Sei sulla pagina 1di 9

Chemical Engineering Science 57 (2002) 3325 – 3333

www.elsevier.com/locate/ces

Hydrodynamics and mass transfer in gas–liquid #ow


through static mixers
A. Heyounia , M. Roustana; ∗ , Z. Do-Quangb
a LIPE EA 833, Dpt GPI, INSA 135 Avenue de Rangueil, 31077 Toulouse, France
b CIRSEE-Lyonnaise des eaux, 38 Avenue du Pr'esident Wilson, 78230 Le Pecq, France

Received 23 April 2001; received in revised form 12 November 2001; accepted 12 November 2001

Abstract

The objective of this study was to characterize the two-phase #ow hydrodynamic behaviour and mass transfer in a static mixer in
a horizontal pipe. Di5erent arrangements of elements of the static mixer were tested and their performances compared. The pressure
drop, bubble diameters and mass transfer coe6cient were measured. The in#uence of operating conditions was also studied. A di5erent
correlations are proposed and compared with other correlations found in the literature. ? 2002 Elsevier Science Ltd. All rights reserved.

Keywords: Static mixer; Hydrodynamics; Pressure drop; Bubble diameter; Interfacial area; Mass transfer

1. Introduction Adrien, & Luck, 2000):

In recent years, static mixers have become an integral and • high mixing between the gas and the water,
basic equipment in the chemical process industries. These • small diameter bubbles,
mixers are found in a wide range of applications. They are • high interfacial area,
used for laminar and turbulent mixing of miscible liquids, • high mass transfer coe6cient,
in laminar-#ow heat exchangers, for laminar and turbulent • plug #ow behaviour.
homogenization, as tubular reactors, for dispersion of im-
miscible phases, and for interphase mass transfer between The objective of the present work was to characterize the
immiscible phases (Mustakis, Strei5, & Schneider, 1986; hydrodynamic behaviour of two-phase #ow and the mass
Baker, 1991; Germain, Strei5, & Juvet, 1996; Myers, Baker, transfer in a static mixer in order to evaluate its capacity to
& Ryan, 1997). transfer certain gases (O2 , O3 ). To this end, the following
A static mixer consists of a series of stationary mixing parameters were determined: pressure drop, bubble diame-
elements inserted end-to-end in a pipe. Each element is a ters, interfacial area and mass transfer coe6cient, as a func-
specially designed rigid structure that divides the #ow and tion of two operating parameters, liquid and gas #ow rates.
recombines it in a geometric sequence. Mixing and contact-
ing take place as the #uids are sheared and directed radially
across the pipe or duct. The only power required for static 2. Materials and methods
mixers is the external pumping power that propels the #uids
through the mixer. Static mixers o5er many advantages as A schematic diagram of the experimental apparatus is
gas–liquid reactors (Cybulski & Werner, 1986; Rader, Mus- shown in Fig. 1. The experiments were conducted in a hor-
takis, Grosz-Roll, & Maugweiler, 1989; Traversay, Bonnard, izontal pipe 9:7 m in length and 52:6 mm in diameter. A
Lightnin static mixer was located at a distance of 4:7 m from
the pipe inlet. The characteristics of this mixer are given
∗ Corresponding author. Tel.: +33-0561-5597-51; in Table 1. Air was injected upstream of the static mixer
fax: +33-0561-55-9760. through a tube having holes of internal diameter 4 mm. Wa-
E-mail address: roustan@insa-tlse.fr (M. Roustan). ter was circulated through the pipe using a centrifugal pump.

0009-2509/02/$ - see front matter ? 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 0 9 - 2 5 0 9 ( 0 2 ) 0 0 2 0 2 - 6
3326 A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333

shown in Fig. 2. The helixes were placed so that the Lrst el-
ement of a left-hand helix straddled the Lnal element of the
right-hand helix. This was obtained by indexing the adja-

cent blade edges 60 from each other. ConLguration 2 was
composed of 4 helixes of 7 elements in which each of the
Lrst 3 helixes was followed by 4 elements. These 4 elements

were placed in such a way that an angle of 60 was formed
between 2 consecutive elements. In conLguration 3, all the

elements were placed in such way that an angle of 60 was
formed between 2 consecutive elements.
Experiments were performed at superLcial gas velocities
ranging from 0 to 0:5 m=s and at superLcial liquid velocities
from 0.1 to 2 m=s. In this study, the following parameters
were determined:

• Pressure drop.
• Bubble diameter, db .
• Mass transfer coe6cient kL a.
Fig. 1. Experimental plant. 1: Water tank, 2-Pump, 3-Rotameter of liquid,
4-Air inlet, 5-Rotameter of gas, 6-Pipe, 7-Static mixer, 8-Pressure drop 2.1. Pressure drop
measurement, 9-Oxygen probe.
Pressure drop was measured using a manometer con-
nected to a computer. Pressure drop measurements were
Table 1
Geometric characteristic of Lightnin static mixer tested taken at points 12:5 mm before the Lrst element at the in-
let and 12:5 mm after the last element at the outlet, result-
Interior diameter, D 52:6 mm ing in an e5ective length of 52 cm for the pressure drop.
Length of static mixer, L 49:5 cm
This pressure drop was measured for di5erent liquid and of
Void fraction, SM 87%
gas #ow rates. For each experimental condition, the instan-
taneous pressure drop was measured for 5 minutes and the
average value deduced.

2.2. Bubble diameter

Bubble size was determined using a video camera (Sony


XC-75=CE) connected to a computer. A program was de-
veloped using Visilog Software to obtain the average value
of the diameter (dbm ) of each bubble. For each gas velocity
and liquid velocity, 200 –220 bubbles were analysed. The
average Sauter diameter, db , was calculated from the fol-
lowing relation:
 
3
m db m
db =  2 :
m db m

For experimental reasons, bubble diameter measurements


Fig. 2. The Lightnin static mixer structure tested.
were done just at the outlet of the mixer. It was assumed that
the diameters of bubbles measured at the exit of the mixer
were identical to those found inside the mixer.
The Lightnin static mixer consisted of 42 elements (21
left-hand elements and 21 right-hand elements). Each ele- 2.3. Gas–liquid mass transfer coe9cient
ment had three blades. The regular assembly of seven ele-
ments produced a helix. The characteristics of the elements The mass transfer coe6cient was measured in an oxygen–
of this static mixer made it possible to adapt the conLgura- nitrogen–water system according to the method of Turunen
tion of the mixer to every speciLc application. In this work, and Haario (1994). Water was saturated with oxygen in a
we tested three di5erent conLgurations (Fig. 2). ConLgu- tank and fed into the static mixer. Just before the reactor
ration 1 comprised 3 right-hand and 3 left-hand helixes as entrance, gaseous nitrogen was fed into the water stream.
A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333 3327

Therefore, inside the static mixer, nitrogen was transferred 0.6


from the gas bubbles into the liquid, and oxygen from the structure1
0.5
liquid into the gas bubbles. Oxygen concentration in the liq- structure2

Pressure drops (bar)


uid was measured on-line with a probe at the inlet and out- 0.4 structure3
let of the static mixer. By a simple balance of mass transfer
between the entrance and the exit of the static mixer, it was 0.3
possible to determine the values of the mass transfer coe6-
0.2
cients kL a from the following equation:
 ∗ ∗
 0.1
QL Ln((CLe − CLe )=(CLs − CLs ))
kL a = (CLe − CLs ) ∗ ) − (C − C ∗ ) :
VL (CLe − CLe Ls Ls 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
UL (m/s)
3. Results and discussions
Fig. 3. Comparison of pressure drops between the di5erent structures
3.1. Gas hold-up tested (Ug = 0 m=s).

The concept of slip velocity proposed by Lapidus and


Elgin (1957) and Wallis (1969) can be used to predict the 0.6
Ug = 0 m/s
Ug = 0.016 m/s
gas hold-up in a homogeneous #ow regime. According to Ug = 0.043 m/s
this concept, the slip velocity Us is the relative velocity of Pressure drop (bar) 0.5
Ug = 0.103 m/s
Ug = 0.185 m/s
the two phases Ug and UL . Us is given by the equation Ug = 0.298 m/s
Ug = 0.437m/s
Ug UL 0.4
Us = − :
g L 0.3
In the case of two-phase #ow in an immersed horizontal pipe
with static mixer, the phases have the same velocity, so the 0.2
horizontal slip velocity is close to zero. In these conditions
0.1
Us = 0 and the gas hold-up is given by
Ug 0
g = :
Ug + U L 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
UL (m/s)
3.2. Pressure drop
Fig. 4. Pressure drop versus liquid and gas velocities (structure 2).

The pressure drop is the most important design criterion


for a static mixer. It permits to quantify the dissipated power Table 2

in the static mixer. The pressure drops is also a decisive Pressure drop correlation parameters, OP = AUL Ug
factor for the estimation of the e6ciency of static mixers.
Structure A   R2
It depends on the following parameters (Zhu, Liu, & Xu,
1989; Traversay et al., 2000): Structure 1 0.18 1.74 0.09 0.99
Structure 2 0.18 1.71 0.07 0.99
• liquid and gas #ow rates, Structure 3 0.21 1.80 0.05 0.99
• type of #uid,
• type, number and arrangements of element of static mixer.
The experimental results show that the pressure drop de-
In this study, the pressure drops were measured for the pends not only on the liquid velocity but also on the gas
3 structures tested for di5erent liquid and gas velocities. velocity. To take the in#uence of the gas on the pressure
Fig. 3 shows a comparison of pressure drop between the drop into account, we propose the following correlation:
di5erent structures. As can be seen from this Lgure, there
OP = AUL Ug
is practically no di5erence between the pressure drop mea-
surements in structures 1 and 2. The pressure drop in struc- with A,  and  parameters to be identiLed for each structure.
ture 3 is higher than the pressure drop in structures 1 and 2. The di5erent parameters relative to the structures and the
Fig. 4 shows an example of the variation of pressure drop correlation coe6cients R2 are given in Table 2.
with gas velocity for structures 2. It can be seen that, for Table 2 shows that the coe6cients A,  and  are iden-
a Lxed liquid velocity, the pressure drop increases slightly tical for structures 1 and 2: the pressure drop is the same
with increasing gas superLcial velocity. The same result has for these 2 structures. For structure 3, the pressure drop is
also been noted for the two other structures. slightly greater. The values of the exponent  are very small
3328 A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333

Dissipated power per unit o fmass (W/kg)


3.5 350
Ug = 0.016 m/s Ug = 0 m/s
Ug = 0.043 m/s Ug = 0.016 m/s
3 Ug = 0.103 m/s 300
Ug = 0.043 m/s
Ug = 0.185 m/s
Ug = 0.103 m/s
2.5 Ug = 0.298 m/s 250 Ug = 0.185 m/s
Ug = 0.437 m/s
Friction factor f

Ug = 0 m/s Ug = 0.298 m/s


2 200 Ug = 0.437 m/s

1.5 150

1 100

0.5 50

0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0 20000 40000 60000 80000 100000 120000
UL (m/s)
Re

Fig. 6. Variation of power dissipated with liquid and gas velocities


Fig. 5. Evolution of friction factor (structure 2).
(structure 2).

compared to . This indicates that the liquid velocity has a


greater e5ect than the gas velocity on pressure drop. Fig. 6 represents an example of the dissipated power ob-
tained for structure 2 according to the superLcial velocity of
3.3. Friction factor the liquid and gas. This dissipated power (20 –200 W=kg)
is distinctly greater than the value for a conventional stirred
The friction factor f is a very important parameter since tank (0.5 –3 W=kg). The power dissipation depends on the
its knowledge enables the pressure drop generated by the liquid velocity and the gas velocity. For a given liquid ve-
static mixers to be evaluated. The friction factor f is deduced locity, an increase of gas velocity results in an increase in
from following equation: the power dissipation.
OP D
f= :
2UL2 L 3.5. Bubble diameter
The friction factor f depends on the Reynolds number and
on the geometry of the static mixer. Fig. 5 shows an ex- The study of bubble diameters is very useful for a better
ample of the variation of the friction factor f with the understanding of the dispersion mechanisms in a gas–liquid
Reynolds number for structure 2 and for di5erent gas veloc- reactor. The diameter of the bubbles combined with the
ities. The same behaviour was observed for the two other gas hold-up enables the interfacial area to be estimated.
structures. The coe6cient of friction f decreases quickly if In addition, the knowledge of the mass transfer coe6-
the Reynolds number increases. For high Reynolds numbers, cient kL a permits the mass transfer coe6cient kL to be
the coe6cient of friction f tended toward a plateau value calculated.
for all 3 structures tested. These limit values for structures 1; The size of the bubbles depends on the energy input, and
2 and 3 were 0.63, 0.67 and 0.97, respectively. Martin and the type and number of mixing elements (Strei5, Mathys,
Galey (1994), working on another type of static mixer, pro- & Fischer, 1997, Grosz-Roll, Battig, & Germain, 1983).
posed a value of about 2. The values of f were greater than The results of bubble diameter experiments for various gas
those obtained in an empty tube, where they are within the and liquid velocities are shown in Figs. 7–9. For a given
range 0.002– 0.006. This was due to the presence of mixers liquid velocity, increasing the gas velocity resulted in an
inside the tube. The friction factor f can be considered as increase in the bubble diameter. Conversely, for a given gas
constant in fully turbulent #ow. velocity, bubble diameter decreased with increasing liquid
velocity. Increasing the liquid velocity means increasing the
3.4. Power dissipation energy input to the system. The higher levels of turbulence
generated due to the higher power input and the #uid–solid
The power dissipated per unit of mass of liquid (W=kg) interaction, lead to greater bubble breakage. It is also worth
constitutes a real criterion for comparing the di5erent gas– noting that:
liquid reactors. The mean value of power dissipation can be
• for lower liquid velocities (UL ¡ 1 m=s), structure 3 bub-
calculated from the #ow rate and the pressure drop in the
ble diameters were smaller than those found for the other
static mixer:
    two structures. On the other hand, for higher liquid veloc-
P QL OP ity (UL ¿ 1 m=s) bubble diameters for structure 3 were
=
M SM (1 − g )VMS higher than those obtained for structures 1 and 2.
A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333 3329

Ug = 0.016 m/s 2.3 Ug = 0.016 m/s


2.3 Ug = 0.043 m/s Ug = 0.043 m/s
2.1 Ug = 0.103 m/s
Ug = 0.103 m/s Ug = 0.185 m/s
2 Ug = 0.185 m/s 1.9 Ug = 0.298 m/s
Ug = 0.298 m/s
1.7
1.7

db (mm)
1.5
db (mm)

1.4 1.3

1.1
1.1
0.9
0.8 0.7

0.5
0.5
0 20 40 60 80 100
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
UL (m/s) Dissipated power per unit of mass (W/kg)

Fig. 7. E5ect of liquid and gas velocities on bubble diameter (structure 1). Fig. 10. E5ect of power dissipated on bubble diameter (structure 2).

Ug = 0.016 m/s Table 3



2.3 Ug = 0.043 m/s Bubble diameter correlation parameters, db = AUL Ug
2.1 Ug = 0.103 m/s
Ug = 0.185 m/s
Structure A × 10−3   R2
1.9
Ug = 0.298 m/s Structure 1 2.09 − 0.9 0.11 0.97
1.7
Structure 2 2.01 − 1.01 0.10 0.96
db (mm)

1.5 Structure 3 1.93 − 0.4 0.09 0.96


1.3
1.1
0.9 Fig. 10 shows a plot of bubble diameter, obtained for struc-
0.7 ture 2, versus power dissipation per unit of mass. For the
given gas velocity of 0:016 m=s, an increase in liquid veloc-
0.5
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
ity from 0.7 to 1:3 m=s, which increased power dissipation
UL(m/s)
from 10 to 65 W=kg, led to a reduction in bubble diameter
from 1:73 to 1:03 mm.
Fig. 8. E5ect of liquid and gas velocities on bubble diameter (structure 2). The bubble diameters can be correlated as db = AUL Ug .
The di5erent parameters relative to each structure and the
correlation coe6cients R2 are given in Table 3.
2.3
The coe6cient  obtained for all structures was about 0.1.
2.1 It is thus clear that the gas velocity has the same in#uence
1.9 on bubble diameter for the three structures tested. The co-
1.7
e6cient  for structures 1 and 2 was higher than the value
obtained for structure 3. This shows that the liquid velocity
db (mm)

1.5
has more e5ect on decreasing bubble diameter in structures
1.3
Ug = 0.016 m/s 1 and 2 than with structure 3.
1.1 Ug = 0.043 m/s In forced convection, with gas dispersion in a horizontal
0.9 Ug = 0.103 m/s pipe or in vertical co-current downward #ow, the bubble
0.7
Ug = 0.185 m/s
size is imposed by the turbulent shear stresses. Hinze (1955)
Ug = 0.298 m/s
proposed a formula for estimating bubble size, namely,
0.5
 0:6  −0:4
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4  P
db = 1:14 :
UL (m/s)  M
Fig. 9. E5ect of liquid and gas velocities on bubble diameter (structure 3). Fig. 11 shows a comparison between the experimental bub-
ble diameters obtained with structure 2 and the relationship
proposed by Hinze. It can be seen that Hinze’s relation-
• for a given gas velocity, the e5ect of liquid velocity on ship does not Lt the experimental result. The bubble diam-
the decrease in bubble diameter seemed to be higher in eters predicted by this relationship are smaller than the ex-
structures 1 and 2 as compared to structure 3. perimental bubble diameters. This shows that the dissipated
3330 A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333

10 1200
Ug = 0.016 m/s
Relationship of HINZE Ug = 0.043 m/s
1000
Ug = 0.103 m/s
Experimental bubble diameters Ug = 0.185 m/s
800
Ug = 0.298 m/s

a(m2/m3)
600
d b(mm)

1
400

200

0
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
UL(m/s)
0.1
1 10 100
Dissipated power per unit of mass (W/kg) Fig. 12. Variation of interfacial area with liquid and gas velocities (struc-
ture 2).
Fig. 11. Comparison between experimental point and law of Hinze (struc-
ture 2).

Ug = 0.016 m/s
1200 Ug = 0.043 m/s
Table 4 Ug = 0.103 m/s

Bubble diameter correlation parameters, db = A(P=M ) g 1000 Ug = 0.185 m/s
Ug = 0.298 m/s
Structure A× 10−3   R2
800
Structure 1 6.45 − 0.25 0.18 0.97
a (m 2/m 3)

Structure 2 7.02 − 0.28 0.17 0.95 600


Structure 3 3.18 − 0.09 0.12 0.96

400

power is not the only parameter to be considered when esti- 200


mating the bubble diameters. For a good estimation, the gas
hold-up must also be taken into account. 0
By analogy with the stirred gas–liquid vessel, we pro- 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
pose the following correlation which takes dissipated power UL (m/s)
(P=M ) and gas hold-up g into account:
  Fig. 13. Variation of interfacial area with liquid and gas velocities (struc-
P ture 3).
db = A g :
M
The di5erent parameters obtained for each structure, and the
correlation coe6cients R2 are given in the Table 4. The co- inside the static mixer. Its value can be calculated knowing
e6cients A,  and  obtained for structures 1 and 2 are very the gas hold-up g and the bubble diameter db :
similar. The coe6cient  obtained for structure 3 is greater 6 g
than that obtained with the other two structures. This shows a= :
db (1 − g )
that the dissipated power has more e5ect on decreasing bub-
ble diameter in structure 1 and 2 than in structure 3. Note The gas hold-up in a static mixer is given by
that the gas hold-up is an important parameter which it is Ug
necessary to take into account for the estimation of bubble g = :
Ug + U L
diameters.
Figs. 12 and 13 show the variation of interfacial area a with
3.6. Interfacial area the liquid velocity for di5erent gas velocities for structure 2
et 3 (The same behaviour of interfacial area was observed
The gas–liquid interfacial area a (m2 =m3 of liquid) is for structures 1 and 2). We see that the values of interfacial
an important parameter for the design of reactors in which area are much larger than those obtained with an empty pipe
dispersion occurs because it in#uences the volumetric mass (15 –40 m2 =m3 ) or a bubble column (20 –100 m2 =m3 ). They
transfer coe6cient kL a. In the case of a static mixer, the are in the range of 100 –1000 m2 =m3 . This is due to the high
value of interfacial area a is greater than that obtained with energy dissipation inside the static mixer. The expression
an empty pipe. This is due to the high energy dissipation of interfacial area a as function of liquid and gas velocities
A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333 3331

Ug = 0.016 m/s Table 6


3.000 
Ug = 0.044 m/s
Ug = 0.103 m/s Correlation parameters of mass transfer coe6cient, kL a = A(P=M ) g
Ug = 0.185 m/s
2.500 Ug = 0.299 m/s
Ug = 0.437 m/s Structure A   R2

Structure 1 0.13 0.73 0.48 0.97


2.000
Structure 2 0.26 0.69 0.58 0.98
kL a (s 1)
_

Structure 3 0.13 0.79 0.47 0.99


1.500

45E-04 Ug = 0.016 m/s


1.000
40E-04 Ug = 0.044 m/s
Ug = 0.103 m/s
0.500 35E-04
Ug = 0.185 m/s
30E-04 Ug = 0.299 m/s
0.000

kL(m/s)
0.60 0.70 0.80 0.90 1.00 1.10 1.20 1.30 1.40 25E-04
UL (m/s) 20E-04

15E-04
Fig. 14. Variation of kL a with UL and Ug (structure 2).
10E-04

05E-04
Table 5

Correlation parameters of mass transfer coe6cient, kL a = AUL Ug 00E+00
0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4
Structure A   R2 UL(m/s)

Structure 1 2.28 1.63 0.57 0.98


Structure 2 3.45 1.50 0.64 0.98 Fig. 15. Variation of kL with liquid and gas velocities (structure 1).
Structure 3 2.88 1.80 0.53 0.99

interfacial areas in the system, thus giving rise to the high


can be obtained from the equation giving the bubble diam- mass transfer.
eter db and hold-up g , namely: Table 5 shows that, for all structures tested, the values of
 were greater than those obtained for . This shows that the
• a = 2870UL−0:1 Ug0:89 , for structure 1, liquid velocity has more e5ect on mass transfer coe6cient
• a = 2990UL0:01 Ug0:9 , for structure 2, kL a than the gas velocity.
• a = 3110UL−0:6 Ug0:81 , for structure 3, By analogy to the stirred vessels, we propose another
correlation according to the dissipated power (P=M ) and the
For all structures, the interfacial area increased when the gas hold-up g :
 
gas velocity increased. For structures 1 and 2, interfacial P
area was little in#uenced by the liquid velocity whereas for kL a = A g :
M
structure 3, the liquid velocity in#uenced the interfacial area.
The values of A, ,  and correlation coe6cients are sum-
marized in Table 6.
3.7. Mass transfer
Table 6 shows that the values obtained for  vary between
0.7 and 0.8. This is in agreement with other results published
Few experimental data on mass transfer coe6cients in
in the literature (Martin, 1992; Middleton, 1978). However
static mixers are available in the literature in comparison
the values obtained for  are lower than the values found
with those for other gas–liquid reactors (bubble columns,
in the literature, which range from 0.6 (Martin, 1992) to 1
stirred vessels). An example of experimental values of kL a
(Roes, Zeerman, & Bukkems, 1984).
obtained as a function of liquid and gas velocities is given for
the structure 2 in Fig. 14. The same behaviour was observed
for the two other structures. 3.8. Prediction of the =lm mass transfer coe9cient kL
Experimental data could be correlated with superLcial
liquid velocity and gas velocity following equation kL a = The values of the mass transfer coe6cient kL can been
AUL Ug . The values of A, ,  and correlation coe6cients estimated from the values of kL a and the speciLc interfacial
R2 are summarized in Table 5. area a, following the equation:
As can be seen from Fig. 14, mass transfer coe6cients kL a
kL = :
increased with increasing liquid and gas velocities. Much a
higher values of kL a were obtained than with other gas– Fig. 15 shows an example of the variation of kL versus liquid
liquid reactors (bubble column, stirred vessels). This e5ect velocity and gas velocity for structure 1. As can be seen,
was mainly due to an ability of such devices to develop large kL increases with increasing liquid velocity and when gas
3332 A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333

velocity decreases. The values of kL obtained are in the range Notation


of 5 –40 × 10−4 m=s and are higher than those obtained with
a interfacial area, m2 =m3
some other gas–liquid reactors such as the bubble column
A correlation parameter
(2.5 –4:3 × 10−4 m=s). The expression of kL as a function
CLe concentration of oxygen in the liquid phase at
of liquid velocity and gas velocity can be obtained from the
the inlet of the static mixer, kg=m3
expression of kL a, and from the expression for the interfacial
CLs concentration of oxygen in the liquid phase at
area a, namely:
the exit of the static mixer, kg=m3
kL = 7:94:10−4 UL1:73 Ug−0:32 ; for structure 1; D diameter of static mixer
db sauter Diameter, m
kL = 12:20:10−4 UL1:51 Ug−0:26 ; for structure 2; kL Llm mass transfer coe6cient, m=s
kL a volumetric coe6cient of mass transfer, s−1
kL = 9:26:10−4 UL2:4 Ug−0:28 ; for structure 3: L length of static mixer, m
QL #ow rate of liquid, m=s
Re reynolds number
4. Conclusion Ug superLcial gas velocity, m=s
UL superLcial liquid velocity, m=s
This study characterizes the hydrodynamics and mass VL volume of liquid, m3
transfer of a static mixer with di5erent arrangements of mix- VMS volume of static mixer, m3
ers inside the contactor. The pressure drop, bubble diame-
ters and mass transfer coe6cient were measured for various Greek letters
structures and operating conditions, and correlated. The fol-
lowing conclusions can be drawn:  density of water, kg=m3
 correlation parameter
• The pressure drop depends on the structure of the static  correlation parameter
mixer, the liquid velocity and the gas velocity. However,  void fraction of static mixer
the e5ect of liquid velocity to increase the pressure drop g gas hold-up
is very much higher than the e5ect of gas velocity. The OP pressure drop between inlet and exit of static
correlations obtained can accurately predict the pressure mixer, bar
drop for every structure tested.
• The bubbles diameters depend on liquid velocity, gas ve-
locity and also on the structure of the static mixer. References
• The e6ciency of the static mixer seems to be higher
than that of other gas–liquid reactors. Indeed, they gener- Baker, J. R. (1991). Motionless mixers stir up new uses. Chemical
ate a high interfacial area as compared with classic gas– Engineering Progress, 87(6), 32–38.
liquid reactors such as bubble columns for example, and Cybulski, A., & Werner, K. (1986). Static mixers—criteria for applications
and selection. International Chemical Engineering, 26(1), 171–180.
therefore give high values of mass transfer coe6cients
Germain, E., Strei5, F. A., & Juvet, J. E. (1996). Les mSelangeurs statiques,
(Fig. 16). des rSeacteurs simples et e6caces. Informations chimie, 374, 141–147.
Grosz-Roll, F., Battig, J., & Germain, M. (1983). Transfert de matiTere
gaz-liquide: emploi des mSelangeurs statiques Informations Chimie,
244, 131–136.
Hinze, J. O. (1955). Fundamentals of the hydrodynamic mechanism of
slipping in dispersion processes. AIChE Journal, 1, 289.
Lapidus, L., & Elgin, J. C. (1957). Mechanics of vertical moving #uidized
systems. AIChE Journal, 3, 163.
Martin, N. (1992). Etude de l’application de l’ozone en traitement
des eaux, strat'egie de quali=cation de cuves industrielles; emploi
des m'elangeurs statiques. Ph.D. thesis, UniversitSe de Rennes France,
ENSCR, No. 929.
Martin, N., & Galey, C. (1994). Use of static mixer for oxidation and
disinfection by ozone. Ozone Science Engineering, 16, 455–473.
Middleton, J. C. (1978). Motionless mixers as gas–liquid contacting
devices. AIChE 71st annual meeting, Miami Beach, USA.
Mustakis, M., Strei5, F. A., & Schneider, G. (1986). Advances in static
mixing technology. Chemical Engineering Progress, 82(7), 42–48.
Myers, K. J., Baker, A., & Ryan, D. (1997). Avoid agitation by selecting
static mixers. Chemical engineering progress, June, 28–38.
Fig. 16. Comparison of interfacial area and mass transfer coe6cient
Rader, G., Mustakis, M., Grosz-Roll, F., & Maugweiler, W., (1989).
between static mixers and others classic gas–liquid reactors.
Better absorption? Try a static mixer. Chemical Engineering, 137–142.
A. Heyouni et al. / Chemical Engineering Science 57 (2002) 3325–3333 3333

Roes, A. W. M., Zeerman, A. T., & Bukkems, F. H. J. (1984). Turunen, I., & Haario, H. (1994). Mass transfer in tubular reactors
High-intensity gas–liquid mass transfer in the bubbly #ow region equipped with static mixers. Chemical Engineering Science, 49(24B),
during co-current up #ow through static mixer. International chemical 5257–5269.
engineering, symposium series, Vol. 38 (pp. 231–238). Wallis, G. B. (1969). One dimensional two-phase Dow. New York:
Strei5, F. A., Mathys, P., & Fischer, T. U. (1997). New fundamentals McGraw-Hill.
for liquid–liquid dispersion using static mixers. R'ecents progr'es en Zhu, Q., Liu, C., & Xu, Z. (1989). A study of contacting systems in
G'enie des proc'ed'es, 11(51), 307–314. water and wastewater disinfection by ozone. 2-Mathematical models
Traversay, C., Bonnard, R., Adrien, C., & Luck, F. (2000). Static mixer: of the ozone disinfection process with static mixer. Ozone Science &
a reactor for the ozone process. Proceedings of the international Engineering, 11, 189–207.
specialised symposium IOA—fundamental and engineering concepts
for ozone design—Toulouse (pp. 155 –161).

Potrebbero piacerti anche