[go: up one dir, main page]

0% found this document useful (0 votes)
33 views15 pages

Hassan 2012

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
33 views15 pages

Hassan 2012

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 15

Chemical Engineering Journal 184 (2012) 42–56

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

A consistent dimensional analysis of gas–liquid mass transfer in an aerated


stirred tank containing purely viscous fluids with shear-thinning properties
Raouf Hassan a,b , Karine Loubiere c,d,∗ , Jack Legrand a,b , Guillaume Delaplace e
a
Université de Nantes, Laboratoire GEPEA, CRTT, 37 boulevard de l’université, BP 406, F-44602 Saint-Nazaire, France
b
CNRS, Laboratoire GEPEA, F-44602 Saint-Nazaire, France
c
Université de Toulouse, INPT, UPS, Laboratoire de Génie Chimique (LGC), 4 allée Emile Monso, BP 84234, F-31432 Toulouse, France
d
CNRS, Laboratoire de Génie Chimique (LGC), F-31432 Toulouse, France
e
INRA, Laboratoire PIHM UR 638, 369 Rue Jules Guesde, BP 39, F-59651 Villeneuve d’Ascq Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: This paper deals with gas–liquid mass transfer in an aerated stirred tank containing Newtonian or shear-
Received 17 June 2011 thinning fluids. The aim is to demonstrate that, for a given mixing system, an unique dimensionless
Received in revised form correlation gathering all the mass transfer rates (150 kl a measurements) can be obtained if and only if the
24 November 2011
variability of the rheological material parameters is correctly considered when implementing the theory
Accepted 14 December 2011
of similarity. More particularly, it is clearly illustrated that a too gross simplification in the relevant list
of the parameters characterizing the dependence of apparent viscosity with shear rates leads to pitfalls
Keywords:
when building the -space set. This is then a striking example showing that a robust predictive correlation
Dimensional analysis
Variable material property
can be established when the non-constancy of fluid physical properties ceases to be neglected.
Gas–liquid mass transfer © 2012 Elsevier B.V. All rights reserved.
Shear-thinning fluids
Aerated stirred tank
Model of Williamson–Cross

1. Introduction including tank geometry and dimensions, impeller type, dimen-


sions and rotational impeller speed, aeration system and gas flow
The dispersion of a gaseous phase in a liquid phase for mass rate, physical and rheological properties of gas and liquid phases,
transfer purposes is involved in many processes, in the field of temperature, and pressure. In an attempt to elucidate the effects
chemical reaction engineering (e.g. for chlorinations, hydrogena- of the latter parameters on the absorption rate coefficient, a great
tions, oxidations, alkylations, ammonolysis and so forth) but also amount of published investigations are encountered in the litera-
in biochemical engineering (including fermentation, waste water ture; some interesting overviews are given in [1–5]. They provide
treatment). The use of agitated tanks is a widespread practice for improved knowledge on gas–liquid mass transfer through experi-
operating such absorption processes as offering the advantages to mental data, empirical correlations, mechanistic analysis or more
generate high interfacial areas and intense mixing of liquid phase. recently numerical simulations.
Understanding and modelling mass transfer between phases is of Most of them deal with the cases when the liquid phase is a
importance, because it may often become the critical step deter- Newtonian fluid with low viscosities. From various sets of experi-
mining the achievement of the application, and thus may give the ments carried out at lab-scale and/or at larger scale, some empirical
main guidelines on which the design and the scale-up of the pro- correlations for kl a are proposed, involving either dimensional or
cess will be based. The transferred mass quantity depends on the dimensionless parameters. At present, the most frequently used
solute solubility in the liquid phase, but above all on the interfacial dimensional correlation remains the one of Van’t Riet [6] or some
area, a, and on the overall liquid-phase mass transfer coefficient, Kl . variants in which the constant and exponents have been modified.
The product of these latter parameters is commonly called absorp- They are expressed such as:
tion rate coefficient or overall volumetric gas–liquid mass transfer  P C2
coefficient (Kl a or kl a for low soluble gases). The factors influ- kl a = C · (Ug )C1 · · ()C3 (1)
Vl
encing of such unit operation (in particular Kl a) are very large,
where the constant C is strongly affected by the geometrical param-
eters of the agitation system, and  is the Newtonian viscosity.
∗ Corresponding author. Tel.: +33 0 5 34 32 36 19; fax: +33 0 5 34 32 36 97. Garcia-Ochoa and Gomez [3] recently summarized the different
E-mail address: Karine.Loubiere@ensiacet.fr (K. Loubiere). exponents associated with Eq. (1) that are available in the literature.

1385-8947/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.12.066
R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56 43

Nomenclature
Sc Schmidt number defined according to gas phase
a specific interfacial area (m−1 ) properties (Eq. (31))
CO2 concentration in dissolved oxygen (kg/m3 ) ∗
tw dimensionless time number issued from the
CO∗ concentration in dissolved oxygen at saturation Williamson–Cross’s model (Eq. (45))
2
(kg/m3 ) Ug∗ dimensionless superficial gas velocity (Eq. (26))
d air sparger diameter (m)
D impeller diameter (m) Subscripts
D oxygen diffusion coefficient in the liquid phase g gas phase
(m2 /s) l liquid phase
G/Vl mass throughout per unit of liquid volume (kg/s/m3 )
g gravity acceleration (m/s2 )
H Henry’s constant (Pa) Another approach is to use correlations with dimensionless
Ht tank height (m) groups; contrary to dimensional correlations, they guarantee firm
kl liquid-side mass transfer coefficient (m/s) basis for process scale-up, provided that they must be established
K consistency index from the model of Ostwald–de- with respect to the theoretical context of the theory of similarity.
Waele (Eq. (14)) (Pa snost ) The pioneer work’s of Zlokarnik [7] has established the relevant
Kl overall mass transfer coefficient (m/s) list of influencing intensive parameters and proposed the following
kl a overall volumetric gas–liquid mass transfer coeffi- dependence between dimensionless numbers:
cient (s−1 )  ∗  ∗ 
∗ P Qg
KMO Metzner–Otto constant (Eq. (5)) (kl a) = f1 , ,  ∗ , Sc, Si∗
Qg Vl
m Henry’s constant ⎧  1/3  ∗  
nost flow index from the model of Ostwald–de-Waele ⎪
⎨ (kl a)∗ = kl a · g2l ,
P
=
P
· [l · (l · g)
2/3 −1
] , (2)
Qg Qg
(Eq. (14)) where  ∗    1/3
nw flow index from the model of Williamson–Cross (Eq. ⎪
⎩  ∗ =  · [l · (4 · g)1/3 ]−1 , Qg = Qg · l
l Vl Vl g2
(15))
N rotational impeller speed (s−1 ) Note that, in Eq. (2), Sc is the Schmidt number and Si* is a material
P/Vl power dissipated per unit of volume (W/m3 ) dimensionless parameter which describes coalescence behaviour
Ps pressure in the system (Pa) of solutions (i.e. ionic strength, electrical charge of ions, . . .). Thanks
Qg gas flow rate (m3 /s) to this approach, Zlokarnik [7] could rigorously distinguish dif-
t time (s) ferent process relationships depending whether the system is
tw time parameter from the model of coalescent or non-coalescent.
Williamson–Cross (Eq. (15)) (s) Few years later, Judat [8] has critically examined the existing
Tl temperature of the liquid phase (K) publications on gas–liquid mass transfer (coalescing systems) in
Tt vessel diameter (m) stirred vessels. Description of experimental data with the aid of
Ug superficial gas velocity (m/s) intensive parameters has leaded this author to (±30% deviation):
V tank volume (m3 )
∗0.40
Vl liquid tank volume (m3 ) ∗ (P/Vl )
(kl a) = 9.8 × 10−5 · (3)
B−0.6 + 0.81 × 10−0.65/B
Greek letters
ε mean standard deviation (Eq. (50)) where B = (Qg /Tt2 ) · (l · g)−1/3 , the others numbers being defined
˙ shear rate (s−1 ) as in [7]. Judat [8] has then shown that a monoparametric represen-
˙ o reference shear rate (s−1 ) tation of (kl a)* versus (P/Vl )* is inadequate, and that only a -space
˙ av average shear rate defined from the Metzner–Otto representation containing both dimensionless power per unit vol-
concept (Eq. (4)) (s−1 ) ume and superficial gas velocity can satisfactorily correlate (kl a)*.
 dynamic viscosity (Pa s) This author has also put forward that another -space, contain-
a apparent viscosity (Pa s) ing non intensive parameters (rotational impeller speed instead
o reference apparent viscosity parameter defined at of power per unit liquid volume) could be used to describe the
reference shear rate (Pa s) measures of (kl a)*, but this i -space is larger than the previous one.
w viscosity parameter from the model of Few authors (for example [9,10]) have conserved the dimen-
∗ 1/3
Williamson–Cross (Eq. (15)) (Pa s) sionless group (kl a) = kl a · (l /g 2 ) defined by these two pioneer
 cinematic viscosity (Pa s) works for modelling absorption processes. Most of them have
 density (kg/m3 ) adopted, with or without theoretical backgrounds, others defini-
 surface tension (N/m) tions for making dimensionless kl a [11–17]. They include notably
a modified Sherwood number (kl a · Tt2 /D) or Stanton number
Dimensionless numbers (kl a · Vl /Qg ) [3].
Fr Froude number (Eq. (27)) When shear-thinning fluids are involved, two additional ques-
kl a* dimensionless volumetric mass transfer coefficient tions arise unfortunately in a point of view of the theory of
(Eq. (25)) similarity:
* dimensionless viscosity (Eq. (29))
i dimensionless number deduced from the theory of - How should we proceed to guarantee that the results obtained
similarity with Newtonian fluids can be extended to these non-Newtonian
* dimensionless density (Eq. (28)) liquids? The spatial distribution of liquid viscosity in the tank, due
* dimensionless surface tension (Eq. (30)) to its dependency with shear rates, constitutes a major difficulty
with regard to the choice of a representative viscosity.
44 R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56

Table 1 The sparger was located 30 mm from the bottom of the tank, in
Dimensions of the experimental setup.
the axis of the impeller. The gas flow rate (Qg ) was regulated by
Tank size Tt = 0.212 m using a manometer (Samson® 47 08-1155) and measured with
Ht = 0.316 m a volumetric flow meter (Brooks® R2-25-C) with an accuracy of
V = 10 L
0.05 L/min. Ranged from 0.33 to 3.33 L/min, these values remained
Hl = 0.212 m
Vl = 7.4 L
quite narrow when compared to the available literature, they were
Baffles wb = Tt /10 (width, in m) initially imposed by the application underlying this work (namely
bb = Tt /50 (distance from walls, in m) the Autothermal Thermophilic Aerobic Digestion of sludge, see
Impeller (six-concave-blade D = 0.4·Tt (impeller diameter, in m) [19]). In terms of gas–liquid regime, it can be noticed (visual obser-
turbine) Ds = 3·Tt /4 (disk diameter, in m)
vations) that the operating conditions under test (N, Qg , fluids)
C = Tt /4 (clearance from the bottom, in m)
b = D/5 (blade height, in m) leaded to a complete dispersion regime, meaning thus that bub-
w = 0.5 D/5 (blade width, in m) bles were almost uniformly distributed throughout the tank. One
l = D/4 (blade length, in m) exception was for N = 200 rpm where the loading regime took place
Sparger d = D (sparger diameter, in m)
(presence of bubbles only in the upper part of the tank).

- How should we proceed to guarantee that the results obtained at 2.2. Methods of overall volumetric gas–liquid mass transfer
lab-scale will be also scalable at industrial scale? coefficient measurement

Until now, most of works encountered in the literature did not 2.2.1. Standard dynamic method
take care about these questions: they simply consist in replacing For implementing the standard dynamic method, the liquid
the Newtonian viscosity by an apparent viscosity defined from the phase was deoxygenated by flushing with nitrogen. Then, after
Ostwald–de-Waele’s model in which an average shear rate is con- replacing nitrogen by air, the variation in dissolved oxygen con-
sidered according to the well-known concept of Metzner–Otto: centrations with time was measured until reaching the saturation.
For that, two probes (InPro-6050, Mettler-Toledo® ) and an acqui-
˙ av = KMO · N (4)
sition card were implemented, as well as the LabView® software
where the constant KMO depends on the agitation system. Such for data acquisition. The positions of the probes are represented in
choice of apparent viscosity is in many cases questionable, in par- Fig. 1: they are located vertically at 4 cm and 18 cm above the bot-
ticular when (i) the flow regime is not laminar (the use of Eq. (4) tom of the vessel, and horizontally at 1.5 cm from walls. Assuming
becomes then quite haphazard) and (ii) the rheological behaviour a well mixed liquid phase, the mass balance in dissolved oxygen
of fluids cannot be described in the whole range of shear rates by concentration is given by
the Ostwald–de-Waele’s model.
The present paper aims at rigorously answering these two lat- dCO2
= Kl · a · (CO∗ − CO2 ) (5)
ter questions, starting from the theoretical background underlying dt 2

the dimensional analysis and extending it to the cases of variable


material properties. More accurately, the objective is to show how where Kl is the overall mass transfer coefficient in the liquid side
to proceed: (i) to construct a complete list of relevant parameters and a is the interfacial area between gas and liquid phases. The two-
able to consider variable rheological parameters, and consequently film theory of Lewis and Whitman [20] assumes that Kl is the result
(ii) to elaborate, without pitfalls, a set of dimensionless numbers of two local mass transfer coefficients (kl and kg ):
characterizing all the factors governing absorption rate coefficients
(kl a) in an aerated stirred tank where purely viscous fluids with 1 1 1
= + (6)
or without shear-thinning properties are involved. To support this Kl kl m · kg
theoretical approach, a set of experiments was carried out to mea-
sure kl a in a stirred tank aerated in volume. Different operating where m is the Henry’s constant (m = H/Ps ). The solubility of
conditions (rotational impeller speed, gas flow rate) were covered oxygen is low: H is equal to 4.05 × 109 Pa (corresponding to
as well as various fluids (seven purely viscous fluids of which four C* = 9.09 mg L−1 ) in deionised water at 293 K in equilibrium with
have shear-thinning properties). air under atmospheric pressure. So, all the resistance to oxygen
mass transfer is located in the liquid film, leading to Kl ≈ kl .
As a consequence, the volumetric gas–liquid mass transfer coef-
2. Materials and methods
ficient, kl a, can be directly deduced from the slope of the curve
relating ln(CO∗ − CO2 ) to time, obtained when integrating Eq. (5).
2.1. Experimental set-up 2
The dynamics of the oxygen probe can be described using a first-
As shown in Fig. 1, the experimental set-up consisted of a cylin- order differential equation [21] as:
drical PMMA vessel of 10 L with a curved bottom. It was equipped
dCp 1
with a square double jacket (27.2 cm × 27.2 cm) and four baffles in = (CO2 − Cp ) (7)
stainless steel mounted perpendicular to the vessel wall. Table 1 dt tp
presents the geometrical details of the tank. The agitation system
was composed of a home-made six-concave-blades disk turbine where Cp is the dissolved oxygen concentration inside the probe.
which dimensions were respectful for the ones implemented in The time constant of the oxygen probe, tp , was measured using a
commercial CD6 Chemineer® impellers. The rotational impeller method based on probe response to negative oxygen steps [22]
speed (N) was regulated by using an electrical motor (Ikavisc MR- and found equal to 16 s. This latter value remained small when
D1 Messrührer, Janke & Kunkel, Ika® ), and varied from 200 to compared to mass transfer characteristic times, 1/kl a. As the tem-
1000 rpm. perature Tl slightly varied (20 ± 3 ◦ C) with power dissipation, the
Gas (air or nitrogen) was fed into the tank using a ring sparger usual temperature correction was applied [23]:
with a diameter equal to the impeller diameter, as recommended by
[18]. The latter was composed by 20 holes of 0.5 mm in diameter. kl a20 = kl aTl · 1.024(20−Tl ) (8)
R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56 45

Fig. 1. Experimental set-up.

For a given operating condition (N and Qg ), the mean kl a in the tank during an aeration time taeration , with the small quantity of Na2 SO3
was calculated by averaging the values measured by the two probes introduced:
and for the three runs (N = 3), such as:
1
Na2 SO3 + O2 → Na2 SO4 (11)
1 |kl atop + kl abottom | 2
kl a =< kl a >= · (9)
N 2 The mass of Na2 SO3 to initially introduce (mt ) must be chosen
N
carefully, as it should enable to keep a zero oxygen concentration
The axial homogeneity (h) of volumetric gas–liquid mass transfer during the aeration time (taeration ) while avoiding an excessive use
coefficients was also evaluated by using the criterion h defined as of Na2 SO3 . Indeed, it is important to guaranty that the coalescing
follows [24]: properties of the liquid phase were not affected by the presence
of Na2 SO3 . A good compromise was to maintain an initial concen-
1 |kl atop − kl abottom |
h=  · (10) tration below 0.5 g/L (i.e. mt < 3.5 g with Vl = 7.4 L) [25]. Note that
N < kl a > to minimize mt , it was also possible to play on the aeration time
N 
(taeration ), which was here typically ranged from 1.5 min to 9 min.
where N was the number of experiments (N = 60 for each liquid The optimization of both mt and taeration was made easier by the
phase). fact that the orders of magnitude of kL a were known thanks to the
measurements issued from the dynamic method.
2.2.2. Chemical method When such conditions are respected, Painmanakul et al. [27]
When applying the latter dynamic method in viscous fluids, the have shown that the overall volumetric gas–liquid mass transfer
impact of the probe dynamics and of the liquid film in front of the coefficient can be deduced from:
membrane on the probe can no more be ignored, as possibly bias- (1/2)(MO2 /MNa2 SO3 ) · (mt − mr )
ing the measurements of kl a [25]. In addition, in such fluids, the kl a = (12)
taeration · Vl · CO∗
gas hold-up structure is known to be very different from the one 2

observed in water and other low-viscosity liquids: many tiny bub- where mt is the total mass of Na2 SO3 initially introduced, mr is
bles appear to accumulate during aeration and circulate with the the mass of Na2 SO3 remaining in the tank after an aeration period
liquid while large bubbles are also observed (bimodal bubble popu- taeration , and CO∗ is the concentration in dissolved oxygen at satura-
lation). These tiny bubbles can actively contribute to mass transfer, 2
tion. For glycerine solutions, CO∗ was by default considered equal to
depending whether they are in equilibrium with the level of dis- 2

solved solute in the liquid phase (high residence times) or not [26]. 8.8 mg L−1 as in deionised water at 20 ◦ C. An identical assumption
For these reasons, it has been chosen to implement a second was made for CMC and xanthan gum solutions in agreement with
method for kl a measurement. It will then enable to test the valid- the data reported by [28] who showed that, in the range of con-
ity and accuracy of the dynamic method in the viscous fluids centrations here involved, no major variation of CO∗ occurs when
2
involved. This alternative method was the chemical method devel- compared to water.
oped by [27], based on a mass balance on sodium sulphite (Na2 SO3 ) At last, for each condition, three samples (10 mL) were taken in
concentrations during a given aeration time. Nitrogen was firstly the tank, immediately mixed with 10 mL of standard iodine reagent
injected into the liquid phase in order to eliminate the dissolved at 0.12 equiv./L. The titration of these samples with a sodium thio-
oxygen present in the tank. When the concentration of dissolved sulphate solution (0.05 equiv./L) and a starch indicator (iodometry
oxygen reached nearly zero, an adequate amount of Na2 SO3 was titration) gave access to the concentration of Na2 SO3 remaining in
introduced; air was then introduced in the tank and will react, the tank after an aeration period taeration , and thus to mr .
46 R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56

2.2.3. Surface- and volume-aerations (a) 100


The overall volumetric gas–liquid mass transfer coefficient mea-
Exp.
sured by the latter methods is in reality the global result of both 10 Ostwald
contributions: Williamson

viscosity (Pa.s)
1
- the surface-aeration: it corresponds to the mass transfer occurring
at the free surface which importance depends strongly on the
0.1
surface motion. It also includes the aeration associated with the
bubbles entrained from the surface into the liquid bulk;
0.01
- the volume-aeration: it is induced by the bubbles generated at the
sparger directly inside the liquid bulk.
0.001
This can be expressed such as: 0.1 1 10 100 1000 10000
shear rate (s-1)
kl a|t = kl a|surf + kl a|vol (13)
(b) 100
Some measurements are made with mechanical agitation and with-
Exp.
out bubbling at the sparger. They enable to get an idea of the relative 10
importance of each contribution. Ostwald
Williamson

viscosity (Pa.s)
1
2.3. Fluids
0.1
The application underlying this study dealt with investigations
on aeration performances in an Autothermal Thermophilic Aero-
0.01
bic Digestion (ATAD) process for treating sludge issued from waste
water treatment plant [19]. As sludge was a very complex mate-
rial, it was decided in a first step to work with model fluids instead 0.001
of sludge. Their formulation was chosen so as to obtain rheologi- 0.1 1 10 100 1000 10000
cal behaviours as close to sludge as possible, in particular in terms shear rate (s-1)
of shear-thinning properties. For these reasons, and with regard 100
to literature, aqueous solutions of carboxymethylcellulose (CMC)
(c)
Exp.
and xanthan gum were selected. The use of two types of fluids
10 Ostwald
offered the advantage to cover a wider range of combinations of
viscosity (Pa.s)

flow and consistency indexes. To ensure their stability in time (dur- Williamson
1
ing several days), NaCl was added at 0.1% (w/v) to the solutions
of xanthan gum solutions [29] and NaHCO3 (0.1 mol/L) + Na2 CO3 ,
10H2 O (0.1 mol/L) to the solutions of CMC [30]. Various concentra-
tions of CMC and xanthan gum were tested, and finally converged 0.1
towards the following ones: 4 and 6 g/L for CMC, and 1 and 2 g/L for 0.01
xanthan gum. In addition to these non-Newtonian fluids, deionised
water and two aqueous solutions of glycerine (50% and 70%, v/v) 0.001
were also chosen as Newtonian fluids. 0.1 1 10 100 1000 10000
The rheology of these fluids was measured, at 20 ◦ C, by a shear rate (s-1)
rotational stress-controlled rheometrer (MCR500, PAAR Physica® )
(d) 100
equipped a cone-plate device (50 mm in diameter, 3 degree in cone
angle). Their density and surface tension were determined using Exp.
a densimeter ERTCO® and a tensiometer involving the Wilhelmy 10 Ostwald
plate method (3S GBX® ). The physical and rheological proper- Williamson
viscosity (Pa.s)

ties of both Newtonian and non-Newtonian fluids are collected in 1


Table 2.
The rheological behaviours of the non-Newtonian fluids were 0.1
firstly characterised by measuring the variation of shear stress ( )
or apparent viscosity (a ) as a function of shear rates ()˙ which 0.01
range varied from 0.1 to 3000 s−1 . When comparing the curves
obtained for increasing and decreasing shear rates, no difference 0.001
was observed whatever the non-Newtonian fluids: no hysteresis 0.1 1 10 100 1000 10000
phenomenon then existed. Shown in Fig. 2, the rheograms obtained
shear rate (s-1)
(issued from several trials) clearly illustrate the shear-thinning
properties of these fluids. For each fluid, the yield stress was also Fig. 2. Rheograms: (a) xanthan gum at 1 g/L, (b) xanthan gum at 2 g/L, (c) CMC at
determined, by applying the method proposed by [31] which con- 4 g/L, and (d) CMC at 6 g/L.
sisted in oscillating stress sweep tests at a constant frequency
(1 Hz); the dynamic yield stress was then defined at the end of the
linear viscoelastic region, namely from the abscissa correspond- was found to vary between 0.1 and 1 Pa, and remained thus neg-
ing to the intersection point between the tangent to the plateau ligible. To complete the rheological characterisation of the fluids,
and the tangent to the inflexion point of the curve linking complex the viscoelastic properties were investigated, by means of creeping
modulus and shear stress. Depending on the fluid, the yield stress tests, relaxation tests and/or dynamic oscillating measurements. In
R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56 47

Table 2
Physical and rheological properties of the fluids.

l (kg/m3 ) l (Pa s)  (N/m) Ostwald–de-Waele’s model: Williamson–Cross’s model:

K (Pa sn ) nost w (Pa s) tw (s) nw

Air 1.18 1.85 × 10−5 – – – – – –


Newtonian fluids
Deionised water 998 0.001 0.0728 – – – – –
Glycerine 50% [Gly50] 1145 0.0109 0.0456 – – – – –
Glycerine 70% [Gly70] 1195 0.0349 0.0503 – – – – –
Non-Newtonian fluids
CMC 4 g/L [CMC4] 997 – 0.0717 0.1914 0.642 0.091 0.029 0.546
CMC 6 g/L [CMC6] 1006 – 0.0771 0.9470 0.527 0.948 0.844 0.514
Xanthan gum 1 g/L [XG1] 1013 – 0.0753 0.0890 0.543 1.885 57.85 0.411
Xanthan gum 2 g/L [XG2] 1032 – 0.0767 0.5084 0.373 29.505 150.36 0.281

the range of shear rate investigated, no major elastic property was describing the physical absorption process according to the two-
highlighted. film theory:
Based on these findings, some rheological models were chosen
G G
to mathematically describe the variation of apparent viscosity (a ) = kl · a · C or kl · a = (16)
Vl Vl · C
with shear rates ranging from 0.1 to 3000 s−1 . Among the large
variability available in the literature, two models were selected: where G/Vl is the mass throughput per unit volume of liquid and
C is a characteristic concentration difference. Eq. (16) implicitly
- the Ostwald–de-Waele’s model assumes that (i) the intensity of gas–liquid contacting is so high
that a quasi-uniform system is produced, (ii) the gas-phase mass
a = K · ˙ nost −1 (14) coefficient kg is negligible when compared to kl (low soluble gases),
(iii) the absorption rate at the interface is extremely fast, resulting in
where K and nost are respectively the consistency and flow indexes an equilibrium concentration of the dissolved gas at the interface C*
respectively. (e.g. C = C* − C). Hence, the establishment of the list of parameters
- the Williamson–Cross’s model influencing the main parameter kl a should respect the following
w rules [7]: (i) kl a must be independent of all geometrical parameters
a = (15)
˙ 1−nw
1 + (tw · ) (i.e. diameters of stirrer and tank, etc.), (ii) kl a must be independent
of the material parameters of gas phase, and (iii) kl a is an intensive
where w is a parameter describing a pseudo-Newtonian quantity because of its volume-related formulation.
behaviour for the smallest shear rates, tw is a time parameter As previously mentioned the number of the parameters influ-
characterizing the transition between the “pseudo-Newtonian” encing kl a, even performed in Newtonian liquids, is large and can
and purely shear-thinning behaviours, and nw is the consistency be decomposed according to:
index.
• The geometric parameters (see legend in Table 1), characterizing
The values of K, nost , w , tw , nw are reported in Table 2 for - the tank: Tt , Hl , curvature radius and angle (for tank’s bottom),
each fluid; they have been obtained by multi-parameter optimiza- ...
tions using the software Auto2fit® . In terms of consistency index, - the impeller: D, Ds , C, w, b, l, . . .
the most shear-thinning fluid appeared to be the solution of xan- - the sparger: d, number, diameter and shape of holes, . . .
than gum at 2 g/L. Based on the latter parameters, the apparent • The material parameters:
viscosities predicted by Eqs. (14) and (15) were compared to the l , l , g , g , ␴, D, C*
experimental ones in Fig. 2. For all the fluids, a better agree- • The process parameters:
ment with experiments was obtained with the Williamson–Cross’s
model, insofar as it enables to describe the most faithfully possible Qg
g, N, Ug = , Tl , Ps , . . .
the shape of the rheograms over the whole range of shear rates. This  · Tt2 /4
demonstrates that the fluids under test were not shear-thinning on
the whole range of shear rates investigated, three parameters being
required to well describe their behaviour.
Note that, in the present study, all the experiments were con-
ducted at room temperature and atmospheric pressure, inducing
3. Dimensional analysis for Newtonian and non-Newtonian thus that temperature and absolute pressure will not be listed. The
Fluids geometry of the tank was also unchanged, as well as the type and
position of both sparger and impeller. As a consequence, the list
3.1. Generation of i -sets governing aeration process for of individual physical quantities could be reduced: Table 3 shows
Newtonian fluids the dimensional matrix obtained with the reduced list of relevant
parameters.
In the present stirred tank, bubbles were directly generating It is voluntarily chosen to list the non-intensive parameters N
inside the liquid bulk by means of a gas sparger. Surface-aeration and D instead of power per unit of liquid volume (P/Vl ). The main
was of course present, but its contribution remained minor when motivation is that P/Vl is an intermediary variable which is not
compared to volume-aeration (see Section 4). Consequently, the always available (in particular at industrial scale), and thus using
overall volumetric gas–liquid mass transfer coefficient, kl a, can be such intensive variable would restrict the field of applications of
considered as the tractable quantity which is significantly influ- the final dimensional correlation which will be established.
enced by aeration conditions: it will be thus taken as target variable. In Table 3, the columns are assigned to the individual phys-
Remind that such choice is based on the following relationship ical quantities and the rows to the exponents appearing when
48 R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56

Table 3
Dimensional matrix of the influencing parameters (Newtonian fluids).

Core matrix Remnant matrix

g g g kl a Ug N D l l  D

Mass, M (kg) 1 1 0 0 0 0 0 1 1 1 0
Length, L (m) −3 −1 1 0 1 0 1 −3 −1 0 2
Time, T (s) 0 −1 −2 −1 −1 −1 0 0 −1 −2 −1

each quantity is expressed as an appropriate power product of D


4 = (20)
the base dimensions (mass, length, time). This table is structured −2/3 2/3
g · g · g −1/3
in a core matrix and a residual matrix. The core matrix regroups
the individual physical quantities put forward by the user to form l
5 = (21)
the dimensionless ratios from other individual physical quantities g
(namely kl a, Ug , N, D, l , l , , D). Depending on the individ- l
6 = (22)
ual physical quantities assigned in the core matrix, several set of g
dimensionless numbers i can be obtained. It has been shown [32]

that all the i -sets obtained from a single and identical relevance 7 = (23)
−1/3 4/3
list are equivalent to each other from a point of view of dimen- g · g · g 1/3
sional analysis, and can be mutually transformed at leisure. The D
final form for the i -set should be laid down by the user so as to be 8 = (24)
g−1 · g
the “best” suitable for evaluating and presenting the experimen-
tal data. Contrary to what commonly found in the literature, the Note that 1 is nothing other that the dimensionless volumetric
gas properties g and g (and not the liquid properties) were here mass transfer coefficient defined by [7], the combination of 4 with
chosen as individual physical quantities; the associated motiva- (3 )2 leads to the Froude number, and 8 is the inverse of a Schmidt
tion was to generate dimensionless numbers dependent of a single number defined according to gas phase properties. By introducing
influencing parameter, the gas phase being kept unchanged in this the gas kinematic viscosity (g = g /g ) and giving explicit nota-
study (air). tions, the latter numbers become:
Generating the set of dimensionless numbers (and possibly their   1/3
g
future transformation) represents an extremely easy undertaking 1 = kl a∗ = kl a · (25)
when compared to the drawing up of a reliable and as accurate as g2
possible relevance list; this can be made by matrix transformation. Ug
The starting point consists in carrying out the so-called Gaussian 2 = Ug∗ = (26)
(g · g)1/3
algorithm in order to obtain a unit core matrix by linear transfor-
mations (zero-free main diagonal, beneath it zeros). Table 4 reports N2 · D
the unit core matrix associated with the dimensional matrix of 3 = Fr = (27)
g
Table 3. The analysis of the unit core matrix leads to the dimen-
l
sionless ratios. Indeed, the rows of residual matrix are assigned to 5 = ∗ = (28)
g
the exponents with whom the elements of the core matrix appear
when the individual physical quantities of the residual matrix are l
6 = ∗ = (29)
expressed as an appropriate power product of the physical quanti- g
ties of the core matrix. Literature [33] offers detailed examples of 
how to handle matrix transformation and recombination in order 7 =  ∗ = (30)
1/3
(g3 · g4 · g)
to quickly obtain the complete set of dimensionless numbers. This
aspect will be then only briefly described in this paper. For New- g
8 = Sc = (31)
tonian fluids, the matrix analysis leads to the eight dimensionless D
numbers, 1 to 8 : Thus, at a given temperature, under a given pressure, for the geom-
etry of the aerated stirred tank under test (in particular for Hl /Tt = 1,
kl a
1 = (17) D/Tt = 0.4, Ds /Tt = 0.75 and for the other geometrical ratios char-
1/3 −1/3
g · g · g 2/3 acteristics of the system), the dimensional analysis states that,
when Newtonian fluids are involved, dimensionless volumetric
Ug mass transfer coefficient (kl a*) is potentially affected by six dimen-
2 = (18)
−1/3 1/3 sionless numbers, respectively describing the effects of superficial
g · g · g 1/3
gas velocity, rotational impeller speed, liquid density, Newtonian
N viscosity, liquid surface tension and oxygen diffusivity:
3 = (19)
−1/3
1/3
g · g · g 2/3 kl a∗ = f Ug∗ , Fr, ∗ , ∗ ,  ∗ , Sc (32)

Table 4
Unit core matrix obtained by linear transformations of dimensional matrix (Newtonian fluids).

Core matrix Residual matrix

g g g kl a Ug N D l l  D
1
M + T + 2A 1 0 0 3
− 13 1
3
− 23 1 0 − 13 −1
3M + L + T + A 0 1 0 − 13 1
3
− 13 2
3
0 1 4
3
1
A = − 13 × (3M + L + 2T ) 0 0 1 2
3
1
3
2
3
− 13 0 0 1
3
0
R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56 49

Such formulation of dimensionless numbers offers the advantage to material function s(p) which is apparent viscosity a (),
˙ the
enable the impact of all influencing parameters to be studied sepa- Pawlowski’s work [34] can be summed up, as follows:
rately, or in others words each dimensionless number is defined for
a single influencing variable. This is not the case in the literature • Firstly, all the dimensional parameters defined in the relevant list
where, for example, most of the authors used the aeration num- with fluids having constant properties (Newtonian case) should
ber, Na = Flg = Qg /(N · D3 ), in which the effect of gas flow rate is not be conserved, except for the variable physical properties in ques-
decoupled from the one of rotational impeller speed. tion (here apparent viscosity a ()).˙
At this state, the dependence of Eq. (32) is all that can be con- • Secondly, the Newtonian viscosity should be replaced by a refer-
tributed by the theory of similarity. The mathematical expression ence apparent viscosity, o , calculated at a reference shear rate,
for the function f, namely for the process relationship, has to be ˙ o . It is important to point out that any value for the reference
determined experimentally. shear rate can be chosen.
• Thirdly, the reference shear rate should be added in the list of
3.2. Extension of the theory of similarity to the cases of variable relevant parameters. At this stage, we can point out that some
material properties exceptions to this rule exist. Indeed, it has been demonstrated
that adding the reference point is not necessary when the mate-
When using the dimensional analysis to model system answers, rial function s(p), here a (),
˙ can be described by the following
it is generally assumed that the material properties remain unal- family of curves:
tered in the course of the process. However, the invariability of
s(p) = (A + B · p)c or s(p) = exp(A + B · p) (33)
material properties cannot be assumed when non-Newtonian flu-
ids are involved. Indeed, at the least one of the material properties, where A, B and C are three independent constants.
the apparent viscosity, can no longer be considered as a constant • Finally, a set of additional dimensionless numbers should be
inside the whole volume of the aerated stirred tank, insofar as the added in the relevant (dimensional) list to take into account the
dependency of this latter with the shear rates () ˙ generates a spa- dependency of material function. These dimensionless parame-
tial distribution of the liquid viscosity. The underlying question ters, called rheol , correspond to all the dimensionless ratios i
addressed to the dimensional analysis is now: how must the space which appear in the expression of the function u, except for the
of dimensionless numbers, i , be built in presence of such variable ratio /
˙ ˙ 0 :
material property?
In the case of materials with constant properties, no special pre-
1
 da

caution should be made to guarantee that a process relationship {rheol } = {i } such as u = (˙ − ˙ 0 ) ·
correlating a set of dimensionless ratios is also applicable to another a (˙ 0 ) d˙ =
˙ ˙ o
material. This is not true for materials with variables properties, as  ˙ 
demonstrated by [34]. In this case, we should first ensure as prior- =g ; {i } (34)
ity that a certain similarity exists for materials in order to extend ˙ 0
the range of validity of the process relationship to other materi- When integrating the previous guidelines, the list of the influ-
als. Despite that, the theory of similarity has little changed since encing parameters established for Newtonian fluids becomes for
its beginning, and the dimensional modelling involving materials non-Newtonian fluids:
with variable physical properties remains treated, in most of the
papers, as the case with constant material properties. The authors {kl a, g , g , g, Ug , N, D, l , , D, a (˙ o ), ˙ o , rheol } (35)
ignore then the fact that the spatio-temporal variability of mate- The next step is to find the type of material function describing the
rial properties influences the course of the process! One exception rheological behaviour of viscous fluids having shear-thinning prop-
is the modelling of the transformation processes where a mate- erties. As presented in Section 2, two models are well adapted for
rial having a temperature-dependence in viscosity is submitted to the investigated fluids: the Ostwald–de-Waele’s model (Eq. (14))
heat transfer condition. Most of attempts made to take into account and the Williamson–Cross’s model (Eq. (15)). As these models are
the variability of product properties in the reactor have consisted able to describe the variation of viscosity with shear rate for all the
in adding a new ratio raised to a certain exponent to character- aqueous solution of CMC and xanthan gum, each of them constitute
ize the system response. This ratio is defined by the ratio between one possible material function.
the viscosities at bulk temperature and at wall temperature. How-
ever, this kind of enlargement of the set of dimensionless ratios 3.2.1. Case No. 1: Ostwald–de-Waele’s model.
is theoretically valid only for few limited cases, that is to say only When the material function corresponds to the Ostwald–de-
if the material function (here viscosity versus temperature) sat- Waele’s model, the function u can be expressed as:
isfies specific criteria [34]. In other words, such method leads to  ˙ 
biased predictions and thus, cannot be generalised when handling u= − 1 (nost − 1) (36)
other fluids. Despite this fact, this ratio remains systematically ˙ 0
used, whatever the products investigated, in most of the studies Consequently,
since [35].
{rheol } = {nost } (37)
The theoretically consistent way of proceeding has been
introduced by Pawlowski in 1971 [34] and remembered by [32,36]. The Ostwald–de-Waele’s model, defining as a = K (Eq. · ˙ nost −1
The method consists in introducing some additional parameters in (14)), verifies Eq. (33), implying thus that the reference shear rate,
the relevance list so as to take into account the variation in the ˙ o , can be removed from the relevance list. As a consequence, for
flow domain of the physical property, noted s (for example viscos- purely viscous fluids having shear-thinning properties, the addi-
ity), as a function of a parameter noted p (for example temperature tional parameters is restricted to nost and the Newtonian viscosity
or shear rate); s(p) is called the material function (for example (T) is replaced by a (˙ o ). Eq. (32) established for Newtonian fluids
or ()).
˙ This implies that the i -space governing the process will becomes then:
be extended in comparison to fluids having constant properties. 
∗ a (o )
The guidelines to introduce the right number of additional dimen- kl a = g Ug∗ , Fr, ∗ ,  ∗ , Sc, ∗ = , nost (38)
g
sional parameters are detailed in [34]. Hence, when dealing with a
50 R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56

Table 5
Dimensional results for kl a (expressed in s−1 ): Newtonian and non-Newtonian fluids.

N (rpm) Qg (L/min) Newtonian fluids Non-Newtonian fluids

Water Gly50 Gly70 CMC4 CMC6 XG1 XG2

200 0.33 9.05 × 10−4 – – – – – –


0.9 1.91 × 10−3 – –– – – – –
1.6 3.35 × 10−3 7.75 × 10−4 3.13 × 10−4 1.64 × 10−3 1.26 × 10−3 1.55 × 10−3 1.31 × 10−3
2.33 4.58 × 10−3 1.05 × 10−3 4.22 × 10−4 2.08 × 10−3 1.85 × 10−3 2.06 × 10−3 1.47 × 10−3
3 5.33 × 10−3 1.45 × 10−3 5.15 × 10−4 2.58 × 10−3 2.01 × 10−3 2.68 × 10−3 2.16 × 10−3
3.33 5.88 × 10−3 1.77 × 10−3 5.75 × 10−4 2.82 × 10−3 2.3 × 10−3 2.85 × 10−3 2.56 × 10−3
400 0.33 2.31 × 10−3 – – – – – –
0.9 4.91 × 10−3 – – – – – –
1.6 8.69 × 10−3 2.90 × 10−3 6.82 × 10−4 3.68 × 10−3 2.54 × 10−3 4.83 × 10−3 3.61 × 10−3
2.33 1.12 × 10−2 3.50 × 10−3 8.00 × 10−7 4.43 × 10−3 4.40 × 10−3 5.93 × 10−3 4.81 × 10−3
3 1.29 × 10−2 4.28 × 10−3 9.99 × 10−4 4.98 × 10−3 4.98 × 10−3 6.59 × 10−3 6.51 × 10−3
3.33 1.31 × 10−2 4.63 × 10−3 1.19 × 10−3 5.19 × 10−3 5.53 × 10−3 7 × 10 × 10−3 8.01 × 10−3
600 0.33 4.30 × 10−3 – – – – – –
0.9 9.19 × 10−3 – – – – – –
1.6 1.62 × 10−2 4.72 × 10−3 1.38 × 10−3 7.51 × 10−3 3.99 × 10−3 9.52 × 10−3 7.68 × 10−3
2.33 1.82 × 10−2 5.73 × 10−3 1.77 × 10−3 8.60 × 10−3 6.90 × 10−3 1.13 × 10−2 9.57 × 10−3
3 2.06 × 10−2 6.39 × 10−3 2.25 × 10−3 9.23 × 10−3 8.20 × 10−3 1.30 × 10−2 1.11 × 10−2
3.33 2.26 × 10−2 6.78 × 10−3 2.55 × 10−3 9.44 × 10−3 9.05 × 10−3 1.39 × 10−2 1.20 × 10−2
800 0.33 4.94 × 10−3 – – – – – –
0.9 1.39 × 10−2 – – – – – –
1.6 2.26 × 10−2 6.87 × 10−3 2.35 × 10−3 1.14 × 10−2 6.40 × 10−3 1.29 × 10−2 1.08 × 10−2
2.33 2.68 × 10−2 7.61 × 10−3 3.31 × 10−3 1.32 × 10−2 8.30 × 10−3 1.51 × 10−2 1.28 × 10−2
3 2.81 × 10−2 8.14 × 10−3 3.97 × 10−3 1.45 × 10−2 9.17 × 10−3 1.79 × 10−2 1.43 × 10−2
3.33 3.21 × 10−2 8.45 × 10−3 4.46 × 10−3 1.48 × 10−2 9.75 × 10−3 1.87 × 10−2 1.57 × 10−2
1000 0.33 1.05 × 10−2 – – – – – –
0.9 1.43 × 10−2 – – – – – –
1.6 2.25 × 10−2 8.36 × 10−3 4.41 × 10−3 1.54 × 10−2 8.90 × 10−3 1.70 × 10−2 1.38 × 10−2
2.33 2.70 × 10−2 8.87 × 10−3 5.21 × 10−3 1.75 × 10−2 1.05 × 10−2 2.06 × 10−2 1.68 × 10−2
3 2.84 × 10−2 9.18 × 10−3 5.74 × 10−3 1.88 × 10−2 1.17 × 10−2 2.36 × 1010−2 1.89 × 10−2
3.33 3.34 × 10−2 9.48 × 10−3 6.21 × 10−3 1.90 × 10−2 1.29 × 10−2 2.61 × 10−2 2.06 × 10−2

When degenerated to the Newtonian case, the material function dimensionless numbers defining aeration process is enlarged by
associated with the Ostwald–de-Waele’s model leads to nost = 1 and two dimensionless numbers:
* = a /g where a = l is the Newtonian viscosity.
1
  1/3
∗ g
9 = nw and 10 = tw = · (45)
3.2.2. Case No. 2: Williamson–Cross’s model tw g2
When the material function corresponds to the
Eq. (32) established for Newtonian fluids becomes then:
Williamson–Cross’s model, the function u can be expressed
as: 
a (1/tw )
 ˙ − ˙  (n − 1) · (t · ˙ )(−nw ) kl a∗ = h Ug∗ , Fr, ∗ ,  ∗ , Sc, ∗ = ∗
, nw , tw (46)
0 w w 0 g
= · (39)
˙ 0 1 + (tw · ˙ 0 )1−nw
Note that the Williamson–Cross’s model leads to a Newtonian
Consequently, behaviour for particular values of nw and tw : namely nw = 1 and
{rheol } = {nw , tw · ˙ 0 } (40) tw = 1. In this case, l = w /2.

The reference shear rate, ˙ o , should be listed as the


Williamson–Cross’s model does not verify Eq. (33). Finally, the new 4. Results
variables to add in the relevant dimensional list are:
4.1. Validation of kl a measurements
{˙ o , nw , tw · ˙ 0 } (41)
In Table 5 are collected the set of experiments which will serve
Nevertheless, as ˙ o can theoretically take any value, it is possible
as database for the dimensional analysis. It is constituted by 150
to choose:
measures of kl a (including 70 values for the Newtonian case), car-
1 ried out at five rotational impeller speeds (200 ≤ N ≤ 1000 rpm)
˙ o = (42)
tw and four flow rates (0.33 ≤ Qg ≤ 3.33 L/min), and for several flu-
By this way, the supplementary variables are restricted to: ids (three Newtonian fluids, four non-Newtonian fluids). Note that
these overall volumetric mass transfer coefficients correspond to
{tw , nw } (43) the values obtained with the dynamic method, and averaged from
As a consequence, for the fluids described by the the measures for both probes and three runs (Eq. (9)).
Williamson–Cross’s model, the additional parameters are tw Whatever the fluids, the criterion for axial homogeneity (defined
and nw , and the Newtonian viscosity should be replaced by in Eq. (10)) has been found varying from 6 to 8% for N = 200 rpm,
1 w
and from 1 to 3% for N = 1000 rpm [19,37], the smallest value being
a = 0 = (44) obtained for the probe located at the top of the tank. Then, no signif-
tw 2 icant spatial heterogeneity takes place in the tank, and these values
After introducing these variables in the core matrix (Table 3) of kl a can be considered representative of the aeration state in the
and applying the linear transformations from Table 4, the set of whole tank.
R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56 51

- from 27 to 32% in glycerine at 50%, and from 9.7 to 20% in glycerine


k l a (s-1)
3.10-2 at 70%,
with chemical method Water 400 (1.6)
- from 43 to 66% in CMC at 4 g/L, and from 38 to 43% in CMC at 6 g/L,
Water 400 (3.33)
- from 50 to 83% in xanthan gum at 1 g/L and from 40 to 66% in
Water 800 (1.6)
2.10
-2
Water 800 (3.33) xanthan gum at 2 g/L.
Gly50 400 (3.33)
Gly50 800 (3.33) The comparison of such values of kl a, or any other attempts
10-2 XG2 400 (3.33) for their modelling, is usually made (see Section 1) by calcu-
k l a (s-1) CMC4 400 (3.33) lating the apparent viscosity basing on the well-known concept
with physical method of Metzner–Otto and the Ostwald–de-Waele’s model (Eq. (14)).
0 When applying this method [19], it is possible neither to explain
0 10-2 2.10-2 3.10-2 nor to understand satisfactorily these results, confirming thus the
requirement to perform more consistent investigations on the
Fig. 3. Comparison between chemical and physical methods for measuring kl a (the
influencing parameters.
dotted lines correspond to a deviation of ±15%; in the legend, the first number is N
in rpm, the number into brackets Qg in L/min).
4.2. Dimensionless results for Newtonian fluids

The oxygen diffusivity, D, was unknown for the viscous


Fig. 3 presents, for some representative cases, a comparison
Newtonian and non-Newtonian fluids under test. Indeed, such
between the physical and chemical methods for measuring kl a. A
information remains unavailable in the literature, and the usual
good agreement between both methods is observed: the deviation
correlations (for example the Wilke–Chang one) cannot be applied
never exceeds 15% which corresponds to the order of magni-
by lack of some required data (for example the association fac-
tude associated with the experimental uncertainty in the chemical
tor of solvent). So, the contribution of Schmidt number in the
method [27]. As a consequence, the values of kl a reported in Table 5
process relationship (Eq. (32)) cannot be rigorously sought. By
can be assumed relevant as validated by two methods.
default, whatever the liquid phases, the oxygen diffusivity, D, will
The contribution of surface aeration to the overall volumetric
be assumed equal to the one in water at 20 ◦ C (i.e. to 2 × 10−9 m s−2 ).
gas–liquid mass transfer coefficient has been estimated by mea-
In this case, the Schmidt number Sc (defined with respect to gas cin-
suring kl a without bubbling at the ring sparger (see Section 2). For
ematic viscosity, Eq. (31)) is then equal to 7850. As a consequence,
200 ≤ N ≤ 1000 rpm and 1.6 ≤ Qg ≤ 3.33 L/min, the following trends
it seems reasonable to assert that the process relationship estab-
have been obtained [37]:
lished will be insured for values of Schmidt numbers close to this
  latter. In addition, the variation of density for the fluids investigated
- for water, kl a < 0.14 × kl a , is not important (Table 2), implying thus that the change in * is
surf  t 
- for glycerine at 50%, kl a < 0.15 × kl a , and for glycerine at weak (847 < * < 1015). As a consequence, for securing the process
 
surf t relationship, it is chosen to ignore the possible alterations of * and
70%, kl a < 0.23 × kl a , Sc in Eq. (32), leading to:
surf  t 
- for CMC at 4 g/L, kl a < 0.08 × kl a , and for CMC at 6 g/L,
  surf t kl a∗ = f  {Ug∗ , Fr, ∗ ,  ∗ } (48)
kl a  kl a ,
surf t   Having no mechanistic indication on the form of the f -relation, the
- for xanthan gum at 1 g/L, kl a < 0.17 kl a , and for xanthan simplest monomial form is looked for:
 surf  t
gum at 2 g/L kl a < 0.14 × kl a .
surf t kl a∗ = ˛ · (Fr)a · (Ug∗ )b · (∗ )c · ( ∗ )d (49)

where ˛, a, b, c and d are respectively the constant and the


This demonstrates that the surface aeration remains small when exponents to which the dimensionless Froude, gas velocity, vis-
compared to volume aeration, confirming thus that kl a|t is the ade- cosity and surface tension numbers are raised. The dimensionless
quate target parameter to tract in the dimensionless analysis for viscosity, *, is calculated using the Newtonian viscosity (l )
qualifying the aeration state in the tank. reported in Table 2. The software Auto2fit® is used to perform
When analysed in detail [19], Table 5 points that, for a given the multi-parameter optimization required to determine ˛, a, b, c
fluid, the overall volumetric mass transfer coefficients logically and d. Different mathematical algorithms are systematically tested
increase for increasing rotational impeller speeds and gas flow (global Levenberg–Marquardt, global Quasi-Newton, standard dif-
rates. The relative contributions of gas sparging (Qg ) and mechan- ferential evaluation, genetic algorithm) to verify the stability of the
ical agitation (N) on the variations of kl a are comparable, even if results and their independency with initial conditions. The mean
the rotational impeller speed plays a more pronounced role. The standard deviation is calculated from:
present investigations are then performed under the intermediary  
1  (kl a)∗exp,i − (kl a)∗mod,i 
condition defined by [13,14], namely the condition ranged between ε=   (50)
the bubbling-controlling condition (at relatively high gas flow N  ∗
(kl a)exp,i 
i=1,N
rates) and the agitation-controlling condition (at relatively high
rotational impeller speeds). More important is the major reduc- In a first time, it is interesting to visualize the effect of Newto-
tion in kl a observed in presence of viscous fluids (Newtonian and nian viscosity (*) on (kl a)* separately from the other variables.
non-Newtonian) when compared to water. To better appreciate The problem is that the graphic representation associated with Eq.
this phenomenon, the following ratio can be defined: (49) requires five dimensions as (kl a)* depends on Fr, Ug∗ , * and
*. A simple way to sidestep this problem is to come down to a 2D
kl a|viscous fluid representation, in which the impact of * on (kl a)* is neglected and
R= (47) an identical exponent is imposed for the Froude and dimensionless
kl a|water
superficial gas velocity numbers. This is illustrated in Fig. 4, where
For instance, at Qg = 3 L/min and for 200 ≤ N ≤ 1000 rpm, the latter (kl a)* is plotted as a function of (Ug∗ · Fr)2/3 for Newtonian fluids.
ratio R varies: The value of 2/3 is chosen as a first approximation for the exponent
52 R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56

10-6 10-5 10-4 10-3


10-3
(k l a )*exp

Water
[µ * = 54, σ * = 73385] 10-4

Gly50
[µ * = 589, σ * = 45066]
10-5
Gly70
[µ * = 1886, σ * = 50704]

(k l a )*mod
10-6

Fig. 4. Effect of dimensionless Newtonian viscosity, kl a∗ = kl a · (g /g 2 )


1/3
, versus Fig. 5. Experimental dimensionless overall gas–liquid mass transfer coefficient
2/3 versus kl a* predicted from Eq. (52) (Newtonian fluids). The dotted lines correspond
(Ug∗· Fr) (the dotted/continuous lines correspond to the values predicted by Eq.
to a deviation of ±20%.
(51)).

on Fr and Ug∗ , as already encountered in the literature (see review of encountered in the literature. Based on that, the effect of surface
1/3
[3]). Thus, Fig. 4 offers the advantage to easily appreciate the neg- tension,  ∗ = /(g3 · g4 · g) , has been added in Eq. (51) (Case
ative effect of * on aeration performances: whatever (Ug∗ · Fr)2/3 , No. 3): the constant is in return increased and a negative exponent
an increase of * from 54 (water) to 1886 (glycerine 70%) leads to appears for * (−0.245). This latter logically confirms the positive
a drastic reduction of (kl a)* (more than 85%). impact of decreasing surface tension on aeration: when surface ten-
When imposing the exponent 2/3 to (Ug∗ · Fr), the best fit- sion is reduced (for example when adding surfactant or alcohol in
ting between experimental data and Eq. (49) without taking into the liquid phase), smaller bubble sizes are generated, leading to a
account * leads to: rise in interfacial area and thus in overall volumetric mass trans-
fer coefficient. In the literature [7], this usually results in positive
kl a∗ = 0.1420 · (Fr · Ug∗ )2/3 · (∗ )−0.591 (51) exponents for the Weber number.
The associated mean standard deviation is equal to 17.1%. To test At last, at a given temperature, under a given pressure, for the
the robustness of such correlation (in particular of the exponents geometry of the aerated stirred tank under test (in particular for
found), several cases are now tested when implementing the multi- Hl /Tt = 1, D/Tt = 0.4, Ds /Tt = 0.75 and for the other geometrical ratios
parameter optimization: characteristics of the system), the dimensional analysis states that
the process relationship for Newtonian fluids is expressed by:
- Case No. 1: all the exponents in Eq. (49) are kept free and * is 2/3 −0.591 −0.245
kl a∗ = 0.2097 · (Fr · Ug∗ ) · (∗ ) · ( ∗ )
neglected; 0.096 < Fr < 2.4, 0.0029 < Ug∗ < 0.029, 54 < ∗ < 1886 (52)
- Case No. 2: the exponents of Fr and of Ug∗ are imposed identical valid for
847 < ∗ < 1015, 50704 <  ∗ < 73385, Sc = 7850
without any specified value, and * is neglected;
- Case No. 3: the exponents of Fr and Ug∗ are imposed equal to 2/3
Fig. 5 compares the experimental data with the dimensionless
(as in Eq. (51)), but now the effect of * is taken into account.
overall volumetric gas–liquid mass transfer coefficients predicted
by Eq. (52). A good agreement is observed, as the mean standard
In Table 6 are collected, for each case, the constant and expo-
deviation remains smaller than 17%. It can been observed that some
nents of Eq. (49) deduced from the multi-parameter optimization.
data corresponding to the smallest kl a* (obtained for low speeds in
In a general point of view, no major difference appears between the
glycerine) tends to move away from the straight lines representing
different cases: the exponent of * remains close to −0.59 (devia-
(kl a∗ )exp = ±20% · (kl a∗ )pred . This should be linked to the fact that,
tion < 1.3%), and the exponents of Fr and Ug∗ do not vary significantly
in such conditions, the gas–liquid regime corresponds to a loading
whether they are imposed identical or not (deviation < 6.5%). These
regime, and not to a complete dispersion regime characterizing all
findings confirm that the orders of magnitude of the exponents
the other operating conditions. The mechanical agitation is not then
found in Eq. (51) are relevant, and thus, that this correlation is
sufficient to disperse uniformly the bubbles in the whole volume of
mathematically robust. Note that, even if Fr and Ug∗ have the same
tank, in particular in the lower part of the tank (below the impeller).
exponent, the contribution of N is two times higher than the one
Hence, the relative contributions of the mechanisms controlling kl a
of superficial gas velocity, as the Froude number is expressed in
(mechanical agitation against sparger aeration) deviates from the
squared rotational impeller speed (Eq. (27)); this is coherent with
ones acting when a complete dispersion regime takes place. When
these observations made on the dimensional results (Section 4.1).
such changes in regime occurs, it becomes then difficult to define
The cases No. 1 and No. 2 leads to a slight improvement (smaller
an unique and accurate process relationship able to describe the
than 1%) of the mean standard deviation (ε); however, the expo-
entire range of operating conditions. Some deviations can be also
nents of Eq. (51) will be thereafter conserved, insofar as they lead to
observed for the highest vales of kl a; they can be here associated
the most simple formulation, and also as the value of 2/3 is already
with an increasing contribution of the surface aeration (mass trans-
fer occurring at the free liquid surface) when rotational impeller
Table 6 speed rises. Indeed, for the highest N, a significant vortex appears
Dimensionless modelling for Newtonian fluids using Eq. (49) (in bold the case in the centre of the tank and entrains many bubbles. In such condi-
retained).
tions, the mechanism (or regime) of aeration  is deviated to a pure
Case ˛ a b c d ε (%) volume-aeration, and thus, the choice of kl a (and not kl a ) for
t surf
No. 1 0.02125 0.747 0.66 −0.605 – 16.1 tracting volume-aeration state is less representative.
No. 2 0.01535 0.683 0.683 −0.592 – 16.8 To conclude, it should be kept in mind that the validity of Eq.
No. 3 0.2097 2/3 2/3 −0.591 −0.245 16.8
(52) can be at present guaranteed only in the range of dimensionless
R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56 53

Table 7
Dimensionless numbers characteristics for liquid properties.

* * nost ∗ost nw ∗
tw ∗w

Water 848 73385 1 54 1 1 108


Gly50 973 45966 1 589 1 1 1178
Gly70 1015 50704 1 1886 1 1 3773
CMC4 861 75905 0.642 1866 0.546 0.201 4944
CMC6 877 77316 0.527 5309 0.513 6.84 × 10−3 5.12 × 104
XG1 847 72276 0.543 539 0.411 9.98 × 10−5 1.02 × 105
XG2 855 77719 0.373 1366 0.281 3.84 × 10−5 1.59 × 106

numbers above mentioned, at given temperature and pressure, and In Fig. 6, (kl a)* is plotted as a function of (Ug∗ · Fr)2/3 for all the
in the geometry defined in Section 2. fluids. Such figure is particularly important as the effects of * and
nost can be visualized separately. Indeed, at a given *, an decrease
4.3. Dimensionless results for non-Newtonian fluids in nost clearly induces an increase in (kl a)*, or in other words,
the shear-thinning character of fluids favours the aeration perfor-
4.3.1. When considering the model of Ostwald–de-Waele mances, probably due to the spatial heterogeneity of viscosity. This
As any value can be considered (see Section 3.2), the refer- is illustrated when comparing either (i) the glycerine at 70% and
ence shear rate, ˙ o , has been arbitrary chosen equal to 120 s−1 . The the aqueous solutions of CMC at 4 g/L which have almost the same
dimensionless numbers describing the rheological properties asso- dimensionless apparent viscosity (1886 against 1866), but differ-
ciated with the Ostwald–de-Waele’s model (Table 2) and such ˙ o ent flow indexes (1 against 0.642), or (ii) the glycerine at 50% and
are collected in Table 7 (fourth and fifth columns). It can be logi- the aqueous solutions of xanthan gum at 1 g/L which have almost
cally observed that the flow index (nost ) decreases when increasing the same dimensionless apparent viscosity (589 against 538), but
the concentration, the smallest value being obtained for the most different flow indexes (1 against 0.543). The impact of nost seems
concentrated solution of xanthan gum. Dimensionless apparent however less pronounced than the one of *.
viscosity (∗ost ) are higher for aqueous solutions of CMC than the Fig. 7 compares the dimensionless overall volumetric mass
ones of xanthan gum. transfer coefficients measured in presence of non-Newtonian fluids
Basing on the reasons mentioned in the case of Newtonian fluids, with the ones predicted using the process relationship estab-
the possible changes in * and in Sc will be neglected in Eq. (38) lished for Newtonian fluids (Eq. (52)). The points related to each
previously established when the model of Ostwald–de-Waele is non-Newtonian fluid are regrouped along individual straight lines
considered. Thus, Eq. (38) becomes: which are parallel to each others and to the curves previously
 obtained for Newtonian fluids. This illustrates that the introduc-
a (o ) tion of nost into the dimensionless modelling (Eq. (54)) could be
kl a∗ = g  Ug∗ , Fr,  ∗ , ∗ = , nost (53)
g a mean for differentiating the non-Newtonian fluids from each
others and from the Newtonian ones. Nevertheless, this group of
The simplest monomial form is here also looked for the g -relation four straight lines (one for each fluid) is not classified according
(no mechanistic information available): to the values of flow index, in particular the associated nost are
    not decreasing when deviating from the Newtonian curve (nost = 1).
kl a∗ = ˛ · (Fr)a · (Ug∗ )b · (∗ )c · ( ∗ )d · (nost )e (54) Indeed, the straight line the closest from the Newtonian curve
corresponds to nost = 0.543 (XG 1 g/L), followed by nost = 0.373 (XG
where ˛ , a , b , c , d and e are respectively the constant and
2 g/L), then nost = 0.642 (CMC 4 g/L), until reaching the most distant
the exponents to which the dimensionless Froude, superficial gas
line, nost = 0.527 (CMC 6 g/L). This finding tends to show that gather-
velocity, viscosity, surface tension and flow index numbers are
ing all the data related to non-Newtonian fluids over the Newtonian
raised (Auto2fit® ). However, it is important to keep in mind that
ones will be difficult by adding only the flow index nost in the
the modelling for Newtonian fluids is a degraded case of the one
relevant list. In others words, this would mean that the material
for non-Newtonian fluids. As a consequence, the exponents of Fr,
function associated with the model of Ostwald–de-Waele could be
Ug∗ , * and * (namely a , b , c , d ) are already defined, and will be
insufficient and/or unsuitable for describing completely the effect
thus taken equal respectively to 2/3, 2/3, −0.591 and −0.245 as in
of the shear-thinning properties on (kl a)*.
the Newtonian case (Table 6, Case No. 3).

-5 -4
10-6 10 10 10-3
10-3
(k l a )*exp
Water
Gly50 n1.0E-04 -4
ost =0.543 10
Gly70
CMC4 n ost =0.373
CMC6
n ost =0.642
XG1 10-5
XG2 n ost =0.527

(k l a )* mod
10-6

Fig. 7. Modelling using the Ostwald–de-Waele’s model: comparison between the


Fig. 6. Effect of dimensionless apparent viscosity and flow index when the experimental (kl a)* and the values predicted by Eq. (52) for Newtonian fluids. The
2/3
Ostwald–de-Waele’s model is considered: (kl a)* versus (Ug∗ · Fr) . dotted lines correspond to a deviation of ±20%.
54 R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56

-6 -5 -4
10-6 10-5 10-4 10-3 10 10 10 10-
10-3 10-
(k l a )*exp (k l a *)exp
Water
Water
Gly50
Gly50
10-4 Gly70 10-4
Gly70
CMC4 CMC4
CMC6 CMC6
XG1 10-5 XG1 10
-5

XG2
XG2
(k l a )* mod
(k l a *)mod
10-6 -6
10
Fig. 8. Modelling using the Ostwald–de-Waele’s model: comparison between the
Fig. 9. Modelling using the Williamson’s model: comparison between the experi-
experimental (kl a)* and the values predicted by Eq. (55). The dotted lines correspond
mental (kl a)* and the values predicted by Eq. (57). The dotted lines correspond to a
to a deviation of ±20%.
deviation of ±20%.

The fitting between Eq. (54) and experimental data (determina-


necessary to visit again the way of characterizing the rheological
tion of the constant ˛ and the exponent of nost , the other exponents
behaviours, namely the choice of the material function.
being the ones of the Newtonian case) leads to:

∗ 2/3 −0.591 −0.245 4.3.2. When considering the model of Williamson–Cross


kl a∗ = 0.2284 ∗
 · (Fr · Ug ) · ( ) · ( ∗ ) · nost −1.341
As explained in Section 3.2.2, the reference shear rate, ˙ o , is
0.096 < Fr < 2.4, 0.0029 < Ug∗ < 0.029, 54 < ∗ < 5309
valid for here chosen equal to 1/tw where tw is a time parameter describing
847 < ∗ < 1015, 50704 <  ∗ < 77719, Sc = 7850, 0.37 < nost < 1
(55) the pseudo-Newtonian behaviour for the smallest shear rates (Eq.
(15)). Using such ˙ o and the parameters found when applying the
Williamson–Cross’s model to experimental rheograms (Fig. 2) leads
The associated mean standard deviation is equal to 25%. In Fig. 8, to the dimensionless numbers collected in Table 7 (sixth, seventh
the experimental dimensionless mass transfer coefficients (kl a)* and eighth columns). It can be observed that: (i) the flow index,
are reported as a function of the ones predicted by Eq. (55). The nw , have the same order of magnitude than the ones found for
exponent of nost is found negative, and is then coherent with Fig. 6 the model of Ostwald–de-Waele nost , (ii) the dimensionless time
where the decrease of (kl a)* with increasing nost has been clearly characteristic number, tw ∗ , is a differentiating parameter between
observed. More important, this figure confirms what already sug- the non-Newtonian fluids as being significantly smaller for aque-
gested in Fig. 7: adding nost as the single dimensionless ratio in ous solutions of xanthan gum than for the ones of CMC, and (iii)
the relevant list (when compared to the Newtonian case) does not the dimensionless viscosity, *, are now ranged between 108 and
enable to gather all the data over the Newtonian curve; in particu- 1.59 × 106 .
lar, the points related to aqueous solutions of CMC strongly deviate As for Newtonian fluids, the simplest monomial form is looked
from the others. Thus, the set of dimensionless ratios involved with for the h-function in Eq. (46), and the possible changes in * and in
the Ostwald–de-Waele’s model is not able to describe all the com- Sc are neglected. This leads to:
plexity of the rheological behaviours encountered in the present
 
aeration experiments (namely the dependency of viscosity with kl a∗ = ˛ · (Fr)a · (Ug∗ )b · (∗ )c · ( ∗ )d · (nw )e · (tw
∗ f
) (56)
shear rate for the range of fluids under test).
For securing the latter process relationship (Eq. (55)) in a math- where the exponents of Fr, Ug∗ , * and * are the ones determined for
ematical point of view, two additional cases are tested when the Newtonian case, namely a = b = 2/3, c = −0.591 and d = −0.245.
implementing the multi-parameter optimization: the first one The coefficient ˛ , the exponents of nw and of tw ∗ (e and f respec-

(Case No. 5) in which * is not taken into account, and the second tively) are determined by the multi-parameter optimization. The
one (Case No. 6) in which * is taken into account with a non- following process relationship is then obtained:
imposed exponent. The results are collected in Table 8 (Case No. 4 kl a∗ = 0.02109 · (Fr · Ug∗ )
2/3
· (∗ )
−0.591
· ( ∗ )
−0.245
· (nw )
−2.399 ∗
· (tw )
−0.168

corresponds to Eq. (55)). Whatever the cases, the exponent of nost , ⎧


e, remains almost constant, about −1.34 (deviation below 2.6%), ⎨ 0.096 < Fr < 2.4, 0.0029 < Ug∗ < 0.029, 108 < ∗ < 1.6.106
(57)

and validates thus the robustness of the modelling. The occurrence valid for 847 <  < 1015, 50704 <  ∗ < 77719,

of a exponent of * (d ) has no major effect on the mean standard Sc = 7850, 0.28 < nw < 1, 3.8.10−5 < tw

<1
deviations obtained (25.6% against 25.0%), showing thus that this
The associated mean standard deviation is equal to 17.5%, and
parameter is neither the one controlling the process relationship
is thus smaller than the one found when applying the model
nor the additional one enabling to gather all the data.
of Ostwald–de-Waele (Eq. (55)). Fig. 9 plots the experimental
To conclude, these findings have clearly demonstrated that the
dimensionless mass transfer coefficients (kl a)* as a function of the
use of the Ostwald–de-Waele model is not relevant for encom-
ones predicted by Eq. (57). All the points are remarkably grouped
passing all (kl a)* values by a unique process relationship. It is then
together around the data of Newtonian fluids, demonstrating thus
the use of the Williamson–Cross’s model is relevant for predicting
Table 8 the effect of the non-Newtonian fluids under test on the aeration
Dimensionless modelling for non-Newtonian fluids when the Ostwald–de-Waele’s
model is considered (Eq. (54)).
performances. Then, when the i -space is enlarged by two dimen-
sionless ratios to describe the material function (namely nw and tw ∗ ),
Case ˛ d e ε (%) all the aeration experiments can be gathered on a unique curve. This
No. 4 0.2284 −0.245 −1.341 25.2 clearly shows that that it is possible to obtain a suitable dimen-
No. 5 0.01505 0 −1.377 25.6 sionless correlation for describing the variations of kl a at various
No. 6 0.04964 −0.107 −1.306 25.0
operating conditions (N, Qg ) and types of fluids, if and only if all the
R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56 55

Table 9 (such as rotational impeller speed or gas flow rate) or material


Dimensionless modelling for non-Newtonian fluids when the Williamson–Cross’s
parameter (such as surface tension, apparent viscosity, flow index
model is considered (in bold the case retained).
or Williamson–Cross time parameter), acts individually on the
Case ˛ d e f ε (%) overall volumetric mass transfer coefficient. The correlation estab-
No. 7 0.02109 −0.245 −2.399 −0.1682 17.5 lished (Eq. (57)) is at present validated in the range of dimensionless
No. 8 0.02090 0 −2.358 −0.1792 17.5 numbers previously defined, for a given temperature, under a given
No. 9 0.3058 −0.2440 −2.470 −0.1732 17.4 pressure and for the agitation/aeration system used (in particular
for Hl /Tt = 1, D/Tt = 0.4, Ds /Tt = 0.75 and for the other geometrical
ratios characteristics of the present system). Further experiments
rheological properties are correctly taken into account (e.g. using are now required to:
the three parameters involved in the model of Williamson–Cross)
instead of neglecting some of them (namely, by simple fitting with
the model of Ostwald–de-Waele). - to evaluate the effect of the Schmidt number on kl a*; for that,
Here also, the mathematical robustness of Eq. (57) is tested by the diffusion coefficient of oxygen in the viscous fluids under test
considering several cases when implementing the multi-parameter have to be experimentally measured.
optimization: in the first case (Case No. 8), the effect of * is - to definitively appreciate the validity of such process relationship
neglected whereas, in the second case (Case No. 9), the effect of when extending out of the domain investigated in the present
* is taken into account with a non-imposed exponent. The results study (for example at larger tank, at higher gas hold-up and for
are collected in Table 9 (Case No. 7 corresponds to Eq. (57)). They different types of agitation systems).
reveal that the occurrence of an exponent of * (d) has a major effect
neither on the mean standard deviations nor on the exponents of
5. Conclusion
nw (e ) and of tw∗ (f ) which remain almost constant whatever the

cases (deviations below 2.4% and 3.2% respectively). Note also that,
The present paper dealt with a consistent dimensionless analy-
when the exponent of * is kept free in the optimization, the value
sis of gas–liquid mass transfer in an aerated stirred tank containing
found (d) is close to the one already obtained for the Newtonian
purely viscous fluids with shear-thinning fluids. More particularly,
case. These tests clearly validate Eq. (57) in a mathematical point
this work showed how to proceed:
of view.
In addition, it can be observed that the exponent of tw ∗ , and

thus the impact of tw on kl a*, is negative. This is illustrated - to construct a complete list of relevant parameters able to build
in Fig. 10 where kl a* is plotted (Ug∗ · Fr)2/3 · ∗ −0.591 · nw −2.399 · an unique -space which keeps unchanged for both Newtonian
 ∗ −0.245 (according to the exponents found in Eq. (57) or Case and non-Newtonian fluids.
No. 7 in Table 9). The points related to each non-Newtonian fluid - and consequently to elaborate, without pitfalls, a set of
are regrouped along individual straight lines which deviate more dimensionless numbers characterizing all the factors governing
or less from the curves previously obtained for Newtonian flu- absorption rate coefficients (kl a) in a stirred tank where shear-
ids. These straight lines are remarkably classified according to thinning fluids are involved.
decreasing tw ∗ as far as going far from the Newtonian points (t ∗ = 1).
w
Thus, at given dimensionless apparent viscosity and flow index, the This theoretical approach was supported by a set of kl a mea-
smallest is the time parameter of the Williamson–Cross’s model surements in a tank stirred by a six-concave-blade disk turbine
(tw ) the highest is the overall volumetric mass transfer coefficient and aerated by a ring sparger, under different operating condi-
(kl a). tions (rotational impeller speed, gas flow rate) and for various fluids
All the findings demonstrate that the rigorous extension of the (water, glycerine 50% and 70%, solutions of CMC at 4 and 6 g/L,
theory of similarity to the case of variable material properties (e.g., solutions of xanthan gum at 1 and 2 g/L). These measures were val-
dependence of viscosity with shear rate) and the choice of an appro- idated notably by means of two methods (physical and chemical).
priate material function (here issued from the Williamson–Cross’s At last, a suitable dimensionless correlation could be obtained for
model) have made possible an accurate dimensionless modelling describing all the variations of kl a if and only if all the rheological
of the impact of the shear-thinning character of fluids on the aer- properties were correctly taken into account (e.g. using the three
ation performances in a stirred tank. Such approach also enables parameters involved in the model of Williamson–Cross) instead of
to understand how each parameter, either operating parameter neglecting some of them (namely, by simple fitting with the model
of Ostwald–de-Waele. It was expressed by:

2/3 −0.591 −0.245 −2.399 −0.168


kl a∗ = 0.02109 · (Fr · Ug∗ ) · (∗ ) · ( ∗ ) · (nw ) ∗
· (tw )
0.096 < Fr < 2.4, 0.0029 < Ug∗ < 0.029, 108 < ∗ < 1.6.106
valid for 847 < ∗ < 1015, 50704 <  ∗ < 77719,
Sc = 7850, 0.28 < nw < 1, 3.8.10−5 < tw∗
<1

In the future, further experiments will be required to


definitively appreciate the effect of Schmidt number and the
validity of such process relationship when extending out of the
domain investigated in the present study (for example at larger
tank).
This paper constitutes then an eloquent example demonstrating
how the variability of physical material parameters in a process
equipment should be integrated. The advantage of such approach
is to be perfectly transposable to any other physical properties
Fig. 10. Effect of dimensionless time parameter of the Williamson’s model: (kl a)* and other unit operation involving material with non-constant
−0.591 −2.399 −0.245
versus (Ug∗ · Fr)
2/3
· (∗ ) · (nw ) · ( ∗ ) . properties.
56 R. Hassan et al. / Chemical Engineering Journal 184 (2012) 42–56

References [20] W.K. Lewis, W.G. Whitman, Principles of gas absorption, Ind. Eng. Chem. 16
(1974) 1215–1220.
[1] J.B. Joshi, A.B. Pandit, M.M. Sharma, Mechanically agitated gas–liquid reactors, [21] J. Ingham, I.J. Dunn, E. Heinzle, J.E. Prenosil, Chemical Engineering Dynamics.
Chem. Eng. Sci. 37 (6) (1982) 813–844. Modeling with PC Simulation, VCH Publishers Inc., New York, 1994, ISBN 3-
[2] E.L. Paul, V.A. Atiemo-Obeng, S.M. Kresta, Handbook of Industrial Mixing: Sci- 527-28577-6.
ence and Practise, J. Wiley & sons, 2004. [22] C.O. Vandu, K. Koop, R. Krishna, Volumetric mass transfer coefficient in a slurry
[3] F. Garcia-Ochoa, E. Gomez, Bioreactor scale-up and oxygen transfer rate in bubble column operating in the heterogeneous flow regime, Chem. Eng. Sci. 59
microbial processes: an overview, Biotechnol. Adv. 27 (2009) 153–176. (2004) 5417–5423.
[4] M. Martin, F.J. Montes, M.A. Galan, Bubbling process in stirred tank reac- [23] J.K. Bewtra, W.R. Nicholas, L.B. Polkowski, Effect of temperature on oxygen
tors. II. Agitator effect on the mass transfer rates, Chem. Eng. Sci. 63 (2008) transfer in water, Water Res. 4 (2) (1970) 115.
3223–3234. [24] F. F. Cabaret, L. Fradette, P.A. Tanguy, Gas–liquid mass transfer in unbaffled
[5] M. Martin, F.J. Montes, M.A. Galan, Mass transfer rates from bubbles in stirred dual-impeller mixers, Chem. Eng. Sci. 63 (2008) 1636–1647.
tanks operating with viscous fluids, Chem. Eng. Sci. 65 (2010) 3814–3824. [25] V. Linek, V. Vacek, P. Benes, A critical review and experimental verification of
[6] K. Van’t Riet, Review of measuring methods and results in non-viscous the correct use of the dynamic method for the determination of oxygen transfer
gas–liquid mass transfer in stirred vessels, Ind. Eng. Chem. Process Des. Dev. in aerated agitated vessels to water, electrolyte solutions and viscous liquids,
18 (3) (1979) 357–364. Chem. Eng. J. 34 (1987) 11–34.
[7] M. Zlokarnik, Sorption characteristics for gas–liquid contacting in mixing ves- [26] A.S. Khare, K. Niranjan, Impeller-agitated aerobic reactor: the influence of tiny
sels, Adv. Biochem. Eng. 8 (1979) 133–157. bubbles on gas hold-up and mass transfer in highly viscous liquids, Chem. Eng.
[8] H. Judat, Gas/liquid mass transfer in stirred vessels - a critical review, Ger. Chem. Sci. 50 (7) (1995) 1091–1105.
Eng. 5 (1982) 357–363. [27] P. Painmanakul, K. Loubière, G. Hébrard, M. Mietton-Peuchot, M. Roustan, Effect
[9] V. Schlüter, W.D. Deckwer, Gas–liquid mass transfer in stirred vessels, Chem. of surfactants on liquid-side mass transfer coefficients, Chem. Eng. Sci. 60
Eng. Sci. 47 (9–11) (1992) 2357–2362. (2005) 6480–6491.
[10] M. Roustan, A. Liné, Rôle du brassage dans les procédés biologiques d’épuration, [28] L.-K. Ju, C.S. Ho, The structure-breaking effect on oxygen diffusion coefficients in
Tribune de l’Eau 49 (1996) 109–115. electrolyte and polyelectrolyte solutions, Can. J. Chem. Eng. 67 (1989) 471–476.
[11] J.F. Perez, O.C. Sandall, Gas absorption by non-Newtonian fluids in agitated [29] T. Espinosa-Solares, E. Brito-De La Fuente, A. Tecante, P.A. Tanguy, Gas dis-
vessels, AIChE J. 20 (1974) 770. persion in rheologically-evolving model fluids by hybrid dual mixing systems,
[12] H. Yagi, F. Yoshida, Gas absorption by Newtonian and non-Newtonian flu- Chem. Eng. Technol. 25 (7) (2002) 723–727.
ids in sparged agitated vessel, Ind. Eng. Chem. Process Des. Dev. 14 (1975) [30] M. Hilal, Etude de l’écoulement d’un fluide non Newtonien à travers les réac-
488–493. teurs à lit fixe: chute de pression, dispersion axiale et transfert de matière,
[13] M. Nishikawa, M. Nakamura, H. Yagi, K. Hashimoto, Gas absorption in aerated Thèse de doctorat, Université of Nantes N 91 NANT 2049, 1991.
mixing vessels, J. Chem. Eng. Jpn. 14 (3) (1981) 219–226. [31] M. Mori, J. Isaac, I. Seyssiecq, N. Roche, Effect of measuring geometries and
[14] M. Nishikawa, M. Nakamura, K. Hashimoto, Gas absorption in aerated mix- of exocellular polymeric substances on the rheological behaviour of sewage
ing vessels with non-Newtonian liquid, J. Chem. Eng. Jpn. 14 (3) (1981) sludge, Chem. Eng. Res. Des. 86 (2008) 554–559.
227–232. [32] M. Zlokarnik, Scale-up in Chemical Engineering Second, Completely Revised
[15] E. Costa, A. Lucas, J. Aguado, J.A. Avila, Transferencia de materia en tanques agi- and Extended edition, Wiley-VCH Verlag GmbH & Co., 2006.
tados: burbujeo de gases en líquidos newtonianos y no newtonianos. I. Turbinas [33] T. Szirtes, Applied Dimensional Analysis and Modelling, McGraw-Hill, New
de 6 paletas y difusor plano, An. Quim. 78 (1982) 387–392. York, 1998.
[16] R.S. Albal, Y.T. Shah, A. Schumpe, N.L. Carr, Mass transfer in multiphase agitated [34] J. Pawlowski, Die ähnlichkeitstheorie in des physikalissch-technischen
contactors, Chem. Eng. J. 27 (1983) 61–80. Forschung- Grundlagen und Anwendungen, Springer Verlag,
[17] F. Garcia-Ochoa, E. Gomez, Mass transfer coefficient in stirred tank reactors for Berlin/Heidelberg/New York, 1971.
xanthan gum solution, Biochem. Eng. J. 1 (1998) 1–10. [35] E.N. Seider, G.E. Tate, Heat transfer and pressure drop of liquids in tubes, Ind.
[18] C. Xuereb, M. Poux, J. Bertrand, Agitation et mélange: Aspects fondamentaux Eng. Chem. 28 (1936) 1429.
et applications industrielles, L’Usine Nouvelle Dunod, 2006. [36] M. Zlokarnik, Scale-up of processes using material systems with variable phys-
[19] R. Hassan, J. Loubière, J. Legrand, Gas–liquid mass transfer in a stirred tank for ical properties, Chem. Biochem. Eng. Q 15 (2) (2001) 43–47.
non-Newtonian fluids. Application to the autothermal thermophilic digestion [37] R. Hassan, Etude expérimentale et modélisation du transfert de matière gaz-
of sludge, in: 9th International Congress of Chemical and Process Engineering liquide en cuve agitée en présence de fluides non-Newtoniens simulant des
(CHISA), Prague, Czech Republic, 2010. boues de station d’épuration, Thesis, University of Nantes, 2011.

You might also like