurs between highly associated vapor phase and a
more normal dense liquid phase and accompanies by extremely small entropy
and enthalpy changes.
Another exciting property of dipolar fluid revealed in simulations is that
dipolar forces alone can create an orientationally ordered liquid state. Starting
from the first demonstration of the phenomenon in molecular-dynamics simula-
tions [32,33], this conclusion was confirmed later by Monte Carlo calculations
[26,34,35,36]. In fact, there is no simple connection of spontaneous order of dipo-
lar fluid and chain association phenomenon of aligned dipoles [35,37]. However
the contraposition of liquid condensation and association feature seemed to be
well established in the numerical experiments [26,27,28,29,30,31]. At the same
time, the data of Monte Carlo simulations made in [38,39,40] with different
conditions at the boundary of simulation cell similarly to old study [5] fully con-
tradict to results [26,27,28,29,30,31,32,33,34,35,36,37] and predict at λ > 3 the
Magnetic Fluid as an Assembly of Flexible Chains 165
The phenomenological theory [47], combined the van der Waals theory for
liquid condensation with the association Jordan’s theory, revealed a competition
between liquid condensation and chain formation driven by anisotropic dipo-
lar interactions. The analogous result was obtained in [46,48,49,50] where some
phenomenological parameters of work [47] were estimated. An interesting con-
clusion of theoretical studies [48] is that a validity of association theory of non-
interacting chains is possibly broader than it was assumed before [42,43] and is
not limited only to the low-density phase of fluid. This result is indicated by
the appropriate evaluation of the steric and dipolar interchain interactions [48].
Finally, authors [45,53] studied the magnetic and rheological properties of MF
assuming the rod-like form of aggregates of magnetic grains.
Summarizing the review of recent theoretical works we see that all of them
deal with only one characteristic of the particle chains - their length. Meanwhile
a genuine understanding of the phenomenon of chain formation remains incom-
plete without studying the spatial and orientational correlations inside dipolar
chain. As shown below, the association theory allows to investigate this aspect
of the problem in detail and often analytically. Developing our approach, we es-
tablish an important role of chain flexibility; this attribute of dipolar chains has
been remained without proper attention in previous works [45,46,48,49,51,52,53].
De Gennes and Pincus [1] were the first who noted an analogy between dipolar
chains and polymer molecules. Actually, our investigation of the pair correla-
tions and the chain flexibility leads us to natural extension of the basic concepts
of polymer physics like the persistent length and Kuhn segment, coil, globule
[54] etc. to the case of dipolar chains. The aim of this paper is study of the spa-
tial and orientational correlations inside dipolar chains, determination of possible
forms of the chains (statistical coil, globule), and investigation of the coil-globule
phase transition. The paper is organized as follows. In Sect. 2 we consider the
basic statements of an association theory of MF, define the concept of chain of
magnetic grains, establish the main property of the proposed model of neighbor-
ing grain interactions and determine the averaged length of dipolar chains. The
statistical properties of a single dipolar chain are studied in Sect. 3 where the
important role of chain flexibility is established and main flexibility character-
istics are determined. The influence of chain flexibility on structure of a chain
and coil-globule transition are considered in Sect. 4. We make some concluding
remarks in Sect. 5.
with numbers i and i+1 is the sum of hard-sphere UHS (ri,i+1 ) and dipole-dipole
Udd interactions:
m2
U (i, i + 1) = UHS − 3 [3(ei ni,i+1 )(ei+1 ni,i+1 ) − ei ei+1 ] , (1)
ri,i+1
where m is the magnetic moment, ei and ni,i+1 are unit vectors along the dipole
of particle i and the interparticle vector r i,i+1 , respectively.
Let gN be the number of N -particle chains per unit volume. The volume den-
sity F/V of the free energy of system includes ideal gas term and the intrachain
contribution [45,46,48]
∞
!
F gN v
= gN ln + gN fN , (2)
V kB T e
N =1
achieve
ZN = Z2N −1 . (6)
The factorization of the problem in the case of infinitely strong magnetic field
is obvious.
The mentioned above setting of the problem is considered in number of works
[46,48,49,51]. As we know, the factorization was noted only in [46]. Unfortunately,
the results of this work are most likely mistaken (see below). Now we can cal-
culate the distribution gN of chains upon their length. Substituting (6) into (2)
and using the condition of chemical equilibrium between chains of different sizes
[42,43] or equivalently the minimization of the free energy of the system [45,48]
we find
gN = xN /vZ2 , (7)
)∞
where parameter x is expressed via normalization condition N =1 N gN = n (n
is the concentration of magnetic grains in MF) as
x
= ΦZ2 . (8)
(1 − x)2
Here Φ = nv is the volume fraction of magnetic particles. From (7) and (8) we
finally determine the average number of particles in the chain
,
1 1
N = + + ΦZ2 . (9)
2 4
We see that owing to factorization the problem of determination of the av-
erage length of the chains is reduced to calculation of the partition function of
dimer. The asymptotic representation at λ 1 of the dimer partition function
in zero and infinite external field is
e2λ 8 23 229 5263 11536 57427
Z2 (0) = 3 1 + + + 3+ + + , (10)
3λ 3λ 3λ2 9λ 54λ4 27λ5 27λ6
e2λ 5 41 155 11195 39235 628145
Z2 (∞) = 2 1 + + + + + + . (11)
3λ 3λ 12λ2 18λ3 432λ4 432λ5 1728λ6
These expansions are given up to terms of order of O(λ−6 ) because of the
slow convergence of the series. The main (first) term in the right side of (10)
and (11) is a result of de Gennes and Pincus [1]. The other terms are new.
We note that in [48] it was informed about mistaken value of the second virial
coefficient b2 (or Z2 ) in [1]. This is however misunderstanding: the authors [48]
used for b2 the different notation than de Gennes and Pincus did. The calculated
dependencies of both partition functions are shown in Fig. 1; note that in the
figure and below the minimal value of λ is 0.5 according to our definition of a
chain.
It is interesting that maximal deviations of calculated values from unity (i.e.,
from de Gennes and Pincus asymptotic values) takes place at small λ ∼ 2 ÷ 3
170 K.I. Morozov and M.I. Shliomis
7
Z 2 (0)
2 Z 2 (∞)
1
0
0 2 4 6 8 λ
Fig. 1. Partition functions of dimer in zero and infinite external field in units of de
Gennes and Pincus asymptotic values e2λ /3λ3 and e2λ /3λ2 , respectively
however even at λ = 10 it achieves 50%. The expansions (10) and (11) describe
the calculated values within the accuracy of 10% starting from λ ∼ 5 ÷ 6.
Let us analyze relation (9) in some limiting cases. When λ 1 we recover the
result for average chain length in zero field [48,49]. Next we consider the case of
low-concentrated magnetic fluid with strong dipolar interaction so as Φb2 1.
As it was noted above, both main theories should have the same limit of low
concentration. From (9) and (5) it follows N = 1 − Φb2 (b2 is negative). At the
same time the theory of de Gennes and Pincus [1] predicts N GP = 1 − 2Φb2 .
The origin of double difference of the coefficient is clear. In the approach [1] the
steric interactions inside the chain were ignored. It results in double exceeding
of contribution of neighboring particles interaction in partition function of a
chain. As we mentioned above the factorization of the considered problem was
noted in [46] where the grand partition function of the system was calculated.
The author obtained for variable equivalent to our x the cubic expression (see
equation (6) in [46]) instead of quadratic one (8). The character of Letter does
not allow to check all the calculations made in [46]. Nevertheless, we argue that
there is a mistake in his equation (5), containing the redundant coefficient 2.
Without this factor the equation (6) in [46] reduces to the correct relation (8).
We can support our point of view on [46] also a posteriori. Indeed, the relation
(8) is a straightforward consequence of our formalism based on calculation of the
partition function of a chain. It is clear that the final result should not depend
on the choice of thermodynamic potential considered.
In this Section we developed the association theory of ideal chains of mag-
netic grains. The size distribution of dipolar chains obeys to equations (7) and
(8). The average length of chains (9) increases ∼ Φ exp(2λ) under small values
Magnetic Fluid as an Assembly of Flexible Chains 171
√
of this parameter and ∼ Φ exp(λ) at high λ. The average value N in zero
and infinite external field can easily calculated according to (9)-(11) and data
shown in Fig. 1. Being the very important integral characteristics of the phe-
nomenon of chain formation, the average length itself however does not contain
an exhaustive information about system. Our understanding of the phenomenon
would be incomplete without study of statistical properties of the chains, with-
out their spatial and orientational correlations. Just these questions we start to
investigate.
Here we will see that the factorization is a general property of the considered
model what allows to express any interparticle correlations in terms of spatial
and orientational correlations inside dimers. First, let us define four orientational
correlation functions
e1 r 12 (e1 r 12 )2
A = e1 e2 , B = , C = (e1 e2 )2 , D = , (12)
d d2
and two spatial characteristics for a dimer:
r12 r12
2
E= , F = 2
. (13)
d d
Averaging in these expressions is fulfilled over variables r 12 and e2 of second
particle whereas the orientation e1 of first grain is taken as polar axis. The polar
angle between e1 and the displacement vector r 12 takes the values from zero to
π/2 according to method of integration mentioned at the beginning of Sect. 2.
Due to such integration over spatial variable the correlation function B is not
zero. It is clear that in considered cases of zero and infinite external field all
the correlations (12) do not depend on the orientation e1 . The correlations (12)
together with (13) are dimensionless functions upon λ and easily calculated. For
the case of zero external field they are depicted in Fig. 2. We note that the curve
B lies higher the curve D. The maximal size of dimer (see curves E and F ) suits
to intermediate values of λ. Indeed, when λ ≤ 1 the cluster of two grains falls
into separate particles even for very small distance between them, r ∼ d. At high
values of λ, the mean interparticle distance r is of the order of d as well due
to strong dipolar interactions. We note also that r never exceeds 2d (in fact,
it is always less than 1.4d) what means the numbering of the grains in a chain
and verifies the chosen in Sect. 2 method of integration over space coordinate.
All the introduced dimer’s variables tend to unity with increasing of coupling
parameter. Their asymptotic representation at λ 1 is
2 2 20 266 4241 77669
A=1− − − 3− 4− − , (14)
λ λ2 3λ 9λ 27λ5 81λ6
172 K.I. Morozov and M.I. Shliomis
2 .0
F
1 .5
E
1 .0
B
D
0 .5
C
A
0 2 4 6 8 λ
Fig. 2. Correlation functions (12) and spatial mean values (13) of dimers in zero field
versus the coupling parameter λ
2.0
F∞
1 .5
E∞
1.0
B∞
D∞
0 .5
0 2 4 6 8 λ
single preferable direction in the system and plays a role of external field for
other particles in the chain. We average over the variables of particles starting
from the “tail” of a chain, i.e., from N th grain and then going to the grains with
number N − 1, N − 2 etc. At each step we take the direction of previous grain
as polar axis. Due to mentioned independence of the integral (4) as well as the
dimer variables (12) and (13) on orientation of previous (first) particle, at such
step of averaging we obtain one of the dimer functions Z2 , A, B, C, D or F .
Just this signifies the factorization of any two-particle correlation.
Let us assume for definiteness that 1 ≤ i < j ≤ N . The integration over variables
of grains N , ..., j + 1 is obvious: it reduces to multiplying by unity, ...N,...,j+1 =
1. Here and further we indicate explicitly the numbers and the order of particles
been integrated. Integration over particle j in the coordinate system connecting
with particle j − 1 gives
Now we consider mean value ei;α ei;β which looks like initial quantity, but with
j = i. Again we connect the coordinate system with particle i − 1 and integrate
over variables of grain i. The result is
1−C 3C − 1
ei;α ei;β i = δα,β + ei−1;α ei−1;β , (26)
2 2
where δα,β is a unit tensor. The last expression is a recurrent one. After averaging
over grains with numbers i − 1, i − 2, ..., 2 we have
1
ei;α ei;β i,i−1,...,2 = δα,β + T i−1 Δα,β , (27)
3
where the notation T = (3C −1)/2 is used and the irreducible second rank tensor
Δα,β is introduced according to relation
1
Δα,β = e1;α e1;β − δα,β . (28)
3
Finally, combining expressions (25) and (28) the considered mean value is found
in the form
1
ei;α ej;β = Aj−i δα,β + T i−1 Δα,β . (29)
3
Magnetic Fluid as an Assembly of Flexible Chains 175
Let us analyze the result (37), (38). In the case of dimers, N = 2, the obvious
result R2 /d2 = D, R⊥ 2
/d2 = (F − D)/2 is recovered. When the chains are
very short, N λ, we have R2 = N 2 d2 , R⊥ 2
= 0. This is just the case
of rod-like aggregates considered in [45,53]. This case, however, is never put
into effect in reality. The fact is that both the energy gain accompanying every
pairing of magnetic grains (∼ λkB T ) and the chain correlation length (∼ λd)
are determined by the same coupling constant λ. Therefore, the mean number of
grains in a chain N as well as its correlation length Lp increases with enlarging
of λ, so that the condition N λ is never satisfied. [From the formal point of
view, N increases with growth of λ exponentially (see (9)) whereas Lp ∼ λ/2
according to (36)]. Statistically, some number of very short chains (N λ) are
certainly existed, but their fraction is negligible. We note that recently authors
[45,53] changed their own previous representations and admitted the deformation
of dipolar chain [58].
For a long chain equations (37) and (38) give (R2 − R⊥ 2
)/R2 = 0 and
2AB 2
R = N d
2 2
F+ , N →∞. (39)
1−A
Such a N -dependence of the mean-squared end-to-end distance of a long chain
is typical of ideal polymer chain [54]. Its local stiffness is often characterized
– along with the persistent length Lp – by the so-called Kuhn segment LK
determined by the relation R2 = N rLK d, where r is the mean distance
between neighboring grains in the chain. Both the quantities, Lp and LK , are
depicted in Fig. 4 as functions of the dipole parameter λ. When it is sufficiently
large, we have
2AB 2 19 2 715 8377 332995
LK = F + E −1 = λ − + − − − . (40)
1−A 6 9λ 108λ2 216λ3 1296λ4
As before, this expansion perfectly describes the calculated values of LK at
λ ≥ 6. It is interesting that starting from λ 5 (i.e., just when the aggregation
Magnetic Fluid as an Assembly of Flexible Chains 177
3
LK
2
Lp
1
0
0 2 4 6 8 λ
Fig. 4. The persistent length Lp and the Kuhn length LK as a function of coupling
parameter λ
or coil of connected monomers and has quasi-spherical form. In fact however this
conclusion proves to be valid even for sufficiently short chains with N ≥ LK (see
Sect. 3.5).
that characterizes the chain scales in two directions. The effective form of the
chain can be judged calculating the parameter of nonsphericity S defined as
S
1.0
ξ=∞
0 .8
0.6
0 .4
ξ=0
0.2
0 .0
0 5 10 15 20 25 N
Fig. 5. Nonsphericity parameter S as a function of number of particles in chain in
zero (white diamonds) and infinite (black ones) external field. The coupling parameter
λ = 10
As it follows from analysis of the figure, even for a very strong dipole inter-
actions the value of S for zero field (lower curve in Fig. 5) quickly decreases with
growth of number of particles in the chain. In fact, any dipolar chain can be
considered as a coil slightly elongated along one direction. The more number of
grains in a chain, the smaller the coil elongation. At the same time in the case
of infinite external field the nonsphericity parameter S increases fast with the
growth of number of particles in a chain and approaches to unity (upper curve
in Fig. 5). Does it mean that dipolar chain becomes stiffer in external field? Of
Magnetic Fluid as an Assembly of Flexible Chains 179
course, it does not. The point is that each Kuhn segment is directed along the
external field and all the Kuhn segments are connected with each other.
Summarizing the results of the Section 3 we can conclude that the ideal
chain of magnetic grains is very flexible formation with short (of the order of
a few particles) persistent length or the Kuhn segment. The chain looks like a
quasi-spherical coil of connected links in the absence of external field. This coil
is disentangled with increasing of the field and transforms into chain randomly
and weakly bent relative to the field direction.
4 Nonideal Chains
Above we have considered the case of ideal chains. We made allowance only for
interactions of neighboring particles belonging to the same chain ignoring in-
teractions between 1) the non-nearest neighbors along the chain, 2) the distant
segments of the same chain which are finding near each other owing to chain
flexibility, and 3) the segments of different chains. The account of the first of
them in the case of straight linear chains is reduced to re-normalization on 20%
of the nearest-neighbors interactions (see Sect. 1). The chain flexibility should
only reduce this value, so that one can regard these interactions as insignificant.
The second type of interactions can form antiparallel side-by-side configurations
of dipoles in zero field. The third type should be taken into account along with
the second one because the friable coils of ideal chains (with typical volume of
∼ N 3/2 in zero field) begin to overlap for strong dipolar interactions (λ ≥ 7) al-
ready at small volume fractions (≤ 1%) of ferroparticles. A due regard for these
interactions can be carried out in the frame of the concept of quasi-monomers
[54] treating every long flexible chain as a system of disconnected segments. We
identify each a segment with the individual grain and take into account the in-
teractions between segments in the approximation of the second virial coefficient
bm . [More precisely, one should examine the Kuhn segment or the segment of
persistent length as an effective monomer. This strict consideration is similar
to above-mentioned accounting for the next-nearest grains in the segment what
re-normalizes slightly the grain-grain interactions. Such an approach, however,
complicates considerably the calculations, what is why we use here more intuitive
rather than the strict arguments]. bm is not the usual second virial coefficient of
interactions of two hard dipoles. The point is, even considering the chain as a
system of disconnected grains, we should remember of the connection of all its
segments. Thus, calculating bm , we assume that the most energetically profitable
positions corresponding to head-to-tail configurations are occupied already by
the grains of main chain and then forbidden for any other interacting segments.
The way of determination of bm can be divided on four steps. First, we right
the usual second virial coefficient b2 (ξ) of interacting dipoles in external magnetic
field H
2
1 ξ
b2 (ξ) = − {e−U (12)/kB T − 1}eξ(e1 h+e2 h) de1 de2 dr , (45)
2v 4π sinh ξ
180 K.I. Morozov and M.I. Shliomis
where ξ = mH/kB T is the Langevin parameter and h is unit vector along field.
Obviously, the equation reduces to (5) in zero field. Second, we approximate (45)
replacing the orientation of one particle by the field direction, e1 → h, so as
ξ
b2 (ξ) ≈ − {e−U (1 2)/kB T − 1}eξ(e2 h) de2 dr , (46)
8πv sinh ξ
where the symbol of averaging denotes the mentioned replacement. Equation (46)
gives the exact values of the second virial coefficient (45) in the limiting cases
of zero and infinite external field. The former follows from the independence
of typical integral I (4) on orientation of first particle, the latter is evident.
So, (46) is an interpolation formula for moderate values of external field. Our
spot check for some values of λ and ξ shows that the difference between (45)
and (46) does not exceed 10%. Third, in order to exclude the occupied states we
restrict the possible values of polar angle θ between vectors r and h by the range
[π/3, 2π/3] what effectively takes into account the presence near particle 1 of two
head-to-tail aligned neighboring dipoles. Fourth, we overcome the nonphysical
logarithmic divergence owing to previous step by the cutting off the region of
integration over interparticle distance r. We assume that r can vary up to its
maximal value r∗ defined by the physical limitation m2 /r∗3 kB T = 1 (within the
spherical volume r ≤ r∗ the dipole potential for the antiparallel orientation of
dipoles exceeds thermal energy. Finally, we write the second virial coefficient of
two interacting quasi-monomers in the form
r∗
2π/3
ξ
bm =− dr dθ de2 {e−U (1 2)/kB T − 1}eξ(e2 h) r2 sin θ . (47)
4v sinh ξ
0 π/3
5 Conclusion
We studied here the spatial and orientational intrachain correlations in zero and
infinitely strong external field and established the important role of flexibility in
the description of conformational properties of dipolar chains. Among the prob-
lems which proved to be out of the scope of the present investigation we would
like to mark the following ones. First, the problem of taking into account of the
interchain interactions which can be significant due to the long-range nature of
dipolar interactions. As it is shown in [58] the interchain correlation can notice-
182 K.I. Morozov and M.I. Shliomis
ably change the mean length of a chain. Second, there is a problem of the unique
description of both effects of dipolar interparticle interactions – the condensa-
tion and association phenomenon. On the language of present work it means
the interconnection between condensation and globule formation. We assume
that direct identification of both phenomena with each other is not appropri-
ate. Most likely the globule formation precedes the liquid-gas phase transition
and the globules themselves are the nuclei of future concentrated phase. An-
other interesting aspect of the problem is the study of both effects in the case of
moderately concentrated magnetic fluids. In our approach it was convenient to
divide mentally the microstructure formation on stages of (i) appearance of ideal
chain and then (ii) its transformation to nonideal one due to chain flexibility. In
reality both processes take place simultaneously. So, it is reasonable to wait for
the formation in sufficiently concentrated magnetic fluids of extended network
[37] instead of globule formation. Finally, there is the problem of description of
thermodynamical and dynamical properties of magnetic fluids with developed
microstructure. This work was begun in [45,48,49,53,58]. We will return to study
of these intriguing questions later.
This work was supported by the Russian Fund for Fundamental Research
(Project 02-03-33003) and the Israel Science Foundation (Grant 336/00).
References
1. P.G. de Gennes, P.A. Pincus: Phys. Kondens. Materie 11, 189 (1970)
2. S.A. Adelman, J.M. Deutch: Anv. Chem. Phys. 31, 103 (1975)
3. M.S. Wertheim: J. Chem. Phys. 55, 4291 (1971)
4. W. Sutherland, G. Nienhuis, J.M. Deutch: Mol. Phys. 27, 721 (1974)
5. K.-C. Ng, J. Valleau, G. Torrie, G. Patey: Mol. Phys. 38, 781 (1979)
6. C.F. Hayes: J. Coll. Int. Sci. 52, 239 (1975)
7. J.-C. Bacri, D. Salin: J. Phys. Lett. 43, 649 (1982)
8. J.-C. Bacri, R. Perzynski, D. Salin, V. Cabuil, R. Massart: J. Magn. Magn. Mater.
85, 27 (1990)
9. J.-C. Bacri, F. Boué, V. Cabuil, R. Perzynski: Colloids Surfaces A 80, 11 (1993)
10. C.F. Hayes, S.R. Hwang: J. Coll. Int. Sci. 60, 443 (1977)
11. E.A. Peterson, D.A. Krueger: J. Coll. Int. Sci., 62, 24 (1977)
12. R.W. Chantrell, J. Sidhu, P.R. Bissel, P.A. Bates: J. Appl. Phys. 53, 8341 (1982)
13. A.F. Pshenichnikov, I.Yu. Shurubor: Bull. Acad. Sci. USSR, Phys. Ser. 51, 40
(1987)
14. A.F. Pshenichnikov, I.Yu. Shurubor: Magnetohydrodynamics 24, 417 (1989)
15. A. Cebers: Magnetohydrodynamics 18, 137 (1982)
16. K. Sano, M. Doi: J. Phys. Soc. Jpn. 52, 2810 (1983)
17. K.I. Morozov: Magnetohydrodynamics 23, 37 (1987)
18. K.I. Morozov: Bull. Acad. Sci. USSR, Phys. Ser. 51, 32 (1987)
19. Yu.A. Buyevich, A.O. Ivanov: Physica A 190, 276 (1992)
20. K.I. Morozov, A.V. Lebedev: J. Magn. Magn. Mater. 85, 51 (1990)
21. A.F. Pshenichnikov, V.V. Mekhonoshin, A.V. Lebedev: J. Magn. Magn. Mater.
161, 94 (1996)
22. A.O. Ivanov, O.B. Kuznetsova: Phys. Rev. E 64, 041405 (2001)
Magnetic Fluid as an Assembly of Flexible Chains 183
Supplementary Glossary
A = e1 e(
2 dimer correlation function
a = d + R2 longitudinal size of chain
B = e1 r 12 /d dimer correlation function
B∞ dimer correlation function B in infinite field
b = d + R⊥ 2 transverse size of chain
b2 second virial coefficient of dipoles
bm second virial coefficient of quasi-monomers
C = (e1 e2 )2 dimer correlation function
D = (e1 r 12 )2 /d2 dimer correlation function
D∞ dimer correlation function D in infinite field
d particle diameter
E = r/d dimensionless interparticle distance
E∞ dimer correlation function E in infinite field
ei unit vector along dipole i
F = r2 /d2 dimensionless square of interparticle distance
F∞ dimer correlation function F in infinite field
gN number of N -particle chains per unit volume
H external magnetic field
h unit vector along field
kB Boltzmann’s constant
LK Kuhn segment length
Lp persistent length
m particle magnetic moment
N number of grains in chain
n number of grains per unit volume
)N −1
R = i=1 r i,i+1 “end-to-end” vector of N -particle chain
R2 longitudinal mean square of “end-to-end” vector
R⊥2
transverse mean square of “end-to-end” vector
r mean interparticle distance
S = (a2 − b2 )/(a2 + b2 ) nonsphericity parameter
T temperature
v grain volume
Z2 dimer partition function
λ = m2 /d3 kB T coupling parameter
ξ = mH/kB T Langevin parameter
Φ = πnd3 /6 volume fraction of magnetic grains
Magnetoviscous Effects in Ferrofluids
Abstract. The appearance of field and shear dependent changes of viscosity in fer-
rofluids opens possibilities for future applications e.g. in damping technologies. To
enhance the effects, it is necessary to understand the observed magnitudes of magne-
toviscosity in commercial ferrofluids from a microscopic point of view. Starting from
experimental results, it is described how the magnetoviscous effects can be explained
by chain formation of a small fraction of large particles in the fluid. With a dedicated
experiment ferrofluids are separated into fractions with high and low amount of such
particles. The rheological characterization of the fractions prove the aforementioned
model. Furthermore it leads to additional information concerning viscoelasticity of the
suspensions in a magnetic field.
1 Introduction
The phenomenon of field dependent changes of viscosity of a suspension of mag-
netic nanoparticles is known since more than 30 years. In 1968 McTague [1]
discovered an increase of viscosity of a ferrofluid containing Co-particles in the
presence of a magnetic field. His experiments showed a dependence of the vis-
cosity increase on the magnetic field strength as well as on the field’s direction
relative to the flow. In his paper as well as in an accompanying paper by Hall
and Busenberg [2] the effect was explained by a hindrance of rotation of the
suspended particles due to the action of the magnetic field. This concept should
be shortly explained here, since it has fundamental importance for almost all
discussions about magnetoviscous effects in suspensions of nanosized particles.
Assume a ferrofluid under influence of a shear flow, as shown in figure 1. In
such a situation the magnetic particles will rotate in the flow due to the action
of mechanic torque produced by viscous friction in the fluid. If a magnetic field
H is applied to the fluid, the magnetic moment m of the particles will align with
the field direction. In a situation where field direction and vorticity of the flow
are collinear (figure 1a) the magnetic alignment will only lead to the fact that
the magnetic moment of the particles becomes collinear with the direction of
vorticity. An influence on the motion of the particle and therefore on the flow
of the fluid as a whole does not appear. The situation changes if vorticity and
field direction are perpendicular. Under these conditions the mechanic torque
will force a misalignment of the magnetic moment of the particle and the field
direction, provided the magnetic moment’s direction in the particle is fixed. An
angle between the mutual directions of the magnetic moment and the field will
immediately give rise to a magnetic torque, trying to realign m and H. This
torque acts opposite to the mechanic torque and causes thus a hindrance of the
free rotation of the particle in the flow. This increases the flow resistance and
thus the fluid exhibits an increased viscosity.
A rigorous theoretical analysis of the phenomenon has been given by M.
Shliomis [3] four years after the experimental discovery. Taking into account
Brownian motion of the particles he derived an expression describing the change
of viscosity - called rotational viscosity ηr - as a function of the strength and
direction of the magnetic field H in the form
3 α − tanh α
ηr = φ η0 < sin2 β >, (1)
2 α + tanh α
where η0 denotes the viscosity of the fluid in the absence of a magnetic field, φ
the volume concentration of the magnetic particles including the surfactant and
β is the angle between vorticity and field direction; < ... > denotes the spatial
average. The parameter α is the ratio of magnetic and thermal energy of the
particles
μ0 mH
α= , (2)
kT
where μ0 denotes the vacuum permeability, m the particle’s magnetic moment,
k Boltzmann’s constant and T the absolute temperature. Equation (1) has been
derived under the assumption of two important limitations. First of all it has
to be assumed, as already discussed in the qualitative description above, that
the magnetic moment of the particles is spatially fixed within the particle. This
assumption leads to a question concerning the relaxation of magnetization of
the fluid. As discussed in P. Fannin’s contribution [4] the magnetization can
either relax by the Néel process, i.e. by a change of the magnetic moment’s
direction in the particle, or by the Brownian process, that means by a rotation
of the particle in the flow. The relaxation takes place by the process with the
shorter characteristic relaxation time. As shown in [4] both times depend on the
size of the particles but the Brownian time scales only linear with the particles’
Magnetoviscous Effects 187
volume while the Néel time depends exponential on the volume. Thus, small
particles will relax by the Néel process while particles above a certain critical
diameter follow the Brownian process. For the discussion of rotational viscosity
this means, that only particles with a diameter above the critical one, usually
called magnetically hard, will contribute to the changes of viscosity.
The second fundamental assumption in [3] has been the restriction to highly
diluted systems, neglecting any interaction between the magnetic particles. For
McTague’s experiments both assumptions were fulfilled. He used Co-particles
with a diameter about 10 nm being sufficiently larger than the critical diameter
of 6 nm. These particles are thus magnetically hard and a respective suspension
is expected to show rotational viscosity. Moreover the volume concentration of
the suspensions in [1] were as low as 0.05 vol% - a high dilution preventing
significant interparticle interaction.
A few month before McTague’s experiment was published, Rosensweig [5]
reported a viscosity change in concentrated magnetite suspensions. The increase
of the fluids’ viscosity found in these experiments was relatively high and reached
about 200 % of the zero field value. A result principally encouraging further inves-
tigations in this direction especially with a focus towards technical applications
using the phenomenon. Nonetheless, these results were not discussed further-
more and especially they were never compared with the theoretical description
in [3].
Fig. 2. The change of viscosity in a commercial magnetite ferrofluid together with the
theoretical prediction following equation (1) calculated for the particle size distribution
shown in the inset (dashed line) and with a fit of equation (1) (solid line).
even when the absolute values can not be explained in terms of individual par-
ticles rotating in the flow. Moreover, the fit parameters leading to an apparent
mean particle size of about 16 nm for the full particle concentration indicate
that interaction of particles, forcing formation of agglomerates, has significant
importance in the concentrated suspension investigated.
To avoid confusion we will call the changes of viscosity in concentrated sus-
pensions, where interparticle interaction plays a significant role “magnetoviscous
effect”, while the term “rotational viscosity” will be reserved for the case of highly
diluted suspensions with negligible particle interaction.
A dominant contribution of interparticle interaction to the magnetoviscous
effect leads immediately to the question what kind of microstructure is responsi-
ble for the observed effects. To get a deeper insight into the behavior of the fluid
under shear and magnetic field influence we developed a specialized rheometer
for the investigation of ferrofluids [7], shown in figure 3.
The core part of the rheometer is a modified cone-plate flow cell with a mov-
ing plate and a cone attached to the torque sensor. The connection between cone
and torque sensor is provided by a long rod guided in an air bearing to prevent
spurious friction due to the bearings. The long rod provides a sufficient spatial
separation between the sensor and the magnetic field region, ensuring that no
disturbing influence of the field on the measurement occurs. The rheometer al-
lows investigations of magnetic fluids under stationary as well as time dependent
load.
Using this device we have investigated [8,9] the dependence of the magne-
toviscous effect on shear rate and fluid composition. Figure 4 shows a typical
change of viscosity with field strength for various shear rates for the ferrofluid
APG513A containing 7.2 vol% of magnetite particles with 10 nm mean diameter
in an ester. Obviously a strong magnetoviscous effect is observed at low shear
Magnetoviscous Effects 189
Fig. 3. The rheometer used for the investigation of magnetic fluids in the experiments
presented here. Details are given in the text.
rate diminishing successively with increasing load. The absolute values of the
magnetoviscous effect are again - as in [6] - significantly higher than expected
from the theory in [3]. Furthermore, the field dependent shear thinning is an
effect which is not predicted in [3] at all. These observations led to the assump-
tion that the formation of chains or clusters of particles under the influence of a
magnetic field has to be a leading component in the description of magnetovis-
cosity. The hindrance of rotation of these agglomerates is assumed to give rise
to the strong increase of viscosity with magnetic field strength. Furthermore the
rupture of the chains in a shear flow can be the basis for an explanation of the
observed shear thinning [8,10].
Within the frame of this model one has to observe that only relatively large
magnetite particles with a diameter of more than about 13 nm show an inter-
particle interaction being strong enough to contribute significantly to such chain
formation. That means that only a small fraction of the overall magnetic con-
190 S. Odenbach and S. Thurm
Fig. 4. The field and shear rate dependence of the viscosity of APG513A
F1 32.41 8.3
F2 32.34 8.8
F3 31.54 9.2
F4 32.17 9.2
F5 32.06 10.1
Fig. 5. The magnetoviscous effect for fluids with different content of large particles
parameters like viscosity, interparticle forces and - as the most important factor
- the particle size distribution. First step of the layout of the separation device
was a numerical simulation of the diffusion process to provide information about
expected diffusion times dependent on the magnetic field gradient, which can
vary in strength and spatial structure. Later this approximation of field strength
and geometry was used as input for a computer simulation to develop the shape
of the pole shoes, which are the key parts of the experimental setup.
The numerical simulation bases on the assumption that the magnetic par-
ticles do not interact with each other. This is a safe approximation, because
interparticle forces cause chain or cluster formation, which in turn accelerates
the process of diffusion. In principle two different theoretical approaches exist to
model the process of diffusion in magnetic fluids.
Close to equilibrium, i.e. for weak driving forces leading to diffusion in the
system, an approach from the point of view of irreversible thermodynamics can
be chosen. Detailed information for this approach can be found in Prigogine [12],
de Groot, Mazur [13], Blums et al. [14] and Odenbach [15].
For strong driving forces - as they are required for the separation process
in focus of our experiment - an approach based on statistical physics has to be
used. The individual velocities and trajectories of all particles are obviously not
important, only the integral change of the spatial concentration distribution with
time has to be calculated. The fundamental equations describing this problem
can be found e.g. in Gerber et al. [16], who simulated numerically the process of
HGMS (H igh Gradient M agnetic S eparation). Starting point is the equation of
continuity, describing the temporal changes of the concentration distribution of
the magnetic particles
∂c/∂t + ∇J = 0 (3)
here the total flux of particles J consists of a diffusive part, J d = −D∇c and a
part J f = vc driven by external forces, where v denotes the drift velocity of the
particles. Substituting these expressions into (3) gives
The first one is neglected since its influence is about two orders of magnitude
smaller than the magnetically induced drift in reasonable field gradients. The
latter is excluded by the assumption that the fluid as a whole is at rest during
the separation process.
Further effects could appear due to van der Waals or magnetic interaction
leading to an agglomeration of the particles. This would increase the magnetic
moment entering the magnetic driving force and would thus enhance the drift
velocity leading to an acceleration of the separation process. Since the simulation
is thought to give an approximation of the diffusion process allowing the design
of a separation device enabling separation of the large particles in the fluid on
reasonable time scales, effects like the mentioned agglomeration will only enhance
the performance of a system based on the above assumptions.
Finally (5) can be written as
∂c/∂t = 1/ (6πηr) kT ∇2 c − ∇ (μ0 Vmag M0 ∇Hc) (6)
where m has been replaced by m = Mo Vmag with Vmag denoting the volume
of the magnetic particle and Mo being the spontaneous magnetization of the
magnetic material – in our case magnetite with Mo = 4.5 · 105 A/m.
Fig. 6. Numerically calculated time dependent concentration profile for three different
particle diameters after a separation time of 1 week. The height of fluid in the container
is 5 cm
Using (6) one can calculate the concentration distribution of magnetic par-
ticles as a function of time for different particle sizes. From a comparison of
the resulting distributions it can be judged whether a separation of a fluid in
fractions with different content of large particles and mainly constant content of
small ones is possible.
Fig. 6 shows the calculated time dependent concentration profile for three
different particle diameters after a separation time of one week for a 5 cm deep
194 S. Odenbach and S. Thurm
container. The field gradient applied in the direction of diffusion is constant over
the depth of the container at a value of 107 A/m2 . As a boundary condition it
has been assumed, that the concentration at the lower wall of the container can
not exceed 150 % of the original concentration.
One can see that the magnetic field gradient acts mainly on the largest par-
ticles. Particles which are smaller than 10 nm hardly react on the field gradient,
thus we are able to concentrate just the fraction of bigger particles in the region
of high magnetic field. The parameters separation time and field gradient control
the accumulation and depletion of bigger particles in the lower and upper region
of the container respectively.
The container is located in the gap between two tailor made pole shoes of an
electromagnet. The pole shoes are designed in a way that they provide a constant
magnetic field gradient of 107 A/m2 over the whole fluid volume. This ensures
a uniform diffusion of particles from the top to the bottom of the container as
already shown in the simulation (see Fig. 6).
The container is equipped with an in-situ concentration measuring system,
consisting of three coils, located in different heights. The inductance of the coils is
directly proportional to the local concentration of magnetic particles. Therefore
Magnetoviscous Effects 195
4 Results
The fluid used in the experiments discussed hereafter has been a sample of
APG513A from Ferrofluidics, with a saturation magnetization of 32 kA/m, an
initial susceptibility of 1.4 and a dynamic viscosity of 150 mPas. The magnetic
separation ran over a period of 30 days. After this time period the separation
was stopped, and the fluid was removed in two fractions from the container as
indicated in Fig. 7.
Table 2. Magnetization data from the original APG513A and both fractions after the
separation process
The discussion of the results obtained with these measurements is splitted into
two parts. First we will focus on the magnetoviscous effect, measured with the
rotational mode of the ferrofluid-rheometer described above. The elemental in-
fluence of bigger particles on the change of viscosity under influence of shear flow
and magnetic fields is the central aspect of this discussion. In the second part
we will present some results obtained in the oscillating mode of the rheometer,
leading to information about magneto-viscoelastic effects.
The magnetoviscous effect. As seen from the experiments in [9], the inves-
tigation of field and shear rate dependent changes of viscosity in a ferrofluid
provide an excellent tool to get an insight into the microscopic reasons for the
rheological properties of the suspensions. The model of chain formation of big-
ger particles – being under test with the experiments described here – leads
to a couple of phenomena being expected to appear in a ferrofluid separated
as described above. First of all, in zero field, the dependence between stress
and shear rate in both fractions should be linear. That means the fluid should
behave Newtonian, since no field induced interparticle interactions leading to
chains or clusters are present. The ferrofluid behaves like a normal suspension
of non-magnetic particles.
Applying magnetic fields, the upper fraction containing only a negligible
amount of large particles is expected to show weak magnetoviscous effects since
the majority of particles in this fraction is too small to cause significant effects
due to interaction.
In contrast the lower fraction should react noticeable on increasing magnetic
field strength. These dependencies are illustrated in Fig. 9.
For the upper fraction - represented by open symbols - the relation between
stress and shear is linear. This fraction shows mainly newtonian behavior and
the existing weak dependence from magnetic field strength and shear rate will
only be seen in a field dependent plot (see Fig. 10). This is a direct validation of
the assumption, that just the bigger particles cause chain and cluster formation,
which in turn gives rise to significant changes of the viscosity.
The lower fraction is represented by the filled symbols in Fig. 9 where different
symbols indicate different stength’ of the applied magnetic field. Starting with
198 S. Odenbach and S. Thurm
Fig. 9. Stress versus shear-rate for different magnetic fields. Both fractions and the
original fluid are plotted
the curve for zero field, the similarity between the lower and upper fraction is
obvious. This is expected for H = 0 since the only difference between the fluids
is a slight change in number and size of the suspended particles. To understand
this one can apply Rosensweig’s equation for the dependence of viscosity on the
volume fraction of magnetic particles in moderately concentrated suspensions
[17],
⎛ ⎞
& '2 −1
η 5 5 φ̃
= ⎝1 − φ̃ + φ̃c − 1 ⎠ , (7)
η0 2 2 φ̃c
where φ̃ denotes the volume fraction of particles including their surfactant and
φ̃c is a critical volume fraction for which the suspension’s viscosity diverges -
this value is usually chosen as φ̃c = 0.74. It is easy to see that a difference of
volume concentration of about 10% - as it was found for the two fractions of our
separation experiment - leads to a viscosity difference of about 7% only.
With increasing magnetic field strength an increase of viscosity - the magne-
toviscous effect - can been seen from the change of the slope of the stress-shear
relation. Furthermore one observes that this difference diminishes with increas-
ing shear rate. So presenting the magnetoviscous effect in the commonly used
form by plotting the change of viscosity as a function of field strength for various
shear rates leads to the well known behavior as shown in Fig. 10. As discussed
earlier these changes of viscosity can be explained by the formation and rupture
of chains formed by the large particles in the suspension in the presence of a
magnetic field. We compare here the change of viscosity for the original fluid
with both fractions for two shear rates as a function of magnetic field strength.
As expected, the magnetoviscous effect increases in all three fluids with de-
creasing shear rate. Furthermore the influence of the concentration of bigger
particles is directly visible. The fluid with a large concentration of bigger parti-
Magnetoviscous Effects 199
Fig. 10. Magnetoviscous effect in original APG513A, lower and upper fraction
cles has a much stronger magnetoviscous effect than the original fluid, and vice
versa in the upper fraction with a small concentration of bigger particles.
Viscoelastic effects. The formation of chainlike clusters gives rise to the as-
sumption, that viscoelastic effects should occur in ferrofluids under the influence
of magnetic fields. A first evidence for such effects was provided from Odenbach
et al. [18] with the investigation of the Weissenberg-Effect in ferrofluids exposed
to magnetic fields. Using rheometrical investigations, the application of an oscil-
lating load gives direct access to information about viscoelastic properties. The
advantage of the oscillating mode is, that measurements with small amplitudes
does not change the “zero-shear-rate” structure of the fluid.
As an example we’ll discuss here the phase shift between shear and stress,
being a characteristic measure for the appearance of viscoelasticity. For an elastic
body this phase shift should be zero, while it is π/2 for a Newtonian liquid.
Correspondingly viscoelastic liquids exhibit phase shifts between 0 and π/2.
It was shown in Fig. 9, that the upper fraction shows a linear relation between
stress and shear rate, indicating newtonian behavior. This is confirmed by the
phase shift δ = π/2, being constant over the whole range of oscillation frequencies
ω as shown in Fig. 11. In contrast the lower fraction with many bigger particles
should have a phase shift between 0 degree and π/2, dependent on the length of
the chains formed by the particles. This length itself depends on shear rate as
well as on the strength of the applied magnetic field.
The lower fraction shows just under zero field no phase shift- i.e. it behaves
Newtonian for vanishing field. If a field is applied, a phase shift between 0 and
π/2 is observed. This phase shift decreases continuously with decreasing shear
rate and thus with increasing length of the particle chains. The phase shift of
the original fluid takes values in between the phase shifts of the upper and lower
fraction. The higher number of bigger particles in the lower fraction obviously
200 S. Odenbach and S. Thurm
Fig. 11. The phase shift between shear and stress for the original fluid and both
fractions obtained by separation as a function of the frequency of the oscillating load
for different shear rates. To obtain the given data points, measured values in a field
range between 3 kA/m and 17 kA/m have been averaged. The error bars give the
standard deviation of this averaging process.
Acknowledgements
References
1. J. P. McTague: Magnetoviscosity of magnetic colloids. J. Chem. Phys., vol.51, no.1
(1969)
2. W.F. Hall, S. N. Busenberg: Viscosity of magnetic suspension. J. Chem. Phys.,
vol.51, no.1 (1969)
3. M. I. Shliomis: Effective viscosity of magnetic suspensioms. Soviet Phys. JETP,
vol.34, no.6 (1972)
4. P. Fannin: Magnetic spectroscopy as an aide in understanding magnetic fluids.
Springer LNP this issue (Berlin, New York 2002)
5. R.E. Rosensweig, R. Kaiser, G. Miskolczy: Viscosity of magnetic fluid in a magnetic
field. J. Colloid Interface Sci., vol.29, no.4 (1969)
6. O. Ambacher, S. Odenbach, K. Stierstadt: Rotational viscosity in ferrofluids. Z.
Phys. B-Condensed matter, vol.86 (1992)
7. S. Odenbach, T. Rylewicz, M. Heyen: A rheometer dedicated for the investigation
of viscoelastic effects in commercial magnetic fluids. J. Magn. and Magn. Mat.,
vol. 201 (1999)
Magnetoviscous Effects 201
1 Introduction
Since Rabinow and Winslow’s [1,2] discoveries in the 1940’s, magnetorheology
(MR) and electrorheology (ER) have emerged as a multidisciplinary field whose
importance has considerably increased these last ten years. The rheology of
these fluids is very attractive since it can be monitored by the application of a
field, either magnetic or electric. The most important advantages of these fluids
over conventional mechanical interfaces is their ability to achieve a wide range of
viscosity (several orders of magnitude) in a fraction of millisecond. This provides
an efficient way to control force or torque transmission and many applications
dealing with actuation, damping, robotics have been patented and are coming
on the marketplace [3]. The basic phenomena in electro or magnetorheology is
the ability to control the structure of a biphasic fluid. The two phases are usually
made on one hand of solid particles in the micrometer range and on the other
hand of a carrier fluid. The application of a field polarizes the particles and
induces their transient aggregation, hence an increase of the viscosity. In some
other cases the two phases can be two immiscible fluids and the application of
the field will change the shape and size of the droplets of the dispersed phase
which will modify the rheology [4,5]. Some intermediate cases can exist where
the carrier phase is itself a suspension of nanoparticles, a typical case being
a ferrofluid. Furthermore new materials where the two phases are solids, like
solids particles embedded in a rubber matrix belong to the same category of
smart composites whose rheology can be controlled by an external field [6,7,8].
All these materials have common features regarding the relation between the
shape and size of the domains of the dispersed phase and their rheology, even
2 Overview of MR Suspensions
where r is the separation vector between the centers of the two particles. This
energy is minimum when the two dipoles are aligned with r and maximum when
they are perpendicular leading to a preferential aggregation as chains of particles
aligned on the direction of the field. The formation of aggregates of particles will
depend on the ratio of this interaction energy to kT . Taking as reference the
energy of two dipoles in repulsive configuration gives:
1 m2 1 πμ0 μf β 2 a3 H0 2
λ= = (2)
4πμ0 μf r3 kT 2kT
units). It means that for usual magnetic fields the magnetic forces dominate
the Brownian forces. The situation is quite different if we consider ferrofluids
which are magnetic fluids with particle diameters of about 100 Angstrom; then
in the same conditions we have λ = 1 for H = 1600 Oersted. It means that for
usual fields the magnetic forces will always be dominated by Brownian forces and
we cannot expect to change significantly the viscosity of a ferrofluid by applying
a magnetic field. Furthermore this criterion on λ has to be reinforced by the fact
that the efficiency of thermal forces to break a chain of spheres increases with the
length of the chain. Despite these restrictions which seem to prevent the finding
of a noticeable ER or MR effect with suspensions of colloidal particles there are
some experimental evidences for large changes of viscosity in some ferrofluids
[18]; we shall discuss this point in more details in the next section. The quantity
λ is the key quantity which, together with the volume fraction, Φ = N vp /V , will
determine the equilibrium structure of a suspension of monodisperse particles
as a function of the applied field. We shall come back to this important point in
section 3.
In order to obtain all the quantities which will rule the behavior of the sus-
pension it is useful to start from the equation of motion of one particle and to
put it in a dimensionless form. This is what we would do to calculate the trajec-
tories of the particles by using Brownian or Stokesian dynamics [19,20,21]. For
a given particle we can write:
dv
m = F H + F ext + F I + F B (3)
dt
The first term F H is the hydrodynamic force on the test particle coming from
the hydrodynamic friction and is proportional to −ξ(v −v 0 (x)), where ξ = 6πμa
with μ the viscosity of the suspending fluid and v 0 (x), the imposed velocity field
at the location x of the particle. The term F ext is the hydrodynamic force due
to the symmetric part of the velocity gradient tensor. In the case of a pure shear
characterized by the shear rate γ̇ this force scales as 6πμγ̇a2 . The third force F I
is the interparticle force coming from the dipole-dipole interaction and given by
minus the gradient of (1). For two particles α and β the force on α will be:
a 4
FαI = 12πμ0 μf a2 β 2 H02 (2 cos2 θαβ − sin2 θαβ )er + sin 2θαβ eθ (4)
r
The meaning of the different vectors is defined in Fig.1. For two spheres
placed side by side (r = 2a, θαβ = 90o ) we have FαI = fd = −(3/4)πμ0 μf a2 β 2 H02
which is repulsive as expected; we shall take −fd as the scaling factor of the
interparticle force. The last term in (3) is the Brownian random force which
scales as kT /a. In general we can neglect the inertial force (the time needed
to reach a constant velocity: τ = m/ξ is 1.7μs for an iron particle of radius 1
micron in water; it is much smaller than the other characteristic times) so we
can neglect the left hand side of (3). Dividing all the terms of (3) by 6πμγ̇a2
and rearranging we obtain:
I B
(v − v 0 ) F F
= + + F ext (5)
γ̇a Mn Pe
Magnetorheology 205
The brackets mean that the forces have been divided by their scaling factor.
Mn is the Mason number:
6πμγ̇a2 8μγ̇
Mn = − = (6)
fd μ0 μf β 2 H02
It expresses the ratio of shear to magnetic forces. Note that in other papers
the definition can differ by a multiplicative factor. P e = 6πμγ̇a3 /kT is the Peclet
number and expresses the ratio of shear forces to Brownian forces. For particles
larger than one micron and reasonable shear rates, the Peclet number is large (for
instance Pe=4.5·106 for a particle of radius one micron in water with γ̇ = 1s−1 )
and the Brownian force can be neglected; nevertheless we have to keep in mind
that the influence of the Brownian motion can still be important even at quite
high Peclet number.
The dimensionless equation of motion depends on two quantities (actually
the three quantities defined above are related by M nλ = 2 P e/3) so, for a given
suspension, all the trajectories and hence all the properties-and in particular
the viscosity - will be the same for the same values of M n and λ. Of course
this equivalence only applies for systems of particles starting in the same initial
conditions, that is to say with the same volume fraction Φ and the same initial
configurations. This last point is usually not critical if we are only interested
by equilibrium properties, so we can say that for monosized hard spheres with
particles having the same magnetic permeability, the viscosity (normalized by
the one of the carrier fluid) will depend only on three quantities which are Φ,
M n, λ.
The efficiency of an MR fluid is firstly judged through its yield stress, τy ,
which measures the strength of the structure formed by the application of the
field. The restoring force per unit surface, which resists to the deformation of the
structure, is given by the derivative of the magnetic energy per unit volume rel-
atively to the strain γ : τ = −dW/dγ. The yield stress represents the maximum
of the stress versus strain: τy = max(τ ) since above a critical shear strain, γc ,
the gel-like structure will break. This definition of the yield stress through the
energy of the field allows to understand why it is easier to get a larger yield stress
with MR fluids than with ER fluids : the vacuum magnetostatic energy density:
μo H 2 for H = 3000 Oe is an order of magnitude larger than the electrostatic
energy density o E 2 for a field E = 3 kV/mm (close to the breakdown field).
206 G. Bossis et al.
It does not mean that ER fluids will never show yield stress as high as in MR
fluids, because it also depends on the efficiency of the polarization mechanisms.
For instance ER fluids having ionic polarizability may show high yield stress if
the ions can remain confined inside or on the surface of the particles. This is an
other issue, let us simply present the state of the art for MR fluids.
2.2 MR Fluids
B = μ0 (M + H) with M = χH (7)
The magnetic induction can be measured with the help of a sensing coil and
of a fluxmeter and is safer to use for comparisons between different fluids because
it is a rather well conserved quantity inside a magnetic circuit. On the contrary
the magnetic field H inside the fluid is not directly measurable; it can be deduced
from the measurement of the field with a Hall gauge placed in an air gap close to
the fluid; we shall come back to this point in section 4.1. This problem as well as
the presence of additives which can modify the state of aggregation or impose a
non zero gap between particles, also the lack of knowledge of the polydispersity,
are factors which make difficult the comparison of the results between different
authors. Density of iron is large (=7.87 g/cm3 ) and a particle of radius one
micron in water has a sedimentation velocity of 5.4cm per hour. In order to
prevent the sedimentation and potential redispersion difficulties, it is usual to
add a gel forming additive. The gel must have a low yield stress in order to flow
easily under a small agitation. Nanometer silica particles can be used in this aim
because they easily form a physical gel at low volume fraction (2-3%) and the
yield stress in the absence of field is only a few Pascal. An other problem-even
more important than sedimentation- is the corrosion of iron particles. It has been
found [25] that under permanent use in a damper valve and in the presence of a
magnetic field, the fluid thickens progressively and can show an increase of the
off state (zero field) viscosity by a factor of 3 after 600000 cycles. This thickening
has been identified as due to the removal of thin oxide fragments from the surface
of the particles. The operating conditions in a damper (high shear rate: 104 to
Magnetorheology 207
105 s−1 , strong attractive forces between the particles in the presence of the field,
possible corrosion due to dissolved oxygen) favor the fragmentation process and
must be prevented by a proper surface protection of the particles. The new fluid
made by Lord Corporation no longer show any appreciable thickening even after
2 million cycles. Progress in the strength of the fluids has been realized by using
materials with a higher saturation magnetization like for instance cobalt-iron
alloys (μ0 Ms = 2.45 Tesla for about 50%Fe, 50%Co) which shows a higher yield
stress than iron: 70kPa instead of 50kPa for H=4000 Oersted [23]. Commercial
fluids available from Lord Corporation have typically yield stresses between 50
and 70kPa for a volume fraction between 0.35 and 0.4 and a field of 300kA/m
(3770 Oersted). The off state viscosity is often higher than 1Pas at shear rate
of 50s−1 . The composition of these fluids should still progress in order to have
a lower off state viscosity and still keeping a good protection against corrosion
and a low sedimentation rate. At last it is worth mentioning that as proved
by R.T.Foister [26] mixing two different sizes of carbon iron particles improves
considerably - for the same volume fraction - the on/off ratio of stresses (they
report an improvement of 2.7 times over the monosized suspension at Φ = 0.55
and B = 1Tesla). This increase is quite understandable since it is known that the
maximum packing fraction is increased by mixing two sizes such that the small
spheres can be placed in the holes between the large spheres; consequently the
viscosity diverges at a higher volume fraction, which means that, at the same
volume fraction, the off state viscosity is lower than the one of a monosized
suspension. More surprisingly they also observe - for the same volume fraction
- an increase of the yield stress in the on state of about 25% for the bidisperse
suspension which has no clear explanation.
Colloidal MR fluids. If iron based particles with radii larger than 1μm, give
fluids with high strength, they have the inconvenience to be abrasive. Since the
shear force, which will push the particles against the walls or against each other,
is proportional to the square of the size of the particles, it would be advanta-
geous to reduce this size, which would reduce both abrasion, sedimentation and
fragmentation. On the other hand a too low value of the parameter λ would
not authorize the formation of clusters, so there is a compromise to find. For a
particle of diameter 0.1μm we need a field of about 50 Oersted in order to have a
value of unity for λ. It seems difficult to go well below this limit but nevertheless
Kormann et al [18] have shown that a suspension of soft ferrites with an average
diameter of about 30nm and a volume fraction of 23% can give a dynamic yield
stress of 2kPa for an induction of 0.2Tesla. If we suppose a permeability of about
2 and so H = 1000 Oersted we obtain λ = 10 which seems to be too low to ex-
plain this quite large effect on viscosity. It is quite likely that the nanometric
particles are partially aggregated even in zero field; it happens quite naturally
since the particles are monodomain and carry a permanent dipole. The aver-
age dipole of these clusters is zero in the absence of a field but the persistence
length of the correlations between the orientations of the dipoles of adjacent
particles can be much larger than the particle diameter and explain the large
208 G. Bossis et al.
change of viscosity with the field (for magnetite the attractive dipolar energy
between two particles at contact overcomes the thermal energy for a diameter
larger than 10nm [27,28]). When the field is turned on, the dipoles align on the
field, the cluster becomes magnetized and, thanks to its larger dimension, will
attract other clusters and form the gel like structure. The zero field viscosity of
this fluid is quite high (about 50Pas at 0.1s−1 [29]) which supports the hypoth-
esis of a zero field aggregation. In any event such a fluid does not sediment and
is not abrasive so it could be useful if not too high yield stress is needed. An
other attempt to use nanoparticles has been presented recently with carbonyl
iron [30]. The particles were obtained from decomposition of vapors of iron pen-
tacarbonyl with an average diameter of 26nm. For a volume fraction of about
16% it gives an average increase of yield stress of 7kPA. The experiment was
conducted with a MR damper and the value of the field is not given so it is not
possible to compare with the ferrite particles but, here too, we see that this fluid
can be interesting for applications. It is worth noting that this nanosized fluid
presents a quite large yield stress at zero field (3.6kPA) indicating a permanent
aggregation. This is not surprising if we suppose that these particles are still
monodomain, then, due to their larger magnetization saturation compared to
ferrite (by a factor of 4) the dipolar interactions will lead to this aggregation.
Concerning the use of colloidal particles it seems more interesting to use
larger particles (for instance 0.1μm) in order to avoid the presence of a single
ferromagnetic domain but still keeping a large enough value of λ.
images drift the chains of particles towards the electrodes and in a shear the weak
point of the structure is in the middle of the cell. On the contrary for magnetic
particles in the presence of a magnetic field and non magnetic electrodes the
chain do not attach on the electrodes and the density of particles is larger at
the center of the cell, then in the presence of the shear the fracture zone is
located on the electrodes. In the presence of both electric and magnetic field
the chain like structure will be more homogeneous throughout the cell without
weak points, hence the synergistic effect. Nevertheless the use of a non magnetic
layer decreases the magnetic force between particles, so it is not clear at all that
the synergistic effect will compensate the decrease of magnetic stress. In [38] the
results are presented for carbonyl iron coated with titanium oxide and a weight
fraction of 40%. The thickness of the titanium layer is unknown but the volume
fraction of iron is in any event larger than 7%. The magnetic stress is found
to be 400 Pa for H = 2000 Oe and close to 1 kPa by adding an electric field
of 1 kV/mm. By comparison for the same field, and without coating, the yield
stress is 5kPa for a volume fraction of 15% [39]. As the proportionality between
the yield stress and the volume fraction is quite well verified in this range of
volume fraction [35], we should expect a yield stress of at least 2.5 kPa for the
magnetic stress alone. It appears that, at least for this fluid, even the synergistic
effect does not compensate the loss of magnetic stress due to the presence of the
titanium oxide layer; progresses can likely be made by decreasing the thickness
of the insulating layer.
The most promising type of MR fluid is simply obtained by using a ferrofluid
as carrier liquid together with carbonyl iron particles. An increase in yield stress
by a factor of four was reported in this situation [40]. The reason for this en-
hancement is easy to understand since the force between two particles, when
mediated by a fluid of relative permeability μf , is multiplied by this perme-
ability (cf. Eq. (4)). Also, due to local alignment and network formation of the
nanodipoles between the ferromagnetic particles, there is no sedimentation. The
remaining problem for industrial use could be chemical stability of the ferrofluid.
Besides MR fluids, new types of magnetic composites are appearing which
are the solid counterparts of ferrofluid and magnetorheological fluids, namely
magnetic gels and magnetic elastomers. Magnetic gels are chemically cross-linked
polymers network swollen by a ferrofluid [41]. Placed in a gradient of magnetic
field they will deform and change of shape; this ability to elongate and contract
makes them looking like artificial muscle [42]. Nevertheless, if they can develop
large change of shape, it is because their zero field elastic modulus is weak; in
practice the energy which is developed is quite low. The second class: magnetic
elastomers, are made by dispersing iron particles in a polymer and applying a
magnetic field prior cross linking. T. Shiga et al [43] reported an increase in shear
modulus of about 10kPa for a 28% volume fraction of iron particles (diameter
about 100μm) at a strain γ = 0.1 and H=59kA/m. At lower strain and with
carbonyl iron particles of average diameter 3-4μm, Jolly et al [6] obtained a much
higher field induced shear modulus: 0.56 MPa for Φ = 0.3, γ = 0.01, ν = 2Hz
and flux density corresponding to saturation (0.8 Tesla). The investigation of
210 G. Bossis et al.
the same kind of material under traction instead of shear gives comparatively
a still higher increase in Young modulus (0.6MPa for γ = 0.05, Φ = 0.25 and
H = 123kA/m) [8]. The applications of these elastomers for semiactive damping
of vibrations have been patented by several companies and their performance
can still be improved by a better control of the column formation before curing.
This remark introduces the next section which deals with the prediction of the
structures formed by the magnetic particles in the presence of a magnetic field.
has been derived by A. Cebers [52] from the calculation of the free energy of
a perfect gas of chains of spheres. Analogous results have been obtained more
recently with the same model but using a simpler derivation which can be easily
extended to interacting chains. The result for the average number of particles
per chain is [53]:
√
1 + 4y − 1
< n >= y √ with y = Φe2λ (8)
1 + 2y − 1 + 4y
A good agreement was found between the predictions of such model and
experiments on an ER fluid made of monodisperse silica particles [48]. On the
other hand, if we do not look for the equilibrium distribution of chains but rather
for their kinetic of formation it is possible to describe reasonably well the first
stage of the chaining process with the help of the Smoluchowski equation [54].
In a monolayer the aggregation process can be followed with a microscope and
the image analysis shows that the average length follows a power law:< n(t) >=
1 + ktν ; an exponent ν = 0.6 was found for λ = 31 and λ = 0.009 [54] but
a range of exponents (0.37 < ν < 0.6) was also obtained depending on the
experimental conditions [45]. Actually the Smoluchowski equation predicts an
exponent ν = 0.5 for a constant cross section of collision and a mobility of the
aggregate decreasing as the invert of its length [55] (which is not exactly verified
since the mobility of a slender body of length L, along its main axis, decreases
as (ln(L/a) − 0.5)/2πηL [56]).
The second step of the aggregation process is the coarsening of chains into
columns which is still not well understood. We just want to point out a few
results to emphasize the nature of the problems raised by this coarsening process
either in ER or MR fluids. First the field of a periodic chain of dipoles of infinite
length dies off exponentially as cos(πz/a) exp(−π/a) where is the distance in
the perpendicular x-y plane and 2a the period along z. So, except at very short
distances, two infinite chains will not feel each other. Now let us see the situation
for two rigid dipolar chains of finite length. If we consider two parallel rigid rods
(Fig.2a) separated by a distance d and carrying a dipolar density: m’=m/2a, the
interaction energy between the two rods is:
& '
1 m2 1 1 1 1
W = + − −
4πμ0 μf 4a2 2
d2 + zAa 2
d2 + zBb 2
d2 + zBa 2
d2 + zAb
(9)
The energy of interaction is the one of equivalent charges (plus and minus)
located at the end of the rods. If we do not consider very large shifts between
chains, this energy corresponds to a repulsive force, but, except for the very
special case of two chains of the same length and at the same height, there is
also a torque which will rotate the rods. As the chains are close to each other
they will touch during the rotation well before shifting enough to connect end
by end. After two particles of each chain have joined the coalescence occurs
by a “zippering motion” [45] because of the short ranged attractive interaction
between two chains shifted by a radius. Taking L as the order of magnitude of
zBa in Eq. (9) and normalizing the torque K by kT we have K/kT ≈ λad/L2 .
212 G. Bossis et al.
Fig. 2. Chains of spheres aligned by the field (a): Shifted chains at a distance d (b):
Body Cubic tetragonal structure
use dipole images in order to cancel the field on the electrodes; then a chain
filling the gap is a finite chain instead of an infinite one as in the case of electric
field. We have seen in Eq. (9) that such finite chains repel each other, so look-
ing from above we could expect to see the cross sections of individual chains.
Actually there will be a compromise between the coarsening coming from phase
separation and this long range repulsive energy which will prevent the fibers to
join each other. It is worth noting that in ER suspensions the equilibrium state
for high values of λ does not consist of individual fibers but of a unique domain
with a crystal like structure where the particles are placed on a Body Cubic
Tetragonal (BCT) lattice. In this lattice one chain of particles is surrounded by
four chains shifted by a radius (Fig. 2b). Such a structure has been proved to
have the lowest energy [60] and has been observed both by laser beam diffrac-
tion [61] end recently by confocal microscopy [62]. In practice, due to transport
limitation, the size of the domains will remain small. Coming back to the case of
magnetic suspensions, the individual fibers of particles can be easily seen with
a microscope. As their shape can be quite irregular - especially at volume frac-
tion larger than 5% - and also because their apparent size can depend on image
thresholding - it is the average distance between fibers rather than their diame-
ter which is measured by microscope observation [46,63]. The other way is to use
the light scattered by the fibers with the beam (Oz) parallel to the field. It will
give rings in the x-y plane and the position of the first ring will give the aver-
age distance between fibers [64]. As the field induced by equivalent end charges
depends on the length of the aggregates - which is equal to the thickness of the
cell at high λ -, then we expect that the distance, d, between domains will also
depends on the thickness of the cell h. It is quite usual to fit the experimental
results with a power law: d = Chv but it appears that the power can vary from
0.37 [64] to 0.67 [63]. The spreading of the results is mainly due to the fact that
there is no power law between d and h as demonstrated by a model taking into
account the repulsive interaction between the fibers and the surface energy. This
model assumes that there are no particles between the domains, also for the sake
of simplicity only two kinds of domains are considered: ellipsoidal aggregates or
stripes of particles; the first case can approximate the domains created in a uni-
directional field, whereas the second one is appropriated for structures created in
a rotating field or in a shear flow. If we call Ma the total dipole of an aggregate,
we have to distinguish the dipoles which are on the surface of the aggregate from
those which are in the bulk because the Lorentz field (coming from the surface
of a virtual cavity around a given particle) will be different on the surface of the
aggregate: indeed on the surface half of the surrounding particles are missing
and so we can divide the Lorentz field, HL , by a factor of 2. We shall have:
Ma = NB mB + NS mS (10)
with mB = α(H0 + HD + HL )
and ms = α(H0 + HD + HL /2)
moments of a given particle in the bulk of the aggregate and on its surface.
H0 is the external applied field, HD = −ND Ia /μ0 the field coming from the
equivalent charges on the surfaces of the aggregates, HL = Ia /3μ0 is the Lorentz
field, and Ia = Ma /Va the magnetisation inside the domain of volume Va . The
demagnetisation factor, ND , depends on the geometry of the domains [65].The
total magnetic energy per unit volume of the suspension is:
These two equations can be solved numerically for the two unknowns, d/h
and ϕ(or Φa ), and the solution depends on the initial volume fraction Φ and
on λ. An example of the structure obtained in unidirectional field is shown in
Fig.3a; the experimental data for d versus h are plotted in Fig.3b together with
the solution of Eq. (12) for Φ = 0.045 and λ = 273. The agreement is quite
good and, in this range, is close to d ∝ h0.5 . In this derivation the surface
energy appears through the change of Lorentz field, δHL , on the surface; it is
proportional to the square of the magnetization and to a/h since the radius of
the particles determines the thickness of the layer where the field differs from its
bulk value [67].
Some detailed calculation, either with dipolar [68] or multipolar [69] inter-
actions on different lattices have shown that the surface energy can be quite
sensitive to the type of lattice, but the magnetic suspensions are not monodis-
perse and we do not expect a perfect ordered surface so this “homogeneous”
approximation is likely appropriated. If the surface energy becomes negligible,
then the energy depends only on d/h and ϕ and the solution of Eq. (12) no
longer depends on a/h; in other words d increases linearly with h for a given Φ
and λ. Although the equations (12) could be used to predict the critical field
Magnetorheology 215
Fig. 3. Field induced phase separation. (a): top view of the structure. (b): average
distance between aggregates versus the thickness of the cell: ◦ Eq. (12); experiments
for phase separation (the value of λ above which the free energy becomes min-
imum for a value of ϕ different from unity) it is not believed to give the right
answer because, as stated above, these equations skip the existence of the in-
termediate phase made of a gas of chains. To our knowledge such a theory is
still missing since theories [70,71,49] rely on the hypothesis of a transition from
a diluted isotropic phase to a condensed one, and use the osmotic pressure of a
hard spheres liquid instead of the one corresponding to ellipsoids or rods.
Before ending with field induced structures it is worth saying a few words
about structures in a rotating magnetic field. Let us consider a magnetic field
rotating in the x-y plane: H0 = H0 (ex cos ωt + ey sin ωt) at a frequency high
enough such that the dipolar force between two particles can be averaged over a
period without significant change of the interparticle distance. Then the interac-
tion energy corresponding to two dipoles mα = mβ = m(ex cos ωt + ey sin ωt)
rotating in phase with the field is still given by Eq. (1) whose average on a period
gives: W = m2 /2r3 (3 cos2 θ − 1) where θ is the angle between Oz and r (cf. Fig.
1). This energy is negative for θ = π/2 and positive for θ = 0. It means that the
particles will tend to gather into disks whose axis will be along Oz; these disks
will also tend to separate from each other along the z axis due the repulsive force
for θ = 0. This disk shaped structure was observed with dielectric particles of
diameter 10μm in a rotating electric field [72]. In the case of a magnetic colloidal
suspension, the experiment made in a thin cell with the rotating field perpen-
dicular to the plane of the cell shows a set of stripes which are the projections
of the disk-shaped aggregates (Fig. 4) [73]. The thickness of these stripes being
an order of magnitude larger than the diameter of the particles it seems that
there are composed of many sheets of particles. If the field is rotating at a low
enough frequency, chains of particles have enough time to form and rotate. This
regime can give interesting informations on the kinetics of chain formation and
chain rupture [74]. In the following section we shall see that in some cases the
information on structure can help to explain the rheology.
216 G. Bossis et al.
Fig. 4. Structures in a rotating field. (a): geometry; (b): top view of stripe structure
Fig. 5. Magnetic circuits for rheometry. (a): parallel plate geometry; (b): cylindrical
Couette geometry
where Bg is the field measured with a Hall probe placed between the poles besides
the magnetic fluid. From Eq. (13) we can see that the induction measured besides
the fluid can be quite different from the one inside the suspension and that the
correction can only be done if we know the permeability of the suspension. A
way to overcome this problem is to place the Hall probe between the suspension
and the yoke in order to measure directly Bs . Also with two Hall probes placed
as shown in Fig. 5a it is possible to measure simultaneously the induction inside
the suspension and its permeability by the mean of Eq. (13). Lastly, if the field is
imposed only with an external coil, the induction is practically the one measured
in the absence of the suspension (and the field inside the suspension is H =
H0 /μs (H)).
The parallel-plate or cone-plate geometries are widely used to study the rhe-
ology of magnetic suspensions. In the cone plate geometry the shear rate is
constant inside the suspension but the variable thickness can induce a variable
structure as discussed in the preceding section. In the parallel plate geometry the
gap is constant and easy to change, allowing to get rid of slipping velocities on
the plates but on the other hand the shear rate is not constant (since γ̇ = ωr/h).
For a fluid characterized by a Bingham law: τ = τy + η γ̇ the torque recorded by
the rheometer should be related to the shear rate γ̇R = ωR/h on the rim of the
disk by:
R
2πR3 πR3
T = rτ (r)2πrdr = τy + η γ̇R (14)
3 2
0
The software of the rheometer will use the apparatus constant 2/πR3 to relate
the stress to the torque independently of the rheological law, so multiplying (14)
by this constant and extrapolating at zero shear rate we obtain τ = 4τy /3. We
see that the apparent yield stress given by the rheometer in a parallel plates
geometry is overestimated by a factor 4/3. This problem does not exist with
a cone-plate geometry neither with a sliding plate geometry [77] but this last
technique only gives the static yield stress. In the parallel plates geometry it is
218 G. Bossis et al.
Also the use of two different thickness allows to check the existence of a
slipping velocity on the wall and to correct it if any. The determination of the
yield stress, τy , as a function of the induction is the main property of a MR
suspension. Experimentally the static yield stress is obtained by extrapolating
at zero shear rate the stress versus logarithm of shear rate curve (there must be
data until 10−3 s−1 ) or by detecting the change of slope on a stress-strain curve.
But the yield stress is not a well defined quantity and, in many cases, the yield of
a suspension does not represent a bulk property of the fluid but the interactions
between the wall and the particle [77,35]. A more reliable quantity is obtained by
1/2
fitting the rheogram with a Bingham law or a Casson law: τ 1/2 = τd + (η γ̇)1/2
or Hershel-Bukley: τ = τd + (K γ̇)p . The values of τd obtained by fitting these
different laws are quite similar and represent what is usually called the dynamic
yield stress. This stress can be viewed as the one which is needed to continuously
separate the particles against the attractive magnetic forces in the low shear limit
[20,79].
Fig. 6. Modelling the yield stress. (a) affine deformation of a chain; (b) gap between
two particles; for < δ, Hg =Ms
permeability [83]. The effect of saturation on the yield stress can be described
by a simple model which can be applied both to ER [84] and MR [82] fluids.
In this model the permeability of the particles is high enough to neglect the
field Hi inside the particles. Then we can distinguish two domains inside the
gap between two particles (Fig. 6b): the pole region with < δ ( is the polar
coordinate) where the field is given by the saturation magnetization: Hg = Ms
and the other with < δ where the field is given by Hg = H(a + 0.5w)/h()
where H is the average field in the suspension, a, the radius of the particles
and h() the distance between the plane of symmetry and the surface of the
sphere: h() ∼
= 0.5 ∗ (w + 2 ) with w the minimum gap between the two spheres
and = w/a. The distance δ is obtained by equating Hg and Ms for = δ. If
H/M s 1 the radial force between the two spheres which is the integral of the
field on the plane separating the two spheres can be developed as:
a
μ0 H H
Fr = (Hg − H) 2πρdρ ≈ πa
2 2
μ0 Ms2 − + πa2 μ0 Ms2 (16)
2 Ms 2 Ms
0
The first part of right hand side of Eq. (16) corresponds to < δ and the
second part to > δ. Knowing the force between the two spheres, the shear
stress as a function of the strain γ = tg(θ) ≈ sin(θ) is given by:
Fr 3 Fr γ
τ (γ) = N 2
sin(θ) cos2 θ = Φ 2 (17)
L 2 πa 1 + γ 2
The quantity N/L2 is the number of chains by unit surface and Φ is the
volume fraction of solid particles. The cos2 (θ) comes from the projection of the
field squared on the unit vector joining the two spheres and the sin(θ) of the
projection of
the radial force
on the direction of shear. From Eqs. (16), (17) and
using = 2 1 + γ − 1 we obtain an expression for the shear modulus as a
2
function of the strain which is valid for γ 1. The slope of this curve for γ = 0
220 G. Bossis et al.
gives the shear modulus and the maximum for γ = γc gives the yield stress which
are respectively:
Fig. 7. Theories for yield stress (a) stress-strain curve for chains of steel spheres ◦ ex-
periment; finite elements; Eqs. (16)-(17) (b) normalized yield stress versus volume
fraction for silica spheres (diameter 40-50μm) at low field domain (10-30 kA/m); solid
curve: experiment; dashed curve: theory
both for saturation yield stress [24,15] and shear modulus [6]. Other models
based on the same assumptions of a saturation zone give similar predictions
[82,91].
Low permeability α < 5
An other quite different situation occurs for low values of the ratio α = μp /μf .
A way to obtain this situation is to use a mixture of silica spheres in a ferrofluid.
Then we have a suspension of magnetic holes (α < 1) whose permeability for
different volume fractions well follow the Maxwell-Garnett law [35]. In this case
the dipolar approximation can be used and the interparticle forces being long
ranged, the total magnetic energy is no longer sensitive to the change of the inter-
particle distance but rather to the deformation and rotation of the macroscopic
aggregates in the field. If the magnetic moment of the anisotropic medium can
be obtained as a function of the shape and orientation of the mesostructure then
the energy and the magnetic stress can be derived since we have: τ = 2V 1
H0 ∂m
∂γ
z
with mz the component of the magnetic moment of the sample along the direc-
tion of the external field. The aggregates formed in the presence of the field are
either cylinders or stripes of internal volume fraction Φa . The volume fraction
ϕ = Φ/Φa represents the part of the space occupied by the aggregates. If H is
the field inside the suspension, then the predicted yield stress is [81,92]:
√
τy 3 3 ∗ 2 ϕ(1 − ϕ)
= (μs ) (20)
μ0 μf H 2 16 Cs + μ∗s (1 − ϕ)
with Cs = 1 if the particles are gathered into stripes and Cs = 2 if they form
cylinders. The permeability μ0 μf is the one of the suspending fluid (here the
ferrofluid) and μ∗s = μs /μf − 1.
A similar derivation was proposed independently for stripes, with a slight
difference on the analytical result [93,81]. We have measured the dynamic yield
stress by extrapolating at zero shear rate the stress versus shear rate curve, as-
suming a Bingham behavior τ = τy + η γ̇. The comparison of the normalized
222 G. Bossis et al.
experimental yield stress with the one predicted by Eq. (20) with Cs = 2 (cylin-
drical aggregates) is shown in Fig. 7b for different volume fractions and a field
H = 7 kA/m. We see that Eq. (20) well captures the experimental behavior,
especially taking into account that there are no free parameters (the internal
volume fraction, Φa , has been set to 0.69 which is the volume fraction of a body
cubic tetragonal structure). Besides the quantitative agreement it is worth not-
ing that the yield stress presents a maximum with the volume fraction; this is
because the difference in volume fraction between the aggregates and the effec-
tive medium is decreasing with the volume fraction and the shear stress is only
a function of this difference. On the contrary for particles of high permeability,
the stress is mainly sensitive to short range interparticle forces and increases
linearly with Φ.
More recently this approach for a striped mesostructure was extended to
cover both low and high values of α [94]. The permeability of the stripes is
obtained from a formula valid for a cubic lattice of particles whatever α, and the
internal permeability of the particles μp is approximated by a Frolich-Kennelly
law in order to take into account the magnetic saturation.
Instead of simple chains arranged on a network (cubic or bct) some more
complicated arrangements of chains have been studied either with dipolar ap-
proximation or with multipolar development which should be considered for high
permeabilities [80,81,95,96]. The main point to remember is that the yield stress
is not sensitive to the structure itself (single chains, double chains, planes or
double planes give approximately the same values) but to the way it deforms.
Besides the affine motion with equally spaced spheres we can consider a model
with the formation of a single gap or with a single gap but starting at non zero
strain. This last case where, due to extra particles, the structure can expand
before breaking, gives a higher yield stress [32]. An other point which is still de-
bated is relative to the position where the ruptures take place. In a chain model,
due to the finite size, the force needed to separate two particles inside the chain
is weaker at its extremity and the first rupture should take place close to the
walls. For ER fluids with dipole images it is not so intuitive but still verified
[97]. This is true if each extremity of the chain sticks to the wall, but if it slips
and rotates the hydrodynamic tension being maximum at the center it will likely
breaks at the center. Before breaking, the suspension can also slip on the walls
at a lower stress which will depend on the roughness of the walls and also on
the pressure exerted by the magnetic particles on the wall in the presence of the
field. In this case a frictional shear stress (which is the normal stress times the
friction coefficient) is observed [35]. Also it has been found that dense aggregates
can stand high normal stresses obtained by moving a wall in the presence of the
magnetic field. In this case very high frictional shear stresses following the law
τye = τy + kNe where τy is the usual yield stress, Ne the external stress (up to
2M P a) and k a friction coefficient close to 0.25 [98].
This presentation holds only for non Brownian particles (λ → ∞ ). For finite
values of λ the Brownian motion will play a role to weaken the structure in
two different ways: by decreasing the length of the aggregates and by generating
Magnetorheology 223
fluctuations of the gaps between the particles inside the chains, which will lower
the rupture force. The first effect can be accounted for with the help of the theory
leading to Eq. (8) or a similar one [99]. Then the only aggregates participating
to the yield stress are those which have a length larger than the thickness of
the cell [100].This model ignore the importance of the fluctuations of positions
inside the chain which can considerably decrease the force needed for the rupture
of a chain. It is possible to estimate the average distance between two particles
inside a chain with the help of the Boltzmann distribution, but the breakdown
of the percolated network is not sensitive to the average distance between the
particles but rather to the largest fluctuation inside the chain. A Monte-Carlo
calculation shows that the largest gap inside a chain of 200 particles at λ = 300
can be an order of magnitude greater than the average one and can explain the
lower yield stress experimentally observed when λ is decreased [33]. On the other
hand when a mechanical stress is imposed, the corresponding energy should be
incorporated in the Boltzman distribution as is usually done in the standard
theory of viscosity of simple liquids (the probability for a particle to jump from
a potential well to an other is e−(Ud ±vτ )/kT with v the volume of a unit cell and τ
the applied shear stress; the + sign being for the direction of the applied stress).
This approach is used directly in [101] and incorporated in a chain model in [102];
no direct comparison with experimental results concerning the yield stress are
done but rather concerning the Mason number dependence which is the subject
of the next section.
two equations relative to the torque and to the radial force give a solution for the
critical angle which does not depend on the size of the chain, but which increases
if we consider multipolar forces rather than dipolar ones [35]. It is worth noting
that this approach gives a Bingham law with a yield stress τd = Cμ0 μf β 2 H 2 ,
where C is a constant of order one. The order of magnitude of this dynamic yield
stress is the same as the one obtained in a dipolar approximation and taking
into account multipolar interactions will increase the critical angle well above
45o which will still decrease the yield stress. The failure of the rotating chain
model in predicting the dynamic yield stress for strong fluids could come from
the interactions between aggregates which prevent them to reach the critical
angle. These interactions are not directly introduced in the models.
Still it is the interaction between chains which leads to mesostructure for-
mation (usually stripes of particles in the plane defined by the field and the
velocity) at quite low strain (γ = 0.15) as observed experimentally in oscillating
shear [67] or also in steady shear flow [106,73]. The average inclination of chains
and their rupture is then a process mediated by the interactions between chains
inside dense structures rather than by individual rotation and rupture.
Another model [107] based on an ellipsoidal shape of aggregates also uses a
balance between the hydrodynamic and the magnetic torque to find the equi-
librium angle θ, but the determination of the size of the aggregates is obtained
by minimizing the total magnetic energy (which includes the surface energy).
The result is a viscosity which, unlike the preceding model, should vary as:
ηr = 1 + CM n−2/3 . This theory is restricted to low angles of inclination of the
aggregates, or in other words to low Mason number. As the experiments show
exponents which are between 2/3 and 1 it can be tempting to conclude that
the two models correspond to some limit (for instance ν = 2/3 for low λ and
ν = 1 for λ → ∞). The simulations made with Stokesian dynamics [101] do
not comfort this point of view since, for 10 < λ < ∞, the exponent is between
0.9 and 1 for low Mason numbers (2 · 10−5 < M n < 10−3 ) instead of varying
between 2/3 and 1.
Introducing the parameter λ in the theory in order to model the effect of
Brownian motion on the rheology is a quite difficult task. In [102] the chain
model developed for λ → ∞ is extended by introducing an activation energy
which contribute to break the chains before they have reached their critical
angle. A life time of the chains which depends on λ and M n is introduced in this
way and compared to 1/γ̇ in order to determine if the chains will break before
reaching their equilibrium angle. This model predicts that the normalized field
induced stress should scales as (λ3/2 M n)1/λ but comparisons with experiments
are lacking. A different model using the Eyring approach of viscosity is used in
[101]. This model predicts a plateau at very low shear for the viscosity (which
diverges as e1.57λ ). Above this low shear regime the apparent viscosity should
only depend on the dimensionless shear rate: (γ̇a2 /Dλ)e1.57λ . This prediction is
well verified by numerical simulations in the range 2.5 < λ < 17. For particles
carrying a constant magnetic dipole, a low Mason number theory is proposed in
[105] where the size distribution of chains as a function of λ is the one obtained at
Magnetorheology 225
zero Mason number. Then the distribution function of orientations of the chains
for each length is obtained from a Fokker-Planck equation balancing the fluxes
due to hydrodynamic and magnetic torques and to Brownian relaxation. The
hydrodynamic and magnetic stresses are then averaged over the distribution of
lengths and over the distribution of orientations. Some results for a low value of
λ-independent of field for constant dipoles - compare the increase of viscosity for
field aligned on the velocity gradient and field aligned on the velocity, showing
a maximum with the intensity of the field in the second case. Here too, no
comparisons with experiments are done. Mixing thermal forces and shear forces
to predict the equilibrium distribution of chains as a function of λ and M n is
difficult because the force associated with the flow is not the derivative of a
potential energy. Nevertheless it is the case if we consider a chain of particles
aligned on the direction of elongation in an elongational flow. Then it becomes
possible to add the potential energy of the flow: Uf l = −1/2ξEij ri rj with ξ the
Stokes coefficient, r, the separation vector between two particles and Eij the
velocity gradient tensor. The use of the thermodynamic theories is then possible
and the evolution of the chain length as a function of λ and M n is given in [108].
The knowledge of viscoelastic properties of MR fluids such as the storage
and loss moduli, G , G is quite important for damping applications but few
systematic studies are reported in the literature. As was first pointed in [109]
the use of a dimensionless frequency which is equivalent to a Mason number in
oscillating flow, and of a modulus normalized by the energy density of the field
allows to collapse the data obtained at different fields on the same master curve.
This is verified by experiments on MR fluids if λ is high enough [106], otherwise,
as for steady flow, Brownian forces introduce an other time scale. The moduli G
and G do not depend markedly on frequencies whereas they strongly decrease
with the strain [87]. The viscoelastic theories use the same approach as for steady
flow: the aggregates are modeled by independent chains or ellipsoids. A kinetic
chain model taking into account aggregation due to attractive dipole forces and
rupture caused by hydrodynamic forces was proposed [110] by introducing a
phenomenological equation for the evolution of the chain length. On the other
hand the solution of the equation of motion of an ellipsoid in an harmonic shear
flow - with a fixed length depending on λ - is an other possible approach [111].
A comparison of these models with experimental data on G , G is lacking.
Most of the models are based on the description of individual aggregates and
ignore the effect of mesostructures on magnetorheology. This effect is obvious
at low Mason number where it explains that the first stress-shear rate curve,
obtained under steady shear flow after applying the field on an homogeneous
suspension, is different from the second one (Fig. 8): the first flow curve is ob-
tained starting from a mesostructure formed of cylinders, whereas the second
and subsequent curves are obtained starting from a sheet like structure [106].
This sheet like structure gradually disappears when the Mason number is in-
creased but at a Mason number around unity a new striped structure appears
associated with a jump in viscosity [112]. This flow induced phase separation is
due to the combination of the rotation of transient aggregates of particles in the
226 G. Bossis et al.
Fig. 8. Stress versus shear rate for suspension of magnetic polystyrene particles
(d=0.5μm); applied field H=29 kA/m. The top-left photo corresponds to the struc-
ture before shearing and the bottom photo to the structure after shearing
5 Conclusion
Careful experiments made with well characterized magnetic suspensions are still
too scarce in order to check existing rheological models, especially for fluids with
high saturation magnetization. It would be very useful for this purpose to syn-
thesize well monodispersed magnetic particles in the range 0.1 to 1μm where it
is possible to explore the whole range of λ without serious problems with sedi-
mentation. Progresses in the long term stability of these fluids are more expected
than a large increase in yield stress, except may be in specially confined geome-
tries. Besides applications using viscosity control which are reviewed elsewhere
[15,16] let us point out that the use of rotating fields or combined electric and
magnetic fields to arrange colloidal particles in special structures, has also very
promising applications in the domain of nanotechnologies.
References
1. W.M.Winslow, J. Appl. Phys. 20, 1137 (1949)
2. J.Rabinow: AIEE Trans. 67,1308 (1948)
Magnetorheology 227
1 Introduction
Ferrofluids have been used in medicine since the 1960es for e.g. the magneti-
cally controlled metallic thrombosis of intracranial aneurysms and magnetically
guided selective embolization of the renal artery in case of a renal tumor. Per-
manent obliteration have been expected by producing an artificial thrombus in
the target artery [1,2]. Iron catheters manipulated by an external magnetic field
are in use since 1951. In literature the successful diagnosis of a congenital heart
disease by the use of a magnetically guided catheter is described [3]. Another
system using the magnetic guidance is the magnetic implant guidance system for
stereotactic neurosurgery. A small magnetic NdFeB capsule within the brain has
been moved from outside the body by six independently controlled coils. Cer-
tainly it could be used to deliver radioactivity, heat or chemotherapeutic drugs
to a tumor in the brain [4]. Magnets have been also used to extract swallowed
needles and nails or to remove iron objects from eyes or from other body parts
[5]. Ferrofluids are also used as a contrast agent for MRI. Superparamagnetic
iron oxide particles are a new class of MR contrast agents that have been shown
to significantly increase the detectibility of hepatic and splenic tumors. No acute
or subacute toxic effects were detected by histological or serologic studies in
rats and beagle dogs who received a total of 3000 μmol Fe/kg, 150 times the
dose proposed for MR imaging of the liver. These results indicate that ferroflu-
ids are fully biocompatible potential contrast agent for MRI [6]. These particles
are selectively taken up by cells of the mononuclear phagocytosing system (e.g.
Kupffer’s cells in the liver and macrophages in the spleen, lymph nodes, or bone
marrow), depending on the particle design (e.g. coating or size). The value of
these particles in the diagnostic evaluation of liver and spleen tumors has already
been shown in animal experiments and clinical studies [7,8]. Ultrasound, CT, and
MR imaging do not allow reliable differentiation between hyperplastic and tu-
morous lymph nodes or detection of small metastases in normal-sized lymph
nodes. Superparamagnetic iron oxide particles are also used for MR lymphog-
raphy for the detection of lymph node metastases. In tumor-bearing rabbits,
different degrees of metastatic displacement of lymph nodes were discernible,
and even small metastases (3 mm in diameter) could be visualized [9].
Magnetic properties of ferrofluids can be combined with targeting capability
of affinity ligands for specific cell targeting and separation. Magnetic cell sorting
is employed for both diagnostic and preparative applications in detection and
isolation of various cell subpopulations including damaged and diseased cells.
Attachment of monoclonal antibodies (immunomagnetic cell separation) and
other specific ligands to magnetic particles enable the separation of different
cell populations with relatively simple equipment, and the technology provides a
relatively rapid and convenient method for processing large quantities of different
cell subpopulations [10].
Ferrofluids are also an important subject in the development of an im-
plantable artificial heart. A ferrofluidic actuator directly drives magnetic fluids
simply by applying a magnetic field to the ferrofluids and does not require a
bearing. Magnetic fluid in a U-shaped glass cylinder was placed in an air gap of
a solenoid. When a magnetic flux of 0.32 T was applied to the interface of the
ferrofluid and air, the ferrofluid was displaced and a pressure was obtained [11].
Ferrofluids coated with starch polymers can be used as biocompatible carriers
in a new field of locoregional tumor therapy called “magnetic drug targeting”
[12,13,14]. It is of great importance to target selective the antineoplastic agent
to its tumor target as precisely as possible, to reduce the resulting systemic toxic
side effects from generalized systemic distribution and to be able to use a much
smaller dose, which could further lead to a reduction of toxicity.
Magnetic Drug Targeting 235
2.1 Ferrofluids
The ferrofluids used in the experiments were obtained from Chemicell (Berlin,
Germany; German patent application no. 19624426.9) and consisted of a biocom-
patible colloidal dispersion formed by wet chemical methods from iron oxides and
hydroxides to produce special multidomaine particles [15]. These ferrofluids are
covered with hydrophilic starch polymers coupled with endstanding functional
groups (e.g. phosphate) allowing ionic binding to many therapeutic drugs. The
hydrodynamic diameter of the whole particle is about 100 nm Fig. 1 [12].
Organic as well as inorganic agents can ionically bound to the functional
group, in this case phosphat. The chemotherapeutic agent mitoxantrone (Novan-
tronr , Lederle, Wolfratshausen, Germany) is ionically bound to the endstanding
groups of the starch coating (Fig. 2).
Mitoxantrone is a synthetic anthracendion that inhibits DNA and RNA syn-
thesis by intercalating in DNA molecules, which causes strand breaks. It is an
anticancer agent used in treatment of e.g. breast cancer, non-Hodgkin’s lym-
phoma and various solid tumors (carcinoma and sarcoma)[16,17,18].
The body surface area and the dose of mitoxantrone were calculated accord-
ing the recommendations in literature [19].
123
2.2 I-labelled nanoparticles
Ferrofluid (MAG-D 14/11) was mixed with 123 iodine. For the calculation of
the activity, a definitive volume of the suspension was measured in the gamma-
counter.
The experiments were performed in tumor-bearing rabbits with a cathe-
ter in the Arteria femoralis under the gamma-camera. The ferrofluid-123 iodine-
suspension was injected into the tumor-supplying artery for 5 minutes, while
the permanent-magnet was focused on the tumor-area. At the same time the
signaldetection of 123 iodine started.
Fig. 4. Dependence of the magnetic flux density on the distance to pole shoe with the
electromagnet [12].
238 Ch. Alexiou et al.
For application of the chemotherapy, the femoral artery was cannulized and
a catheter was placed approximately 2 cm distal to the inguinal furrow. The
mitoxantrone bound ferrofluid was administered slowly over 10–15 minutes.
below a value of 0.05. Between 0.01–0.05 the result was considered significant
and highly significant if below 0.01.
3 Results
3.1 Tumor Volume
In the control group without treatment tumor volume increased and metastases
appeared. The animals of group 1a, treated intra-arterially with 20% FF-MTX,
had a complete tumor remission between the 15th and 36th day following treat-
ment. The animals of group 1b (50% FF-MTX) had a decrease in tumor volume
similar to group 1a. In group 2 (intra-arterial MTX alone, no magnetic field),
lower dosages (20% and 50% of the systemic dose) did not result in tumor re-
mission and enlarged, palpable inguinal lymph nodes were found after 48 days.
At higher doses (75% and 100%), complete remission of tumor occurred.
The two group 3 animals (intra-arterial FF alone with the magnetic field,
at 20% and 50%) demonstrated a progressive increase in tumor volume with
palpable, enlarged inguinal lymph nodes (metastases).
The six animals of group 4 (intravenous injection via the ear vein of 20%
and 50% FF-MTX with magnetic field) showed a slight tumor remission, but
the reduction of volume was not statistically significant in comparison to the
control group (not shown in figures).
The two animals of group 5 (intra-arterial FF-MTX 20% and 50%, without
a magnetic field) showed a discontinuation of tumor growth and no evidence
of metastases, but no remission of the tumor was seen (not shown in figures).
Fig. 5. Group 1a: Effect of i.a. application of FF-MTX [20% of the regular systemic
dose (•)] on relative tumor volume after magnetic drug targeting compared with control
group [group 6 (Δ), control(no treatment)]. Symbols, the median tumor volume; bars,
the maximum and minimum; metastases, onset of metastases; treatment, the day of
treatment (singular treatment)[12].
240 Ch. Alexiou et al.
Fig. 6. Group 1b: Effect of i.a. application of ferrofluids bound to mitoxantrone (FF-
MTX, 50% of the regular systemic dose),(•) on relative tumor volume after magnetic
drug targeting compared with control group [group 6 (Δ), control (no treatment)].
Symbols, the median tumor volume; bars, the maximum and minimum values; metas-
tases, onset of metastases; treatment, the day of treatment (singular treatment) [12].
Fig. 7. Group 2: Effect of i.a. application of MTX - 20% and 50% (♦) and 75% and
100% (•) of the regular systemic dose - on relative tumor volume compared with control
group [group 6 (Δ), control (no treatment)]. Symbols, the median tumor volume; bars,
the maximum and minimum value; metastases, onset of metastases; alopecia, onset of
alopecia; treatment, the day of treatment (singular treatment) [12].
At the time of treatment less then 5% of the animals showed a small necrotic
fraction in the area of the tumor area [12].
The general condition of the control group animals (limited to 2 animals for eth-
ical reasons) worsened during the observation period and the animals developed
Magnetic Drug Targeting 241
Fig. 8. Group 3: Effect of i.a. infusion of FF (50%) alone with magnetic field (). Δ,
control (no treatment). Symbols, the median relative tumor volume; bars, the maximum
and minimum values; metastases, onset of metastases; treatment, the day of treatment
(singular treatment).
Fig. 9. Values of the WBC before treatment (day 0) and on days 3-15 (early period),
18-48 (middle period), and 51-81 (late period)after the respective treatment regimes.
Columns, the median values; bars, the maximum and minimum values. The description
of the bars, i.e. the respective group is mentioned beside the graphs [12].
Fig. 10. Same as Fig. 12 for other groups mentioned beside the graphs.
urine, activity, fur condition) remained normal during the whole 3 month obser-
vation period compared to the physiological data of healthy animals (statement
by Charles River, Sulzfeld, Germany). No significant changes in serum iron or
leucocyte values were seen in this group (Fig. 9).
The urine of one animal in group 2 (50% MTX) showed blue-green discolo-
ration and this animal developed mild alopecia in the region of the digits after
48 days. Both animals with low dose MTX (20% and 50%) had a decrease in
leucocyte values, but this was not statistically significant (p=0.29). Both group 2
animals with high dose (75% and 100%) MTX had temporary blue-green urine
discoloration, as well as a unilateral alopecia (palmar region of the digits to
the knee joint) of the limb in which the tumor was implanted developing after
33 days. This hairless area developed cutaneous inflammation and ulceration,
followed by mild alopecia of the fore limbs and head. The musculature of the
treated limb became atrophic and the circumference was noticeably smaller (3
cm) at the end of the 3 month observation period. There was no marked differ-
ence in the severity of the side effects between the two animals, except for the
fact that the animal with the higher MTX dose (100%) developed the changes
several days sooner. Group 2 animals with 50%, 75% and 100% MTX steadily
lost weight after an initial lag-phase and were underweight at the end of the
observation period (mean value 1800 mg below the lower reference values ac-
cording to the breeder’s statement, Charles River, Sulzfeld, Germany). These
animals became leucocytopenic (≤ 2.95 x103 /μl) in the early phase (highly sig-
nificant drop, p=0.004, Fig. 10), but recovered slightly in the middle and late
periods [12].
None of the animals of groups 3 or 4 showed any significant changes in serum
iron (not shown in figures) or leucocyte counts (group 3: Fig. 9; group 4a and b:
Fig. 9) during the observation period when compared with initial values.
Figure 11 up to 14 show cross sections of a VX2 tumor that was excised just after
treatment with “magnetic drug targeting”. Ferrofluids can be seen as blue pig-
ment (Prussian Blue, Fig. 11, 12, 13) or as brown-black pigment (H.E., Fig. 14).
Immediately after magnetic drug targeting, ferrofluids are evident in the vessels
supplying the tumor and in macrophages. After the 3 month observation period,
no viable tumor tissue was histologically evident in the animal group 1, with only
fibrosis seen in the tumor implantation site. No metastases were found in the
regional lymph nodes or in any other organs. Some FF particles were found in
the spleen of the animals, but none were evident in the liver, lungs, brain or the
implantation site and surrounding musculature and skin. No other macroscopic
or histopathological changes were found in any of the investigated organs [12].
In group 2, the VX2 tumors of the 2 low dose animals were 8.644 mm3
(50% MTX) and 2.497 mm3 (20% MTX) in size with a large area of central
necrosis and viable tumor at the periphery. The 2 animals with high dose MTX
(75% and 100%) had no viable tumor at the implantation site. None of the other
Magnetic Drug Targeting 243
Fig. 11. Cross section of VX2 squamous cell carcinoma of the rabbit immediately after
the magnetic drug targeting stained with Prussian Blue. Ferrofluids are visible in the
vessels of the peripheral area of the tumor (original magnification x 100).
Fig. 12. Cross section of VX2 squamous cell carcinoma of the rabbit immediately after
the magnetic drug targeting stained with Prussian Blue. Ferrofluids are visible in tumor
supplying vessels in the connective tissue of the VX2 tumor (original magnification x
200).
investigated organs in the animals of group 2 (liver, kidneys, spleen, lungs, brain)
had any pathological changes.
The tumors of both animals of group 3 measured 13.324 mm3 and 17.649
mm3 , with a large area of central necrosis and viable tumor at the periphery. No
FF particles were found within the tumor or in the surrounding musculature and
skin. Some FF were found in the spleen. Metastases were found in the inguinal
lymph nodes and liver of both animals. None of the other investigated organs
(kidneys, spleen, lungs, brain) had any pathological changes [12].
Fig. 13. Cross section of VX2 squamous cell carcinoma of the rabbit immediately after
magnetic drug targeting stained with Prussian Blue. Proof of iron in macrophages and
endothelium (original magnification x 1000).
Fig. 14. Cross section of VX2 squamous cell carcinoma of the rabbit immediately
after magnetic drug targeting stained with haematoxilin-eosin. Ferrofluids are visible
intravascular and in a macrophage (Original magnification x 1000).
Fig. 15. (left) MRI of tumorous hind limb after i.a. application of ferrofluids; dotted
circle: tumor region; f: head of the femur.
(right) MRI of tumorous hind limb after i.v. application of ferrofluids; dotted circle:
tumor region; f: head of the femur [12].
Magnetic Drug Targeting 245
Fig. 15 shows the left hind limb (implantation site) of two rabbits receiving
intra-arterial or intravenous 50% FF-MTX, respectively, 6 hours after treatment.
The tumor is situated at the medial portion of the hind limb (dotted circle), and
the concentration of FF is seen by extinction of signal. Fig. 15(left) (intra-arterial
FF-MTX) shows definite extinction of signal and Fig. 15(right) (intravenous FF-
MTX) only a very discrete signal extinction. The area marked “f ” is at the head
of the femur and appears to be hypodense.
The tumor is situated at the medial portion of the hind limb (dotted cir-
cle), and the concentration of FF is seen by extinction of signal. Fig. 15(left)
(intra-arterial FF-MTX) shows definite extinction of signal and Fig. 15(right)
(intravenous FF-MTX) only a very discrete signal extinction. The area marked
“f” is at the head of the femur and appears to be hypodense [12].
123
3.5 Biodistribution of Iodine-Ferrofluids
In order to assess the distribution of the used nanoparticles under the influence
of the magnetic field, 123 iodine was storaged in the starch-cover and was then
injected into the tumor-supplying artery while the magnetic field was focused
on the tumor for 60 minutes.
At the same time, the signal-detection of 123 iodine started. Around the
tumor-area there was a “region of interest” (ROI) determined and the activ-
ity was measured over the time.
The injected whole-body-activity of the rabbit could not be measured with
the gamma-camera, so the activity in the tumor-area was measured at differ-
ent times and was related to the initial activity-maximum (relative activity in
percent of the initial activity).
The scintigraphically detected signal after intra-arterial application has been
shown to be at least twice as much higher in the magnetically focused region
compared to the application without external magnetic field.
Enrichment of 123 I-labelled nanoparticles in the area of interest (VX2 tumor)
after magnetic drug targeting:
The activity in the tumor region scintigraphically measured by gamma-
camera and focused by the external magnetic field was after 30 minutes 38.2
%, after 40 minutes 31.8 %. and after 60 minutes 22.3 % (n=1). Without ex-
ternal magnetic field, the activity was 14.7 % after 30 minutes, 13.0 % after 40
minutes and 11.8 % after 60 minutes (n=1).
4 Discussion
systemic distribution and to be able to use a much smaller dose, which would
further lead to a reduction of toxicity.
Regional chemotherapy via a regional artery administers a more concentrated
dose of the active agent directly into the tumor [20]. The advantage of this
approach is limited, however, by drain-off via the venous blood which limits
exposure time and reduces the overall efficacy.
At present, i.a. delivery of chemotherapeutic agents is approved and well
accepted for treatment of liver metastases [21] and has occasionally been used
for other tumor types also (e.g., inoperable head and neck tumors); but it has
often necessitated complicated, time-consuming operative procedures, including
general anaesthesia [22].
“Magnetic drug targeting” means holding the chemotherapeutic agent due to
the carrier (ferromagnetic particles) at the desired site of activity, thus increasing
efficacy and diminishing systemic toxicity.
The goal of the present study was to concentrate ferrofluids coupled with
chemotherapeutic agent in a desired target area using a magnetic field. The
principle of “magnetic drug targeting” consists of two steps: The first step is
the delivery of the mitoxantrone bound magnetic nanoparticles to the desired
Magnetic Drug Targeting 247
body compartment (tumor), the second is the release of the drug from its carrier
[12,13,14]. An additional helpful factor is that microvascular permeability in
neoplastic tissues is increased (8-fold compared with normal tissue) as is diffusion
(33-fold) [23], which allows chemotherapeutic agents much easier to penetrate
into tumor tissue. The metabolism of ferrofluid particles takes place in the liver
and spleen in analogy to iron metabolism.
In this investigations, the authors found that this approach led to complete
tumor remission with reduced doses of 20% and 50% MTX bound to FF (Figs. 5
and 6). The application was well tolerated by the animals and no signs of toxicity
were detected. On the contrary, intra-arterial infusion of the same doses, 20% and
50%, of MTX alone (group 2, Fig. 7) resulted in no reduction of tumor volume
and the animals developed metastases and suffered from chemotherapeutic side
effects. Only when the dose of MTX alone was increased to 75% and 100%,
a tumor remission was seen (Fig. 7), but this resulted in severe side effects
(alopecia, ulcers, leucocytopenia), as seen in Fig. 10. Intravenous infusion of the
FF-MTX complex was also ineffective since only a slight tumor remission that
was not statistically significant resulted. The same result was observed after
intra-arterially infused FF-MTX without an external magnetic field, since the
tumor remained at the same size, without remission. Thus, the combination of
intra-arterial infusion with a magnetic field was safe, effective, well tolerated by
the animals and was very effective in treating the tumor although the dose of
chemotherapeutic agent was markedly reduced [12,13,14].
At present, intra-arterial delivery of chemotherapeutic agents is approved
and well accepted for treatment of liver metastases [22], and have occasionally
also been used for other tumor types (e.g. inoperable head and neck tumors), but
often necessitate complicated, time consuming operative procedures, including
general anesthesia [22]. Experimentally, Swistel et al. [24] described encouraging
results using intra-arterial chemotherapy for VX2 squamous cell carcinoma. They
achieved complete tumor remission after intra-arterial application of adriamycin
in 4 of 6 animals, whereas intravenous infusion of adriamycin caused severe
toxicity and resulted in complete remission in only two cases.
A potential complication that could arise with the use of FF compounds is
the fact that with larger particles embolization could occur, preventing a suffi-
cient concentration of the chemotherapeutic agent from reaching the tumor. On
the other hand, if the particles are too small, the external magnetic field might
not provide sufficient attraction so that the particles are drawn into the tumor.
The particles used in this study had a size of 100 nm (nanometer). No emboliza-
tion was seen in the vascular system of the tumor. Our histological findings
showing distribution of FF particles throughout the tumor strongly support the
concept that large molecular weight substances such as chemotherapeutic agents
or monoclonal antibodies can be effectively targeted to tumor tissue. In addi-
tion, the fact that the FF alone with a magnetic field failed to cause tumor
remission (Fig. 8) indicates that therapeutic effect resulted from the action of
the chemotherapeutic agent itself, rather than intratumoral embolization by the
particles [12].
248 Ch. Alexiou et al.
The electromagnet used for this study produced a magnetic flux density
of a maximum of 1.7 Tesla which decreased depending on the distance to the
pole shoe (Fig. 4). The magnetic gradient can be seen as a collection of vec-
tors which point in the direction of increasing values as shown in Fig. 3 (yellow
arrows). The arrow sizes correspond to the strength of the magnetic gradient.
Both factors (direction and magnitude) reflect the inhomogenous character of
the magnetic field, which is of key importance for “magnetic drug targeting”. In
previous studies it was suggested that a magnetic field strength of 8000 Gauss
(0.8 Tesla) is sufficient to exceed linear blood flow in the intratumoral vasculature
and allow 100% localization of magnetic carrier containing 20% magnetite [25].
In contrast, Goodwin and coworkers applied magnetic-targeted carriers (MTCs)
intra-arterially in a swine model, focusing a magnetic field of only 250 to 1000
Gauss (0.025-0.1 Tesla, permanent neodymium magnet) to the desired compart-
ments in the liver and lungs. The depth of this MTC targeting was 8 - 12 cm
and the particle size 0.5 μm - 5 μm. With this model, MTCs with a predefined
activity had a concentration of 67% in the liver and 50% in the lung localized
by the magnet [26].
The magnetic field strength with a maximum of 1.7 Tesla used in the present
investigation was the strongest ever applied for magnetic drug targeting. We
achieved a high concentration of FF within the tumor after intra-arterial infu-
sion of FF which was seen by histological (Figs. 11 to 14) and MRI methods
(Fig. 15). The VX-2 squamous cell carcinoma was superficially exposed and had
no migratory motion, as was the case with the liver and lung targets (breathing
fluctuations) in the swine model of Goodwin and coworkers [26]. In addition
these organs lie deeply in the body cavity (8-12 cm from the body surface),
greatly complicating focusing a maximum of the magnetic flux density onto the
tumor area. Two approaches to overcome this problem are possible,
The particles (FF-MTX) used in the present study were 100 nm in size and
have shown good therapeutic results in smaller animals (mouse, rat) as well
[28,29].
The strong magnetic field was very efficacious in combination with these
particles, but further experiments (which we have already begun) should be
performed using marked FF to clarify the optimal magnetic field strength and
particle size [12].
A remarkable feature of using ionically bound pharmaceuticals is that the
low molecular weight substances (molecular weight of mitoxantrone 517) is able
to desorb from the carrier (FF) after a defined time span and can then pass
through the vascular wall or interstitium into the tumor cells. This is important
since once the FF-MTX complex has been directed to the tumor by the magnetic
field, the drug must dissociate in order to act freely within the tumor. As shown
in Fig. 2 MTX desorbs from the FF after 30 min (half life) and so 50% of the
drug is free to act on the tumor after 30 min.
Magnetic Drug Targeting 249
Dextran coated iron oxides have been shown to produce signal loss by MRI
and have been used as a contrast medium for the detection of metastatic lymph
nodes (negative contrast) [30]. We found total signal loss and therefore a very
high concentration of FF by MRI after focusing by the magnetic field (Fig. 15
(left)) [12]. Recent studies have shown that intra-arterial application of radio-
active-labeled, magnetic carriers with an external magnetic field resulted in re-
tention of at least 50% in the target site [31]. In comparison, after intravenous
injection, only very slight signal loss was seen, indicating a very low concen-
tration (Fig. 15(right)). This underscores the advantage of intra-arterial versus
intravenous infusion in magnetic drug targeting.
Biodistribution was studied by the use of 123 iodine-nanoparticles. The scinti-
graphically detected 123 iodine- signal after intra-arterial (artery leading the tu-
mor, femoral artery) application has been shown to be significantly higher in
the magnetically focused region compared to the application without external
magnetic field (Fig. 16). 60 minutes after application there is still twice as much
concentration of 123 iodine in the magnetically focused region compared to the
application without the magnetic field [14].
Previous studies by Bacon et al. concerning FF with a particle size of 0.5-1.0
μm found no acute or chronic toxicity after the intravenous infusion of 250 mg
iron/kg body weight in rats [32] and 1-3 mg iron/kg body weight in humans
has been shown to be safe as well [33]. This agrees with our findings, since FF
infusion was not associated with any signs of toxicity [12,13,14].
Magnetic microspheres loaded with the gamma-emitting radioisotope 90Y
have also successfully been used as a form of radionuclide therapy. In one study,
this compound was maneuvered within the body of a mouse to a subcutaneous
lymphoma, resulting in eradication of the tumor [33]. Magnetic fluids have also
been used for the so-called “magnetic fluid hyperthermia” that has been used to
control the local growth of murine mammary carcinoma [35]. Additional modifi-
cation of the magnetic particles, so that they could bind monoclonal antibodies,
lectins, peptides or hormones could make delivery of these compounds more
efficient and also highly specific. Targeted drug delivery using nanotechnology
opens a new field in oncology science and improves basis science and clinical
practice [36].
Acknowledgements
Supported by the Margarete Ammon Foundation, Munich, and grants from the
Technical University of Munich, Germany.
References
1. Alksne JF, Fingerhut A, Rand R. Magnetically controlled metallic thrombosis of
intracranial aneurysms. Surgery 1966; 60: 212-218.
2. Hilal SK, Michelsen WJ, Driller J, Leonard E. Magnetically guided devices for
vascular exploration and treatment. Radiology 1974;113: 529-534.
250 Ch. Alexiou et al.
21. Link, K.H., Kornmann, M., Formenti, A. et al. Regional chemotherapy of non-
resectable liver metastases from colorectal cancer – literature and institutional
review. Langenbecks Arch. Surg., 384: 344-353, 1999.
22. v. Scheel, J.: Invasive procedures for antineoplastic chemotherapy. In: Naumann
et al. (eds.): Head and Neck Surgery. Stuttgart, New York: Thieme, 1998.
23. Gerlowski, L.E. and Jain, R.K. Microvascular permeability of normal and neoplas-
tic tissues. Microvascular Research 31:288-305, 1986.
24. Swistel AJ, Bading JR and Raaf JH Intraarterial versus intravenous Adriamycin
in the VX2 tumor system. Cancer (Phila.) 53: 1397-1404, 1984
25. Senyei, A., Widder, K., and Czerlinski, C. Magnetic guidance of drug carrying
microspheres. J. Appl. Physiol., 49: 3578-3583, 1978.
26. Goodwin, S., Peterson, C., Hoh, C., Bittner, C. Targeting and retention of magnetic
targeted carriers. J. Magn. Magn. Mater., 194 : 132-139, 1999.
27. Bergemann, C., Müller-Schulte, D., Oster, J., à Brassard, L., Lübbe, A.S. Mag-
netic ionexchange nano- and microparticles for medical, biomedical and molecular
biological applications. J. Magn. Magn. Mater.,194 : 45-52, 1999.
28. Lübbe, A. S., Bergemann, Ch., Huhnt, W., Fricke, T., Riess, H., Brock, J. W.,
and Huhn, D. Preclinical experiences with magnetic drug targeting: tolerance and
efficacy. Cancer Research, 56: 4694-4701, 1996.
29. Lübbe, A. S., Bergemann, Ch., Riess, H. et al. Clinical experiences with magnetic
drug targeting: a phase I study with 4’-epidoxorubicin in 14 patients with advanced
solid tumors. Cancer research, 56: 4686-4693, 1996.
30. Taupitz, M., Wagner, S., Hamm, B., Dienemann, D., Lawaczeck, R. and Wolf, K.J.
MR Lymphography using iron oxide particles. Detection of lymph node metastases
in the VX2 rabbit tumor model. Acta Radiologica, 34:10-15, 1993.
31. Widder, K.J., Senyei, A.E. and Scarpelli, D.G. Magnetic microspheres: a model
system for site specific drug deleivery in vivo. Proc. Soc. Exp. Biol. Med. 58: 141-
146, 1978.
32. Bacon, B.R., Park, D. D., Saini, S., Groman, E. V., Hahn, P. F., Compton, C. C.,
and Ferrucci, J. T. Ferrite particles: A new magnetic resonance imaging contrast
agent. Lack of acute or chronic hepatoxicity after intravenous administration. J.
Lab. Clin. Med., 110: 164-171, 1987.
33. Rummeny, E., Weissleder, R., Stark, D. D., Elizondo, G. and Ferrucci, J. T. Mag-
netic resonance tomography of focal liver and spleen lesions. Experiences using
ferrite, a new RES-specific MR contrast medium. Radiologe 28: 380-386, 1988.
34. Häfeli, U.O., Pauer, G.J., Roberts, W.K., Humm, J.L. and Macklis, R.M. Mag-
netically targeted microspheres for intracavitary and intraspinal Y-90 radiotharpy.
In:U. Häfeli and W. Schütt (eds.), Scientific and clinical applications of magnetic
carriers, pp 501-516, New York and London: Plenum press, 1997.
35. Jordan, A., Scholz, R., Wust, P. et al. Effects of magnetic fluid hyperthermia
(MFH) on C3H mammary carcinoma in vivo. Int. J. Hyperthermia, 13: 587-605,
1997.
36. Partridge, M., Phillips, E., Francis, R., Li, S.R. Immunomagnetic seperation for
enrichment and sensitive detection of disseminated tumour cells in patients with
head and neck SCC. J. Pathol., 189:368-377, 1999.
Magnetic Unit Systems
Historically numerous different magnetic unit systems are used. In this volume of
Lecture Notes in Physics mainly the SI-system (chapters by Rosensweig, Blums,
Zubarev, Thurm & Odenbach and Bossis) and the Gaussian system (chapters
by Shliomis and Morozov & Shliomis) are employed.
To reduce the problems with the comparison of statements in the various
chapters, the major units and their conversion are given in the table below.