You are on page 1of 16

Computers and Chemical Engineering 31 (2007) 334–345

CFD phenomenological model of solid–liquid mixing in stirred vessels



Louis Fradette a , Philippe A. Tanguy a, , Franc¸ois Bertrand a , Francis Thibault a ,
Jean-Benoˆıt Ritz b, Eric Giraud b
a
URPEI, Department of Chemical Engineering, Ecole Polytechnique, P.O. Box 6079, Montreal H3C 3A7,
Canada b SME, BP 57, 33166 St-Me´dard-en-Jalles Cedex, France
Received 17 April 2006; received in revised form 18 July 2006; accepted 21 July 2006
Available online 7 September 2006

Abstract
Particle migration in a concentrated viscous suspension subjected to a non-homogeneous shear field was computed using a 3D extension of
the diffusion model of Phillips et al. [Phillips, R. J., Armstrong, R. C., Brown, R. A., Graham, A. L., & Abott, J. R. (1992). A constitutive
equation for concentrated suspensions that accounts for shear-induced particle migration. Physics of Fluids A, 4, 30–40]. The numerical results
were compared to experimental data from the literature for simple flows and to our own data in the case of two helical ribbon based mixing
systems. It is shown that this type of diffusion model, which was developed to predict the behavior of concentrated suspensions dominated by
particle–particle interactions, can predict migration trends but definitely requires additional improvements.
© 2006 Elsevier Ltd. All rights reserved.

Keywords: Mixing; Particle migration; Concentrated suspensions; CFD; Simulations

1. Introduction comparable to that obtained in many polymerization


processes, it would be tempting to use helical ribbon
The makedown of viscous suspensions at high solids con- impellers. In practice, however, this approach is not used by
tent is a unit operation involved in the production of coatings, fear of particle segre- gation in poorly mixed regions and
specialty chemicals, drugs and space fuels. This operation is particle migration under non-uniform shear, which might lead
generally carried out in batch or semi-batch mode in stirred to suspension demixing and therefore mixture quality
vessels. The very viscous nature of the media and the homo- problems. As an example of indus- trial impact, shear-induced
geneity requirements for the resulting mixture impede the use migration of conductive particles is sometimes responsible for
of classical impeller mixers. Hence, adapted mixing technolo- the loss of electrical conductivity in polymer compounds
gies such as planetary mixers with kneading blades (Zhou, (Jana, 2003).
Tanguy, & Dubois, 2000) and multishaft mixers that combine in Since the first observations of particle diffusion under shear
a single apparatus a high-speed dispersing turbine, a pumping deformation by Gadala-Maria and Acrivos (1980), several con-
impeller and a wall scraping blade (Tanguy, Bertrand, Thibault, tributions have led to a better understanding of the hydrody-
& Galy-Jammou, 2001) are common options. From an industrial namic phenomena that induce demixing within concentrated
perspective, however, these mixers are costly to acquire, suspensions (Altobelli, Givler, & Fukushima, 1991; Leighton
operate and maintain. & Acrivos, 1986, 1987a,b). It has been shown in particular
In the polymer industry, it is well known that viscous that the diffusion is related to the particle collision frequency
mixing is best achieved with close-clearance impellers like as well as the variations in the suspension viscosity. In 1992,
helical rib- bon impellers. For the makedown of high solids Phillips, Armstrong, Brown, Graham, and Abott (1992) pro-
content viscous suspensions, as the level of viscosity induced posed a conservation equation for the solid phase that includes
by the high solids concentration and sometimes by the convective transport, collision-induced diffusion, and
suspending matrix itself is viscosity effects. These authors used the Krieger and
Dougherty model (Krieger & Dougherty, 1959) to express the
variation of the sus- pension viscosity as a function of the

Corresponding author. Tel.: +1 514 340 4017; fax: +1 514 340 4105. solid volume fraction. They combined this equation with the
E-mail address: philippe.tanguy@polymtl.ca (P.A. Tanguy). momentum conservation equation for the continuous phase to
investigate the diffusion of

0098-1354/$ – see front matter © 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compchemeng.2006.07.013
L. Fradette et al. / Computers and Chemical Engineering 31 (2007) 334– 33
345
showed that the model predictions are in very good agreement
Nomenclature with the analytical solution of the flow between parallel
aparticle diameter (m) plates. However, the results on the solid volume fraction in the
cimpeller distance to the bottom of the tank core of the eccentric cylinder problem has not been validated.
Dimpeller diameter The model of Phillips et al. (1992) is not the only model that
DTtank diameter can be used to simulate the flow of concentrated suspensions.
Hliquid height in tank In a recent paper by Tanguy, Thibault, Ascanio, and Brito-
Kconsistency index (Pa sn) De La Fuente (2006), the performance of the network-of-zone
Kcempirical constant approach (Mann & Hackett, 1988) used in combination with the
Kμempirical constant virtual finite element method (Bertrand, Tanguy, & Thibault,
nshear-thinning or power-law index 1997) to analyze the complex flow and suspension
Nrotational speed (rev/s) mechanisms in a coaxial mixer was investigated. Experiments
Nmigmigration number carried out at lab scale confirmed the validity of the
predictions. However, this model could not capture the shear-
ppressure (Pa)
induced particle migration, and as a result, the agreement
ttime (s)
between the numerical and the experiments in the intensely
Ttorque (N m)
sheared zones was unsatisfactory.
Vvelocity (m/s)
In all the above approaches, the solid phase is considered
as a continuum. However, it is possible to use direct numerical
Greek letters
sim- ulation to describe the motion of each individual particle
γ˙rate-of-strain
η ρ φ φm Ω tensor (s−1) viscosity (Pa s)
in the liquid phase. The strategy consists of solving the
density (kg/m3) Navier-Stokes equations in the liquid phase while taking into
account the pres- ence of the particles through its physical
particle concentration (v/v)
properties and their motion through Newton’s second law of
maximum particle concentration (v/v) computational domain motion for each par- ticle (Glowinski, Pan, & Perieux, 1994;
Hu, 1996; Hu, Joseph, & Crochet, 1992; Maury & Glowinski,
1997). This approach allows for particle–particle and particle–
fluid interactions within the suspension to be modeled in a
concentrated monodisperse suspensions of PMMA physically correct manner. In practice, due to computer
(polymethyl methacrylate) particles in silicon oil (45 < φ < 55 limitations, the number of particles is limited to a few
vol.%). The particles were non-colloidal (a = 675 µm) and thousand, whereas in a typical mixing applica- tion this
neutrally buoy- ant. Couette and Poiseuille flows were studied. number is several orders of magnitude larger.
It was shown that for these two ‘academic’ flows, a The objective of the present study is to assess the accuracy
concentration gradient was generated in the radial direction. of the model proposed by Phillips et al. (1992) and augmented
Model results were shown to compare well with experimental as per Zhang and Acrivos (1994), to capture the particle
particle concentration pro- files obtained by nuclear magnetic suspend- ing phenomenon and particle migration in solid–
resonance (NMR). liquid mixing. We call this model the shear-induced migration
The phenomenon of particle migration was also modeled in model (SIMM). Three mixing systems will be considered: a
the case of particle resuspension in horizontal ducts, in the propeller and two helical ribbon geometries. Since these
lami- nar and transition regimes (Zhang & Acrivos, 1994). The mixing systems lead to complex three-dimensional flows,
authors complemented the model of Phillips et al. (1992) with a there are no analytical solu- tions to the corresponding flow
flux term to account for particle settling and thus make it fields so that experimental results will be used for validation.
usable for non- neutrally buoyant particles. Here again, the The present work can be consid- ered as a real, asymmetrical
prediction of the solids concentration was compared with three-dimensional application of the model of Phillips et al.
NMR measurements from Altobelli et al. (1991). This study (1992) to a mixing situation. To our knowledge, there does not
highlighted the effect of the Reynolds number and the exist a similar application in the lit- erature, the closest being
suspension concentration on the resuspending mechanism. the one presented by Subia, Ingber, Mondy, Altobelli, and
From a qualitative standpoint, the solid–liquid interface was Graham (1998).
correctly captured by the numerical model.
Particle migration along solid walls was also investigated 2. Experimental systems
(Jana, Kapoor, & Acrivos, 1995). An apparent wall slip veloc-
ity coefficient was determined experimentally, and was shown The first experimental set-up (Fig. 1) consists of a small
to be insensitive to the shear rate while being rather strongly reser- voir (glass beaker) with diameter DT = 7.2 cm and
influenced by the solids content of the suspension. height H = 6.5, and a down-pumping marine propeller with
Finally, a few studies have used the model of Phillips et diameter D = 1.8 cm that provides agitation. The propeller is
al. (1992) to simulate particle migration within concentrated located at a distance c = 2.6 cm from the bottom.
suspensions flowing between parallel plates and in the gap of The investigation of the particle motion inside this mixer
off- centered cylinders. In particular, Fang and Phan-Thien involved a Newtonian solution of corn syrup with a viscosity
(1995) of 1.05 Pa s and a density of 1360 kg/m3. The particles were
red Ballotini glass beads. The spherical beads had a uniform
33 L. Fradette et al. / Computers and Chemical Engineering 31 (2007) 334–
diame- 345
Fig. 1. Dimensions of the marine propeller mixing system (in cm).

ter of 1 mm and a density of 2500 kg/m3. The maximum


random packing concentration (φm) for this particle type is
0.68.
In order to investigate the suspending phenomenon, two
parameters were varied: the propeller rotational speed and the
particle volume concentration. The conditions were as
presented in Table 1.
In each experiment, the particles were initially at rest at the
vessel bottom. Once the agitator was started up, the particles
motion was recorded with time until equilibrium was obtained
on the basis of torque and visual observation of the particles
dispersion.
The other two mixing systems make use of the same
D = 27 cm diameter ribbon. Two tank diameters were used in
the study, providing a wide clearance mixing system (impeller-
to- tank diameter ratio, D/DT = 0.4) and a close clearance
mixing system (D/DT = 0.87). Impeller pitch was equal to 1 in
both cases. A schematic of the close clearance system is given
in Fig. 2. The wide clearance system differs only by the tank Fig. 2. Small clearance mixing system with an impeller-to-tank diameter ratio
diameter which was 65 cm compared to the 31 cm of the close of 0.87.
clearance system. The counterclockwise rotational speed was,
respectively, 50 and 16 rpm for the wide and close clearance
3. Numerical model
impellers. The close clearance mixer was instrumented with a
torque meter and a speed encoder directly mounted on the
For the prediction of shear-induced particle migration, the
actu- ated shaft to allow the real-time measurement of torque
SIMM was used, which states that, in a viscous concentrated
during the mixing process. The fluid was a suspension of
sus- pension, small but non-Brownian particles migrate from
monodisperse spherical glass beads in polybutadiene, a
regions of high shear rate to regions of low shear rate, and
slightly shear-thinning fluid obeying a power-law model:
from regions of high concentrations to regions of low
−1 concentrations in addi- tion to which settling by gravity is
η(γ˙ ) = K|γ˙ |n , (1)
added. In the case of a mixing process, owing to the action of
where γ˙ is the shear rate, K the consistency index, and n is the shear and inertia, the particles may segregate and demix,
power-law index. For the entire experimental program, several thereby generating concentration gra- dients in the vessel. This
glass bead diameters ranging between 0.1 and 3 mm were con- shear-induced migration phenomenon can be simulated either
sidered. The properties of these glass beads and those of the at the microscopic scale, in which case the particles are
suspending solution of polybutadiene are given in Table 2. considered as interacting entities making up the dispersed
phase (Brady & Bossis, 1985) or at the macroscopic
Table 1
Experimental conditions for the glass beaker agitation system
Table 2
Physical properties of the solid–liquid system used in the helical ribbon mixers
Experiment Rotational speed (rpm) Bead concentration (vol.%)
Specific gravity Rheological parameters
1 173 2.8
2 230 7.1 Glass beads 2.48 –
3 350 11.9 Polybutadiene (PBHT) 0.9 K = 500 Pa sn, n = 0.8
scale, where the suspension is modeled as one continuous where α equals 1.82 and where
phase through a viscosity law (Phillips et al., 1992). If
microscopic scale simulations should yield more accurate Dφ ∂φ
results, they are still today very costly and out of reach for = + v · grad φ (8)
Dt ∂t
predicting the behav- ior of industrial process suspensions. For
is the material derivative.
this reason, the motion of particle suspensions is accounted for
In Eq. (7), the solid and liquid phases are treated as a
in this work using a macroscopic model.
con- tinuum and the viscosity helps keep the particle
Because it was only slightly shear-thinning as shown in
concentration φ below the selected maximum concentration
Table 2, the fluid was assumed to be a Newtonian suspension
φm. As noted by Fang and Phan-Thien (1995), numerical
of monodisperse spherical particles in the numerical model.
problems are some- times encountered in the low shear
Note that, as evidenced by the work of Kim (2001), the
zones with particle con- centrations close to − φM. Indeed, the
rheological properties of a dispersing medium can sometimes
have a signif- icant effect on the flow behavior. The governing term (φM φ) then tends
equations were as follows: towards zero, which can generate numerical instabilities in
these zones. These problems can be further magnified if, as
∂v noted by Rao, Mondy, Sun, and Altobelli (2002), and Rao,
ρm . + v · grad vΣ + grad p − div(ηm grad v) = 0, (2)
∂ Mondy, Baer, Al t obell i , a nd S t e ph en s ( 20 02 ) , the
t hindered settling velocity does not go down to zero when the
div v = 0, (3)
volume fraction reaches φM. The model in Eq. (7) is empirical
where ρm = φρparticle + (1−φ)ρliquid denotes the density of the since constants Kc and Kμ have been determined from
suspension and where ηm, its viscosity, is modeled through the experimental measurements of the concentration profiles in
Krieger–Dougherty correlation (Krieger & Dougherty, 1959): Couette and Poiseuille flows (Hampton, Mammoli, Tetlow, &
Altobelli, 1997; Phillips et a l . , 1992). It must be
remembered that they have been obtained for concen- trations
ranging between 45 and 55 vol.%. Since their respective

ηm = ηliquid . Σ
φ −1.82
values are almost identical for both flow types, they are
1− m . (4) generally considered to be constant. These same values were
φ used in all the simulations in this work, that is Kc = 0.41 and
Kμ = 0.62. Never-
In this equation, ηliquid represents the viscosity of the theless, Tetlow et al. (1998) showed that Kc could be related to
suspending fluid, φm the maximum particle concentration in the local concentration and physical characteristics of the solid
the suspension and φ the particle concentration that can be particles. These authors also observed a dependence of the
computed by solv- ing: con- centration as the cube of the particle diameter. Because
dφ they were experimental data developed in one dimension in
2 2 the case of a wide-gap Couette apparatus, these results were
+ v · grad φ = a Kc div(φ grad|γ˙ | + φ|γ˙ | grad φ)
dt . 2 Σ not included in the present work.
2 φ |γ˙ | ∂ηm Boundary conditions are required for mathematical well-
+ a Kμ div grad φ , (5) posedness. As no particle can penetrate or exit the domain, the
ηm ∂φ
sum of fluxes through the wall and the free surface is zero, that
where |γ˙ | denotes the magnitude of the rate-of-strain tensor, a is: Σ
the particle diameter, and Kc and Kμ are model constants that Σ. αφ2
2
account for the particle/particle and fluid/particle interactions, n· a2 Kμ | + a Kc φ| grad φ
respectively. γ˙ | φM − γ˙ |
φ
Σ
Following along the lines of Zhang and Acrivos (1994), an
n 2 2
extra term is added to express the sedimentation flux by means + vsφ(1 − φ) RZ + a Kcφ grad| . (9)
of the settling velocity of the non-buoyant particles owing to γ˙ |
gravity: As for the initial conditions in the tank volume (Ω), two
options were considered: all the particles are at rest and
vφ = vs(1 − φ) , nRZ
(6)
located at the tank bottom (Ωbottom) or are evenly dispersed
where vφ is the classical hindered settling velocity as defined according to the average concentration. In the first case, the
by Richardson and Zaki (1954). Unlike in Zhang and Acrivos solid volume fraction corresponds to the maximum volume
(1994), the exponent nRZ was kept constant in the computations. fraction (φm):
Its value depends on the size of the settling particles and
ranges from 4.5 for a = 100 µm, to 3 for a = 700 µm. φt=0 = φm in Ωbottom, (10)
Combining (4) and (5) leads to the following mass and
conservation equation for
Ω
the solid phase: φt=0 = 0 in Ωbottom , (11)
Σ Σ
Dφ 2 αφ2 2 In the second case:
φt=0 = φavg in Ω, (12)
+ div[a Kc φ grad|γ˙ |] + div[v
2 2
s φ(1 − φ)
nRZ
], (7)
so that the solids mass conservation condition over the whole
domain Ω reads as:
nelt
Σ Vj
φ¯ = j
(13)
V
φt=0 t

j=
1
where Vt represents the total volume, Vj the volume of the
finite element Ωj, φt=0 the initial concentration in this element,
j
and ‘nelt’ the number of elements.
Problems (2), (3) and (7) were discretized using the finite ele-
ment method. Velocity and pressure were approximated with
the stable quadratic discontinuous-pressure P2+–P1 tetrahedral
ele- ment (Bertrand, Gadbois, & Tanguy, 1992). Such a second-
order finite element was required∇so that γ˙ in (7) could be
computed accurately. Concentration φ was approximated by
means of lin- ear P1 tetrahedral elements.
The solution of coupled Eqs. (2), (3) and (7) was obtained in
a decoupled fashion using a Picard-like successive substitution
algorithm. Given a concentration field, mixed problem (2)
and
(3) was solved for velocity and pressure by means of a
conjugate- gradient based Uzawa algorithm (Bertrand &
Tanguy, 2002). Given the velocity field, the advection-
diffusion Eq. (7) was solved for concentration using the
GMRES algorithm (Saad, 1985). This strategy was iterated
upon until a stable solution was reached. Transient simulations
Fig. 3. Contours of the particle concentration in an axially symmetric 4:1
were performed in all cases and, whenever a steady-state
expan- sion. Concentration maps of the exit flow for (a) 500 µm and (b) 100
solution existed, it was obtained as the asymptotic time
µm particles at a 50 vol.% average concentration. (Experimental data from
solution. Altobelli et al., 1997, reproduced with permission from Journal of Rheology.)
The solution of (7) requires some discussion since in some
convection-dominated situations, reaching a numerically satis-
In the present work, the validation is extended to the case
fying solution may be difficult. The Pe´clet number (Pe)
of a 4:1 sudden expansion, as shown in Fig. 3. For this simple
is a dimensionless group that expresses the importance of
convec- tive mass transfer with respect to conductive mass geometry, no analytical solution exists to describe the velocity,
transfer. For the problem at hand, it is the inverse of the so- pressure, and concentration profiles. In the expansion, the flow
called migration number (Ritz, Bertrand, Thibault, & Tanguy, is directed from right to left. The NMR experimental concen-
2000):
Nmig a trations are those of Altobelli, Fukushima, and Mondy (1997)
= . Σ2, (14)
and correspond to an average particle concentration of 50 vol.
D %.
which shows that shear-induced particle migration (diffusion) are successfully compared to the analytical solutions of Couette
is proportional to the square of the particle diameter compared and Poiseuille flows (Ritz et al., 2000).
to a characteristic length. It must be noted that Nmig appears
explicitly in the right hand side of Eq. (7) when expressed in
dimensionless form.
When diffusion dominates (Nmig 1),≥ the problem is ellip- tic
and a standard Galerkin method is adequate for solving the
conservation equation. However, when transport by convec-
tion becomes predominant, Eq. (7) becomes hyperbolic and
the Galerkin method is then known to generate solutions with
spu- rious oscillations (Johnson, 1992). In this work, a stable
and consistent SUPG method (Grygiel & Tanguy, 1991) was
used to compute the solution of (7) for all values of Nmig.

4. Validation

A quantitative validation of the above described method


has been published elsewhere, where the numerical solutions
This imaging technique outputs bitmap images that cannot
be made more precise than what is presented in their paper.
The par- ticle diameter is 500 µm in Fig. 3a, and 100 µm
in Fig. 3b. One can readily note that the model can capture
the salient features of such flow. In the case of the 500 µm
particles, the depletion in concentration in white and light
grey along the solid wall where the shear is important and
the transport of this low concentration downstream from
the expansion are clearly predicted by the sim- ulation.
This is revealed by the drop-shaped pale zone extending
from the corner after the expansion. Moreover, the increase
in concentration along the perpendicular wall behind this
depleted
zone is also captured. This increase in concentration results
from the circulation between the wall and the drop-shaped
zone. A similar particle compaction phenomenon was also
noted both experimentally and numerically in a journal
bearing apparatus (Fang & Phan-Thien, 1995; Tetlow et al.,
1998).
For the 100 µm particles, the drop-shaped zone is clearly
more slender than in the previous case. The very thin
depleted layer at the wall in the reduced-radius section,
again indicated by the pale coloration, is transported
through the expansion and
follows the same path as in the experiment. The different of particles remain dormant at the bottom where the volume
shapes representing the depleted zones generated with the fraction is at its maximum (φm). This proportion of quiescent
small and larger particles are a good indication that the model particles decreases as the rotational speed is increased.
correctly scales according to the size of the particles. Note that The mesh used for the simulations is shown in Fig. 5,
the changes in concentration are fairly steep, which renders the where only the elements on the impeller and the tank surfaces
solution pro- cedure more difficult but feasible with the are dis- played. It contains 19,879 P2+–P1 finite elements for
algorithms described in the previous section. velocity and pressure (249,626 velocity equations) and 19,879
P1 finite elements for the concentration. A Lagrangian frame
5. Particle migration in mixing systems of reference was used for the simulations, which simplified
the imposition of boundary conditions and allowed for the
5.1. Beaker solution of a steady- state problem. Convergence of the
computations was obtained after six coupling iterations of the
For this experimental mixing system, the following qualita- momentum, continuity and concentration equations. The
tive observations about the resuspending mechanism were simulated conditions are identical to the experimental
made and are illustrated in Fig. 4. At start-up, the top layer of conditions presented in Table 1. The phys- ical properties of
solid particles starts sliding. Under the action of the impeller, the dispersed phase are those of Table 2. The 1 mm particle
the par- ticles pile up and are eventually dragged upward by size combined to the characteristic dimension of the mixing
the flow before being propelled to the wall of the beaker. system in the order of 1 cm leads to a migration num-
Some of these particles get suspended in the upper part of the ber in the order of Nmig = 10−2. For this value of the migration
fluid while others settle to the bottom. At low rotational speed, number, it is expected that shear induced migration will play a
a large proportion significant role on the equilibrium concentration field.

Fig. 4. Resuspension mechanism at 230 rpm and 7.1 vol.% of glass beads: (a) t =0 s, (b) t = 25 s, (c) t = 50 s, (d) t = 90 s, and (e) equilibrium.
experimentally. It is interesting to analyze these results based
on the fact that the SIMM predicts that the migration of the
parti- cles occurs from high shear regions towards low shear
regions. The first comment is that the model does not predict
with good accuracy the packing at the bottom of the beaker for
the lower rotational speeds (173 and 230 rpm). As a matter of
fact, the max- imum concentration calculated ranges between
15 and 25 vol.%, far from the theoretical value of 68 vol.%.
However, the model is able to predict the movement of the
particles towards the low shear zone under the impeller. The
calculated concentration field at 350 rpm appears to be in
closer agreement with the experi- mental results. In all cases,
the particles move under the impeller in the low shear zone.
The model also predicts a close to zero concentration at the
free surface, which is also in accordance with our
Fig. 5. Finite element mesh for the axial marine propeller system. observations. Finally, the model gives the correct trend in
predicting a more effective circulation when rpm is increased.
Fig. 6 displays the equilibrium values of the volume frac-
tion and the norm of the rate-of-strain tensor on a mid-plane, 5.2. Close clearance helical ribbon
both obtained from simulations. Also appearing in this figure
are snapshots of the particle distribution in the beaker The mesh of the helical ribbon mixing apparatus is
obtained presented in Fig. 7. The mesh comprises 12,532 P2+–P1 finite
elements for

Fig. 6. Volume fraction of solids at equilibrium for three sets of operating conditions in the beaker: (a) 173 rpm—2.8 vol.% solids; (b) 230 rpm—7.1 vol.% solids;
(c) 350 rpm—11.9 vol.% solids.
velocity and pressure (154,923 velocity equations), and
12,532 P1 finite elements for the concentration. The simulated
case cor- responds to an initial non-uniform particle volume
distribution with 5 vol.% located at the bottom, 21 vol.%
elsewhere in the volume for an average concentration of 26
vol.%. The rotational speed was constant at 16 rpm. The
simulations were performed with neutrally buoyant particles
but gravity was nevertheless considered in the simulations.
The simulated beads had a diam- eter of 100 µm to mimic the
real particles.
The motion of the solid particles is presented in Fig. 8. It
can be seen that the particles are transported by the fluid and
that there are no zones where the particles are trapped. Also
of interest is the resuspending mechanism of the solid particles
that are initially located at the bottom of the tank. These parti-
cles are dragged by the fluid up along the central shaft and
back to the bottom of the tank in the gap region between the
ribbon tip and the tank walls. Migration, if any, in this system
is unim- portant owing to the absence of dead zones and the
Fig. 7. Finite element mesh for the close clearance helical ribbon mixing efficient top-to-bottom pumping features associated with this
appa- ratus. impeller. This situation is totally different from that of the
beaker, and Fig. 8 shows the efficiency of this close-clearance
impeller sys- tem for mixing a viscous suspension. As a matter
of fact, the migration number in the case of the helical ribbon
is in the order

Fig. 8. Contours of the particle concentration in the tank equipped with a close clearance ribbon impeller (D/DT = 0.87; N = 16 rpm; particle size, 100 µm; average
concentration, 25 vol.%).
Fig. 10. Finite element mesh for the large clearance helical ribbon mixing
appa- ratus.
Fig. 9. Experimental and numerical torques vs. time for a close clearance heli-
cal ribbon (D/DT = 0.87; 16 rpm; particles size, 100 µm; average
before stabilizing, the numerical value then being close to the
concentration, 26 vol.%).
experimental one. Two aspects of the model should indeed be
considered here. First, the solid phase is considered as a
of 10−8, a value a million times lower than the one calculated contin- uum and nothing allows for the creation of clusters of
for the beaker. This very low value indicates that shear-induced particles that may arise from strong particle–particle
migration is extremely low and, consequently, only the global interactions. It is likely that the real system is in fact not
circulation of the fluid generates noticeable particle movement. uniform in concentration immediately after start-up and that
F i g. 9 shows the evolution of experimental and numerical some particle agglomer- ates are initially present and partly
torques with respect to time at start-up. In order to avoid the responsible for the torque variations. Moreover, the way
complicated and often inaccurate task of summing the forces on torque measurements are made, without short time averaging
the surface of the impeller, the numerical torque was computed for instance, may also have an influence on the torque versus
by a macroscopic energy balance: time curve shape and generate oscillations like the ones
∫ observed here.
(ηm |γ˙ |2 ) dΩ
T= Ω . (15)
2πN 5.3. Large clearance helical ribbon
It can be noticed that, notwithstanding the oscillations in the In this experiment, pumping is expected to be less efficient
experimental signal, the model is capable of predicting with than with the close clearance ribbon, leaving poorly mixed
good accuracy the torque with time and its final value after regions along the tank walls because of the reduced size of the
300 s. This is a good indication that, for this transport- impeller. Consequently, particles should migrate from the
dominated flow, the rheological behavior of the suspension is high- est shear rate region swept by the helical ribbon towards
well described by the Krieger–Dougherty model (3), all the the periphery of the tank where the lowest shear rates are
more so within the homogeneous layer located initially at the experi- enced.
bottom of the tank. The mesh of the large clearance geometry is presented in
The notable difference in the torque behavior after the start- Fig. 10. It contains 18,602 P2+–P1 finite elements for velocity
up is an indication of some limits in the way the viscosity is and pressure (233,779 velocity equations), and 18,602 P1
dealt with in the model. While the numerical curve flattens, finite elements for the concentration (3714 equations). The
the experimental torque shows a decreasing oscillatory pattern simulated

Fig. 11. Contour of the particle concentration generated with a large clearance helical ribbon impeller rotating at 50 rpm after 1, 20, and 100 s (D/DT = 0.4; average
concentration, 50 vol.% of neutrally buoyant particles; a, 3 mm).
Fig. 12. Evolution of the envelope circumscribing the region within which the concentration reaches 45 vol.% in solids at times 1, 20, and 100 s and 50 rpm in the
large clearance ribbon impeller apparatus (D/DT = 0.4; average concentration, 50 vol.%; a, 3 mm).

Fig. 13. Concentration profiles in the large clearance helical ribbon apparatus (D/DT = 0.4; 50 rpm; average concentration, 26 vol.%). Top, 300 µm; bottom, 3 mm.
case corresponds to an initial uniform particle volume distribu- The results obtained in this work revealed that the model is
tion of 50 vol.%. The rotational speed was constant at 50 rpm. adequate for predicting the transport-dominated flow in a
The simulations were performed with neutrally buoyant parti- small- gap (close-clearance) system. On the other hand,
cles but gravity was however included in the calculations. The simulations performed in the case of a large gap between a
simulated beads had a diameter a of 3000 µm. helical ribbon and the tank walls have shown that the SIMM
Shear induced particle migration is evidenced in Fig. 11, need to be improved. Rao, Mondy, Sun, et al. (2002), and Rao,
which depicts concentration contours at three different times. Mondy, Baer, et al. (2002) expressed some concerns about the
The resulting concentrations vary from 25 to 59 vol.% in the capability of this model to provide acceptable solutions based
most concentrated regions. This is close to the maximum on the observation that, in wide-gap Couette geometries, the
pack- ing limit of 68 vol.%. At equilibrium, the particles are shear rate and the vis- cosity terms are both large in magnitude
located in low shear regions, that is near the walls and around and of opposite signs. The predicted results become very
the agitation shaft leaving a cavern-like region with a lower sensitive to errors, which may lead to large discrepancies with
average concen- tration where mixing is intense. In fact, it can respect to the experimental flow and concentration fields. To
be observed that the lowest concentration occurs in the overcome this weakness in the model, Fang and Phan-Thien
vicinity of the impeller, as expected. (1995) used a ‘flow aligned ten- sor’ to render the particle
Fig. 12 shows the variation with time of the envelope migration more anisotropic. However, this addition did not
within which the concentration is equal to 45 vol.% in solids. improve the behavior of the model close to the walls where
One can readily see that the particles escaping the high shear migration is known to be the most intense. It is also in these
rate are near the impeller and migrate towards the lower shear regions that the simple particle–particle collision model of the
rate zones so that the envelope becomes larger and larger until SIMM hits its validity limit due to the intense multi- body
it reaches its maximum size at equilibrium. collisions taking place. This limitation was also noted in
Fig. 13 shows the effect of the particle size (3 and 0.3 mm, Hampton et al. (1997). The second aspect of the model that needs
respectively) on the migration phenomena for an average con- to be discussed is that related to its potentially erratic behavior
centration of 26 vol.% and the same impeller speed of 50 rpm. in close-to-zero shear rate and high concentration zones.
The 3 mm particle size was chosen so that the migration num- Based on the results shown here in which the maximum
ber would be a hundred times larger than for the 0.3 mm case. packing is never reached while the experimental results show
The results point out that shear-induced migration also occurs the opposite, it is evident that the correct description of the
with smaller particles although with much less significance. maximum pack- ing remains a problem. Future investigations
The upper part of the figure (a = 300 µm) reveals a should point at the viscosity model, the only way maximum
relatively packing can interfere with the concentration field calculations.
uniform concentration ranging from 24 to 28 vol.%, the low- Finally, it was noticed in this work that the numerical
est and highest concentrations being located at the same place torque values could be made closer to the experimental ones
as in the case of large particles (a = 3 mm) presented on the by arbitrar- ily changing the values of the empirical constants
lower part of the figure. In this latter case, the concentra- Kc and Kμ. However, doing so resulted in concentration
tions vary from 13 up to 33 vol.%, which gives a range that profiles that did not comply with the experimental
is five times wider than that with smaller particles. The com- observations. Moreover, and despite the very interesting work
parable torque values obtained from the two cases seem to of Tetlow et al. about the vari-
indicate that torque is not significantly affected by the migra- ation of the parameters with respect to concentration (Tetlow
tion phenomenon. This question will need to be investigated et al., 1998), changing of these parameters in order to better fit
thoroughly. the experimental values is not a satisfying solution from a
scientific standpoint. Rao, Mondy, Sun, et al. (2002), and Rao,
6. Additional remarks and conclusion Mondy, Baer, et al. (2002) observed recently that the shear
parameter (Kμ in Eq. (5)) is the one that must be modified to
This paper aimed at simulating the flow of suspensions in simulate more accurately the flow of a suspension in a large
stirred tanks using an implementation of the model of Phillips Couette device.
et al. (1992). For this purpose, a numerical method based on a In addition, it must be said that the presence of zones of
macroscopic model and the finite element method was very low velocity combined to concentrations close to
validated and used. maximum packing fraction may generate important numerical
The model was developed to take into account the convergence problems leading to non-physical results. The
migration of particles due to particle collisions and viscosity maximum con- centration can indeed be exceeded in those
variations. From this investigation, it can be concluded that the regions because the model does not prevent the concentration
model is able to fairly predict the average behavior of a from going above the selected maximum packing fraction.
suspension in a three dimensional situation. The only protection against concentration overshoots comes
The calculations performed during this research required from the rheological model that however becomes ineffective
CPU times in the order of days and typically ran on machines when the velocities are very small. The settling of particles by
like or comparable to IBM RS/6000 pSeries 630 with AIX 5.1 gravity in low velocity zones only augments this problem.
as OS. The problems size was always in the order of 150MB Another aspect that will need to be addressed following the
of RAM for solving 150,000 equations. work of Kim (2001) is the influence of the rheological
behavior of the dispersing medium on the migration
phenomena. Two additional issues are worth investigating,
namely the values of
the constants Kc and Kμ in the SIMM – Tetlow reported better Jana, S. C., Kapoor, B., & Acrivos, A. (1995). Apparent wall slip velocity
modeling results with a linear variation of their ratio (Tetlow coef- ficients in concentrated suspensions of non-colloidal particles.
et al., 1998) with concentration – and the formation of particle Journal of Rheology, 39(6), 1123.
clusters on the migration phenomenon. Johnson, C. (1992). Numerical solution of partial differential equations by the
finite element method. Cambridge University Press.
Kim, C. (2001). Migration in concentrated suspension of spherical particles
Acknowledgements dis- persed in polymer solution. Korea-Australia Rheology Journal, 13, 19–
27. Krieger, I. M., & Dougherty, T. J. (1959). A mechanism for non-
The financial assistance of NSERC is gratefully acknowl- Newtonian flow in suspension of rigid spheres. Transactions of the Society of
edged. The authors also want to thank Dr. Steve Altobelli for Rheology, 3,
137–152.
kindly providing the original pictures from his work.
Leighton, D., & Acrivos, A. (1986). Viscous resuspension. Chemical
Engineer- ing Science, 41(6), 1377–1384.
References Leighton, D., & Acrivos, A. (1987a). The shear-induced migration of particles
in concentrated suspensions. Journal of Fluid Mechanics, 181, 415–439.
Altobelli, S. A., Givler, R. C., & Fukushima, E. (1991). Velocity and Leighton, D., & Acrivos, A. (1987b). Measurement of shear-induced self-
concentra- tion measurements of suspensions by nuclear magnetic diffusion in concentrated suspensions of spheres. Journal of Fluid
resonance imaging. Journal of Rheology, 35(5), 721–735. Mechan- ics, 177, 109–131.
Altobelli, S. A., Fukushima, E., & Mondy, L. A. (1997). Nuclear magnetic Mann, R., & Hackett, L. A. (1988). Fundamentals of gas–liquid mixing in a
res- onance imaging of particle migration in suspensions undergoing stirred vessel: An analysis using network of backmixed zones. In
extrusion. Journal of Rheology, 41, 1105–1115. Proceed- ings of the sixth European conference on mixing (pp. 321–328).
Bertrand, F., & Tanguy, P. A. (2002). Krylov-based Uzawa algorithms for the Maury, B., & Glowinski, R. (1997). Fluid-particle flow: A symmetric
solution of the stokes equations using discontinuous pressure tetrahedral formula- tion. Me´thode Nume´riques de la Me´canique, 1–6.
finite elements. Journal of Computational Physics, 181, 617–638. Phillips, R. J., Armstrong, R. C., Brown, R. A., Graham, A. L., & Abott, J. R.
Bertrand, F., Gadbois, M., & Tanguy, P. A. (1992). Tetrahedral elements for (1992). A constitutive equation for concentrated suspensions that accounts
fluid flow. International Journal for Numerical Methods in Engineering, for shear-induced particle migration. Physics of Fluids A, 4, 30–40.
33, 1251–1257. Rao, R., Mondy, L., Sun, A., & Altobelli, S. (2002). A numerical and experi-
Bertrand, F., Tanguy, P. A., & Thibault, F. (1997). A three-dimensional mental study of batch sedimentation and viscous resuspension.
fictitious domain method for incompressible flow problems. International International Journal for Numerical Methods in Fluids, 39, 465–483.
Journal for Numerical Methods in Fluids, 25, 719–736. Rao, R. R., Mondy, L. A., Baer, T. A., Altobelli, S. A., & Stephens, T. S.
Brady, J. F., & Bossis, G. (1985). The rheology of concentrated suspensions (2002). NMR measurements and simulations of particle migration in non-
of spheres in simple shear flow by numerical simulation. Journal of Fluid Newtonian fluids. Chemical Engineering Communications, 189(1), 1–22.
Mechanics, 155, 105–129. Richardson, J. F., & Zaki, W. N. (1954). Sedimentation and fluidisation: Part 1.
Fang, Z., & Phan-Thien, N. (1995). Numerical simulation of particle Transactions of the Institution of Chemical Engineers, 32, 35–53.
migration in concentrated suspensions by a finite volume method. Journal Ritz, J.-B., Bertrand, F., Thibault, F., & Tanguy, P. A. (2000). Shear-induced
of Non- Newtonian Fluid Mechanics, 58, 67–81. particle migration in a short-dwell coater. Chemical Engineering Science,
Gadala-Maria, F., & Acrivos, A. (1980). Shear-induced structure in a 55, 4857–4867.
concentrated suspension of solid spheres. Journal of Rheology, 24, 799– Saad, Y. (1985). GMRES: A generalized minimal residual algorithm for
814. solving non-symmetric linear systems. SIAM Journal of Scientific and
Glowinski, R., Pan, T.-W., & Perieux, J. (1994). A fictious domain method for Statistical Computing, 7, 850–869.
Dirichlet problem and applications. Computer Methods in Applied Subia, S. R., Ingber, M. S., Mondy, L. A., Altobelli, S. A., & Graham, A. L.
Mechan- ics and Engineering, 111, 283–303. (1998). Modeling of concentrated suspensions using a continuum constitu-
Grygiel, J. M., & Tanguy, P. A. (1991). Finite element solution for advection- tive equation. Journal of Fluid Mechanics, 373, 193–219.
dominated thermal flows. Computer Methods in Applied Mechanics and Tanguy, P. A., Bertrand, F., Thibault, F., & Galy-Jammou, P. (2001).
Engineering, 93, 277–289. Simulation of kneading mechanisms in a three blade planetary mixer,
Hampton, R. E., Mammoli, A. A., Tetlow, N., & Altobelli, S. A. (1997). Enyclopedia of materials science and technology. Elsevier, pp. 7425–
Migra- tion of particles undergoing pressure-driven flow in a circular 7428.
conduit. Journal of Rheology, 41, 621–640. Tanguy, P. A., Thibault, F., Ascanio, G., & Brito-De La Fuente, E. (2006).
Hu, H. H. (1996). Direct simulation of flows of solid–liquid mixtures. Solid- liquid mixing: Numerical simulation and physical experiments. In
Interna- tional Journal of Multiphase Flow, 22(2), 335–352. Sunggyu Lee (Ed.), Enc. Chem. Proc. (pp. 2753–2768). M. Dekker.
Hu, H. H., Joseph, D. D., & Crochet, M. J. (1992). Direct simulation of fluid Tetlow, N., Graham, A. L., Ingber, M. S., Subia, S. R., Mondy, L. A., &
particle motions. Theoretical and Computational Fluid Dynamics, 3, 285– Altobelli,
306. S. A. (1998). Particle migration in a Couette apparatus: Experiment and
Jana, S. C. (2003). Loss of surface and volume electrical conductivities in modeling. Journal of Rheology, 42, 307–327.
poly- mer compounds due to shear-induced migration of conductive Zhang, K., & Acrivos, A. (1994). Viscous resuspension in fully developed lam-
particles. Polymer Engineering and Science, 43(3), 570–579. inar pipe flows. International Journal of Multiphase Flow, 20(3), 579–591.
Zhou, G., Tanguy, P. A., & Dubois, C. (2000). Power consumption in a double
planetary mixer with non-Newtonian and viscoelastic materials. Chemical
Engineering Research and Design, 78, 445–453.

You might also like