You are on page 1of 31

Chemical Engineering Science 61 (2006) 1627 – 1657

www.elsevier.com/locate/ces

Magnetoviscous control of wall channeling in packed beds using magnetic


nanoparticles—Volume-average ferrohydrodynamic model and
numerical simulations
Faïçal Larachi∗,1 , Damien Desvigne1
Department of Chemical Engineering, Laval University, Qué., Canada G1K 7P4

Received 26 June 2005; received in revised form 20 September 2005; accepted 27 September 2005
Available online 16 November 2005

Abstract
A theoretical framework for analyzing ferrofluid magnetoviscosity in anomalous non-magnetic porous media is constructed. Upscaling the
pore-level ferrohydrodynamic model resulted in a volume-average model with 15 closure equations. A simplified zero-order three-dimensional
axisymmetrical model for Forchheimer non-turbulent flows is cast for steady-state isothermal incompressible Newtonian ferrofluids through a
Müller porous medium of spherically shaped grains and subjected to external constant gradient (bulk-flow oriented) magnetic fields, ferrofluid
self-consistent demagnetizing field and induced magnetic field in the solid. The model reveals interesting potential chemical engineering
applications of ferrofluids in contexts where wall flow maldistribution due to low column-to-particle diameter ratios exists. Judicious magnetic
field arrangement mitigates wall channeling by inflating wall flow resistance through magnetoviscothickening and Kelvin body force density
rerouting thus a fraction of wall flow towards bed core. Depending on whether the Kelvin force density acts against or in favor of the flow
direction, the pressure drop is, respectively, larger or smaller than that without magnetic field, though, in all cases, magnetoviscothickening is
the prevailing mechanism.
䉷 2005 Elsevier Ltd. All rights reserved.

Keywords: Ferrofluid; Magnetosviscosity; Porous medium; Volume-average ferrohydrodynamic model; Kelvin body force; Spin–vorticity coupling

1. Introduction gives rise to strong and coherent magnetization which, for


motionless ferrofluids, aligns along the direction of the total
Ferrofluids are non-naturally occurring magnetic liquid magnetic field, and which obeys the equilibrium Langevin
suspensions obtained by seeding surfactant-wrapped single- magnetic-field-locked magnetization law. In this study, we will
domain superparamagnetic nanoparticles in appropriate organic make the assumptions that the Brownian relaxation outweighs
or aqueous carrier matrices (Rosensweig, 1997; Rinaldi and the Néelian relaxation (i.e., magnetic moment spatially locked
Zahn, 2002a; Odenbach, 2003). Such stabilized ferrocolloidal to nanoparticle) and that the magnetic dipole interaction energy
suspensions exhibit a number of intriguing behaviors when is unable to oppose neither the deconstructing thermal agita-
subjected to external—uniform, spatially inhomogeneous or tion nor shear flow, thus maintaining the nanoparticles afloat
oscillating—magnetic fields. Owing to a giant single-domain and unassociated. In addition, should the ferrofluid motion-
magnetic moment, ca. 104 Bohr magnetons, each nanoparti- less state be restored, the non-zero magnetization relaxation
cle stands as a small permanent magnetic dipole. Collective time will impede instantaneous directional readjustment of
response of these nanoparticles to magnetic field stimulation magnetization vector towards equilibrium.
When ferrofluids are brought to motion in shear flows,
∗ Corresponding author. Tel.: +1 418 656 3566; fax: +1 418 656 5993. misalignment between “dynamic” magnetization M and to-
E-mail address: faical.larachi@gch.ulaval.ca (F. Larachi). tal magnetic field H vectors arises due to an asynchrony
1 Current address: Total S.A.—Centre Technique et Informatique de Lyon, between the nanoparticle spin  and the ferrofluid vorticity,
2 ∇ × v (Rosensweig, 1997). For illustration purposes, the total
1
ARKEMA, Chemin de la Lône-BP32 69492 Pierre-Bénite cedex, France.

0009-2509/$ - see front matter 䉷 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2005.09.019
1628 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Fig. 1. Illustration of the magnetoviscosity effect: (a) magnetoviscothickening, (b) magnetoviscothinning.

magnetic field vector and the fluid vorticity vector are assumed tostatic and hydrodynamic torques. This cancels the internal
orthogonal as depicted in Fig. 1, and in first approximation, angular momentum equation and eliminates spin coupling with
both angular acceleration and diffusive couple dyadic contribu- linear momentum balance.
tions are ignored. Hence, this spin–vorticity asynchrony gives Pioneering works by the Soviet school provided a large body
rise to a mechanical torque that will tend to pull the dynamic of knowledge confirming the special behavior of magnetic flu-
magnetization vector out from the direction of equilibrium (or ids experiencing inhomogeneous or oscillating magnetic fields
equivalently, the direction of the local total magnetic field at in Hagen–Poiseuille capillary flows or in Darcy regime with
motionless state). Non-collinearity between dynamic magneti- very restrictive conditions on the filtration law in porous me-
zation and magnetic field in return results in a counteracting dia (Ivanov and Taktarov, 1990; Taktarov, 1980, 1981, 1983).
magnetostatic torque, which will try to pull in the magnetiza- Other recent works of porous media ferrofluid flows highlights
tion vector back to its equilibrium position. The mechanical interesting geoapplications of migration of magnetic fluids in
torque is proportional to a vortex viscosity, which proportion- remote locations which preclude direct physical control such
ates the magnetostatic torque response. Two magnetoviscous as for environmental subsurface contexts (Borglin et al., 2000;
behaviors, as a consequence, could arise depending on mag- Oldenburg et al., 2000).
netic field and ferrofluid vorticity magnitudes and mutual ori- We propose to explore in this paper potentially new ap-
entations with respect to wall. plications of magnetoviscosity in chemical engineering. The
Magnetoviscothickening, a magnetically driven apparent in- demonstration example will highlight the control of wall chan-
flation in local ferrofluid suspension viscosity, occurs when neling in porous media exhibiting low column-to-particle di-
the fluid vorticity and the magnetized nanoparticle spin are in ameter ratio. It is well known that such low-ratio columns
contrarotation or are in corotation with fluid vorticity faster may occasion severe maldistribution due to large radial per-
than the nanoparticle spin velocity (Fig. 1a). The retarding ef- meability contrast with preferential flow short-circuiting along-
fect by the nanoparticle spin velocity results locally in an in- side the high-permeability wall area. The challenge will be to
creased dissipation in the flow and, eventually, a slow down in identify which configurations are propitious to creating suffi-
fluid linear velocity. Magnetoviscothickening has been reported cient magnetoviscosity near the vessel wall to hammer the by-
for Hagen–Poiseuille and turbulent pipe flows with steady and pass and to reroute maximum fraction of fluid flow towards
oscillating spatially homogeneous magnetic fields (McTague, bed core.
1969; Schumacher et al., 2003). Similarly, Magnetoviscothin- High fluid vorticity is encountered near the vessel wall where
ning, tantamount to an apparent drop off in local ferrofluid an unobstructed gross wall flow develops. Conversely, bed core
suspension viscosity, occurs when fluid vorticity and magne- vorticity is the lowest due to porosity distribution damping,
tized nanoparticle spin are in corotation with the nanoparticle which implies weaker local velocity variations. It is anticipated
spin velocity faster than the fluid vorticity (Fig. 1b). Part of that application of magnetic fields to ferrofluids would yield
the nanoparticle internal angular momentum is transmitted to skin magnetoviscothickening mainly located in the wall region.
the fluid deflating fluid dissipation as in drag reduction rem- It is hoped that nanoparticle spin retardation would therefore
iniscent of polymeric behaviors. This phenomenon has been artificially increase resistance to flow in this region thus yield-
observed in laminar pipe flow with high-frequency oscillating ing lower wall bypass fraction. One simple arrangement to
linearly polarized magnetic fields (Bacri et al., 1995; Zeuner induce magnetoviscosity corresponds to axisymmetrical mag-
et al., 1998). An intermediate case, corresponding to classical netic field and geometry to have magnetostatic torque, vortic-
fluid mechanics situation, also exists when no spin lag between ity and spin velocity orthoradial (Fig. 1). To sharpen the con-
fluid element and magnetized nanoparticle yields null magne- trasts in the velocity profiles owing to the Kelvin body force, a
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1629

divergenceless external magnetic field with positive and nega- Γγ κ centroid


tive constant-gradient streamwise component is chosen. ∂Vκe Vκ
∂Vγ e Vγ
Magnetoviscosity is analyzed by formulating the general

local ferrohydrodynamic model, the general upscaled volume-
average model in porous media with the closure problem, Γγ κ
R
and solution and discussion of a simplified zero-order steady- d yγ
state isothermal incompressible axisymmetrical model for yκ
non-Darcy Forchheimer (1914) flow of a Newtonian ferrofluid ∂ V κe
x nγκ
in a Müller (1991, 1992) porous medium subject to a total
magnetic field accounting for bulk-flow oriented external field Vγ
L
H0 , ferrofluid self-consistent demagnetizing field, and in- ∂ Vγ e
duced magnetic field in the non-magnetic granular phase. The
Shliomis (1972) magnetization relaxation equation has been (b) V
written in the fluid compressible form according to Felderhof
(2001). Both external and total magnetic fields are taken to 1

be irrotational (zero free-current density in packed bed, i.e., (a)


Flow 0.8
non-conducting ferrofluid and porous medium).
0.6

ε
2. Mathematical tools and properties 0.4

0.2
For the derivation of the volume-average equations from the
pore-level ferrohydrodynamic equations, we will recall some 0
basic properties regarding the phase characteristic function, the 0 0.2 0.4 0.6 0.8 1

superficial and intrinsic volume-average quantities, the fun- (c) r/ R


damental theorem for spatial averaging and its accompany-
ing corollaries, the Gray (1975) spatial decomposition around Fig. 2. Porous medium details: (a) packed bed details, (b) averaging volume
details, (c) radial profile of porosity of a Müller (1991, 1992) porous medium
the intrinsic averages, and the general derivative-integral inter- of spherical grains of 2 mm in size in a 2 cm ID diameter tube.
change theorem.

2.1. Averaging theorems boundary. Let |V | be the volume of the averaging region V
defined by |V | = V d.
Fig. 2b shows the averaging volume region V = V ∪ V ∪ To prevent the notational misinterpretations contingent to
 which is a bounded open set inclusive of V , the union weak symbolization of the field variables during the manipu-
of a finite number of open sets representing the granular - lation of the mathematical concepts, more specifically the dis-
phase inside V. Complementarily; the open set V represents continuities in materials properties that occur at the interfaces,
the region within V occupied by the ferrofluid -phase. The set use will be made of the characteristic (or phase indicator) func-
V  is the adherence of the ferrofluid phase, and jVe and jVe tion (Drew and Passman, 1999). We write, for instance, for any
designate, respectively, the -phase and the -phase associated point r = x + y in V:
entrance/exit contours. Therefore, the ferrofluid–solid interface 
1 if r ∈ V ∪  ,
inside V is defined as  = (V  /V )/jVe .  (r) = (1)
0 otherwise.
For the sake of brevity, the boundary conditions on jV =
jVe ∪ jVe of the vector and scalar fields and the initial con- Hence, the -associated superficial average of a bounded
ditions are omitted, even though these are usually not known a piecewise continuous field variable  in the compact set V and
priori. We will postulate, without proof, that there exists for the function of vector x + y is defined as
model Müller porous medium, a spatially periodic representa-  
1 1
tive elementary volume REV of the order of the macroscopic  |x+y |x =  |x+y d = |x+y d. (2)
volume, such that ignorance of these boundary conditions does |V | V |V | V
not harm the closure problem of the volume-average equations Restriction of the normalization volume to the -phase vol-
(Whitaker, 1999) to be developed later. ume yields the interstitial average
We will borrow some of the notations used by Whitaker  
(1999) and Rosensweig (1997). Let x be the position vector 1 1
 |x+y  |x =  |x+y d = |x+y d. (3)
locating, in the macroscale frame, the centroid of the averaging |V | V |V | V
volume region V, and y (or y ) be the relative position vector in
Putting  = 1 in Eq. (2) gives the obvious result
the microscale frame of a -phase (or -phase) point (Fig. 2b). 
Let n be the normal canonical unit vector associated with  , 1 |V |
 |x+y |x =  |x+y d = =  . (4)
i.e., unit normal vector directed outwards from the ferrofluid |V | V |V |
1630 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Therefore, combining Eqs. (2)–(4) yields the known relation- 2.2. Gray spatial decomposition
ship between the superficial and the interstitial volume averages
of a given field variable At this stage, we will introduce the well-known Gray (1975)
length-scale decomposition of the local scalar and vector fields
 |x+y |x = |x+y  |x  . (5) that are split up into an intrinsic average and a spatially fluctu-
Assuming in addition that ∇  is integrable in V, the spatial ating term reminiscent of the traditional Reynolds decomposi-
averaging theorem formally states, in terms of the superficial tion in turbulent flows. We will make, upfront, the assumption
quantities, as (Whitaker, 1999; Drew and Passman, 1999) of scale separation between pore-level microscale (the realm of
 y) and porous medium macroscopic scale (the realm of x).
1 For a scalar field for instance, such spatial decomposition is
∇ |x+y |x = ∇ |x + |x+y n d. (6)
|V |  written as

Note that the spatial derivative of Eq. (6) LHS is performed,  |x+y =   |x +  |
˜ y (12)
prior to averaging, on the microscale relative location vector,
y, whereas the RHS left-term spatial derivative concerns the Here also, for ease of notation, the intra-subindex x+y in the
macroscale location vector, x, of the already spatially averaged RHS first term in Eq. (12) has been omitted. The tilde tagged
quantity within the averaging volume region. For ease of nota- variable in Eq. (12) represents the spatial fluctuation (or the
tion, the intra-sub index x + y in the RHS first term in Eq. (6) deviation) around the intrinsic average in V ∪  . Applying
has been dropped. the superficial average operator on both sides of Eq. (12) yields
Corollaries (see Appendix A) to Eq. (6) concern vector/tensor  |x+y |x =   |x |x +  |
˜ y |x , (13)
field gradients and divergences, and vector field del cross which
can be written, respectively, as  |x =  |x  |x +  |
˜ y |x . (14)

1 Eq. (14) restitutes the relationship between the intrinsic and the
∇ |x+y |x = ∇ |x + |x+y ⊗ n d, (7)
|V |  superficial averages defined in Eq. (5) provided the average of
 the spatial fluctuations in V ∪ is zero, which we assume true
1
∇ |x+y |x = ∇ |x + |x+y ⊗ n d, (8) in our treatment of the ferrohydrodynamic problem by analogy
|V |  with the Reynolds decomposition:

1  |
˜ y |x = 0. (15)
∇ ·  |x+y |x = ∇ ·  |x + |x+y · n d, (9)
|V | 
 2.3. General derivative-integral interchange theorem
1
∇ ·  |x + y |x = ∇ ·  |x + |x+y ·n d, (10)
|V |  Another useful theorem to be used in the volume-average
∇ ×  |x+y |x = ∇ ×  |x formulation is the general theorem of derivative-integral inter-
 change, which, for stationary  interface and constant aver-
1
+ n × |x+y d (11) aging volume, for a scalar field is written as
|V |    
j  j
in which the operators ⊗, ·, × stand, respectively, for the tensor    =  |x (16)
jt x jt
product, the dot product or contracted product and the curl.
For ease of deciphering the type of the field variables in play, and which requires that   (i) be differentiable with respect to
the order of a tensor is tagged with a matching number of time in IR+ , (ii) integrable with respect to r = x + y in V, (iii)
underscore lines (0: scalar, 1: vector, etc.). j
is piecewise continuous in V × IR+ , (iv) and verifies the
j t  
The volume-averaging theorems represent a multidimen-
sional version of the Leibniz rule for interchanging between domination hypothesis: ∃ V → IR+ , intergrable on V, ∀r ∈
j
differentiation and integration operators (Whitaker, 1999; Drew V , ∀t ∈ IR+ , | jt  |r,t  |r .
and Passman, 1999). Physically speaking, Eqs. (6)–(11) repre-
sent some important forms of the averaging theorem that will 3. Formulation of the microscale Cauchy problem
allow us to express the ferrofluid transport laws with respect
to the two-phase averaging volume region V using only the An isothermal and incompressible single-phase Newtonian
volume averaged (scalar, vector or tensor) field properties as- ferrofluid upflow through a non-magnetic rigid Müller (1991,
signed to the element centroid, x, and the field average values 1992) porous medium of spherical grains (Fig. 2c) is assumed.
along the contour  when the functions take non-zero val- As alluded to earlier regarding wall magnetoviscosity effects,
ues on it. The usefulness of these theorems will become clear the choice of a Müller porous medium model is motivated by the
in the section Upscaling devoted to obtaining an equivalent fact that a zero-order Bessel function of the first kind restores
set of volume-average equations for describing the ferrofluid in a realistic manner the damped oscillations induced by the
macroscopic behavior in the porous medium. size and shape of the particles from the wall to the bed core
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1631

in columns that exhibit a low column-to-particle diameter ratio. 3.3. Internal angular momentum at pore scale
Also, a pseudo-Coulombic approach of the coupled magneto-
static problem is used in which the magnetic flux induction is j

I  + ∇ · 
I  ⊗ v
directly expressed in terms of the magnetic field. The Cauchy jt
(or boundary-value) problem describing the ferrofluid flow in = ∇ ·  C − E :  T +  G in V . (19)
the averaging volume region V, needed for spatial averaging as
defined in the earlier section, consists of the continuity, linear In our context, it is the internal rather than the total angular
momentum, internal angular momentum, magnetization relax- momentum that is of interest for highlighting the magnetovis-
ation, Maxwell-flux law and Ampère–Maxwell law local equa- cosity phenomenon. Details on the derivation of Eq. (19), which
tions in V, along with  boundary conditions, the constitutive is relevant to polar fluids such as ferrofluids, can be found in
tensor and vector equations, and a number of additional ancil- Rosensweig (1997).
lary relationships. In Eq. (19), I is the moment of inertia density of the
magnetic nanoparticle;   is the spin density vector (also
referred to as spin velocity, internal spin rate, angular spin
3.1. Continuity of ferrofluid flow at pore scale rate or simply angular velocity) of the magnetic nanoparticle;
 C is the symmetric couple stress dyadic cast in terms of
∇ ·  v = 0 in V . (17) an internal strain rate formally reminiscent of the symmetric
(pressure–viscous) tensor taking part in the linear momen-
This notation results from the incompressibility and the isother- tum. This term conveys the effect of the short-range diffusive
mal conditions (j
/jt =0, ∇
(T )=0); where
is the ferrofluid exchanges of angular momentum being exerted by the im-
suspension mass density, T the absolute temperature, and  v mediate surroundings; the double-contraction term −E :  T
is the ferrofluid local instantaneous mass velocity. between the pressure–viscous-spin/vorticity stress tensor and
the unit antisymmetric orientation Euclidian triadic is nothing
3.2. Linear momentum at pore scale else than the rate of conversion of external angular momentum
into internal angular momentum, and expressed, as seen later,
j as a proportionality relation with respect to the asynchrony

v + ∇ · 
v ⊗ v between the ferrofluid vorticity and the nanoparticle spin den-
jt
sity. As explained in the introduction, this term is a measure
= ∇ ·  T + ∇ ·  M̂ + 
g in V , (18)
of the mechanical torque the magnetic nanoparticle must exert
to oppose the magnetostatic torque associated with the direc-
where  T is the pressure–viscous-spin/vorticity stress tensor tional mismatch between the local dynamic magnetization and
which gathers the classical symmetric component due to pres- the local total magnetic field vectors;  G is the body couple
sure and viscous forces being exerted by the contiguous sur- density or the magnetostatic torque and represents transport of
roundings, in addition to an antisymmetric tensor contribution intrinsic angular momentum due to distant sources.
arising from the spin/vorticity asynchrony between the mag-
netic nanoparticle and the ferrofluid;  M̂ is the magnetic stress 3.4. Magnetization relaxation at pore scale
tensor whose divergence retrieves, under some assumptions,
the well-known Kelvin magnetization body force density that j
 M + ∇ ·   M ⊗ v
results from both long-range external magnetic field as well jt
as self-consistent ferrofluid-borne demagnetizing field and in- =   × M − −1  (M(H) − M (He )) in V (20)
duced field developing in the non-magnetic material (i.e., gran-
ular porous medium); and 
g is the usual gravitational body The Shliomis (1972) constitutive magnetization relaxation
force density in the ferrofluid. Note that the acceleration term equation has been chosen and written in the fluid compress-
in the RHS of the linear momentum equation is written as the ible form according to Felderhof (2001). Other constitutive re-
vector divergence of the tensor product between mass flux and lations could also have been used, e.g. the Martensyuk et al.
local velocity of the ferrofluid. By virtue of the solenoidal con- (1974) equation based on Brownian motion approach in an ef-
dition Eq. (17), this notation is neutral with respect to the con- fective field approximation, or the Felderhof and Kroh (1999)
ventional “Lagrangian” derivative notation in the linear mo- relaxation relation derived from irreversible thermodynamics.
mentum balance equation. However, as will be seen later, it will Magnetization in the ferrofluid phase changes, in the one
be very useful in casting the conservation equations wherein hand, because of the magnetic nanoparticle rotation (first term
fields other than mass fluxes are being advected by  v. It is RHS Eq. (20)) that is responsible for misalignment between
also noteworthy that the convective flux term ∇ ·
v ⊗v is ex- the dynamic magnetization and the total magnetic field vec-
pected to play a crucial role for the“abnormal” porous medium tors (Fig. 1). Magnetization changes, in the other hand, due
assumed in this study as suggested from the shape of the radial to a defect term (second term RHS Eq. (20)) between the dy-
porosity profile in Fig. 2c which will give rise to strong radial namic magnetization feeling the flow-conditions magnetic field
dependence of the radial profile of the axial (or streamwise)  H and the equilibrium Langevin magnetization  M be-
velocity component. ing experienced by the ferrofluid in the magnetohydrostatic
1632 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

conditions, i.e.,  v = 0 and  He . The last phenomenon takes equation. This issue will be settled down when addressing the
place via a global relaxation whose characteristic time  is closure of the drag force on  in the volume-average equa-
the resultant between the Brownian and the Néelian relaxation tions, which can be modeled either as a Darcy creeping flow
times. Note that the convective flux magnetization term (sec- term, a Burke–Plummer inertial term or a general Ergun-like
ond term LHS Eq. (20)) plays a role similar to ∇ · 
v ⊗ v. term (Coulson et al., 1991).
It will be confirmed later to exhibit important spatial depen- To solve the complete internal angular momentum equation,
dences for a Müller porous medium in linear-gradient external two scalar boundary conditions are required for each one of
magnetic fields and therefore cannot be neglected. the spin density components if the symmetric “spin–diffusion”
couple stress dyadic is accounted for. However, the spin bound-
3.5. Maxwell flux law at pore scale ary conditions are less intuitive and are still stirring debate.
Five possibilities have been described in the literature: (i) lin-
∇ ·  0 (H + M) = 0 in V , (21) ear velocity-like boundary conditions, (ii) vorticity matching
on  tantamount to cancellation of the antisymmetric stress
∇ ·  0 (1 +  )H = 0 in V , (22) tensor component such as for non-polar fluids, (iii) linear com-
∇ · 0 H0 = 0 in V , (23) bination of (i) and (ii) boundary conditions, (iv) Eq. (19) is
satisfied at the wall, i.e., equation extension (Rinaldi and Zahn,
where 0 is the vacuum permeability (=4 × 10−7 N/A2 ), 2002a; Schumacher et al., 2003). In a similar manner to the ve-
 is the magnetic susceptibility of the non-magnetic material locity field, we will see later on that these boundary conditions,
making up the granular phase, and H0 is the externally applied except boundary condition (iv), are of little help to close the
magnetic field before the granular phase and the ferrofluid are volume-average internal angular momentum equation so that
being placed in the column. The total local magnetic fields, different assumptions will be made to deal with the closure
respectively, in the ferrofluid and the granular phase, decom- terms associated with the spin density vector when Eq. (19) is
pose as  H =  (H0 + h) and  H =  (H0 + h) in which satisfied wallwise
 h and  h are, respectively, the self-consistent demagnetiz-
ing field induced inside the magnetized ferrofluid matter and ∇ · 
I  ⊗ v = ∇ ·  C − E :  T +  G on  . (29)
the induced magnetic field in the non-magnetic granular phase.
The magnetization relaxation equation requires one single
boundary condition on the magnetization vector. For commod-
3.6. Ampère–Maxwell law at pore scale ity, an extension of Eq. (20) can be used
∇ ×  (H0 + h) = 0 in V , (24) ∇ ·  M ⊗ v

∇ ×  (H0 + h) = 0 in V , (25) =   × M − −1  (M(H) − M (He )) on  . (30)

∇ × H0 = 0 in V . (26) Also, only one boundary condition is required for each of


the ferrofluid-side self-consistent demagnetizing field and the
These equations mean that both external and total magnetic granular-side induced magnetic field. The tangential compo-
fields are taken to be irrotational, i.e., zero free-current density nents of the self-consistent demagnetizing field  h and the in-
in packed bed is assumed. duced magnetic field  h are continuous in their way across
the ferrofluid–granular interface (du Trémolet de Lacheisserie,
3.7. Boundary conditions on fluid–solid interface  inside 2000):
averaging volume region lim h × n − lim h × n
 
Although it is not our intent to solve the ferrohydrody- = h × n − h × n = 0

on  . (31)
namic boundary value problem before any preliminary aver-
aging process, initial and boundary conditions are required The normal components of the demagnetizing field  h and
for completeness as some of these conditions can be usefully the induced magnetic field  h present a discontinuity in their
taken advantage of in the derivation of the volume-average way across the ferrofluid–granular interface due to the non-zero
ferrohydrodynamic model. In this study, we will restrict our- ferrofluid-side magnetization normal component (du Trémolet
selves with an analysis of the steady-state model solutions, i.e., de Lacheisserie, 2000):
j/jt ( or  ()) = 0(0). lim h · n − lim h · n
 
For the ferrofluid velocity field, the usual slipless no-
penetration condition can be used: = h · n − h · n = M · n on  (32)

n · v = 0 on  , (27) 3.8. Constitutive relationships


n × v = 0 on  . (28)
For completing the formulation of the Cauchy problem,
A further boundary condition is required to be able to solve a set of constitutive relations must be supplemented for the
the second-order partial differential linear momentum vector pressure–viscous-spin/vorticity stress tensor, the magnetic
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1633

stress tensor, the external–internal exchange angular momen- in which m is the suspension volume fraction of the mag-
tum, the couple stress dyadic, the body couple density, and the netic material embedded in the nanoparticles; m̄ the magnetic
Langevin magnetic-field-locked magnetization law. dipole moment of an individual nanoparticle; dm the nanopar-
If a Newtonian behavior of the ferrofluid suspension is as- ticle diameter of the magnetic core; k the Boltzmann constant
sumed, the pressure–viscous-spin/vorticity stress tensor as writ- (=13.8 yJ/K); and  he is the equilibrium demagnetizing field
ten as the sum of three symmetric tensors and an antisymmetric prevailing in the ferrofluid phase at magneto hydrostatic con-
tensor due to the spin/vorticity asynchrony between the mag- ditions.
netic nanoparticle and the ferrofluid (Rinaldi and Zahn, 2002a): In the low magnetization limit (argument of hyperbolic
cotangent smaller than unity in Eq. (38)), the Langevin equi-
 T =  (−pI + (∇v + t ∇ v) + (∇ · v)I librium magnetization becomes proportional to the local total
+ E · (∇ × v − 2)) in V (33) magnetic field with the proportionality constant corresponding
to the ferrofluid initial magnetic susceptibility 0 :
in which I is the unit isotropic dyadic, p is the pressure, and
, , and  are, respectively, the dynamic viscosity, the second m 0 m̄2 dm
3
0 = . (39)
coefficient of viscosity, and the vortex viscosity. 18 kT
Neglecting the magnetostrictive term and ignoring the dipo-
In this, limit it ensues that
lar interactions between magnetic nanoparticles yields the fol-
lowing expression for the magnetic stress tensor (Rosensweig,  M =  0 (H0 + he ) in V . (40)
1997):
 Conversely, beyond certain magnetic field intensity, the fer-
 M̂ =  − 0 [(H0 + h) · (H0 + h)]I + 0 (H0 + h) rofluid magnetization levels off to yield asymptotically magne-
2  tization saturation when virtually all the body-locked magnetic
⊗ (H0 + h + M) in V . (34) dipoles point at diapason collinearly in the direction of the total
magnetic field:
Under the two assumptions just mentioned, the divergence of
the magnetic stress tensor reduces to the familiar Kelvin body H 0 + he
 M =  m m̄ in V . (41)
force magnetization density given by H0 + he

∇ ·  M̂ = 0 ∇ (H0 + h) · M in V . (35) In this study, the Langevin magnetization is calculated using
the general expression Eq. (38).
Note that we took the convention to write in a hierarchical
manner the contracted product with first the tensor of highest 3.9. Ancillary relationships
rank contracted on the tensor of lower rank. This convention
will be followed throughout the entire article. 3.9.1. Magnetic properties
The couple stress dyadic is formulated symbolically as the The magnetization relaxation time in Eq. (20) needs to be
symmetric pressure–viscous stress tensor as a function of the computed. Magnetization relaxation takes place mainly accord-
spin density (Rosensweig, 1997): ing to two incoherent relaxation mechanisms when, in response
 C =  ( (∇ + t ∇ ) +  ∇ · I) in V , (36) to a change in the local magnetic field direction, the nanoparti-
cle readapts the direction of its magnetic moment. The nanopar-
where  and  are the shear and bulk coefficients of the ticle changes its magnetic moment through external Brownian
spin viscosity, respectively. The viscosity quintet (, , ,  ,  ) rotation whose characteristic time is given by
takes positive values although the last two viscosity values are
dh3 0
often taken to be zero. B = (42)
Neglecting the gravitational body-couple density associ- 2kT
ated with eventual anisotropic mass distribution effects in the and through a rotation of the magnetic moment inside the
nanoparticles, the only body-couple density to be accounted nanoparticle also known as Néel relaxation whose characteris-
for in the internal angular momentum balance equation is the tic time is
magnetostatic torque defined as
N = f0−1 e(K/kT )( dm /6) ,
3
(43)
 G =  0 M × (H0 + h) in V . (37)
where dh is the hydrodynamic diameter of the coated nanoparti-
The magnetization vector for the ferrofluid at rest in a mag- cle, including twice the thickness of the surfactant layer serving
netic field (magnetohydrostatics) is written as to the stabilization of the ferrofluid suspension (Fig. 3); 0 is

the dynamic viscosity of the liquid carrier; K is the anisotropy
3 0 m̄
 M =  m m̄ coth d H0 + h e constant of the single-domain uniaxial ferromagnetic magnetic
6 m kT
nanoparticle (Rinaldi and Zahn, 2002a); and f0 is the Lar-
6 kT H 0 + he
− 3 H0 + he −1 in V (38) mor frequency of the magnetization vector in the anisotropic
dm 0 m̄ H0 + he nanoparticle field (Odenbach, 2003).
1634 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

these constraints the solenoidal condition Eq. (23), the follow-


Surfactant
ing external magnetic field is proposed in this study:
 r 
H0 = t − 0 z +  (47)
2
in which  and  are, respectively, the gradient programming

d
and the offset constant. This latter is chosen for canceling H0

h
dm

z-component in the bed entrance ( > 0,  = 0) or at the bed


exit ( < 0,  = −L).

3.9.2. Suspension viscous properties


The Newtonian ferrofluid suspension dynamic viscosity is
estimated from the carrier liquid viscosity, the magnetic vol-
Magnetic core
ume fraction, the hydrodynamic and magnetic diameters of the
(coated) nanoparticle defined earlier, according to a relation
Fig. 3. Schematic structure of a surfactant-dressed magnetic nanoparticle in proposed for ferrofluid suspensions by Rosensweig (1997)
a stabilized ferrofluid suspension.  
 5 d h − dm 3
= 1 − m 1+
An effective relaxation time estimated from the sum of the 0 2 dm
two incoherent relaxation processes (defect magnetization term   −1
d h − dm 6
in Eq. (20)), one from Brownian and the second from Néelian −1.552m 1 + . (48)
dm
relaxations, is calculated as (Henjes, 1995)
1 1 1 Simple geometrical arguments enable to establish the fol-
= + . (44) lowing relationship between the volume fraction of the naked
 B N
magnetic material, m , and the hydrodynamic volume fraction,
When two materials exhibiting close magnetic permeabilities h , of the surfactant-dressed nanoparticles:
are put into contact and are subjected to a low magnetic field so
that their linear magnetic material (LMM) behavior is preserved h dm
3
= m dh3 . (49)
(such as described by Eq. (40)), the Maxwell–Garnett theory,
assuming a local Lorentz field for the ferrofluid–granular phase This enables in turn to express the vortex viscosity as a
assemblage, can be very helpful in decoupling the magneto- function of the dynamic viscosity of the ferrofluid suspension
static problem from the hydrodynamic problem of the ferro- (Rinaldi and Zahn, 2002a; Schumacher et al., 2003):
hydrodynamic Cauchy problem (Khuzir et al., 2003). Simple
 = 1.5h . (50)
calculations in the LMM approximation show that neglecting
the magnetic susceptibility of the granular phase with respect The moment of inertia density of the surfactant-dressed mag-
to that of the ferrofluid yields the following equality between netic nanoparticle is calculated assuming spherical magnetic
the intrinsic volume-average demagnetizing field in ferrofluid core and surfactant mantle (Fig. 3):
and induced magnetic field in granular phase:
5 +
(d 5 − d 5 )
1
p dm s h m
h = (1 + )h , (45) I= (51)
10
p dm
3 +
(d 3 − d 3 )
s h m
where, the LMM correction factor is defined as a function only
of the ferrofluid initial magnetic susceptibility in which
p and
s are, respectively, the densities of the mag-
netic core and the dressing surfactant.
 = 0 (3 + 2 0 )−1 . (46)
We will reexamine later in some detail the LMM approxi- 3.9.3. Porous medium properties
mation issue during the implementation of the volume-average We will assume that the porosity field in the bed exhibits
formulation of the ferrohydrodynamic model. only radial dependence and that it is axially invariant. The mul-
The external magnetic field, H0 , is purposely chosen to ex- tiparameter porosity law proposed by Müller (1991, 1992) to
hibit a constant gradient. Linear-gradient magnetic fields are model the voidage left by an assemblage of spherical grains in
the simplest fields to give rise to a Kelvin magnetization body low column-to-grain ratio geometries is given by the following
force. It is assumed azimuthally invariant and exhibits a null relationship:
tangential magnetic field component. It is chosen here to point  
a(D − 2r) −b(D−2r)/2d
in the bulk-flow direction (H0 z-component always positive), (r) = b + (1 − b )J0 e (52)
2d
whereas it is allowed to take positive and negative gradients
by imposing H0 z-component equal zero at the entrance in the where J0 , expressed as a function of the local polar radius, r,
former case and at the exit in the latter case. To satisfy with in the bed, is the zero-order Bessel function of the first kind
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1635

defined as Distribution of divergence in tensor product



 (−1)i r 2i ∇ ·  ⊗  = (∇ · ) + ∇ · . (58)
J0 (r) = . (53)
22i (i!)2
i=0 Idemfactor divergence–gradient equivalence
The other local bed porosity parameters are expressed as a ∇ · I = ∇. (59)
function of the ratio of the particle diameter, d, and the column
diameter, D: Distribution of divergence in summation of tensor and its
transpose
d
b = 0.365 + 0.22 , (54)
D ∇ · (∇ + t ∇ ) = ∇ · ∇ + ∇(∇ · )
⎧ d D =  + ∇(∇ · ). (60)
⎨ 7.45 − 3.15 ∈ [2.02 − 13.0],
a= D d (55) Laplacian equivalence
⎩ d D
7.45 − 11.25 13.0,
D d  = ∇(∇ · ) − ∇ × ∇ × . (61)
d
b = 0.315 − 0.725 . (56) Antisymmetric triadic divergence–gradient equivalence
D
It is well known that low-ratio columns may occasion se- ∇ · (E · ) = −∇ × . (62)
vere maldistribution due to large radial permeability contrast
with preferential flow short-circuiting alongside the high- Contracted tensor product with antisymmetric triadic
permeability wall area. The challenge will be to identify
E : (E · ) = −2. (63)
configurations that are propitious to creating sufficient mag-
netoviscosity near the vessel wall to quench the bypass and Distribution of gradient of scalar function product
to reroute maximum fraction of fluid flow towards bed core.
Hence, to assess the comparative performances among the ∇() = ∇ + ∇. (64)
simulated configurations, a wall bypass fraction, BPw , is de-
fined as a measure of the amount of ferrofluid flux escaping Distribution of gradient of vector/scalar function product
alongside the wall region over a distance corresponding to
∇() = ∇ +  ⊗ ∇. (65)
half the grain diameter when the fluid velocity exhibits an in-
homogeneous radial profile with respect to the ideal situation Distribution of divergence of scalar/tensor product
when the radial profile is flat. Such departure from ideality is
expressed following the relationship: ∇ · () = ∇ ·  +  · ∇. (66)
  −1 Distribution of curl of scalar/vector product
D/2 D/2
BP w = r(r)vz (r) dr rU 0 dr , (57)
D/2−d/2 D/2−d/2 ∇ × () = ∇ ×  + ∇ × . (67)

where U0 is the ferrofluid superficial velocity. Distribution of divergence of tensor product

∇ · ( ⊗ ) = ∇ ·  + (∇ · ). (68)
4. Upscaling
Distribution between tensor and scalar product
Upscaling is an inescapable prerequisite step allowing the
passage from complex and still out-of-reach pore-level so- ( ⊗ ) ·  = ( · ). (69)
lutions of the above physical microphenomena. Upscaling
Putting =1 in Eq. (6) and using Eq. (4), yields the following
consists in formulating an equivalent set of volume-average
useful geometrical theorem (Whitaker, 1999):
equations describing the ferrofluid macroscopic behavior at the

porous medium level from a description using the local ferro- 1
∇ = − n d. (70)
hydrodynamic equations. To develop these upscaled equations, |V |  
we will use the spatial averaging theorem and its corollaries,
the theorem of derivative-integral interchange, and the Gray In what follows, we will develop the rigorous volume-
(1975) spatial decomposition around the intrinsic averages, average forms of the ferrofluid conservation equations for con-
introduced earlier. tinuity, linear momentum, internal angular momentum, mag-
We will recall first some important tensor and vector iden- netization relaxation, Maxwell flux law, and Ampère–Maxwell
tities that are useful for the derivation of the volume-average law. The closure issue, the simplified form of the ferrohydro-
equations. Such equations can be established using tensor cal- dynamic model, the corresponding boundary conditions for
culus properties. They can be found in Rosensweig (1997) or the axisymmetrical case and the model implementation will be
established by consulting tensor textbooks (Garrigues, 2001). detailed after.
1636 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

4.1. Volume-average continuity equation Finally, after invoking Eq. (5) for the time-derivative term,
the volume-average substantial derivative term is written as
Applying the spatial averaging corollary, Eqs. (9)–(17) gives  
j
∇ ·  v|x+y |x 
v + ∇ · 
v ⊗ v
 jt
1 j
= ∇ ·  v|x + v|x+y · n d = 0. (71) =
 v +
 ∇v · v +
∇ ·  ṽ ⊗ ṽ . (78)
|V |  jt

Invoking the adherence condition Eqs. (27), (28) on wall and The last term in Eq. (78) designates the inertial dispersion
the relationship between the intrinsic and superficial averages, contributed by the spatial fluctuating velocity in the -phase.
Eq. (5), and after dropping the indices, the volume-average
continuity equation becomes 4.2.2. Divergence of Cauchy stress tensor ∇ ·  T
∇ ·  v = 0. (72) Invoking fulfillment of the local continuity Eq. (17) enables
to drop one term in the Cauchy stress tensor, Eq. (33). After
distributing the volume-average operator, it gives
4.2. Volume-average linear momentum equation
∇ ·  T = ∇ · (− pI) + ∇ · (∇ v + t ∇  v)
We will proceed with a term-wise treatment of the linear
+ ∇ · (E · (∇ ×  v − 2 )). (79)
momentum balance equation (Eq. (18)) which consists of the
following terms: the two substantial derivative terms, the three
pressure–viscous-spin/vorticity terms, the magnetic stress term, Using relations (59), (60) and (61) in Eq. (79), and invoking
and the gravitational body force term. The reader can refer to Eq. (17) in (60) yields
the above microscale Cauchy formulation for a discussion of ∇ ·  T = − ∇ p + ( + )∇ · (∇( v))
the physical meaning of the different terms appearing in the
+ 2∇ ×  . (80)
linear momentum balance equation.
Applying to the pressure gradient term the spatial averaging
4.2.1. Substantial derivative term (j/jt)
v + ∇ · 
v ⊗ v theorem Eq. (6) and using Eqs. (5) and (12) gives
By linearity and by invoking the general derivative-integral

interchange theorem: 1
−∇ p = −∇( p ) − (p + p̃)n d. (81)
  |V | 
j j

v + ∇ · 
v ⊗ v = 
v + ∇ · 
v ⊗ v.
jt jt Invoking the gradient distribution property Eq. (64) and the
(73) geometrical theorem Eq. (70):
Applying Eq. (10) spatial averaging corollary to the second 
1
term of Eq. (73) RHS and invoking again the adherence con- −∇ p = − ∇p − p̃n d. (82)
|V |  
dition Eqs. (27), (28) on wall yields
 
j Unlike pure Newtonian fluids, the ferrofluid suspension vis-

v + ∇ · 
v ⊗ v cous term involves in Eq. (80) a summation of the dynamic vis-
jt
j cosity and the vortex viscosity coming through the asynchrony
= 
v + ∇ · 
v ⊗ v. (74) term between fluid vorticity and nanoparticle spin density. This
jt
distinction is important because it will affect the formulation of
Taking advantage of the Gray spatial decomposition and the the fluid-grain drag using Darcy or Ergun-like equations as will
fluid incompressibility, the inertial term can be further decom- be seen later. Using successively Eq. (10) then Eq. (7) spatial
posed averaging theorem corollaries, then exploiting the wall adher-
ence velocity condition Eqs. (27), (28), and then using Eq. (5),
∇ · 
v ⊗ v =
∇ · ( v +  ṽ) ⊗ (v + ṽ). (75) yields

Since  ṽ = 0 (Eq. (15)) and by virtue of Eqs. (4) and (5), ( + )∇ · (∇( v))
we have
= ( + ) ∇ · ∇( v )
∇ · 
v ⊗ v =
∇ ·  v ⊗ v +
∇ ·  ṽ ⊗ ṽ (76)
 
which can be further simplified using Eq. (58) and invoking the 1
+ ∇( v) · n d . (83)
solenoidal condition Eq. (72) on the volume-average continuity |V | 
equation
Using the Gray decomposition Eq. (12) in the integrand of
∇ · 
v ⊗ v =
∇v ·  v +
∇ ·  ṽ ⊗ ṽ . (77) the surface integral in Eq. (83), distributing and using Eq. (70),
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1637

gives Eqs. (82), (87) and (90):

∇ · (∇( v)) = ∇ · ∇( v ) − ∇(v ) · ∇ ∇ ·  T = −  ∇p + ( + )


  
1 ∇ 
+ ∇( ṽ) · n d. (84) ×  v + ∇v · + v
|V |   

1
+ 2 ∇ ×  + 2 n × ˜ d
Exploiting the splitting expression Eq. (65): |V |  

1
∇ · (∇( v)) = ∇ · ( ∇v + v ⊗ ∇ ) + {−p̃I+(+)∇( ṽ)}·n d. (91)
|V | 
− ∇(v ) · ∇
 Note that in our context, the Brinkman correction term v
1
+ ∇( ṽ) · n d. (85) may not be negligible considering the severe voidage gradients
|V | 
in the porosity law describing the Müller (1991, 1992) porous
medium. In addition, the integrand of the last closure term in
Distributing the divergence operator and invoking Eqs. (66)
Eq. (91) contains the sum of the ferrofluid dynamic and vortex
and (58) gives
viscosities multiplied by the gradient of the fluctuating velocity
∇ · (∇( v)) =  v + ∇v · ∇ + v  and the fluctuating pressure on the contour  . This suggests
 that unlike classical drag formulations, the drag function here
1
+ ∇( ṽ) · n d. (86) must account for the occurrence in the laminar Darcy or Ergun
|V |  equation term of the total viscosity rather than only the dynamic
viscosity. To model this closure term, use will be made of the
This gives for the total viscous term usual Ergun equation (see for example Coulson et al., 1991):

( + )∇ · (∇( v)) 1
  {−p̃I + ( + )∇( ṽ)} · n d
∇  |V | 
= ( + ) v + ∇v · + v   
  150( + ) 1 −  2 1.75
1 − 
 =− + v
( + ) d2  d 
+ ∇( ṽ) · n d. (87)
|V |  ×  v . (92)

Let us know turn our attention to the last term of Eq. (80).
4.2.3. Divergence of magnetic stress tensor  M̂
Using the spatial averaging theorem corollary Eq. (11):
The volume-average expression for the Kelvin body force
2∇ ×   = 2∇ ×   density given by Eq. (35), is written as after spatial decompo-
 sition of the total magnetic field and the magnetization vectors,
1
+ n ×   d. (88)
|V |   ∇ ·  M̂ = 0 ∇( H +  H̃) · (M + M̃). (93)

Substituting the superficial average by the intrinsic average Distributing the volume-average operator induces four terms
using Eq. (5), splitting the spin density vector according to the that need to be determined:
Gray (1975) spatial decomposition Eq. (12), and invoking the (II)
geometrical theorem Eq. (70), gives 
(I)
    
∇ ·  M̂ = 0 ∇( H ) · M  + 0 ∇( H̃) · (M )

(III) (IV)
2∇ ×   = 2 ∇ ×   − ∇ ×       
+ 0 ∇( H ) · (M̃) + 0 ∇( H̃) · (M̃) .
 
1 (94)
+ n ט d . (89)
|V |   Applying Eq. (7) to (I), then using successively Eqs. (4), (70)
and (65):
Using finally the splitting relation Eq. (67):
(I) = ∇( H ) · M , (95)
2∇ ×    
  1
1 (I) = ∇ H  + H ⊗ n d · M , (96)
= 2  ∇ ×  + n ט d . (90) |V | 
|V |  
(I) = (∇( H ) − H ⊗ ∇ ) · M , (97)
To restore the volume-average expression of the
divergence of the Cauchy stress tensor, it suffices to add up (I) =  ∇H · M . (98)
1638 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Hence term (I) represents the Kelvin force density con- Distributing the averaging operator and applying Eq. (10) to
tributed through the average magnetic field and the average the first term of Eq. (109) RHS and using Eq. (5):
magnetization.
Applying Eq. (7) to (II), and using the result of Eq. (15) (IV) = ∇ · ( H̃ ⊗ M̃ )

1
+ (H̃ ⊗ M̃) · n d
(II) = ∇( H̃) · M , (99) |V | 
  −  H̃∇ · ( M̃) . (110)
1 
(II) = ∇ H̃ + H̃ ⊗ n d · M , (100) The fluctuating magnetization force can be decomposed into
|V | 
three fluctuating force terms. Hence, the volume-average Kelvin
  body force is written finally as
1 
(II) = H̃ ⊗ n d · M . (101)  
|V |  1 
∇ ·  M̂ =  ∇H + H̃ ⊗ n d · M
|V | 
Term (II) accounts for the contribution of the fluctuating
magnetic field along the ferrofluid–grain interfaces inside V, + ∇ · ( H̃ ⊗ M̃ ) −  H̃∇ · ( M̃)

which, contracted on the volume-average magnetization, yields 1 
+ (H̃ ⊗ M̃) · n d. (111)
an additional magnetization force that needs to be closed. |V | 
Substituting (III) using Eq. (68):
4.2.4. Gravitational body force density 
g
(III) = ∇( H ) · (M̃) (102) Similarly, the volume-average gravitational body force
density in Eq. (18) is written as
(III) = ∇ · ( H ⊗ M̃) − (∇ · (M̃)) H  (103)

g = 
g. (112)
Distributing the averaging operator and applying Eqs. (10)
The general volume-average linear momentum balance equa-
and (9) to the first and second terms of Eq. (103) RHS:
tion in a porous medium takes finally the following form by
 adding up Eqs. (78), (91), (92), (111) and (112):
1
(III) = ∇ · (H ⊗ M̃) + (H ⊗ M̃) · n d
|V |  j
 
 v +
 ∇v · v +
∇ ·  ṽ ⊗ ṽ
1 jt
− ∇ ·  M̃ + M̃ · n d H . (104) = − ∇p + 
g
|V |   
∇ 
+ ( + ) v + ∇v · + v
Invoking Eq. (15) for the fluctuating magnetization vector  
simplifies the previous equation as   
150( + ) 1 −  2 1.75
1 − 
 − + v
1 d2  d 
(III) = (H ⊗ M̃) · n d
|V |  ×  v + 2 ∇ ×  +  ∇(H0 + h ) · M
  
1 1
− M̃ · n d H . + 2 n × ˜ d
|V | 
(105) |V |  
 
1 
Invoking Eq. (69), + (H̃ + h̃ ) ⊗ n d · M
|V |  0
 
1 +  ∇( H̃0 +  h̃) · (M̃) . (113)
(III) = H (M̃ · n ) d
|V | 
 Here we have decomposed the total magnetic field vector into

1 an externally applied component H0 and a demagnetizing field
− M̃ · n d H . (106) (or induced magnetic field) component. Each component is in
|V | 
turn decomposed according to a fluctuating term and an intrin-
Therefore, sic average. We have finally made the assumption that the -
phase intrinsic average of the external magnetic field coincides
(III) = 0 (107) with the local external magnetic field located at the centroid of
the averaging volume region. This approximation seems plau-
Substituting (IV) using Eq. (68): sible if the materials of the -phase and -phase are assumed
homogeneously distributed. Note that the last term in Eq. (113)
(IV) = ∇( H̃) · (M̃) = ∇( H̃) · ( M̃), (108) is written as  ∇( H̃0 +  h̃) · (M̃) rather than the three
RHS terms of Eq. (110) as these procure no simplification to
(IV) = ∇ · ( H̃ ⊗  M̃) − (∇ · ( M̃)) H̃. (109) the closure problem.
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1639

4.3. Volume-average internal angular momentum equation one obtains



We will proceed with a term-wise treatment of the internal
∇ ·  C =  ∇ · ∇  −  ⊗ ∇
angular momentum balance equation (Eq. (19)) in a similar
way as the one exemplified previously for the linear momentum  
1
balance terms. + ˜ ⊗ n d

|V | 
 
4.3.1. Substantial derivative term (j/jt)
I +∇ ·
I ⊗v 1

The volume-average expression for this term derives exactly −  ∇ ·∇ +  ˜  d
(∇ )·n
|V | 
following the steps used in determining its linear momentum
balance homologue. After calculations, this term is written as
+ ( +  )∇ ∇ ·   −  · ∇
 
j 

I  + ∇ · 
I  ⊗ v 
jt 1
 + ˜ · n d
j |V | 
=
I   +  ∇ · v
jt − ( +  )(∇ ·  )∇ + ( +  )
 
+∇ ·  ˜ ⊗ ṽ . (114) × ˜  d.
(∇ ·  )n (119)

The last term in Eq. (114) designates the dispersion in angular
Invoking Eqs. (65) and (66)
momentum contributed by the spatial fluctuating velocity and
spin density in the -phase. ∇ ·  C =    + ( +  ) ∇(∇ ·  )
 
4.3.2. Couple stress dyadic term  C 1
+∇ · ˜ ⊗ n d

∇ ·  C =  ∇ · (∇  + t ∇  ) |V | 

+  ∇ · ((∇ ·  )I). (115) 
+ (∇ )˜ · n d
|V | 
Invoking Eqs. (59) and (60) in Eq. (115) and distributing the  
1
averaging operator + ( +  )∇ ˜ · n d

|V | 
∇ ·  C =  ∇ · (∇ ) + ( +  )∇(∇ ·  ). (116) 
( +  )
+ ˜  d.
(∇ ·  )n (120)
Using Eqs. (10) and (6) on, respectively, the first and the |V | 
second terms
∇ ·  C 4.3.3. External–internal exchange angular momentum term
  −E :  T
1
=  ∇ · ∇  + (∇ ) · n d The double contraction of the unit antisymmetric orientation
|V |  triadic in −E :  T cancels all the symmetric tensor compo-
 
1 nents in the Cauchy stress tensor. It is easy to verify that
+ ( +  ) ∇∇ ·   + (∇ ·  )n d .
|V | 
E :  (−pI + (∇v + t ∇ v) + (∇ · v)I) = 0. (121)
(117)
Using Eqs. (7) and (9) and (5) yields Thus,
  −E :  T = −E : (E · (∇ ×  v − 2 )). (122)
1
∇ ·  C =  ∇ · ∇  +  ⊗ n d
|V |  Invoking Eq. (63) and distributing the dot product in the RHS
 
1 of Eq. (122) gives
+ (∇ ) · n d
|V |  −E :  T = −E : (E · (∇ ×  v)) + 2E : (E · ( )), (123)
 
1
+( + ) ∇ ∇·  + ·n d −E :  T = 2(∇ ×  v − 2 ). (124)
|V | 
 
1 The volume-average expression of the mechanical torque (or
+ (∇ ·  )n d . (118) the external–internal exchange angular momentum term) can
|V | 
now be developed. Hence, we have
Using the spatial decomposition of the spin density vec-
tor (Eq. (12)) and invoking the geometrical theorem Eq. (70), −E :  T = 2(∇ ×  v − 2 ). (125)
1640 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Utilizing the spatial averaging corollary given by Eq. (11) 4.4.1. Substantial derivative term (j/jt) M + ∇ ·  M ⊗ v
and using Eq. (5) for the volume-average velocity and spin  
j
density  M + ∇ ·   M ⊗ v
jt
j
−E :  T = 2 ∇ ×  v =  M + ∇ M · v
jt
 
1 +  M (∇ · v ) + ∇ ·  M̃ ⊗ ṽ . (131)
+ n × v d − 2  . (126)
|V |   The last term in Eq. (131) represents the change in magne-
tization due to the rotational fluctuation component of the spin
Exploiting the vector identity Eq. (67) and the velocity
density in the -phase. Note that we used the decomposition
adherence condition Eqs. (27), (28), finally gives
  Eq. (58) in a slightly different manner for splitting the tensor
∇ product between magnetization and velocity field. As will be
−E :  T = 2 ∇×v + ×v −2 . (127)
 seen later, because of the retained choice to represent the mag-
netization vector normalized with respect to unit bed volume
4.3.4. Body couple density term  G rather than to unit ferrofluid volume, the solenoidal condition
From Eq. (37) and the spatial decompositions of the mag- of the continuity equation is not used in Eq. (131).
netic field and the magnetization vector, the volume-average
magnetostatic torque or body-couple density term is written as 4.4.2. Term   × M
Distributing and exploiting the properties given by Eqs. (4),
 G = 0 (M + M̃) × ( H +  H̃). (128)
(5) and (15), one obtains
Distributing and exploiting the properties given by Eqs. (4),
˜ × ( M +  M̃),
  × M = ( + ) (132)
(5) and (15), one obtains
 G = 0  M × H + 0  M̃ × H̃ . (129) ˜ × M̃ .
  × M =   × M +   (133)

In which the last term can be thought of as the body couple Besides the misalignment due to volume-average spin den-
density contributed by the fluctuating magnetization and mag- sity, the fluctuating term arising form the curl between the spa-
netic field. This term is the counterpart of the term  ∇( H̃) · tial fluctuation of spin density and spatial fluctuation magneti-
(M̃) arising in the volume averaging process of the Kelvin zation adds up another contribution to Eq. (133).
body force in the linear momentum balance equation, Eq. (113).
The general volume-average internal angular momentum bal- 4.4.3. Relaxation term −1  (M(H) − M (He ))
ance equation in a porous medium takes finally the following −1  (M(H) − M (He )) = −1  (M − M  ). (134)
form by adding up Eqs. (114), (120), (127) and (129):
Finally using Eqs. (131), (133), (134) yields the volume-
j average magnetization relaxation equation
˜ ⊗ ṽ

I   +
I  ∇ · v +
I ∇ ·  
jt j
 
∇  M + ∇ M · v
= 2 ∇ × v + × v − 2 jt

+  Mv (∇ · vv ) + ∇ ·  M̃ ⊗ ṽ
+ 0  M × (H0 + h )
=   × M
+ 0  M̃ × (H̃0 + h̃) +   
˜ × M̃ − −1  (M − M  ).
+   (135)
+ ( +  ) ∇(∇ ·  )
  The volume-average Langevin magnetization vector is cal-
1
+∇ · ˜ ⊗ n d
 culated after taking the volume average of Eq. (38). As a first
|V |  approximation, it was calculated using, as an argument, the lo-

 cal volume-average total magnetic field.
+ (∇ )˜ · n d
|V | 
  4.5. Volume-average Maxwell flux law
1
+ ( +  )∇ ˜ · n d

|V |  Splitting the magnetic fields into the external and induced

( +  ) contributions
+ ˜  d.
(∇ ·  )n (130)
|V |   H =  (H0 + h), (136)

4.4. Volume-average magnetization relaxation equation  H =  (H0 + h) (137)


and substituting Eq. (23) into Eqs. (21) and (22) gives
Eq. (20) undergoes similar term-wise volume-averaging
treatment. ∇ ·  0 (H + M) = 0 ∇ ·  h + 0 ∇ ·  M = 0, (138)
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1641

∇ ·  (1 +  ) 0 H = 0 (1 +  )∇ ·  h = 0. (139) Using the spatial decomposition Eq. (12) for the induced
magnetic field in the granular phase and invoking the geomet-
Using the corollary Eq. (9) of the averaging theorem: rical theorem Eq. (70) in Eq. (148), one gets
 
1 1 
∇ ·  (h + M ) + (h + M) · n d = 0, (140)  ∇ × h − h̃ × n d = 0. (150)
|V |  |V | 

1 Eqs. (26), (149) and (150) represent the volume-average curl
∇ ·  h − h · n d = 0. (141)
|V |  equations to be used in the general volume-average ferrohy-
drodynamic model to represent, respectively, (i) the volume-
Adding Eqs. (140) and (141): average external magnetic field, (ii) the volume-average de-
magnetizing field in the ferrofluid phase, and (iii) the induced
∇ · ( h +  M +  h )
 magnetic field in the granular phase.
1
+ (h + M − h ) · n d = 0. (142) To complete the model, the volume conservation equation is
|V |  required, this is written obviously as
Invoking the discontinuity of the normal components of the  +  = 1. (151)
magnetic field at the ferrofluid–granular interface, boundary
condition Eq. (32), Eq. (142) finally reduces to 5. Closure problem
∇ · ( h +  M +  h ) = 0. (143)
For the sake of compactness, the macroscopic ferrohydrody-
Using the spatial decomposition Eq. (12) for the induced namic model with the unknown closure terms is summarized in
magnetic field in the granular phase and invoking the geomet- Table 1. The resulting volume-average equations contain sev-
rical theorem Eq. (70) in Eq. (141) one obtains eral integrals of the spatially fluctuating microscale terms that
 need to be approximated regarding velocity, spin density, mag-
1  netization, magnetization forces, external magnetic field, and
 ∇ · h − h̃ · n d = 0. (144)
|V |  ferrofluid demagnetizing field and induced magnetic field in
granular phase. Note that this model applies to the points be-
Eqs. (23), (142) and (144) represent the volume-average longing to the packed bed interior domain excluding the vessel
divergence equations to be used in the general volume-average peripheral boundaries, which need appropriate boundary con-
ferrohydrodynamic model to represent, respectively, (i) the ditions to be discussed later. Besides the drag force formula-
volume-average external magnetic field which is equal to the tion (Eq. (92)), formulation of the 14 remaining model closure
local field value at the averaging volume centroid provided terms is a task of formidable complexity, and to the authors’
the materials of the -phase and -phase are homogeneously best knowledge, such closures have not been yet formulated for
distributed, (ii) the volume-average demagnetizing field in the porous media ferrofluid flows.
ferrofluid phase, and (iii) the induced magnetic field in the Assuming as in Schumacher et al. (2003) that spin viscosity
granular phase. shear and bulk coefficients are zero relaxes four closure terms
in the internal angular momentum balance, Eq. (130). Some
4.6. Volume-average Ampère–Maxwell law authors, such as Rinaldi and Zahn (2002a,b) proposed formu-
lations of Couette–Poiseuille flows including the spin viscosity
Using Eq. (26) in Eqs. (24) and (25) yields effects. Therefore, there is no reason that such effects could not
be accounted for in future analyses of the ferrohydrodynamic
∇ ×  H = ∇ ×  h = 0, (145) model.
∇ ×  H = ∇ ×  h = 0. (146) The closures Eqs. (143) and (150) hide in the granular in-
duced magnetic field vector a monumental complexity associ-
Using the corollary Eq. (11) of the averaging theorem: ated with the hydrodynamic perturbation of the ferrofluid dy-
 namic magnetization vector as well as the non-linear character
1
∇ ×  h + h × n d = 0, (147) of the Langevin equation between equilibrium magnetization,
|V |  external and equilibrium demagnetizing fields (38). In the gen-
 eral case, this virtually opens onto a deadlock when attempt-
1
∇ ×  h − h × n d = 0. (148) ing to infer a mean-field effective magnetic permeability of
|V |  the composite ferrofluid–granular system. When fluid and solid
Invoking the continuity of the tangential components of the magnetic permeabilities do not differ much and a LMMs be-
magnetic field at the ferrofluid–granular interface, boundary havior is valid (case of low magnetic field linear-Langevin limit
condition Eq. (31), adding Eqs. (147) and (148), it finally given by Eq. (40)), the effect of contacts between non-magnetic
reduces to grains is not very important and the Maxwell–Garnett theory
can be used (Khuzir et al., 2003). This theory assumes random
∇ × ( h +  h ) = 0. (149) distribution of grains in a carrier medium subject to a local
1642 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Table 1
Complete volume-average ferrohydrodynamic model

Conservation of volume

 +  = 1 (151)

Continuity

∇ ·  v = 0 (72)

Linear momentum
 
j ∇  

 v +
 ∇v · v +
∇ ·  ṽ ⊗ ṽ = −  ∇p + 
g + ( + ) v + ∇v · + v
jt  
 2 
150( + ) 1 −  1.75
1 − 
− + v  v + 2 ∇ × 
d2  d 

1
+  ∇(H0 + h ) · M + 2 n × ˜ d
|V |  
 
1 
+ (H̃ + h̃ ) ⊗ n d · M +  ∇(H̃0 + h̃) · (M̃)
|V |  0 (113)

Internal angular momentum

j

I   +
I  ∇ · v +
I ∇ ·  
˜ ⊗ ṽ
jt
 
∇ 
= +2 ∇ × v + × v − 2 + 0  M × (H0 + h )

+ 0  M̃ × (H̃0 + h̃) +    + ( +  ) ∇(∇ ·  )
  
1 
+  ∇ · 
˜ ⊗ n d + (∇  
˜ ) · n d
|V |  |V | 
  
1 ( +  )
+ ( +  )∇ 
˜ · n d + (∇ ·  
˜ )n d
|V |  |V |  (130)

Magnetization relaxation

j
 M + ∇  M · v +  M (∇ · v ) + ∇ ·  M̃ ⊗ ṽ
jt
=   × M +  ˜ × M̃ − −1  (M − M  ). (135)

Maxwell flux law

∇ · ( h +  M +  h ) = 0 (143)


1
 ∇ · h − h̃ · n d = 0
|V |  (144)

∇ · 0 H0 = 0 (23)

Ampère–Maxwell law

∇ × ( h +  h ) = 0 (149)


1
 ∇ × h − h̃ × n d = 0
|V |  (150)

∇ × H0 = 0 (26)
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1643

Lorentz field. For spheres in densely packed beds with LMMs, Laplacian of an axisymmetrical scalar field:
the previous closures can be replaced by Eq. (45) which relates,
in the volume-average sense, the ferrofluid demagnetizing field j2  1 j j2 
and the induced magnetic field in the granular phase. Due to  = + + 2. (156)
jr 2 r jr jz
lack of better alternatives, we extend in this work the LMM
approximation even to the non-linear region. This artifice al-
lows ignoring the closures associated with Eqs. (143) and (150). Laplacian of an axisymmetrical vector field:
In addition, the solid magnetic susceptibility,  , is neglected
with respect to the initial ferrofluid magnetic susceptibility, 0 , ⎛ ⎞
j2 r 1 jr r j 2 r
+ − +
yielding the LMM approximation correction Eq. (46). In addi- ⎜ jr 2 r jr r2 jz2 ⎟
⎜ 2 ⎟
tion, since the ferrofluid magnetization is the chief determinant ⎜ j  1 j  j2  ⎟
 = ⎜
⎜ jr 2 + − + ⎟. (157)
jz2 ⎟
of both solid induced magnetic and ferrofluid demagnetizing
⎜ r jr r2 ⎟
fields, these should remain small with respect to the external ⎝ j2 z 1 jz j2 z ⎠
field to attempt further simplifications to be discussed next. + +
Closure of the eight remaining terms entails mathematical com- jr 2 r jr jz2
plications and physical efforts beyond the scope of this study.
Rather, we will restrict to solve the zero-order formulation and In what follows and for alleviating the notation, the phase
leave the complete formulation as an open problem for future indicators will be dropped in the balance equations. The sub-
research. scripts r,  and z refer, respectively, to the radial, azimuthal and
axial (or streamwise) projections of the ferrofluid field vari-
6. Steady-state zero-order axisymmetric volume-average ables unless otherwise stated. Furthermore, boundary condi-
model tions need to be specified at the four peripheral boundaries of
the porous medium: z = 0, z = L, r = 0, r = D/2.
We shall now need to formulate the Cauchy problem of the
zero-order macroscopic ferrohydrodynamic flow for a Müller 6.1. Continuity equation
porous medium to allow numerical implementation and discus-
sion of model outcomes in terms of the effects conveyed by The pressure field has been assumed to depend only on the
the lowest-order average terms. We shall assume without proof, axial coordinate. Therefore, instead of solving the divergence
uniqueness of solution for the system of equations describing equation of the volume-average continuity, it was found to be
the steady-state zero-order volume-average ferrohydrodynamic numerically more tractable to implement the cross-sectional
model. version of the continuity equation since the column walls are
Furthermore, we shall assume that all the variables in the impermeable to fluid flow. Eq. (72) turns thus into
physical problem are steady-state and independent of the az-
imuthal coordinate (axisymmetry). The last hypothesis entails 
that the gradient, divergence, curl and Laplacian operators sim-
D/2 U0 D 2
vz r dr = . (158)
plify to the following forms needed for the projection of the 0 8
equations summarized in Table 1:
Gradient of an axisymmetrical scalar field: Relevant boundary conditions are therefore required for the
  pressure field. We have assumed that the pressure gradient at
j j
∇ = t 0 . (152) the vessel outlet (z = L) is equal to the pressure gradient at the
jr jz contiguous internal node just before outlet. Also, to produce
Gradient of an axisymmetrical vector field: monotonous pressure gradient distribution in the entrance sec-
⎛ j tion, the slope of the pressure gradient curve at the vessel inlet
 jr ⎞
r
−  (z = 0) is assumed equal to the slope at the vicinal bed internal
⎜ jr r jz ⎟ node:
⎜ j r j ⎟
⎜  ⎟
∇ = ⎜ ⎟. (153)
⎜ jr r jz ⎟ d dp
⎝ j jz ⎠ (L) = 0, (159)
z
0 dz dz
jr jz
Divergence of an axisymmetrical vector field:
d2 dp
(0) = 0. (160)
1 jrr jz dz2 dz
∇ ·= + . (154)
r jr jz
Curl of an axisymmetrical vector field: 6.2. Linear momentum, radial component
 
j jr jz 1 jr After dropping the unknown closures in Eq. (113), the radial
∇ ×= − t
− . (155)
jz jz jr r jr projection of the volume-average momentum balance equation
1644 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

becomes 6.4. Linear momentum, azimuthal component



jvr v2 jvr Similarly, for the azimuthal direction

vr − + vz
jr r jz    
jv v jv
 2
j vr 1 jvr vr j 2 vr
vr +  + vz 
= ( + ) + − + jr r jz
jr 2 r jr r2 jz2  2
j v 1 jv v j 2 v
 2
vr d  1 d

1 d jvr
 = ( + ) + − 2 +
+ + + jr 2 r jr r jz2
 dr 2 r dr  dr jr  2  
    v d  1 d 1 d jv
j 0 dH0r jhr jhr +  + +
− 2 + mr + + mz  dr 2 r dr  dr jr
jz  dr jr jz    
 jr jz 0 H0r hr
vr 1 −  + 1− + 2 − + m +
− 150 2 jz jr  r r
d  d  
 v 1 −  + 1−

− 150 2
+1.75 vr2 + vz2 + v2 . (161) d  d 
d 

+1.75 vr2 + vz2 + v2 . (171)


An obvious choice for the boundary conditions for the radial d
velocity component is With the corresponding boundary conditions on the tangential
velocity component:
vr (r, 0) = 0 r ∈ [0; D/2], (162)
v (r, 0) = 0 r ∈ [0; D/2], (172)
jvr
(r, L) = 0 r ∈ [0; D/2], (163) jv
jz (r, L) = 0 r ∈ [0; D/2], (173)
jz
vr (0, z) = 0 z ∈]0; L[, (164)
v (0, z) = 0 z ∈]0; L[, (174)
vr (D/2, z) = 0 z ∈]0; L[. (165) v (D/2, z) = 0 z ∈]0; L[. (175)

6.3. Linear momentum, axial component 6.5. Internal angular momentum, radial component

Similarly, for the axial direction After dropping the unknown closures in Eq. (130), the radial
  projection of the volume-average internal angular momentum
jvz jvz

vr + vz balance equation gives


jr jz  
 2   jr  jr
j vz 1 jvz j 2 vz vz d 2  1 d
I vr − v  + vz
= ( + ) + + + + jr r jz
jr 2 r jr jz2  dr 2 r dr  
   jv
1 d jvz j  = −2 + 2r + 0 m (H0z + hz ). (176)
+ + 2 +  jz 
 dr jr jr r
   
0 dH0z jhz jhz With the corresponding boundary conditions on the radial spin
+ mz + + mr density component:
 dz jz jr
 
vz 1 −  + 1−
r (r, 0) = 0 r ∈ [0; D/2], (177)
− 150 2 + 1.75 vr + v z + v 
2 2 2
d  d  d
r (r, L) : continuity extension of Eq. (176)
dp
− −
g. (166) r ∈ [0; D/2], (178)
dz
r (0, z) = 0 z ∈]0; L[, (179)
With the corresponding boundary conditions on the axial
velocity component: r (D/2, z) = 0 z ∈]0; L[. (180)

vz (r, 0) = Uo r ∈ [0; D/2], (167) 6.6. Internal angular momentum, axial component
jvz  
(r, L) = 0 r ∈ [0; D/2], (168) jz jz
jz
I vr + vz
jr jz
 
jvz v jv v d
(0, z) = 0 z ∈]0; L[, (169) = 2  +  +  − 2z
jr r jr  dr
0
− m (H0r + hr ). (181)
vz (D/2, z) = 0 z ∈]0; L[. (170) 
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1645

With the corresponding boundary conditions on the axial spin 6.9. Magnetization relaxation, axial component
density component:  
jmz jmz vr jvr jvz
z (r, 0) = 0 r ∈ [0; D/2], (182) vr + vz + mz + +
jr jz r jr jz
−1
z (r, L) : continuity extension of Eq. (181) = − mr + r m −  (mz − mz ). (196)
r ∈ [0; D/2], (183) With the corresponding boundary conditions on the axial mag-
netization component:
jz
(0, z) = 0 z ∈]0; L[, (184)
jr mz (r, 0) = 0 r ∈ [0; D/2], (197)
z (D/2, z) = 0 z ∈]0; L[. (185) mz (r, L) : continuity extension of Eq. (196)
r ∈]0; D/2], (198)
6.7. Internal angular momentum, azimuthal component jmz
(0, z) = 0 z ∈]0; L], (199)
  jr
j r j

I vr + v + vz
jr r jz mz (D/2, z) : continuity extension of Eq. (196)
 
jvr jvz vz d z ∈]0; L[. (200)
= 2 − − − 2
jz jr  dr
0 6.10. Magnetization relaxation, azimuthal component
+ (mz (H0r + hr ) − mr (H0z + hz )). (186)
  
jm mr jm vr jvr jvz
With the corresponding boundary conditions on the azimuthal vr + v + vz + m + +
jr r jz r jr jz
spin density component:
= z mr − r mz − −1 m . (201)
 (r, 0) = 0 r ∈ [0; D/2], (187) With the corresponding boundary conditions on the azimuthal
magnetization component:
j2  
(r, L) = 0 r ∈ [0; D/2], (188)
jz2 m (r, 0) = 0 r ∈ [0; D/2], (202)

 (0, z) = 0 z ∈]0; L[, (189) m (r, L) : continuity extension of Eq. (201)


r ∈]0; D/2], (203)
 (D/2, z) : continuity extension of Eq. (186)
z ∈]0; L[. (190) m (0, z) = 0 z ∈]0; L], (204)
m (D/2, z) : continuity extension of Eq. (201)
6.8. Magnetization relaxation, radial component z ∈]0; L[. (205)

Eq. (135) for the magnetization relaxation gives the following 6.11. Maxwell flux law of the demagnetizing field in the
projections: ferrofluid phase
 
jmr m jmr vr jvr jvz
vr − v   + vz + mr + + Combining Eq. (135) with LMM approximation Eq. (45) and
jr r jz r jr jz taking the projections gives
−1
=  mz − z m −  (mr − mr ). (191)
mr jmr jmz
+ +
With the corresponding boundary conditions on the radial mag- r jr jz
netization component:  
hr jhr jhz d
= − + + + (H0r + hr )
mr (r, 0) = 0 r ∈ [0; D/2], r jr jz dr
(192)   
H0r jH0r jH0z
mr (r, L) : continuity extension of Eq. (191) − (1 − )  + +
r jr jz
 
r ∈]0; D/2], (193) hr jhr jhz
+ ( + 1) + + . (206)
mr (0, z) = 0 z ∈]0; L], (194) r jr jz

mr (D/2, z) : continuity extension of Eq. (191) With the corresponding boundary conditions on the radial
component of the demagnetizing field in the ferrofluid phase:
z ∈]0; L[. (195)
hr (r, 0) = 0 r ∈ [0; D/2], (207)
Note the magnetization vector is expressed per unit reactor
volume unlike the rest of the field variables, which are expressed hr (r, L) : continuity extension of Eq. (206)
per unit volume of the phase wherein they are defined. r ∈]0; D/2], (208)
1646 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

hr (0, z) = 0 z ∈]0; L], (209) boundary conditions from the ferrohydrodynamic equations
Eqs. (158), (161), (166), (171), (176), (181), (186), (191), (196),
hr (D/2, z) : continuity extension of Eq. (206)
(201) with their corresponding boundary conditions and above
z ∈]0; L[. (210) simplifying assumptions. Strictly speaking, this additionally re-
quires that the norm condition ∇ H0 · M  ∇ h · M due
6.12. Ampère–Maxwell law, azimuthal
to the Kelvin force terms in Eqs. (161) and (166), must also
be verified; which is often the case except very near the wall,
Combining Eq. (149) with LMM approximation Eq. (45) and
the entrance (positive gradient) and the exit (negative gradi-
taking the projections gives
ent) regions at low magnetic fields. Therefore, we make the
 
jhr jhz d assumption that the dominant Kelvin force contribution comes
(1 + (1 − )) − + (H0z + hz ) = 0. (211) from the applied magnetic field H0 allowing the neglect of the
jz jr dr
demagnetizing self-induced field contribution:
With the corresponding boundary conditions on the axial com-
ponent of the demagnetizing field in the ferrofluid phase: dH0r jhr jhr
mr  mr + mz , (223)
dr jr jz
hz (r, 0) = 0 r ∈ [0; D/2], (212)
dH0z jhz jhz
hz (r, L) : continuity extension of Eq. (211) mz  mz + mr (224)
dz jz jr
r ∈]0; D/2], (213)
in addition to the neglect in the internal angular momentum
jhz equations of the induced fields with respect to the externally
(0, z) = 0 z ∈]0; L], (214)
jr applied field, H0 :
hz (D/2, z) : continuity extension of Eq. (211) H0z  hz , (225)
z ∈]0; L[. (215)
H0r  hr , (226)
6.13. Other simplifications mz H0r − mr H0z  mz hr − mr hz . (227)

The external magnetic field has been chosen to exhibit only The approximations issue will be discussed at length in the
radial and axial dependences (Eq. (47)) while fulfilling the di- end of the next section. For the sake of pedagogy and for ease
vergenceless condition (Eq. (23)), thus of implementation of the model by others, the final 3D ax-
isymmetrical volume-average model in its decoupled hydrody-
H0 = 0. (216) namic/magnetostatic blocks is gathered in Table 2. The bound-
This entrains that the azimuthal component of the equilib- ary conditions for the eight resulting field variables P, vr , vz ,
rium magnetization vector is zero (see also Eq. (201)):  , mr , mz , hr , and hz , are given, respectively, by Eqs. (159),
(160); Eqs. (162)–(165); Eqs. (167)–(170); Eqs. (187)–(190);
m = 0. (217) Eqs. (192)–(195); Eqs. (197)–(200); Eqs. (207)–(210) and Eqs.
(212)–(215).
In addition, the boundary conditions Eqs. (172)–(175) for
v , Eqs. (177)–(180) for r , Eqs. (182)–(185) for z , as well 7. Implementation and discussion
as Eqs. (202)–(205) for m are verified by the trivial set:
r = 0, (218) Table 3 contains the simulation parameters of the simplified
zero-order axisymmetrical volume-average ferrohydrodynamic
z = 0, (219) model. Note that three superficial velocities are chosen such that
the ferrofluid flow is (i) completely controlled by inertia, i.e.,
m = 0, (220)
Burke–Plummer type of flow (U0 = 0.2 m/s); (ii) mostly dom-
v = 0. (221) inated by viscous forces, i.e., Darcy flow (U0 = 2 × 10−5 m/s),
(iii) mixed with comparable inertial and viscous contributions,
These are also solutions for our model, as the set of i.e., Forchheimer flow (U0 =2×10−3 m/s). The Reynolds num-
Eqs. (171), (176), (181) and (201) is fulfilled. As for the equi- ber values, Re, corresponding to these three cases are, respec-
librium magnetization on e direction (Eq. (217)), Eq. (220) is tively,
U0 dp / = 2110, 0.211 and 21.1. The consensual val-
also coherent with H0 = 0, and therefore it is not unrealistic ues for the laminar and inertial Ergun constants are used, i.e.,
to assume that respectively, 150 and 1.75; yielding the modified Ergun drag
h = 0. (222) function in Eqs. (161), (166).
Solution of the coupled partial differential and integral equa-
Preliminary simulations reveal that under certain circum- tions is performed using a finite difference scheme on an Aspen
stances (to be discussed later) h  H0 suggesting a fur- Custom Modeler platform. Discretization meshes along r and
ther simplification of the model towards decoupling the mag- z directions are, respectively, 0.2 and 4 mm. To ensure conver-
netostatic equations Eqs. (206), (211), (45) along with their gence of the numerical scheme, both fast-Newton method for
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1647

Table 2
Simplified zero-order axisymmetrical volume-average model

Decoupled hydrodynamic submodel


Continuity
 D/2
U0 D 2
vz r dr =
0 8 (158)

Linear momentum (radial)


   
jvr jvr j2 vr 1 jvr vr j2 vr vr d 2  1 d 1 d jvr

vr + vz = (  + ) + − 2 + + + +
jr jz jr 2 r jr r jz 2  dr 2 r dr  dr jr
 
j 0 dH0r vr 1 −  + 1−

− 2 + mr − 150 2 + 1.75 vr2 + vz2


jz  dr d  d  d (161)

Linear momentum (axial)


   
jvz jv z j2 vz 1 jvz j2 v z vz d 2  1 d 1 d jvz

vr + vz = (  + ) + + + + +
jr jz jr 2 r jr jz 2  dr 2 r dr  dr jr
   
j  0 dH0z vz 1 −  + 1−
dp
+ 2 + + mz − 150 2 + 1.75 vr2 + vz2 − −
g
jr r  dz d  d  d dz (166)

Linear momentum (azimuthal)

v = 0 (171/221)

Internal angular momentum (radial)

r = 0 (176/218)

Internal angular momentum (axial)

z = 0 (181/219)

Internal angular momentum (azimuthal)


   
j j jv r jv z vz d 

I v r + vz = 2 − − − 2 + 0 (mz H0r − mr H0z )


jr jz jz jr  dr  (186)

Magnetization relaxation (radial)


 
jmr jm r vr jv r jv z
vr + vz + mr + + =  mz − −1 (mr − mr )
jr jz r jr jz (191)

Magnetization relaxation (axial)


 
jmz jm z vr jv r jv z
vr + vz + mz + + = − mr − −1 (mz − mz )
jr jz r jr jz (196)

Magnetization relaxation (azimuthal)

m = 0 (201/220)

Decoupled magnetostatic submodel


Maxwell–flux law (radial component of demagnetizing field, hr )
 
mr jmr jm z hr jh r jhz d
+ + = − + + + (H0r + hr )
r jr jz r jr jz dr
    
H0r jH0r jH0z hr jhr jh z
− (1 − )  + + + ( + 1) + +
r jr jz r jr jz (206)

Ampère–Maxwell law (axial component of demagnetizing field, hz )


 
jhr jhz d
(1 + (1 − )) − + (H0z + hz ) =0
jz jr dr (211)

Azimuthal component of demagnetizing field


h = 0 (222)
1648 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Table 3
Selected numerical and property values used in ferrohydrodynamic model

Anisotropy constant K (kJ/m3 ): 23 Ferrofluid viscosity  (mPa s): 0.2 Relaxation, Néelian N (s): 18.4
Bed height L (cm): 10 Ferrofluid vortex viscosity  (mPa s): 0.07 Second coefficient of viscosity  (Pa s)–
Boltzmann constant k (yJ/K): 13.8 Initial magnetic susceptibility 0 (–): 1.28 Shear/bulk spin viscosity  ,  (kg m s−1 rd−1 ): 0
Demagnetization LMM correction  (–): 0.23 Larmor frequency f0 (GHz): 1 Surfactant density
S (kg/m3 ): 500
Diameter of column D (cm): 2 Magnetic particle density
p (kg/m3 ): 5300 Temperature T (K): 300
Diameter, hydrodynamic (coated) dh (nm): 28 Liquid carrier viscosity 0 (mPa s): 0.1 Volume fraction, hydrodynamic h (%): 23.4
Diameter, magnetic (uncoated) dm (nm): 15 Magnetic dipole moment m̄(MA/m): 0.45 Volume fraction, magnetic m (%): 3.6
Diameter of packing d (mm): 2 Moment of inertia density I (am2 ): 45 Vacuum permeability 0 (N/A2 ): 1.26
Ferrofluid density
(kg/m3 ): 1056 Relaxation, Brownian B (s): 0.83

non-linear solver with convergence criterion on residuals, and (Fig. 5b). The blocking effect of the magnetostatic torque at the
a MA48 linear solver are used. Typical simulation duration is bed exit is more spectacular under positive gradient in compar-
30 min on a 3 GHz CPU and 1.5 Go RAM computer with 10−5 ison to negative gradient (Fig. 5a). In addition, the higher the
for the absolute equation tolerance. magnetic field the more dramatic the reduction in angular ve-
It is noteworthy that the ferrofluid velocity and nanoparti- locity at bed outlet. For example under a positive magnetic field
cle spin density fields are invariant to switching H0 into −H0 gradient of 60 000 Oe/m, the nanoparticle angular velocity un-
because the induced magnetic field in the solid and the demag- dergoes up to a factor 30 reduction near the wall as opposed to
netizing field in the ferrofluid have been neglected in the fer- the H0 = 0 case (Fig. 5b). For negative gradient configurations,
rohydrodynamic block of equations. For consistency purposes, since the magnetic field cancels at bed exit (but not its gradi-
the H0 z-projection is always positive, i.e., bulk-flow oriented ent), the ferrofluid state approaches, for sufficiently long beds,
external magnetic field (invariant polarity in the encompassing the one occurring in the absence of magnetic fields (H0 = 0).
solenoid surrounding the bed). Furthermore, parameter  of the As alluded to in the introduction, the most dramatic reductions
magnetic field Eq. (47) is chosen such that H0z = 0 at the en- in nanoparticle angular velocity take place in the wall region
trance (positive gradient) or at the exit (negative gradient). for r/R approximately between 90% and 100%. This annular
For the simulated conditions, the dot product between the layer corresponds to the region where the unobstructed gross
volume-average fluid vorticity and the spin density vector is wall ferrofluid flow develops because of the largest prevailing
found to be positive (not shown here) regardless of magnetic porosity giving rise to high ferrofluid vorticity. Farther in the
gradient sign. Thus fluid vorticity and nanoparticle spin are bed towards the axis of symmetry, fluid vorticity as well as
corotative. It is convenient to determine now whether the nanoparticle spin velocity are the lowest due to the damping
nanoparticle spins slower (magnetoviscothickening) or faster of the porosity distribution. Therefore, the magnetoviscous ef-
(magnetoviscothinning) than fluid vorticity. Figs. 4a–c plot the fect is expected to act mostly as a skin mechanism in the wall
ratio of the spin density module to the fluid vorticity module region to produce the searched effect of slowing down chan-
for positive (Figs. 4a,b) and negative (Figs. 4c,d) gradients neling alongside the wall.
alongside the wall (r = D/2) and at the bed exit (z = L) for Figs. 6a–c illustrate the effect of Re number and magnetic
maximum applied magnetic field values ranging between 200 field on the radial profile of the axial velocity component of the
and 5000 Oe and Re = 2110. In all cases, as can be seen, mag- ferrofluid at z = L. For Re = 2110, regardless of the magnetic
netoviscothickening is the prevailing mechanism, especially field gradient sign, a damping of velocity in the wall region is
alongside the wall (Figs. 4a,c) which will be used later to ex- systematic. Interestingly, it is the negative gradient magnetic
plain the reduction of the wall flow channeling when the mag- field, which brings about the highest reductions in wall flow.
netic field is on. Also, gradient reversal of the magnetic field This result seems in opposition with the damped behaviors ob-
does not seem to affect the magnetoviscous mechanism (Figs. served earlier for the angular velocity where positive gradients
4c,d). Obviously, when the magnetic field is off, the model were responsible for the largest blockages in spin. To interpret
restores the expectedly equal fluid vorticity and nanoparticle the reduction of wall bypass with negative gradient magnetic
spin. field, one has to invoke rather the Kelvin magnetization body
Figs. 5a,b show the impact of positive and negative magnetic force term appearing in the z-projection linear momentum bal-
field gradients on the nanoparticle angular velocity for varying ance Eq. (166). Since by convention the applied magnetic field
maximum magnetic fields at the selected locations r =D/2 and points in the direction of flow, i.e., upwards, then so does the
z = L, respectively (Re = 2110). As expected, when the mag- magnetization vector. However, the negative gradient magnetic
netic field is off, the nanoparticle spin is the highest in com- field yields a Kelvin force oriented downwards and then adds
parison to the configurations with activated magnetic field due up to the gravitational body force giving rise to hypergrav-
to the damping behavior of magnetoviscothickening. Further- ity conditions in the bed. Here the term hypergravity refers
more, if nanoparticle spin attains fully developed profiles after to the case where the resultant of the body force densities is
a streamwise dimensionless distance z/L=10.20% for H0 =0, larger than the contribution of the gravitational force only. On
the same cannot be said when magnetic field gradients prevail the contrary, hypogravity conditions correspond to a total body
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1649

1 1

0.1 200(+)
0.1
400(+)
600(+)
800(+) 2 ωθ 2 ωθ
1000(+)
∂vr ∂v z v z dε ∂vr ∂v z v z dε
3000(+) − − − −
4000(+)
∂z ∂r ε dr r =D / 2 Re = 2110 ∂z ∂r ε dr z =L Re = 2110
0.01 0.01
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(a) z /L (b) r /R

1 1

0.1 200(-) 0.1


400(-)
600(-)
2 ωθ 800(-)
2 ωθ
3000(-)
∂vr ∂v z v z dε ∂vr ∂v z v z dε
− − 4000(-) − −
Re = 2110 ∂z ∂r ε dr r =D / 2 5000(-) Re = 2110 ∂z ∂r ε dr z =L
0.01 0.01
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(c) z /L (d) r /R

Fig. 4. Ratio of spin density to ferrofluid vorticity (a) positive gradient magnetic field along wall r = D/2, (b) positive gradient magnetic field at bed exit
z = L, (c) negative gradient magnetic field along wall r = D/2, (b) negative gradient magnetic field at bed exit z = L. Re = 2110. Box legend syntax:
A(+) ⇒ H0z = Az/L Oe; A(−) ⇒ H0z = A (1 − z/L) Oe.

force resultant lower than the contribution from the sole grav- symmetry of the bypass with respect to the vertical axis in Fig.
itational force. The artificial hypergravity state gives rise to a 7, which is more severe for negative gradient than for positive
resistive work the ferrofluid must fight against to be able to get gradient, is ascribed to a combinatorial effect of magnetovis-
out of the bed with a result of an improved mitigation of the cosity and Kelvin body force density. Because of the gross ap-
bypass in comparison with the positive gradient case (which proximation made as discussed earlier regarding the neglect of
by opposition yields hypogravity state, see discussion later on the demagnetizing field and its gradient with respect to the ex-
bed pressure drop and wall bypass fraction). It is interesting ternal magnetic field and its gradient under low external mag-
to note that the magnetic fields exert greater influence on the netic fields, the curve in the vicinity of dH0 z/dz = 0 has to be
velocity profiles in the case of Darcy and mixed flows (Re = considered with caution. This point will be revisited later in the
0.21, Re = 21.1) than for inertia-controlled flow (Re = 2110). discussion of the validity of the solution of the magnetostatic
In the Darcy flow case (Fig. 6c), stronger magnetic fields are block of equations (Table 2).
even capable of reversing the flow in the wall region at the Fig. 8 shows the outlet pressure gradient normalized with
bed outlet. respect to the pressure gradient at H0 = 0 for the conditions
Fig. 7 illustrates the evolution of the fractional bypass nor- simulated in Fig. 7. Although the negative gradient region co-
malized with respect to the bypass when H0 = 0 (i.e., external incides with the expected magnetoviscothickening, the positive
magnetic field off) as a function of the z-component external gradient region seems to contradict Fig. 6 findings since pres-
magnetic field positive and negative gradients for Re = 2110 sure drops in the positive gradient branch are predicted to be
and 21.1. Recall that the bypass is calculated using Eq. (57). lower than when H0 = 0. This result is easily explained by
Fig. 7 is a confirmation of a magnetoviscothickening effect at the fact that spatially inhomogeneous magnetic fields give rise
the expense of magnetoviscothinning, which has not been ob- to a Kelvin body force, which produces work. In the positive
served for the configuration simulated in this study. Hence, gradient case, the Kelvin body force z-component is oriented
magnetoviscosity tends to reduce the preferential flow along- streamwise thus creating conditions of hypogravity. As a re-
side the wall. The maximum reduction could be as low as 18% sult of this magnetic levitation, pressure gradient diminishes.
at the highest negative gradient for Re = 2110 and up to 60% at This reduction in pressure gradient can be dramatic, especially
Re = 21.1 in the positive portion (inset in Fig. 7). At Re = 0.21 at low Reynolds numbers where up to a sevenfold reduction
(not shown here), because of flow reversal the drop in bypass could be attained (inset in Fig. 8). On the contrary, when the
can be as low as 760% of the bypass value at H0 = 0. The non- magnetic field gradient is negative, the Kelvin body force acts
1650 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

4000

3000

2000

ωθ rd / s 1000

6000(+)
-1000
1000(+)
0
-2000 1000(-)
6000(-) Re = 2110
-3000
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(a) r /R

4000

3500
Re = 2110
3000 6000(+)
1000(+)
2500 0
ωθ rd / s

1000(-)
6000(-)
2000

1500

1000

500

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(b) z /L

Fig. 5. Spin density profiles and positive/negative gradient magnetic field at (a) bed exit z = L, (b) along wall r = D/2. Re = 2110. Box legend syntax:
A(+) ⇒ H0z = A z/L Oe; A(−) ⇒ H0z = A (1 − z/L) Oe.

counterstreamwise to the flow like gravitation and opposes to nent, mr , also points inwardly (Fig. 9c). Qualitatively, hr vari-
the ferrofluid from exiting the bed. ations are to put into perspective in comparison with the vari-
A set of simulations has been carried out for Re = 2110, a ations exhibited by mr , (Fig. 9c) in the one hand, and H0r in
positive magnetic field gradient of 2000 Oe/m, and contour plots the other hand. Furthermore, the radial velocity component is a
over the whole bed domain have been obtained for the radial two-dimensional field that evolves crescendo and in a wave-like
(Fig. 9a) and axial (Fig. 9b) components of the self-consistent manner from bed center to wall thus discharging the ferrofluid
demagnetizing field in the ferrofluid, the radial (Fig. 9c) and material outwardly with the less desirable effect of worsening
axial (Fig. 9d) components of the dynamic magnetization, and the bypass as more fluid is being advected near the wall region.
the radial (Fig. 9e) velocity component. The axial component of the induced magnetic field, hz , fol-
The radial component of the induced magnetic field, hr , at- lows the trend of the z-projection imposed magnetic field H0z
tains successively its maximum (outwardly directed) and min- (Fig. 9b). Further, the magnetization z-component (Fig. 9d)
imum (inwardly directed) values in the neighborhood of the points as expected in the direction of the imposed magnetic
wall where the most significant spin and vorticity gradients take field. Interestingly, while hr and mr may exhibit different signs,
place. Since the radial component, H0r , of the external magnetic hz and mz are of the same (positive) sign. This result seems
field is always inwardly directed (Eq. (47) with =2000 Oe/m), odd since demagnetization has a tendency to attenuate the im-
the induced magnetic field enhances the total radial magnetic pact of applied magnetic fields. The fact that magnetization and
field in the negative maxima region. However, since the mag- induced magnetic field components are involved with an atypi-
netization vector follows more or less the magnetic field direc- cal strongly radially varying porosity function via the Maxwell
tion, it is not surprising that the radial magnetization compo- flux and the Ampère–Maxwell equations (Eqs. (206), (211))
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1651

1.6 0.02
6000(+) Re = 2110 H0z ≥ 0
1.4 2000(+)
0
0
1.2 2000(-)
6000(-) -0.02
1 Re = 2110

[ BPw ( H ) - BPw (0)]/ BPw (0)


νz m / s

0.8 -0.04

0.6 -0.06
0
0.4
-0.08
-0.1
0.2 Re = 21.1
-0.1 -0.2
0 -0.3
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 -0.12
(a) r/ R -0.4

-0.14 -0.5
0.014
0
Re = 21.1 -0.6
200(+) -0.16
0.012 1000(+) -0.7
5000(+) dH 0z / dz (Oe /m) 0 20000 40000 60000
-0.18
0.01 -60000 -40000 -20000 0 20000 40000 60000

0.008 Fig. 7. Algebraic reduced bypass fraction for various z-component external
νz m / s

magnetic field gradients. Re = 2110 and 21.1.


0.006

0.004
0.6 0
0.002 -1
0.5 Hypergravity Re = 21.1
-2
[-dP/dz (H) + dP /dz (0)] / (-dP/dz (0))

0 0.4 -3
0 0.2 0.4 0.6 0.8 1
(b) r/ R -4
0.3 -5
0.0018 -6
0 0.2
Re = 0.211 -7
200(+)
1000(+) 0.1 Re = 2110 -8
0.0012
5000(+) 0 20000 40000 60000
0

0.0006 Hypogravity
-0.1
νz m / s

-0.2
0
-0.3
dH 0z / dz (Oe /m)
H0z ≥ 0
-0.4
-0.0006
-60000 -40000 -20000 0 20000 40000 60000

-0.0012
Fig. 8. Algebraic reduced outlet pressure gradient for various z-component
0 0.2 0.4 0.6 0.8 1 external magnetic field gradients. Re = 2110 and 21.1.
(c) r/ R

Fig. 6. Axial velocity component radial profiles and positive/negative gradient


magnetic field at bed exit z = L and various ferrofluid Reynolds numbers (a) Figs. 10a,b are, respectively, the radial (at z = L) and the
Re =2110 (b) Re =21.1, (c) Re =0.21. Box legend syntax: A(+) ⇒ H0z =A axial (at r = D/2) profiles plotting the ratios between ap-
z/L Oe; A(−) ⇒ H0z = A (1 − z/L) Oe.
plied and induced magnetic field components, |H0z / hz | and
|H0r / hr |, and the associated applied and induced Kelvin body
may be one reason for that behavior. Another reason could be force projections |mr dH0r /dr/(mr jhr /jr + mz jhr /jz)| and
the different normalizing volumes for the magnetization vector |mz dH0z /dz/(mz jhz /jz + mr jhz /jr)| at a positive linear-
(unit reactor volume) and induced magnetic field vector (unit gradient magnetic field (50 000 Oe/m) and Re = 2110 for
ferrofluid volume). illustration.
Let us now turn our attention to a discussion of the validity Except alongside the wall over the first 20% steamwise dis-
of the approximations given by Eqs. (223)–(227) that permit tance, it can be seen that the neglect of the induced magnetic
to decouple in the one hand the problem into two separated field components with respect to the corresponding applied
hydrodynamic and magnetostatic sub-models (Table 2), and to magnetic field counterparts is usually a good approximation
procure, in the other hand meaningful estimates of the velocity, (ratios between 10 and 100 both for |H0z / hz | and |H0r / hr |.
spin, magnetization and induced fields. This procures a posteriori validation of the inequalities
1652 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Fig. 9. Typical (r, z) contour plots of (a) radial component induced magnetic field, (b) axial component induced magnetic field, (c) radial component dynamic
magnetization, (d) axial component dynamic magnetization, (e) radial velocity component. Re = 2110, H0z = 2000 z/L in Oe (positive gradient).
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1653

10000
1000
z /L = 1, 5000(+)
r /R = 1, 5000(+)

1000
100

100 10

10 1
H0z/hz
H0z/hz
H0r/hr H0r/hr
axial component:Kelvin(external)/Kelvin(induced) axial component:Kelvin(external)/Kelvin(induced)
radial component:Kelvin(external)/Kelvin(induced) radial component:Kelvin(external)/Kelvin(induced)
1 0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(a) r/ R (b) z/L

10000 10000
H0z/hz H0z/hz
H0r/hr H0r/hr
axial component : Kelvin(external)/Kelvin(induced) axial component : Kelvin(external)/Kelvin(induced)
radial component : Kelvin(external)/Kelvin(induced) radial component : Kelvin(external)/Kelvin(induced)
1000 1000
z /L = 1, 200(+) r/R = 1, 200(+)

100 100

10 10

1 1

0.1 0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(c) r/R (d) z/L

Fig. 10. Illustration of orders of magnitudes of induced versus applied magnetic fields, and induced and applied Kelvin body forces at (a) bed exit z = L, (b)
along bed wall r = D/2. H0z = 5000 z/L in Oe (positive gradient) (c) bed exit z = L, (d) along bed wall r = D/2. H0z = 2000 z/L in Oe (positive gradient).

Eqs. (225)–(227) that enabled simplification of the internal lationship between total magnetic field and magnetization vec-
angular momentum azimuthal projection equation to the form tors (Eq. (40)) as well as a simple proportionality between the
given by Eq. (186). However, one has to notice that these induced fields in the granular phase and in the ferrofluid phase
approximations become less and less valid the lower the mag- (Eq. (45)). These conditions, if met, are to occur under low
netic field especially along the wall and in the entrance region magnetic fields in the low-magnetization limit (linear Langevin
(positive gradient) or exit region (negative gradient). equation Eq. (40)). This exercise clearly shows that the approx-
Regarding the approximations made with respect to the imations given by Eqs. (223)–(227) are justifiable at best under
Kelvin body force, it can be seen that at the bed exit or along high magnetic fields where unfortunately the LMM approxima-
the wall, for positive gradient and strong magnetic field, the tion becomes questionable, indeed. Therefore, at high magnetic
condition given by Eq. (224) is largely fulfilled for the axial fields (and therefore high magnetic field gradient here), our de-
Kelvin force component proving the assumption post facto. coupling approach provides reasonable estimates of the veloc-
However, the approximation given by Eq. (223) for the radial ity and spin density fields as well as perhaps the magnetization
Kelvin force component is not met systematically and is par- vector field, but surely not for the induced magnetic fields. At
ticularly difficult to achieve near the wall (Figs. 10a,b). This low magnetic field, the gradients of the induced magnetic fields
situation may worsen the lower the magnetic fields jeopardiz- are no longer negligible with respect to the gradients of the
ing thus the validity of the decoupling between hydrodynamics applied magnetic fields turning the ferrohydrodynamic model
and magnetostatics (Figs. 10c,d). into a strongly coupled problem. Therefore even the obtained
Recall that illustration of the induced magnetic field effects hydrodynamic solution is likely not to be valid here. Since the
is a pure academic gesture and at most of qualitative value due magnetization vector field solution of the hydrodynamic sub-
to the use of the LMM approximation that supposes a linear re- model may in particular be inaccurate, this propagates into the
1654 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

solutions of the Maxwell equations corrupting the estimation Notation


for the induced magnetic fields in the low-magnetic field do-
main. a coefficient in the Müller porosity model, dimension-
Therefore, as a follow up to this study, additional work on less
the topic is needed with more focus on more rigorous, and b coefficient in the Müller porosity model, dimension-
likely far more demanding, calculations of the demagnetizing less
field effects at low external magnetic fields under the LMM ap- BPw wall by-pass fraction (Eq. (57)), dimensionless
proximation valid conditions. At stronger magnetic fields the C symmetric couple stress dyadic, kg/s2
problem is even harder since better closures for Eqs. (144) and
d grain diameter, m
(150) are required for the fluctuating granular and ferrofluid
dh hydrodynamic diameter of coated nanoparticle, m
phase induced magnetic fields. In addition, proper formulation
dm nanoparticle diameter of magnetic core, m
of the closure terms involving the fluctuations of spin density
D column diameter, m
and magnetization fields is needed. Finally, experimental data
E antisymmetric orientation triadic, dimensionless
will be required to corroborate or invalidate the modeling ap-
proach chosen in this study to highlight the wall bypass prob- F0 Larmor frequency, Hz
lematic in low column-to-grain diameter ratio porous media g Gravity vector, m/s2
systems. G body couple density vector, Pa
h self-consistent demagnetizing field vector induced in-
side the magnetized ferrofluid matter and the induced
magnetic field in the non-magnetic granular phase,
8. Concluding remarks A/m
he equilibrium demagnetizing field vector prevailing in
We have developed a theoretical framework based on the the ferrofluid phase at magnetohydrostatic conditions,
application of the volume-averaging theorems in multiphase A/m
porous media to analyze the flow of ferrofluids in a special class H dynamic total magnetic field vector,A/m
of porous media presenting pronounced effect of wall bypass He magnetohydrostatic total magnetic field vector, A/m
flows. A simplified upscaled zero-order three-dimensional ax- H0 externally applied magnetic field vector, A/m
isymmetrical ferrofluid model for general Forchheimer flows I moment of inertia density of the magnetic nanoparti-
has been built and analyzed assuming isothermal incompress- cles, m2
ible Newtonian suspension of unassociated magnetic nanopar- K Boltzmann constant, J/K
ticles experiencing bulk-flow oriented linear-gradient magnetic K anisotropy constant of the single-domain uniaxial fer-
fields. Regardless of the gradient sign, the ferrofluid flow has romagnetic nanoparticle, J/m3
been shown to exhibit only magnetoviscothickening with coro- L column length, m
tative vorticity and spin density. The presence of a Kelvin body m̄ magnetic dipole moment of an individual nanoparticle,
force due to the magnetic field gradients yielded both hyper- A/m
gravity and hypogravity conditions with an ensuing effect on M ferrofluid dynamic magnetization vector, A/m
the pressure drop of being, respectively, enhanced beyond or M ferrofluid magnetohydrostatic magnetization vector
reduced below the pressure drop when the magnetic field is (Langevin law), A/m
off. In all cases, however, owing to magnetoviscothickening, M̂ magnetic stress tensor whose divergence retrieves the
a damping of the wall bypass has always been observed. Sev- Kelvin magnetization body force density, Pa
eral simplifying assumptions were used to decouple the prob- n normal canonical unit vector associated with contour
lem into a hydrodynamic sub-model and a magnetostatic sub-  , dimensionless
model, including (i) the neglect of several closure terms of the p pressure, Pa
spatial fluctuations of spin density, magnetization and induced r radial coordinate in the column (axisymmetry), m
magnetic fields in the ferrofluid and the granular phase, (ii) the r position vector locating a point in V, m
use of the linear magnetic material approximation. Post facto R column radius, m
analysis of the resulting order of magnitudes allowed conclud- t time, s
ing that the version of the problem solved provides reasonable T absolute temperature, K
estimates of the velocity field, the spin density and magnetiza- T pressure–viscous-spin/vorticity stress tensor, Pa
tion fields only under high magnetic fields. Estimation of the in- U0 ferrofluid superficial velocity (inlet), m/s
duced magnetic fields through Maxwell equations under LMM v ferrofluid local instantaneous velocity vector, m/s
approximation could be inaccurate. At low magnetic field, the V bounded open set representing the averaging region,
simulation indicates that none of the field variables can be es- dimensionless
timated properly and therefore additional conceptual as well as V̄ adherence of the open set V, dimensionless
experimental work is needed to clarify the value of the approx- |V | volume of the averaging region V, m3
imation and to highlight more pertinent closure equations for V open set region with in V occupied by the ferrofluid,
the fluctuating terms. dimensionless
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1655

V̄ adherence of the open set V , dimensionless 0 initial magnetic susceptibility, dimensionless
V̄ union of a finite number of open sets within V occupied  magnetic susceptibility of the non-magnetic material
by porous media (granular), dimensionless making up the granular phase, dimensionless
|V | volume of V , m3  spin density vector, rd/s
x position vector locating in the macroscale frame, the Subscripts
centroid of the averaging volume region V, m
y relative position vector in the microscale frame of a r radial projection of variable
point in V, m z axial projection of variable
z axial coordinate in the column, m  phase related to ferrofluid
 phase related to granular porous media
Greek letters  azimuthal projection of variable

 gradient programming of the applied magnetic field, Tensor operations


A/m2 t transpose
 offset constant of the applied magnetic field, A/m √
· Euclidian norm (for vectors):  =  =  · 
 ferrofluid-solid interface inside V, dimensionless ⊗ tensor product
jV V associated entrance/exit contours, dimensionless : double contracted product
jVe -phase associated entrance/exit contours, dimension- · dot product or contracted product
less × curl product
jVe -phase associated entrance/exit contours, dimension- ∇ gradient
less ∇· divergence
b coefficient in the Müller porosity model, dimension- ∇× curl
less  Laplacian
 porous media holdup, dimensionless
 vortex viscosity, Pa s Tensor
 dynamic viscosity, Pa s  scalar
 shear spin viscosity, kg m s−1 rd−1  vector
0 dynamic viscosity of the liquid carrier, Pa s  dyadic
 second coefficient of viscosity, Pa s  triadic
 bulk spin viscosity, kg m s−1 rd−1
 -phase characteristic function, dimensionless
 -phase characteristic function, dimensionless
 demagnetization LMM correction, dimensionless

ferrofluid suspension density, kg/m3 Appendix A.

s dressing surfactant density, kg/m3

P magnetic core density, kg/m3 Assuming that ∇  is integrable in V, the spatial averaging
 effective relaxation time of magnetization, s theorem formally states, in terms of the superficial quantities,
B Brownian relaxation time, s as (Whitaker, 1999; Drew and Passman, 1999):
N Néelian relaxation time, s
h hydrodynamic volume fraction of the surfactant- 
dressed nanoparticles, dimensionless 1
∇ |x+y |x = ∇ |x + |x+y n d
m suspension volume fraction of magnetic material em- |V | 
bedded in the nanoparticles, dimensionless
 bounded piecewise continuous scalar field variable,
dimensionless 
˜ spatial fluctuation of the bounded piecewise contin- (7): ∇ |x+y |x = ∇ |x + 1
|V |  |x+y ⊗ n d
uous scalar field  around the intrinsic average in
V ∪  , dimensionless
Proof. Let us write vector  in the Cartesian canonical base
 bounded piecewise continuous vector field variable,
(e1 , e2 , e3 ) of IR3 as
dimensionless
 -phase-side values of the variable  on  , dimen-
sionless
 = t (1 2 3 ) where ∀i ∈ {1, 2, 3},
 -phase-side values of the variable  on  , dimen-
sionless
i verifies
˜ spatial fluctuationof the bounded piecewise continuous 
vector field  around the intrinsic average in V ∪  , 1
∇ i |x+y |x = ∇ i |x +  |x+y n d
dimensionless |V |  i
1656 F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657

Hence,
(9) : ∇ ·  |x+y |x

∀(i, k) ∈ {1, 2, 3}2 , =∇ ·  |x + 1
|x+y · n d
|V | 
∇ i |x+y |x · ek

1
= ∇ i |x · ek +  |x+y n d · ek Proof. According to Eq. (A.1), we obtain
|V |  i
∀(i, k) ∈ {1, 2, 3}2 , ∀(i, k) ∈ {1, 2, 3}2 ,
∇ i · ek |x+y |x !  "  
 j i   j i   1
  =  +  |x+y n,k d
= ∇ i  · ek |x +
1
 |x+y n · ek d jxk x+y  jxk x |V |  i
|V |  i x

∀(i, k) ∈ {1, 2, 3}2 ,


!  " Hence,
j i  
 
jxk x+y  ∀i ∈ {1, 2, 3},

x
 !  "  
j i   1 j i   j i   1
=  +  |x+y n,k d. (A.1)   =  +  |x+y n,i d.
jxk x |V |  i jxi x+y  jxi x |V |  i
x

These nine (i, k)-indexed equations can be written in a tensor


form as in Eq. (7).  Writing previous equation for i = 1, 2 and 3, and adding, we
obtain
 ! 
(8): ∇ |x+y |x = ∇ |x + 1
|x+y ⊗ n d  "
|V |  
3
j i   
3
j i  
  = 
jxi x+y  jxi 
i=1 x i=1 x
Proof. Let us write tensor  as a (3, 3) matrix in the Cartesian 

3
1
canonical base (e1 , e2 , e3 ) of IR3 as +  |x+y n,i d
|V |  i
 i=1
1,1 1,2 1,3
 = 2,1 2,2 2,3 ! 3  " 
3,1 3,2 3,3  j i   
3
j i  
  = 
where ∀i, j ∈ {1, 2, 3}2 , i,j verifies jxi x+y  jxi 
i=1 x i=1 x
  
3
1 1
∇ i,j |x+y |x = ∇ i,j |x +  |x+y n d + i |x+y n,i d,
|V |  i,j |V | 
i=1

Hence,
which demonstrates the result for Eq. (9). 
∀(i, j, k) ∈ {1, 2, 3}3 ,
∇ i,j |x+y |x · ek (10) : ∇ ·  |x+y |x = ∇ ·  |x
 
1 + |V1 |  |x+y · n d
= ∇ i,j |x · ek +  |x+y n d · ek
|V |  i,j
∀(i, j, k) ∈ {1, 2, 3}3 , Proof. According to Eq. (A.2), we get
∇ i,j · ek |x+y |x

1 ∀(i, j, k) ∈ {1, 2, 3}3 ,
= ∇ i,j  · ek |x +  |x+y n · ek d !
|V |  i,j  "  
j i,j   j i,j   1
  =  +  |x+y n,k d
∀(i, j, k) ∈ {1, 2, 3}3 ,
!  " jxk x+y  jxk x |V |  i,j
x
j i,j  
 
jxk x+y  Hence,
x
 
j i,j   1
=  +  |x+y n,k d. (A.2)
jxk x |V |  i,j ∀(i, j ) ∈ {1, 2, 3}2 ,
!  "  
j i,j   j i,j  
These 27 (i, j, k)-indexed equations can be written in a tensor   =  + 1  |x+y n,j d.
form as in Eq. (8).  jxj x+y  jxj x |V |  i,j
x
F. Larachi, D. Desvigne / Chemical Engineering Science 61 (2006) 1627 – 1657 1657

Writing previous equation for j = 1, 2 and 3, and adding, we Borglin, S.E., Moridis, G.J., Oldenburg, C.M., 2000. Experimental studies
obtain of the flow of ferrofluid in porous media. Transport in Porous Media 41,
61–80.
∀i ∈ {1, 2, 3}, Coulson, J.M., Richardson, J.F., Backhurst, J.R., Harker, J.H., 1991. Particle
! 
 " Technology and Separation Processes, vol. 2, fourth ed. Pergamon, UK.
 3
j i,j   Drew, D.A., Passman, S.L., 1999. Theory of Multicomponent Fluids. Applied
 

jxj x+y  Mathematical Sciences, vol. 135. Springer, New York, USA.
j =1 du Trémolet de Lacheisserie, E., 2000. Magnétisme. I—Fondements, vol. 1.
x EDP Sciences; Collection Grenoble Sciences, Les Ulis cedex A, France.
 
 j i,j  
3  3
1 Felderhof, B.U., 2001. Flow of a ferrofluid down a tube in an oscillating
=  +  |x+y n,j d
jxj 
magnetic field. Physical Review E 64 (021508), 1–7.
|V |  i,j Felderhof, B.U., Kroh, H.J., 1999. Hydrodynamics of magnetic and dielectric
j =1 j =1
x
fluids in interaction with the electromagnetic field. Journal of Chemical
! 3 " 
 j i,j 
 Physics 110, 7403–7411.
  3
j i,j   Forchheimer, P., 1914. Hydraulik. Teubner, Leipzig, Berlin, Germany.
  = 
jxj x+y  jxj  Garrigues, J., 2001. Éléments d’algèbre et d’analyse tensorielle à l’usage des
j =1 j =1 x x mécaniciens. Lecture Notes, École Supérieure de Mécanique de Marseille,
 
3
France, <http://esm2.imt-mrs.fr/gar/tenshtml/>.
1 Gray, W.G., 1975. A derivation of the equations for multiphase transport.
+ i,j |x+y n,j d.
|V |  Chemical Engineering Science 30, 229–233.
j =1 Henjes, K., 1995. Maxwell’s equations and vorticity: a note on the viscosity
of magnetic fluids. Journal of Magnetism and Magnetic Materials 146,
These three i-indexed equations can be written in a tensor form L236–L240.
as in Eq. (10).  Ivanov, A.B., Taktarov, N.G., 1990. A study into the filtration of magnetic
fluids. Magnetohydrodynamics 26, 390–392.
Khuzir, P., Bossis, G., Bashtovoi, V., Volkova, O., 2003. Flow of
(11): ∇ ×  |x+y |x = ∇ ×  |x magnetorheological fluid through porous media. European Journal of
 Mechanics B/Fluids 22, 331–343.
+ |V1 |  n × |x+y d Martensyuk, M.A., Raikher, Y.L., Shliomis, M.I., 1974. On the kinetics of
magnetization of suspensions of ferromagnetic particles. Soviet Physics
JETP 38, 413–416.
Proof. According to Eq. (A.1), we obtain McTague, J.P., 1969. Magnetoviscosity of magnetic colloids. Journal of
Chemical Physics 51, 133–136.
∀(i, k) ∈ {1, 2, 3}2 , Müller, G.E., 1991. Prediction of radial porosity distributions in randomly
!  "  
j i   j i   1 packed fixed beds of uniformly sized spheres in cylindrical containers.
  =  +  |x+y n,k d Chemical Engineering Science 46, 706–708.
jxk x+y  jxk x |V |  i Müller, G.E., 1992. Radial void fraction distributions in randomly packed
x
fixed beds of uniformly sized spheres in cylindrical containers. Powder
Hence, Technology 72, 269–275.
Odenbach, S., 2003. Magnetic fluids—Suspensions of magnetic dipoles and
∀(i, k) ∈ {1, 2, 3}2 : their magnetic control. Journal Physics: Condensed Matter 15, S1497–
!  " !  " S1508.
j i   j k   Oldenburg, C.M., Borglin, S.E., Moridis, G.J., 2000. Numerical simulation
  −  
jxk x+y  jxi x+y  of ferrofluid flow for subsurface environmental engineering applications.
x x Transport in Porous Media 38, 319–344.
 
j i   j k   Rinaldi, C., Zahn, M., 2002a. Effects of spin viscosity on ferrofluid flow
=  −  profiles in alternating and rotating magnetic fields. Physics of Fluids 14,
jxk x jxi x 2847–2870.
  Rinaldi, C., Zahn, M., 2002b. Effects of spin viscosity on ferrofluid duct flow
1 1
+  |x+y n,k d −  |x+y n,i d profiles in alternating and rotating magnetic fields. Journal of Magnetism
|V |  i |V |  k and Magnetic Materials 252, 172–175.
!  "
j i j k   Rosensweig, R.E., 1997. Ferrohydrodynamics. Dover Publications Inc.,
 
−   New York, USA.
jxk jxi x+y x Schumacher, K.R., Sellien, I., Knoke, G.S., Cader, T., Finlayson, B.A., 2003.
  Experiment and simulation of laminar and turbulent ferrofluid pipe flow
j i  j k  
= −  in an oscillating magnetic field. Physical Review E 67 (026308), 1–11.
jxk jxi  Shliomis, M.I., 1972. Effective viscosity of magnetic suspensions. Soviet
x
 Physics JETP 34, 1291–1294.
1 Taktarov, N.G., 1980. Motion of magnetizable liquids in porous media.
+ ( |x+y n,k − k |x+y n,i ) d.
|V |  i Magnetohydrodynamics 16, 251–255.
Taktarov, N.G., 1981. Convection of magnetizable fluids in porous media.
Considering previous equation for (i, k) ∈ {(2, 1), (3, 2), (1, 3)} Magnetohydrodynamics 17, 333–335.
Taktarov, N.G., 1983. Flow of magnetizable fluids in conducting porous
demonstrates the result for Eq. (11). 
media. Magnetohydrodynamics 19, 250–255.
Whitaker, S., 1999. Theory and Applications of Transport in Porous Media.
References Kluwer Acadmic Press, Dordrecht, The Netherlands.
Zeuner, A., Richter, R., Rehberg, I., 1998. Experiments on negative and
Bacri, J.C., Perzynski, R., Shliomis, M.I., Burde, G.I., 1995. Negative viscosity positive magnetoviscosity in an alternating magnetic field. Physical Review
effect in a magnetic fluid. Physical Review Letters 11, 2128–2131. E 58, 6287–6293.

You might also like