Вы находитесь на странице: 1из 8

Full Paper

Full Papers
Power Requirement when Mixing a Shear-Thickening Fluid with a
Helical Ribbon Impeller Type
By Guillaume Delaplace, Jean-Claude Leuliet, and Gilles Ronse*
The optimal design of close clearance impellers requires the knowledge of the power demand of the mixing equipment. In nonNewtonian mixing, this can be readily obtained using the Metzner and Otto concept [1]. In this work, this concept and the
determination of the Ks value for an atypical helical agitator (PARAVISC system from Ekato firm) have been revised in the case
of shear-thinning fluids and a shear-thickening fluid. For poor shear-thinning fluids, it has been shown that for our mixing system
the Ks value does not vary strongly with the flow behavior index, and may be regarded as a constant for the mixing purpose
design. By contrast, for the shear-thickening fluid, power consumption measurements indicate that the relationship between the
Ks values and the flow behavior index is much more complex due to a partial solidification of the product around the impeller.

1 Introduction
Mixing highly viscous fluids is a common unit operation in
chemical and food industries. To predict the power demand of
a mixing system to achieve such operations, the knowledge of
two parameters usually named Kp and Ks is required.
Kp is the product of the power number Np by the Reynolds
number Re when mixing Newtonian fluids under laminar
regime.
Ks is the constant of proportionality which links the
effective shear rate (defined by the Metzner and Otto concept
[1]) and the rotational speed of the impeller when mixing nonNewtonian fluids under laminar regime.
Numerous experimental and theoretical works have been
carried out to predict the Kp factor only from the knowledge of
geometrical parameters of the mixing system. As a result,
nowadays we can estimate without experimental studies and
with satisfactory accuracy the Kp factor for a wide range of
mixer geometries.
Although the determination of the Ks factor for various
mixing systems using shear-thinning fluids and the Metzner
and Otto concept [1] have been widely covered in the
literature, the prediction of this factor is still critical. Indeed,
even for a given mixing system, lots of disparities exist about
the values of the Ks parameter. For instance, contradictory
results exist concerning the Ks factor dependence on the
rheological properties of the fluid.
To point out the level of existing confusion, one may
mention the experimental works on helical ribbon impellers
using shear-thinning fluids of Hall and Godfrey [2], Nagata et
al. [3], Rieger and Novak [4], Takahashi et al. [5] and Shamlou
and Edwards [6] who concluded that the Ks factor is a pure
geometrical parameter and the works of Yap et al. [7], Brito-de
la Fuente et al. [810], Leuliet et al. [1112], Carreau et al. [13]

[*]

G. Delaplace (author to whom correspondence should be addressed), J.-C.


Leuliet, G. Ronse, Food Process Engineering Laboratory, INRA-LGPTA,
369, Rue Jules Guesde, B. P. 39, 59651 Villeneuve d'Ascq Cdex, France.

Chem. Eng. Technol. 23 (2000) 4,

and Cheng and Carreau [1415] who concluded that Ks


increases with the flow behavior index of shear-thinning fluids.
An exhaustive review about the use of the Metzner and Otto
concept to predict the power demand of a mixing system
highlights another deficiency on the determination of the Ks
factor. Most of the Ks values have been obtained by applying
the Metzner and Otto concept using shear-thinning liquids
(0 < n < 1) and little is known about effective or average
shear rate values when mixing a shear-thickening fluid (n > 1).
This lack of information is quite embarrassing since dilatancy
is the rheological behavior of highly concentrate solid/liquid
suspensions (the concentration of solid is usually greater than
40 % per volume) and mixing operations involving these kinds
of products are often encountered in the pulp and paper
industries.
The aimof thispaper isto fill thisgap.Hence, the Metznerand
Otto concept has been revised for a power-law fluid with flow
behavior index higher than one (n > 1). The impeller used is a
helical agitator (PARAVISC system from Ekato firm) which is
well adapted to mix highly viscous fluids (Tatterson [16]).
Due to the geometrical particularities of our helical agitator
(it is equipped with an anchor) and the existence of many
conflicting results in the literature with shear-thinning fluids,
we have decided in a previous step to check for our mixing
system the dependence or not of Ks values with flow behavior
index on pseudoplastic fluids.
Finally, eight pseudoplastic fluids and a simple dilatant fluid
were prepared and power consumption measurements were
carried out. Using the rheological data obtained by various
viscosimeters and the average effective shear rate concept, the
values of the Ks factor for our helical mixing equipment were
then determined and discussed.

2 Theoretical Aspects
Every mixing system is characterized by a power curve. This
power curve is the plot of the power number Np versus the

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000 0930-7516/00/0404-0329


0930-7516/00/0404-000329
$ 17.50+.50/0
$ 17.50+.50/0

329

Full Paper
Reynolds number Re and is described by the following
relationship1):
Np f Re

(1)

3 Materials and Methods


3.1 Equipment
A picture of the helical impeller tested is shown in Fig. 1.

where

Np P   N 3  d 5

and


Re   N  d2 m

for a Newtonian fluid.


Here, m is the Newtonian viscosity, r is the density of the
fluid, N is the impeller speed, d is the impeller diameter and P
is the power requirement.
When mixing Newtonian fluids under laminar regime
(provided that no vortex is present), it is readily shown that
the power curve can be described by the following simple
relationship:
.
Np Kp Re

(2)

where Kp is a geometrical parameter depending on the


mixing system.
Eq. (2) characterizing the power consumption of the mixing
system, can easily be used to calculate, for a given impeller
speed, the power requirement for similar vessel/impeller
arrangements of any scale containing any Newtonian viscous
liquid.
For non-Newtonian fluids, the shear rate varies through the
vessel, so there is an apparent viscosity profile. Unfortunately,
the lack of theoretical knowledge about the velocity profile
within the vessel prevents us from determining the analytical
shear rate distribution and knowing theoretically or experimentally the value of the process or representative
viscosity which must be taken into account in the Reynolds
number.
In an attempt to solve this problem, Metzner and Otto [1]
proposed to define an effective shear rate _ e in the vessel, in
such a way that the power curve in the laminar flow for both
Newtonian and non-Newtonian fluids is quite similar.
This effective shear rate value is classically linked to the
impeller rotational speed as follows:
_ e KS  N

(3)

This effective shear rate value is used to evaluate an


average or representative apparent viscosity in the Reynolds number and the power input to any non-Newtonian fluid
can easily be evaluated from the Newtonian power curve using
Eq. (2).
Even if the Metzner and Otto concept does not allow us to
obtain local rheological properties of the fluid in the whole
volume of the vessel, it has been widely used and accepted by
the majority of authors, since it is a useful approach for the
estimation of power consumption in a mixing vessel.

1)

330

List of symbols at the end of the paper.


WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

Figure 1. Picture of the helical impeller used (PARAVISC system, EKATO) and
geometrical parameters of the mixing system.

The vessel was a transparent glass cylinder with a rounded


bottom and a flat lid. During all the experiments, the volume
of the liquid was kept constant and equal to 34 liters. The
geometrical parameters of the mixing system studied are
summarized in Fig. 1. The mixing equipment used in this
investigation was driven by an electric motor equipped with a
variable speed drive and a strain gauge torquemeter in the
range of 0 to 20 N m in order to measure power consumption.
The motor speed-drive system allowed us to obtain variable
speeds. Depending on the range of the rotational speed
investigated, either a stopwatch or a digital tachometer was
used to measure the rotational speed of the impeller (the
stopwatch was used when the rotational speed was less than
20 rev/min).

3.2 Fluids and Rheology


3.2.1 Newtonian and Shear-Thinning Fluids
In order to follow the Metzner and Otto procedure for
power prediction for non-Newtonian liquids, it is necessary to
obtain first the Newtonian power curve of the mixing system.
This was done by using solutions of pure or dilute glucose
syrups of varying viscosities (3.5 Pa s < m < 61 Pa s).
Eight different pseudoplastic fluids covering a wide range of
shear-thinning behavior were prepared for this study (0.11 < n
< 0.63). The pseudoplastic fluids tested were aqueous
solutions of carboxymethylcellulose CMC (Rhne Poulenc
LTD), of alginate (SBI Cecalgum S1300), of guar gum (SBI
Viscogum HV 3000 A) and adragante gum (Powder 427,
Alland & Robert).
The viscous properties of the Newtonian and shear-thinning
fluids were measured in classical controlled rotational speed
concentric cylinders (Contraves Rheomat 30).
The dependence of the Newtonian viscosity and density
with respect to temperature was expressed by following
relationships:
m a   b and  c d  

0930-7516/00/0404-000330 $ 17.50+.50/0

(4)
Chem. Eng. Technol. 23 (2000) 4

Full Paper
The values of the parameters a,b,c and d are given in Tab. 1.
Table 1. Rheological and physical properties of the Newtonian fluids used.

A power-law model (Ostwald de Waele's model) was


chosen to describe the shear-thinning behavior of the
pseudoplastic fluids. This model provided a satisfactory fit of
the viscosity data for the range of the shear rate encountered
in the vessel (shear rate values were determined by the
Metzner and Otto method). The resulting values obtained by
linear regression of the power-law constants (n and k) are
given in Tab. 2.
Table 2. Power law parameters of the shear-thinning and shear-thickening fluids
used.

then deduced from flow rate and pressure drop measurements. A sketch of the experimental platform used to make
the in-line rheology is shown in Fig. 2.

Figure 2. Rheology in line: Sketch of the experimental platform.

For the dilatant fluid, the flow curve could not be described
by a simple power-law model. A local power-law model was
used to fit the flow curve. This method consists in approximating each point of the flow curve by its tangency (Wilson and
Thomas [17]). The final relationship is:

ma kg  g ng1

For the pseudoplastic fluids used, a very small increase of


temperature due to energy dissipation has been observed in
the vessel. As a result, the parameters n and k and the density
of the fluid have been considering as constants during each
experimental set of power measurements.

3.2.2 The Shear-Thickening Fluid


This fluid was elaborated by mixing a concentrate suspension of starch (CLEARAM CH20 made by Roquette Firm) in
water (50 % w/w). The flow curves of that fluid were obtained
at the same temperature as that encountered in the mixing
experiments by using different kind of viscosimeters:
a controlled rotational speed concentric cylinder (Mettler,
Rheomat 260, concentric cylinders DIN 145)
a controlled shear stress viscosimeter (Carri-Med CS100,
cone-plate system)
our in-line viscometer
The last set of rheological data was performed by considering the flow of the starch suspension under laminar regime in a
cylindrical duct. The rheological properties of the fluid were

Chem. Eng. Technol. 23 (2000) 4,

(5)

In Eq. (5), the parameters n(g)and k(g)are not constant but


vary with the shear rate. For the starch suspension, the
evolution of the two parameters n(g) and k(g) with the shear
rate values is reported in Tab. 2.
For the laminar flow of a power-law fluid in a pipe, the
analytical relationship between the pressure drop DP and the
flow rate of the fluid Q is:

n
4 L k 3 n1
8
P

Qn
(6)
D
n
D3
D and L are respectively the diameter and the length of the
pipe.
The plot of log DP versus log Q allowed us to obtain the
local rheological properties n and k of the fluid. Using the
fact that the shear rate at the wall _ w and the stress at the wall w
in a pipe are respectively defined by:
_ w

3 n1
Q
and
 32 
4n
D3

(7)

w

PD
4L

(8)

The plot of w versus _ w allows us to compare the


rheological data obtained with the cylindrical duct and with
classical rotational viscosimeters. This was done in order to
evaluate sedimentation effects due the mixing of a high
concentrate suspension in a concentric cylinder at low shear
rate values.

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

0930-7516/00/0404-000331 $ 17.50+.50/0

331

Full Paper
0:29

4 Results

Kp 66  Nr0:84 1=2  t=d 1

4.1 Newtonian Fluids

For our impeller/vessel arrangements, the use of Eq. (9)


gives a value of Kp equal to 318, which is very close to the one
obtained experimentally (Kp = 315).

Fig. 3 shows the Newtonian power curve (log Np vs log Re)


of the mixing system investigated.

l=d

(9)

4.2 Shear-Thinning Fluids


The experimental values of Ks by use of the Metzner
and Otto method for shear-thinning fluids are reported in
Fig. 4.

Figure 3. Power curve of the mixing system studied with Newtonian fluids.
Solutions A, B, C and D are dilute glucose syrups with different amounts of
water.

In this figure, it can be seen that the onset of the transition


regime starts for a Reynolds number value around 60. This
critical value of the Reynolds number is in agreement with
those reported in the literature for typical double helical
ribbon impellers. Indeed, the end of the laminar regime for
these mixing systems occurs at Reynolds number values
between 50 and 100 (Cheng and Carreau [14]).
A linear regression on the power curve data, for Reynolds
number values less than 60, allowed us to obtain the power
constant Kp of the mixing system: Kp = 315.
For typical helical ribbon agitators (not equipped with an
anchor at the bottom of the impeller), lot of correlations have
been proposed to predict the Kp factors only from the
knowledge of geometrical parameters of the mixing systems
(clearance wall size, pitch and width of ribbon ...). In order to
compare the power consumption of our mixer with a similar
typical helical ribbon, the predictive correlations suggested by
Chavan and Ulbrecht [18] and Delaplace and Leuliet [19]
were used. These empirical correlations have been chosen
since they were validated on numerous different helical
ribbon mixing systems (more than 50 impellers set-up for
Chavan and Ulbrecht [18] and 145 for Delaplace and Leuliet
[19]) and the average absolute standard deviation between
predicted values and experimental values was found to be less
than 10 %.
The predicted values of Kp using these models are
respectively Kp = 236 and Kp = 249. Thus, we can assume
that the excess of power consumption measured for our helical
mixer (Kp = 315) is due to the anchor at the bottom of the
impeller.
Experimental studies show that for the PARAVISC mixing
system, a better value of the Kp factor can be predicted using
the following relationship:

332

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

Figure 4. Evolution of the Ks values with the flow behavior index of shearthinning fluids for the mixing system investigated.

From this figure, we can note that:


The Ks factor slightly increases with the flow behavior
index. This result is in agreement with the trends obtained
by recent works on typical helical ribbon impellers (Britode la Fuente et al. [810], Leuliet et al. [1112], Carreau et al.
[13] and Cheng and Carreau [1415]).
For practical purpose, the Ks values of our helical mixing
system may be regarded as independent of the flow
behavior index. Then, the Ks factor is a pure geometrical
constant and its average value is equal to 32.2. The order of
magnitude of this constant is similar with the Ks values
classically obtained for typical helical ribbon agitators when
mixing weak shear-thinning fluids (0.3 <n< 1). Indeed, in
this case, the Ks values are between 20 and 45. Note that an
approximate value of the average Ks factor for our mixing
equipment can be obtained from the correlation proposed
by Delaplace and Leuliet [19]:
Ks

Kp
2


Nr 2  l 
d

(10)

where l* is the height of the helical ribbon (l* = 0.280 m).


Indeed, by using this empirical correlation the absolute
standard deviation, between the predicted (Ks = 33.7) and
average experimental values (Ks = 32.9), is less than 10 %.

0930-7516/00/0404-000332 $ 17.50+.50/0

Chem. Eng. Technol. 23 (2000) 4

Full Paper
4.3 The Shear-Thickening Fluid
4.3.1 Viscosimeter Measurements
Before analyzing the rheological measurements obtained
for our starch suspension, note that the flow behavior of such
product is very complex and it is very difficult to obtain
meaningful data for this suspension. Indeed, this suspension
keeps a purely viscous behavior while the product has not
been subjected to important shear rate values. If the shear rate
values provided to the fluid are too strong, the product
becomes a time-dependent fluid. To avoid such difficulties,
cares have been taken in order to mix the suspension at low
shear rate values.
Fig. 5 shows the in-line rheology of the product before and
after power measurements.

Figure 5. Comparison of pressure drop/flow rate measurements obtained in


cylindrical ducts before and after a trial of power consumption in the mixing
vessel.

As suggested by this figure, it is possible with such


precautions to obtain a stable rheological behavior of the
product with time.
the rheological data of the concentrate starch suspension
obtained with various viscosimeters are presented in Fig. 6.
Although the product is complex, we can observe a good
agreement between the three devices investigated to analyze
the product.

Fig. 6 reveals the complexity of the concentrated starch


suspension. This suspension can exhibit Newtonian, pseudoplastic or dilatant behavior depending on the shear rate range.
The starch suspension shows weak shear-thinning behavior at
low shear rates (< 5 s1), Newtonian behavior at 5 s1 and finally
becomes strongly dilatant at higher shear rates (> 20 s1). For
example the local values of the flow behavior index increase
from 1 at 10 s1 to 2.75 at 80 s1. Simultaneously, the local flow
consistency index decreases from 0.01 to 9.4 104 Pa sn.

4.3.2 Power Consumption


Fig. 7 shows the power consumption results obtained with
the starch suspension (evolution of the torque acting on the
shaft with the rotational speed of the impeller).

Figure 7. Evolution of the measured experimental torque with the impeller


rotational speed when mixing the starch suspension 50 % (w/w).

The concavity of the curve suggests that the agitated fluid is


dilatant. Indeed, in laminar flow, for a Newtonian fluid the
torque is proportional to the rotational speed of the impeller
and for pseudoplastic fluids the concavity is reversed.
A typical determination of the Ks factor following the
Metzner and Otto method obtained with starch suspension for
our system is shown in Fig. 8.

Figure 8. Evolution of the effective shear rate with the rotational speed of the
impeller obtained when mixing a dilatant fluid with helical impeller (PARAVISC system).
Figure 6. Comparison of rheological data of the starch suspension 50 % (w/w)
obtained with different viscosimeters.

Chem. Eng. Technol. 23 (2000) 4,

As expected by Eq. (3), the effective shear rate is


proportional to the rotational speed of the impeller.

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

0930-7516/00/0404-000333 $ 17.50+.50/0

333

Full Paper
A linear regression of these data gives

the helical agitator changes (due to an increase or decrease of


the thickness of the solidification layer around the impeller).
This assumption allows us to write the following system of
equations:

_ e 136  N
The disagreement between this Ks value (Ks = 136) and the
values obtained with shear-thinning fluids (the average value
of Ks is in this case equal to 32.2) is quite obvious and will now
be discussed.
For classical impellers (Ruhston Turbines, propellers),
Calderbank and Moo Young [20] and Metzner et al. [21] have
first obtained significant differences between the Ks factor
obtained with dilatant fluids and pseudoplastic fluids.
In recent studies, Edwards et al. [22], Edwards and Jomha
[23] and Ahmad et al. [24] using anchor and helical ribbon
impellers have also noticed this phenomenon (predicted
power requirements for dilatant fluids, using the effective
shear rate concept and the Ks factor obtained with pseudoplastic fluids, were largely less than the power consumption
experimentally measured).
We think that there is little probability that for dilatant
fluids, for a given impeller rotational speed, the representative
shear rate in the vessel will be 4 or 5 times higher than for
shear-thinning fluids if there is no important change of the
profile shear rate in the vessel. We are convinced that the
physical meaning of such an increase of the average shear rate
is due to geometrical modification of the mixing system caused
by shear-thickening effects of the suspension.
As it was already suggested by previously mentioned
authors, we proposed the following mechanism to explain
the observed phenomenon: for a given impeller rotational
speed, in the vicinity of the impeller, the suspension is
subjected to a very high shear rate. As underlined by Edwards
et al. [22], dilatancy is clearly favored by high shear rate values.
As a consequence, the rheological behavior of the suspension
in the neighborhood of the impeller becomes close to a solid
one and there is a formation of a solid layer close to the wall of
the impeller. This means that the impeller diameter and so the
clearance wall of the mixing system are widely altered when
the product is agitated.
These assumptions are in agreement with the experimental
works of Forresti and Liu [25] who noted a surprising increase
of power requirement when mixing under laminar regime
dilatant fluids in a baffled and nonbaffled vessel. Indeed, this
mechanism of partial solidification around the impeller for
dilatant substances means that the region where most shearing
occurs is shifted away from the impeller and towards the vessel
walls. If we admit that the total energy dissipation (power
consumption) is only due to the shearing of the fluid in the gap.
This effect explains why baffles increase the power consumption in the agitation of dilatant fluids.
To modelize the mechanism of this partial solidification at
the impeller wall, we have suggested that for each impeller
rotational speed, the impeller keeps the same shape (pitch,
width and height of helical ribbon) and only the diameter da of

334

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

da

2 C
k Kp  Ksn1  N n

(11)

Ks

Kp
2


Nr 2  l 

(12)

da da



0:29
Kp 66  Nr0:84 1 2  t da 1
l=da

(13)

n f1 _ eq

(14)

k f2 _ eq

(15)

_ e Ks  N

(16)

The six unknowns are Kp, Ks, n, k, da et _ e .


Eq. (11) derives from the following relationship: Np =
Kp/Rea with Rea = r/N.d2.ma. Eqs. (12) and (13) are predictive
models which give the evolution of the Kp and Ks factors with
the geometrical parameters of the PARAVISC mixing system.
Eqs. (14) and (15) are known from the knowledge of the local
power-law parameters used to fit the rheological behavior of
our suspension. Finally, Eq. (16) is the well-known Metzner et
Otto concept.
The resolution of this system, for each rotational speed of
the impeller, allowed us to obtain on the one hand the
evolution of the fictitious diameter da with the rotational
speed of the impeller (Fig. 9) and on the other hand the
evolution of the Ks factor of the PARAVISC system with the
flow behavior index n of the fluid (Fig. 10).

Figure 9. Evolution of the fictitious diameter of the impeller da with the


rotational speed of the impeller when partial solidification around the impeller
occurs.

If our assumptions are valid, Fig. 9 shows us that


the clearance wall size effectively decreases with the
rotational impeller speed which means that the solid part
around the impeller is all the more important than the shear
rate values are high;
the clearance wall is very quickly entirely filled by a solid
layer.

0930-7516/00/0404-000334 $ 17.50+.50/0

Chem. Eng. Technol. 23 (2000) 4

Full Paper

Figure 10. Evolution of the Ks factor with the flow behavior index for the
PARAVISC mixing system. For dilatant fluid, each symbol refers to a different
fictitious impeller diameter and as a result to different mixing systems.

Fig. 10 seems to support the mechanism of partial


solidification. Indeed, this figure shows that for fluids with a
flow behavior index slightly higher than one, the Ks values are
very close to those obtained for pseudoplastic liquids.
Note that before analyzing this figure care must be taken
since each symbol refers to a different diameter da of the
impeller and as a result to a different mixing system (the gap
size decreases when increasing the rotational speed of the
impeller).

Kp

[]

Ks

[]

k
_
k

[Pa sn]
[Pa sn]

l

l
n
_
n

[m]
[m]
[]
[]

N
Np
Nr
P

[rev/s]
[]
[]
[W]

Re
Rea

[]
[]

s
t
w
L

[m]
[m]
[m]
[m]

proportionality constant of the power


number defined by Eq. (2)
Metzner and Otto constant defined by
Eq. (3)
consistency index of power-law fluid
consistency index of local power-law
fluid
impeller height
height of the helical ribbon
flow behavior index of power-law fluid
flow behavior index of local power-law
fluid
rotational speed of impeller
power number, P/rN3 d5
number of ribbons
power consumption for an agitated
vessel
rotational Reynolds number, r N d2/m
rotational apparent Reynolds number,
rN d2/ma
pitch of the helical ribbon impeller
tank diameter
ribbon width
length of the cylindrical duct

Greek symbols

5 Conclusions
This present work has pointed out the fact that we can
obtain, for a high concentrate suspension, similar flow curves
with both classical viscosimeters (controlled shear rate and
shear stress rotational viscosimeters) and in-line rheology.
This experimental evidence is very interesting for industrial
suspensions applications.
Finally, we have pointed out that a shear-thickening fluid
widely alters the geometrical shape of the mixing system, due
to partial mechanism of solidification of the fluid around the
impeller. This purely geometrical modification of the mixing
system enhances drastic changes of the performance of the
mixer and highlights the limitations of the Metzner and Otto
concept to predict power consumption when dilatancy
phenomena occur.
Received: November 27, 1998 [CET 1063]

[]
[N m]
[m]
[m]

HL

[m]

[s1]


a

[Pa s]
[Pa s]

e

[Pa s]



_


[Kg/m3]
[Pa]
[s-1]
[C]

effective shear rate in the vessel,


defined by Eq. (1)
Newtonian viscosity
apparent viscosity of a non-Newtonian
fluid
process or representative viscosity,
n1
k  _ e
liquid density
shear stress
shear rate
temperature

Subscript
w

at wall

References

Symbols used
c
C
d
da

_ e

clearance wall, (td)/2


torque acting on the shaft
impeller diameter
fictitious impeller diameter (diameter
of the impeller plus thickness of partial
solidification around the impeller)
height of liquid in the vessel

Chem. Eng. Technol. 23 (2000) 4,

[1] Metzner, A. B.; Otto, R. E., AIChE J. 3 (1957) No. 1, pp. 310.
[2] Hall; K. R.; Godfrey, J. C., Trans. Instn. Chem. Engrs. 48 (1970) pp. 201
208.
[3] Nagata, S.; Nishikawa, M.; Hisayuki, T.; Gotoh, S., J. Chem. Eng. Japan 4
(1971) No. 1, pp. 7276.
[4] Rieger, F.; Novak, V., Trans. Inst. Chem. Eng. 51 (1973) pp. 105111.
[5] Takahashi, K.; Yokota, T.; Konno, H., J. Chem. Eng. Japan 17 (1984) No. 6,
pp. 657659.
[6] Shamlou, P. A.; Edwards, M. F., Chem. Eng. Sci. 40 (1985) No. 9, pp. 1773
1781.

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

0930-7516/00/0404-000335 $ 17.50+.50/0

335

Full Paper
[7] Yap, C. Y.; Patterson, W. I.; Carreau, P. J., AIChE J. 25 (1979) No. 3, pp.
516521.
[8] Brito-de la Fuente, E.; Leuliet, J. C.; Choplin, L.; Tanguy, P. A., Trans.
Instn. Chem. Engrs. 69 (1991) pp. 334331.
[9] Brito-de la Fuente, E.; Leuliet, J. C.; Choplin, L.; Tanguy, P. A., in: Process
Mixing: Chemical and Biochemical Applications (G. B. Tatterson, R. V.
Calabrese, Eds.), and AIChE Symp. Ser. 286 (88) (1992) pp. 2832.
[10] Brito-de la Fuente, E.; Choplin, L.; Tanguy, P. A., Trans. Inst. Chem. Eng.
75 (1997) Part A, pp. 4552.
[11] Leuliet, J. C.; Brito-de la Fuente, E.; Choplin, L., Rcents Progrs en Gnie
des Procds 5 (1991) No. 14, pp. 6974.
[12] Leuliet, J. C.; Brito-de la Fuente, E.; Choplin, L., Entropie 28 (1992) No.
171, pp. 5358.
[13] Carreau, P. J.; Chhabra, R. P.; Cheng, J., AIChE J. 39 (1993) No. 9, pp.
14211430.
[14] Cheng, J.; Carreau, P. J., The Can. J. Chem. Eng. 72 (1994) pp. 418430.
[15] Cheng, J.; Carreau, P. J., Industrial Mixing Fundamentals with Applications, AIChE Symp. Ser. 91 (1995) No. 305, pp. 116122.
[16] Tatterson, G. B., AIChE Ann. Meeting, Miami Beach 1986, pp. 146.

[17] Wilson, K. C.; Thomas, A. D., The Can. J. Chem. Eng. 63 (1985) pp. 539
546.
[18] Chavan, V. V.; Ulbrecht, J., Ind. Eng. Chem. Proc. Des. Dev. 12 (1973) No.
4, pp. 472476 and corrigenda Ind. Eng. Chem. Proc. Des. Dev. 13 (1974)
No. 3, pp. 309309.
[19] Delaplace, G.; Leuliet, J. C., Rcents Progrs en Gnie des Procds 11
(1997) No. 53, pp. 331336.
[20] Calderbank, P. H.; Moo Young, M. B., Trans. Inst. Chem. Eng. 39 (1961) pp.
337347.
[21] Metzner, A. B.; Feehs, R. H.; Ramos, H. L.; Otto, R. E.; Tuthill, J. D.,
AIChE J. 7 (1961) No. 1, pp. 39.
[22] Edwards, M. F.; Jomha, A. I.; Macsporran, W. C.; Woodcock, L. V., Inst.
Chem. Eng. Ann. Research Meeting, Bradford, England, 1986, pp. 97105.
[23] Edwards; M. F.; Jomha, A. I., Inst. of Chem. Eng. Delft, Netherlands, Pt/
Procestechniek 42 (1987) No. 10, pp. 7377.
[24] Ahmad, I. J.; Edwards, M. F.; Woodcock, L. V., Chem. Eng. Sci. 45 (1990)
No. 5, pp. 13891396.
[25] Foresti, R.; Liu, T., Ind. Eng. Chem. 51 (1959) No. 7, pp. 860864.

_______________________

336

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

0930-7516/00/0404-000336 $ 17.50+.50/0

Chem. Eng. Technol. 23 (2000) 4

Вам также может понравиться