Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/222721502

Power Consumption, Mixing Time and Homogenisation Energy in Dual-


Impeller Agitated Gas–Liquid Reactors

Article  in  Chemical Engineering and Processing · February 2001


DOI: 10.1016/S0255-2701(00)00128-8

CITATIONS READS

66 2,047

2 authors, including:

Mounir Bouaifi
Solvay
31 PUBLICATIONS   430 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Experimental Analysis and Multiscale Modeling of Three‐Phase Flows in Industrial Reactors View project

All content following this page was uploaded by Mounir Bouaifi on 05 February 2018.

The user has requested enhancement of the downloaded file.


Chemical Engineering and Processing 40 (2001) 87 – 95
www.elsevier.com/locate/cep

Power consumption, mixing time and homogenisation energy in


dual-impeller agitated gas–liquid reactors
Mounir Bouaifi 1, Michel Roustan *
Département Génie des Procédés Industriels, Institut National des Sciences Appliquées, 31077 Toulouse Cedex 4, France

Received 7 May 1999; received in revised form 16 October 1999; accepted 31 May 2000

Abstract

The study relates to the power consumption, mixing time and mixing efficiency defined as the product of the mixing time and
the power consumption in a non standard vessel equipped with various axial and mixed dual-impeller configurations. The effects
of the rotational speed, gas flow rate, impeller type and diameter are investigated. The mixing mechanisms are found strongly
dependent on flow patterns, impeller type and diameter. Dimensionless correlations are proposed to predict power consumption
and the mixing time beyond complete dispersion regime. © 2001 Elsevier Science B.V. All rights reserved.

Keywords: Dual-impeller; Power consumption; Mixing time; Homogenisation energy

1. Introduction mechanism of the process. In multiphase reactors, the


perfect mixing of the liquid phase is often assumed.
Multi-impeller gas – liquid reactors have been found However, the mixing is far from ideal and the hypothe-
ensuring higher efficiency of gas utilisation and longer sis of the perfect mixing is not realistic. Information on
retention time of gas than single impeller systems. the degree of mixing and mixing efficiency should be
Agitated reactors are equipped traditionally with radial considered for an accurate modelling and cost optimisa-
turbines, however, axial impellers recently found big tion of the reactor. Because of the complexity arising
interest in various industrial processes. Some of these from the complexity of the hydrodynamics and the use
impellers are able to disperse gas in liquids and yield a of different types of impellers, the available literature
good mass transfer performance and high hydraulic data are insufficient to design and control gas –liquid
efficiency at reduced energy consumption. Unfortu- reactors. In order to estimate the mixing efficiency, it is
nately, most of the previous investigations have been necessary to define the objectives of the processes. For
carried out only with radial Rushton turbines [1–9] many processes where the temperature must be the
while minimal data has been obtained with axial im- same everywhere, the liquid circulation is very impor-
tant in order to avoid local temperature gradients. The
pellers [10 –21].
characteristics of flow patterns are strongly influenced
The mixing time is an important parameter for reac-
by impeller configuration. A quantitative definition of
tor design. It can be compared with mass transfer time
the mixing performance and efficiency is possible only
or reaction time in order to evaluate the controlling
by setting a model, which takes into account the inter-
nal hydrodynamics of the reactor. Little data is found
* Corresponding author. Tel.: +33-561-559751; fax: + 33-561- concerning mixing time in dual and multi-impeller reac-
559760. tors [7,8,16,21] and aerated systems [7,22 –26].
E-mail addresses: mounir@cre.chalmers.se (M. Bouaifi), rous-
The objectives of this study are to evaluate the power
tan@insa-tlse.fr (M. Roustan).
1
Present address: Department of Chemical Reaction Engineering, consumption, the mixing time, the mixing efficiency and
Chalmers University of Technology, Kemivägen 4, SE-41296 the axial flow impellers performances in dual-impeller
Göteborg, Sweden. Tel.: + 49-317-723034; fax: + 49-317-723035. gas –liquid reactors.

0255-2701/01/$ - see front matter © 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 2 5 5 - 2 7 0 1 ( 0 0 ) 0 0 1 2 8 - 8
88 M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95

2. Material and methods diameter T= 0.43 m and ratio HL/T=2 (Fig. 1),
equipped with various double impeller combinations
The experiments are carried out in a non standard (Table 1). Four wide longitudinal baffles (0.1 T) are
hemispherical bottomed cylindrical tank with internal mounted symmetrically. Four types of impellers were
used (Fig. 2) — the Lightnin axial flow impellers A-315
(Np0 = 0.75) and A-310 (Np0 = 0.35), the four blade 45°
Pitched Turbine pumping Down (PBTD) (Np0 =1.30)
and the Rushton Disk Turbine (RDT) (Np0 =5.10 –
5.40). The different tested configurations are given in
Table 1. The gas phase was air and the liquid phase was
tap water. Compressed air is introduced via a ring
sparger with 2-mm holes. Gas rates ranged from 0.54×
10 − 3 to 2.62× 10 − 3 m3 s − 1 (0.2 –1.1 vvm) and rota-
tional speed ranged from 1.66 to 11.67 s − 1. Torque is
measured by strain gauges mounted on the impeller’s
shaft. To estimate mixing times, two kinds of experi-
ments were performed.

2.1. Discoloration method

It includes the neutralisation of an acid (100 ml of


Fig. 1. The experimental apparatus. 1, Vessel; 2, impellers; 3, sparger; H2SO4 (95% in weight, z= 1830 kg m − 3) by an alkali
4, baffles; 5, micropropeller; 6, conductimetric probe; 7, conductime- (157 g of NaOH (95% in weight)) in the presence of
ter; 8, recorder; 9, magnetic counter; 10, strain gauges; 11, variable phenolphthalein. In this case, liquid is discoloured pro-
speed motor; 12, transducer; 13, computer; 14, camera; 15, video tape gressively, video recordings show clearly dead
recorder; 16, screen; 17, water manometer; 18, flowmeter; 19, oxygen
zones. Unfortunately, this technique cannot be used in
probe; 20, oxymeter. Characteristic dimensions (in meters): T= 0.43;
HL =0.86; C= 0.43; h = 0.10; B= 0.043; r= 0.09. the gas –liquid systems because of the visibility prob-
lems.
Table 1
The various configurations
2.2. Conducti6ity method
Configuration Lower impeller Upper impeller D/T
In this case, a small quantity of tracer (100 ml of
A-315+A-315 A A-315 A-315 0.33 NaCl solution (250 g l − 1)) is added on the top of the
A-315+A-310 B A-315 A-310 0.33–0.44
dispersion surface and the detector is placed in the
A-315+PBTD C A-315 PBTD 0.33–0.44
RDT+A-315 D RDT A-315 0.33 plane of the lower impeller near the wall. The position
RDT+A-310 E RDT A-310 0.33 of the probe and the injection point were diametrically
RDT+PBTD F RDT PBTD 0.33 opposed. An output signal is transmitted to a conduc-
timeter, which is connected to a recorder. The mixing
time is the time required for the distribution of the
tracer throughout the entire volume of the reactor. This
parameter tm99 is defined as the time that it takes for
electrical conductivity fluctuations to be less than 9
1%. On the plot of the evolution of the electrical
conductivity as a function of time, about 50 points are
chosen in order to evaluate the mixing time.

3. Results and discussion

3.1. Power consumption

The cost of a process depends on the power con-


sumption. In order to ensure an optimum productivity,
Fig. 2. The various studied impellers. energy consumption must be optimised. As a conse-
M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95 89

and lower impellers. When the three axial configura-


tions are compared, the power reduction depends both
on the lower and the upper impellers. That can be
explained by the difference of the geometrical charac-
teristics, especially the number and the area of the
upper impeller blades, which are more important for
the PBTD impeller. This impeller has a higher power
number compared with the A-310. It is well known that
the impellers work in different regimes, so cavities
formed behind impeller blades are different because of
the different quantities of gas received. The quantity of
gas received by the upper one is less than that received
by the lower one. When PBTD is used, the fluid density
in the upper half of the tank is further reduced and/or
the cavities, which are formed behind the upper im-
peller blades, are more important. These results imply
Fig. 3. The relative power as a function of rotational speed at
that the upper impeller geometric characteristics have a
Qg = 2.41 ×10 − 3 m3 s − 1. significant effect on the overall power reduction of axial
flow combinations and as a consequence on the hy-
draulic efficiency. If axial and radial configurations are
compared, we observe that for a given gas flow rate, the
ratio Pg/P0 is found less for the radial configurations
when they are compared with the axial ones. It follows
from the results of experiments made, that it is possible
to increase the mixing efficiency at less energy with
axial configurations A and B. This means that more
effective power can be used for the axial configurations
to disperse the gas in the liquid. For a given power
input, the axial configurations have a better hydraulic
efficiency than the radial ones. The axial configurations
are more able to conserve their pumping capacities in
the presence of gas.
Based on Pharamond’s [1] correlation obtained for a
mono-stage system equipped with a Rushton turbine
Fig. 4. 1 − (Pg/P0) vs. the Froude number of the gas Frg for the axial
configurations A, B and C. Pg
1− = 1.6Qg(vvm)D 0.63 (1)
P0
quence, it is important to understand the hydrodynamic
and mass transfer mechanisms. if Qg(vvm)D 0.63 B 0.3, we obtain,
For non aerated systems, the total input power mea- Pg Ug
sured for multiple impellers is found to be the sum of 1− = Cte D 0.13 (2)
the power consumed by the individual impellers, pro-
P0
gD
vided there is a sufficient clearance between impellers. D 0.13 varies in a little way
The distance between impellers is chosen equal to the
Pg Ug
tank diameter in order to avoid interaction between 1− = Cte = C%Frg (3)
them [20]. P0
gD
In aerated systems and for a given gas flow rate, the Frg is the Froude number of the gas.
rotational speed corresponding to the complete disper- The results obtained with the different configurations
sion regime depends on the lower impeller type and are reported in Figs. 4 and 5.
diameter [20]. As shown in Fig. 3, the Pg/P0 ratio has For the axial configurations A, B and C (Fig. 4)
been found decreasing and remaining constant beyond 1−Pg/P0 varies linearly with Ug
gD. The slope of the
the complete dispersion point. These results agree with curve is evaluated to be 38.2. For the mixed configura-
Roustan et al. [3,11] and Abrardi et al. [7] findings. tions D, E and F (Fig. 5), the slope is estimated to 60.4
When the aeration takes place, the power consumption for gas Froude numbers less than 5.2× 103. For Frg \
is decreased (Fig. 3). However, the global power con- 5.2× 103, Pg/P0 is a constant equal to 0.37. It should be
sumption and its reduction depend both on the upper noted that the constant depends on the scale.
90 M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95

Fig. 5. 1 − (Pg/P0) vs. the Froude number of the gas Frg for the axial
Fig. 8. tmg =f(1/N, configuration B, D/T= 0.33).
configurations C, D and E.

3.2. Mixing time and mixing efficiency

3.2.1. Mixing time in non-aerated systems


The measurements of decolorisation times are com-
pared with those obtained by the conductivity technique.
According to video recordings, the probe was placed near
the bottom impeller where the last volume of colour
disappears when the decolorisation technique is used.
Manna et al. (1991) [27] reported that the comparison
between the mixing time values obtained with different
experimental techniques should be very careful because of
the influence of the injecting and measuring position. They
found that the mixing times obtained using the decolorisa-
tion technique (redox reaction) are lower than those
obtained using the conductivity technique. The authors
assumed that this behaviour is due to the probe volume.
Effectively, Thyn et al. [28] found that the mixing time is
strongly dependent on the size of the probe. The mixing
time decreases by increasing the volume of the probe.
Fig. 6. The comparison between the conductivity and the decolorisa- However, in opposition to Manna et al. [27] findings and
tion techniques in non-aerated for configurations b and c with
D/T =0.44.
in agreement with John et al. [25], the results obtained in
this work show that the conductivity and decolorisation
techniques give similar mixing time values for the same
operating conditions (Fig. 6).
It was observed that for all the tested configurations, the
mixing time decreases when the rotational speed increases
(Figs. 7–12).
The visualisation of flow patterns helped in the interpre-
tation of the mixing time results. The mixing mechanisms
are found a strong function of the flow patterns structure
created by the impeller configuration. In non-aerated
systems and at constant D/T, mixing times obtained in the
turbulent regime are inversely proportional to the rota-
tional speed. This confirms Nocentini et al. [5] hypothesis,
which supposes that in the case of a multicellular
behaviour, the mixing time is proportional to the ex-
change flow between compartments ND 3. As a conse-
quence, tm0 is proportional to 1/N. The results obtained
show that the mixing number Ntm0 is a constant depend-
Fig. 7. tmg = f(1/N, configuration A, D/T =0.33). ing on the configuration type and diameter (Table 2).
M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95 91

For D/T= 0.33 and a given rotational speed, the


configurations C and F give the highest values of
mixing time while the configuration E gives the lowest
ones. According to video recordings and local liquid
measurements (not reported here), the A-315, which
has a particular geometry and an important projected
blade area (Fig. 2), is characterised by high pumping
capacities. For the various studied configurations, the
qualitative flow patterns are reported in Fig. 13. The
combination of two A-315 impellers gives a strong axial
component, and a big loop has a tendency to be formed

Fig. 12. tmg =f(1/N, configuration F, D/T =0.33).

Table 2
The turbulent mixing number obtained in non aerated systems

Configuration D/T Ntm0 conduc. Ntm0 decolorisation

A-315+A-315 A 0.33 62 59
A-315+A-310 B 0.33 80 65
0.44 40 38
A-315+PBTD C 0.33 105 –
0.44 75 77
RDT+A-315 D 0.33 65 62
RDT+A-310 E 0.33 50 49
RDT+PBTD F 0.33 85 66
Fig. 9. tmg =f(1/N, configuration C, D/T=0.33).

(Fig. 13a). When the A-310 is placed as upper impeller,


the flow patterns generated by this impeller are similar
to the A-315. For axial dual impellers configurations A
and B, each impeller pumps the liquid axially. Video
recordings show that a radial liquid velocity appears
near the lower impeller. In this case again, a big loop
has a tendency to be formed (Fig. 13a).
The flow patterns generated by the RDT are radial.
After impact with the wall, the streamlines curve and
an axial liquid velocity component appears. The as-
cending axial flow generated by the upper impeller
counters the one generated by the lower impeller. For
Fig. 10. tmg =f(1/N, configuration D, D/T=0.33). the configurations D and E (Fig. 13c), the reactor
operates in one loop. There is no creation of compart-
ments. The PBTD used as upper impeller generates a
mixed flow pattern. Independently of the lower impeller
type, two loops have a tendency to form. Each impeller
establishes its own zone of influence and the reactor
operates in two compartments (Fig. 13b and d). The
dual-impeller systems C and F can be regarded as two
independent loops created by each impeller with recip-
rocal mass exchange (Abrardi et al. [7]). The systems A,
B, D and E can be regarded as one loop. For a given
rotational speed, the mixing time obtained with
configurations A, B, D and E is less than that of the
systems C and F. This behaviour is due to the compart-
mental flow pattern created only by the mixed configu-
Fig. 11. tmg =f(1/N, configuration E, D/T=0.33). rations C and F.
92 M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95

In non-aerated systems, the mixing number can be the impeller’s pumping capacities, which depend di-
correlated to the ratio D/T. The following relationship rectly on the power consumption.
is proposed for the configurations B and C. For a given rotational speed, configuration E gives

Ntm0 =Cte

D 2
(4)
the shortest mixing time, while the configuration C
gives the highest values of mixing time when it is
T compared with the other configurations. It should be
noted that both conductivity and discoloration tech-
For these configurations, the constants, which de-
niques give and idea about the performances of the
pend on the scale, are, respectively, 8.3 and 13.0.
various tested configurations.
The increase of mixing time in the presence of gas is
3.2.2. Mixing time in aerated systems more important for the configurations corresponding to
Mixing time measurements were also made in aerated a one-loop behaviour (A, B, D and E, Figs. 7, 8, 10 and
systems. The mixing time reported as a function of the 11 while the increase of tmg is shorter with the configu-
rotational speed velocity shows the evolution of the rations corresponding to a two-compartment behaviour
different hydrodynamic regimes (Figs. 7 – 12). (configuration C and F, Figs. 9 and 12). It can be
For a given configuration, it was observed that the explained by the reduction of the gas effect due to the
value of tmg depends principally on the lower impeller’s cells. These results confirm Myers et al. [21] findings
hydrodynamic regime. In order to describe and under- that the mixing time provided by a four-impeller mixed
stand the behaviour of the axial configurations, video configuration is less than one half of that of radial one.
recordings were done at constant gas flow rate and For a given rotational speed, when the relative di-
increasing rotational velocity. In the flooding condi- ameter D/T is increased, the mixing time decreases. It is
tions (N BNCD), the agitation has no effect on bubble a consequence of the increase of the impellers pumping
dispersion. The ascending gas flow imposes the flow capacities and the liquid circulation. However, a reduc-
patterns and tmg varies in little way with N. For the tion of the mixing time is achieved by increasing the
conditions situated beyond the complete dispersion power consumption.
point, tmg is proportional to 1/N. It should be noted A dimensionless relationship is proposed to predict
that the complete dispersion regime corresponds to the the mixing time in the aerated systems for operating
situation, in which bubbles reach the bottom of the conditions situated beyond the complete dispersion
reactor. For axial configuration A, B and C (Fig. 13a point,

 n 
and b), the A-315 impeller has sufficient pumping ca-
pacities for the bubbles to arrive at the bottom of the Ntmg Qg 0.56
D 0.50

reactor and the gas is well dispersed. For configurations − 1= Cte (5)
Ntm0 D
Dg
2 T
D, E and F (Fig. 13c and d), the complete dispersion is
reached when bubbles arrive at the bottom of the (N] NCD, Ug 5 0.011 m s − 1, R 2 = 88%).
reactor after impact with the wall. For a given configu- For the studied combinations A–F, the values of the
ration and constant rotational speed, the mixing time constants, which depends on the scale are, respectively,
increases when the gas flow rate increases. The presence 4.13, 5.90, 1.60, 1.21, 3.54 and 1.26. This correlation is
of gas reduces the impeller pumping capacities. The in good agreement with that proposed by Rewatkar
influence of the gas can be described by the decrease of and Joshi (1991) [24]. These authors found that the

Fig. 13. The qualitative flow patterns for the various studied configurations.
M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95 93

give similar gas hold-up performances for a given


power input and superficial gas velocity. However, as
shown in Fig. 15 and for a fixed mixing time, the
configurations C and F consume more energy than the
other configurations.
In order to get a more complete view about the
mixing efficiency, we have calculated the homogenisa-
tion energy, which is defined as the product of mixing
time and power consumption. The mixing efficiency
Pgtmg is considered as the criterion of comparison of
mixing performances of the different studied configura-
tions. The highest mixing efficiency corresponds to the
minimum energy consumption Pgtmg. For a rotational
Fig. 14. Gas hold-up vs. the volumetric power consumption at speed, N=8.33 s − 1 (N\NCD), the mixing results are
Qg = 0.78 ×10 − 3 m3 s − 1. reported in Table 3. As shown, the axial configurations
B and A present the best mixing performance, which
corresponds to lowest homogenisation energy. The
configurations C and F presenting a two-compartment
behaviour and those having a Rushton turbine as lower
impeller (D, E, and F) are those, which present the
worst performance. These observations show that the
choice of the upper impeller is very important. Axial
configurations A and B give very interesting results. It
should be noted that the complete dispersion regime is
reached at lower power input with axial configurations
A and B (Bouaifi et al. [20]). As a consequence, the use
of an axial upper impeller consuming moderate power
is an interesting choose. It is the case of many processes
and operations, which need mass and heat transfer and
Fig. 15. Mixing time vs. the volumetric power consumption at particle suspension in order to avoid local temperature
Qg = 0.78 ×10 − 3 m3 s − 1.
and concentration gradients, which is the case of many
physical and chemical processes like catalytic chloration
exponent of the Froude number of the gas is a function and hydrogenation, in which the mixing must be homo-
of the aeration system type and proposed a value of
geneous and the temperature maintained within very
0.58 for a tank equipped with a single Pitched Blade
narrow limits in order to obtain an optimum
Turbine and a concentric ring sparger.
productivity.
The effect of gas flow rate on mixing efficiency is
3.2.3. Mixing efficiency shown in Table 3 and Fig. 16. The configuration E
It was also observed that the power consumption is presents better mixing efficiency at low gas flow rates
not the best criterion to evaluate mixing time perfor- than the configuration D. However, at high gas flow
mances. Fig. 14 shows that the studied configurations rates, both configurations give similar performances.

Table 3
The mixing time results for the configurations with a ratio D/T= 0.33 at N = 8.33 s−1 (N\N00)

Qg = 0 m3 s−1 Qg = 0.78×10−3 m3 s−1 Qg =1.56×10−3 m3 s−1

tm0 (s) P0 (W) tm0P0 (Ws) tmg (s) Pg (W) tmgPg (Ws) tmg (s) Pg (W) tmgPg (Ws)

A 6.6 65.7 434 10.8 60.3 651 16.2 40.2 651


B 10.2 43.0 439 13.3 37.9 504 16.2 30.5 494
C 12.0 85.1 1021 15.3 70.4 1077 17.3 48.7 842
D 8.8 211.4 1860 9.5 153.7 1460 10.2 100.7 1027
E 5.0 187.6 938 8.3 136.9 1136 11.4 90.1 1027
F 10.2 229.3 2339 15.0 167.8 2517 15.0 117.8 1767
94 M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95

Acknowledgements

We wish to thank Dr Jean-Claude Pharamond for his


help.

Appendix A. Notation

D impeller diameter (m)


Frg Froude number of the gas (= Ug/
gD)(−)
g acceleration due to gravity (m s−2)
Fig. 16. Mixing efficiency vs. gas flow rate at N= 8.33 s − 1. Qg gas flow rate (m3 s)
N rotational speed (s−1)
Ntm mixing number (−)
P0 power consumption in non aerated (W)
4. Conclusions
Pg power consumption in aerated (W)
T tank diameter (m)
The objectives of this work were to study the be-
tm0 mixing time in non aerated, (s)
haviour and the performances of different axial and
tmg mixing time in aerated (s)
mixed configurations. The dual-impeller gas –
Ug superficial gas velocity (m s−1)
liquid reactor behaviour has been found to be very
Vl liquid volume (m3)
complicated. Their design must be considered with a
great care. Axial configurations showed very Indices
interesting energetic and mixing performance. They al- g gas
low the dispersion of the gas at reduced energy con- l liquid
sumption. The ratio Pg/P0, which is constant beyond CD complete dispersion
the complete dispersion, has been found inferior for
mixed configurations than the one for axial ones. This
means that more effective power can be used for the
axial configuration to disperse the gas in the liquid. For
a given power input, the axial configurations are more
References
capable to conserve their pumping capacities in the
presence of gas. As a consequence, the axial configura- [1] J.C. Pharamond, Contribution à l’Etude de l’Agitation en milieu
tions ensure good mixing efficiency with less energy. A Aéré’’, Thése de docteur-ingénieur, Université Paul Sabatier,
dimensionless correlation is proposed to predict the Toulouse, France, 1973.
power consumption beyond the complete dispersion [2] C.M. Chapman, A.W. Nienow, M. Cooke, J.C. Middleton,
point. Particle– gas– liquid mixing in stirred vessels. II. Gas–liquid
mixing, Chem. Eng. Res. Des. 61 (1983) 82 – 95.
The conductivity and decolorisation techniques give [3] M. Roustan, Power consumed by Rushton turbines in non
similar mixing results for the same operating condi- standard vessels under gassed conditions, Proceedings of the
tions. In non-aerated systems, the mixing number is a Fifth International Conference on Mixing, Wurzburg, Germany,
constant dependent on configuration type and 1985, pp. 127– 141.
impeller diameter. In the aerated systems, the presence [4] V. Abrardi, G. Rovero, S. Sicardi, G. Baldi, R. Conti, Sparged
vessels agitated with multiple turbines, Proceedings of the Sixth
of the gas reducing the impellers pumping capacities International Conference on Mixing, Pavia, Italy, 1988, pp.
increases the mixing time. However, beyond the com- 329– 336.
plete dispersion point, it was shown that the mixing [5] M. Nocentini, F. Magelli, G. Pasquali, D. Fajner, A fluid
number is a constant dependent on the gas flow rate, dynamic study of a gas– liquid, non-standard vessel stirred by
the configuration type and diameter. The power con- multiple impellers, Chem. Eng. J. 37 (1988) 53 – 59.
[6] V. Hudcova, V. Machon, A.W. Nienow, Gas– liquid dispersion
sumption is not the best criterion for evaluating the with dual Rushton turbine impellers, Biotech. Bioeng. 34 (1989)
mixing performances. The mixing time depends on the 617– 628.
flow patterns structure. The energy of homogenisation [7] V. Abrardi, G. Rovero, S. Sicardi, G. Baldi, R. Conti, Hydrody-
is lower with the axial configurations B and A, which namics of gas– liquid reactor stirred with multi-impeller system,
present the best mixing efficiency performances. The Trans. Inst. Chem. Eng. 68 (1990) 516– 522.
[8] S.M. Mahmoudi, M. Yianneskis, The variation of flow pattern
use of the axial configurations can be very useful for and mixing time with spacing in stirred vessels with two Rushton
three-phase systems and non-isotherm catalytic pro- turbines, Proceedings of the Seventh International Conference
cesses. on Mixing, Brugge, Belgium, 1991, pp. 17 – 24.
M. Bouaifi, M. Roustan / Chemical Engineering and Processing 40 (2001) 87–95 95

[9] J.M. Smith, Simple performance correlations for agitated vessels, mécaniquement par des mobiles axiaux et radiaux, Thèse de
Proceedings of the Seventh International Conference on Mixing, Doctorat, INSA de Toulouse, France, 1997.
Brugge, Belgium, 1991, pp. 233–241. [19] M. Bouaifi, M. Roustan, R. Djebbar, Hydrodynamics of multi-
[10] J. Oldshue, T. Post, R. Weetman, C. Coyle, Comparison of mass impeller agitated gas– liquid reactors, Proceedings of the Ninth
transfer characteristics of radial and axial flow impellers, Pro- European Conference on Mixing, Paris, France, 1997, pp. 137–
ceedings of the Sixth International Conference on Mixing, Pavia, 144.
Italy, BHRA Fluid Eng., Cranfield UK, 1988, pp. 345–350. [20] M. Bouaifi, M. Roustan, Bubble size and mass transfer coeffi-
[11] M. Roustan, D. Auban, M. Renaudin, Hydrodynamique et cients in dual-impeller agitated reactors, Can. J. Chem. Eng. 76
transfert de matière gaz/liquide dans les cuves agitées mécanique- (1998), pp. 390– 397.
ment, Récents Progrès en Génie des Procédés 3 (8) (1989) 72 – [21] K.J. Myers, J.B. Fasano, A. Bakker, Gas dispersion using mixed
79. high efficiency/disc impeller systems, Proceedings of the Eighth
[12] A. Bakker, Hydrodynamics of stirred gas-liquid dispersions, International Conference on Mixing, Cambridge, UK, 1994, pp.
Ph.D. thesis, Delft University, Netherlands, 1991. 65 – 72.
[13] C.M. McFarlane, Gas–liquid mixing studies with A-315 and [22] A. Einsele, R.K. Finn, Influence of gas flow rates and gas
prochem hydrofoil agitators, Ph.D. thesis, Birmingham Univer- hold-up on blending efficiency in stirred tanks, Ind. Eng. Chem.
sity, 1991. Process Des. Dev. 19 (1980) 600– 603.
[14] H. Desplanches, Y. Gaston-Bonhomme, R.J.M. Arnaud, J.L., [23] A.B. Pandit, J.B. Joshi, Mixing in mechanically agitated gas–liq-
Chevalier, Hydrodynamiques comparées de dispersions gaz– liq- uid contactors, bubble columns and modified bubble columns,
uide Newtoniens et non Newtoniens réalisées par deux agitateurs Chem. Eng. Sci. 38 (1983) 1189.
axiaux à pompage descendant, Proc. 4ème Congrès Français de [24] V.B. Rewatkar, J.B. Joshi, Role of sparger design in mechani-
Génie des Procédés, Grenoble, France, 1993, pp. 67–68. cally agitated gas– liquid reactors, Chem. Eng. Technol. 14
[15] A.W. Nienow, R.J. Weetman, A fluid Dynamic study of the (1991) 386– 393.
retrofitting of a large pilot scale agitated bioreactor, Proceedings [25] A.H. John, W. Bujalski, A.W. Nienow, A novel reactor with
of Third International Conference on Bioreactor and Bioprocess independently-driven dual impellers for gas– liquid processing,
Fluid Dynamic, Cambridge, UK, 1993, pp. 505–519. Proceedings of the Ninth European Conference on Mixing,
[16] V. Schlüter, W.-D. Deckwer, Power input, mixing and mass Paris, France, 1997, pp. 169– 176.
transfer in bioreactors with single and multiple stirrers, Proceed- [26] J.M.T. Vasconcelos, S.S. Alves, A.W. Nienow, W. Bujalski,
ings of the Third International Conference on Bioreactor and Scale-up of mixing in gassed multi-turbine agitated vessels, Can.
Bioprocess Fluid Dynamic, Cambridge, UK, 1993, pp. 117– 133. J. Chem. Eng. 76 (1998) 398– 404.
[17] D. Pinelli, M. Nocentini, F. Magelli, Hold-up in low viscosity [27] L. Manna, F. Chiampo, G. Rovero, S. Sicardi, Mixing time in
gas– liquid systems stirred with multiple impellers. Comparison vessels stirred by mutiple impeller, Proceedings of the Seventh
of different agitators types and sets, Proceedings of the Eighth International Conference on Mixing, Brugge, Belgium, 1991, pp.
International Conference on Mixing, Cambridge, UK, 1994, pp. 111– 119.
81 – 88. [28] J. Thyn, V. Novak, P. Pock, Effect of the measured volume size
[18] M. Bouaifi, Etude de l’hydrodynamique et du transfert de on the homogenisation time, Chem. Eng. J. 12 (1976), pp.
matière gaz– liquide dans des réacteurs multi-etagés agités 197– 211.

View publication stats

You might also like