You are on page 1of 9

chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Effect of different contaminants on the ˛-factor:


Local experimental method and modeling

Pisut Painmanakul a , Gilles Hébrard b,∗


a Department of Environmental Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok, Thailand
b Laboratoire d’Ingénierie des Systèmes Biologiques et des Procédés, UMR INSA-CNRS 5504, UMR INSA-INRA 792,
Institut National des Sciences Appliquées de Toulouse, 135 Avenue de Rangueil, F-31077 Toulouse, Cedex 4, France

a b s t r a c t

In this present paper, the effect of different contaminants (single and mixture) on the variation of ˛-factor (a ratio of
processing water to clean water volumetric mass transfer coefficients, or kL aProcess water /kL aClean water ) was investigated.
Tap water and aqueous solutions with surfactants, NaCl, glucose and inert microorganism were chosen as liquid
phases and an elastic membrane with a single orifice as gas sparger.
This study has clearly shown that the presence of these contaminants affects the bubble generation phenomenon,
the interfacial area (a) and the different mass transfer parameters, such as the volumetric mass transfer coefficient
(kL a) and the liquid-side mass transfer coefficient (kL ). The contamination of surfactants, even in small quantities,
has shown the most significant effects. The determination of the ˛-factor in terms of interfacial area (˛a ) and kL
coefficient (˛kL ) was applied to understand details of the differences of the overall ˛-factor. The liquid surface tension
and the ˛-factor as ˛a and ˛kL prove to be crucial for predicting the changes of ˛-factor in the mixture of various
contaminants.
© 2008 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords: Aeration; ˛-Factor; Contaminants; Interfacial area; Volumetric mass transfer coefficient; Liquid-side mass
transfer coefficient

1. Introduction In order to yield a large surface for oxygen transfer, the


gas is released in the form of small bubbles by using the gas
Concerning the different biological processes such as sparger or bubble diffused aerator (perforated plate, porous
wastewater treatment, fermentation and chemical oxidation disk diffuser and membrane gas sparger). With the punc-
processes, the energy consumption related with aeration sys- tured flexible rubber membranes, a uniform size distribution
tems represents the highest operation cost that ranges from of small bubbles is produced and leads to a large mass transfer
45% to 75% of overall expenditures (Reardon, 1995). From Côté area (Rice and Lakhani, 1983). Consequently, this flexible gas
et al. (1989) and Rosso and Stenstrom (2006), in aeration sys- sparger has been widely used for the last decade in the urban
tem, the oxygen can be transferred into a liquid phase, not wastewater treatment. Several works have been carried out on
only by diffusing gas through a gas–liquid interface, but also the membrane characterization (physical properties) and the
by dissolving gas into the liquid solution by using a semi- bubbles generation phenomena (Rice and Howell, 1986; Rice et
permeable membrane. Normally, the interfacial gas transfer al., 1981; Bischof and Sommerfeld, 1991; Loubière and Hébrard,
created by either shearing the liquid surface with a mechani- 2003; Hébrard et al., 1996; Couvert et al., 1999). Painmanakul
cal aerator (mixer or turbine) or releasing air through spargers et al. (2004) have proposed the methods for comparing several
or porous materials, is applied as the classical environmental membranes and evaluating their performances by experimen-
technologies. tally determining the interfacial area (a) based on the values of


Corresponding author.
E-mail address: gilles.hebrard@insa-toulouse.fr (G. Hébrard).
Received 24 April 2008; Received in revised form 17 June 2008; Accepted 25 June 2008
0263-8762/$ – see front matter © 2008 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.cherd.2008.06.009
1208 chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215

detached bubble diameters (DB ), bubble formation frequencies


Nomenclature (fB ) and terminal bubble rising velocities (UB ). Recently, Rosso
et al. (2008) have analyzed the membrane sparger properties
a interfacial area (m−1 ) change in terms of the fouling and scaling problems obtained
A cross-sectional area of reactor (m2 ) with the real operating conditions.
AB bubble area (m2 ) According to the importance of aeration system’s selection,
C solute concentration in liquid phase (kg/m3 ) the aeration device capacities given by the constructor are
CL dissolved oxygen concentration (kg/m3 ) usually established for standard conditions, i.e. clear water,
C∗L oxygen concentration at saturation (kg/m3 ) with zero dissolved oxygen concentration, at a pressure of
CMC critical micelle concentration (kg/m3 ) 760 mmHg and a temperature of 10 or 20 ◦ C for European or
DB bubble diameter (m) U.S. standards, respectively. Therefore, it is essential to correct
DOR equivalent hole diameter (m) this aeration capacity in respect to the real operating con-
E bubble eccentricity dition. Generally, the impact of water or wastewater quality
fB bubble formation frequency (s−1 ) on the aeration capacity is quantified by the ˛-factor (a ratio
HL liquid height (m) of processing water to clean water volumetric mass trans-
kL liquid-side mass transfer coefficient (m/s) fer coefficients, or kL aProcess water /kL aClean water ). From Kessener
kL a volumetric mass transfer coefficient (s−1 ) and Ribbius (1935), the effects of aeration devices (coarse bub-
K adsorption constant at the equilibrium ble, fine pore, surface, etc.) on the ˛-factor have been studied.
(m3 /mol) Rosso and Stenstrom (2006) have shown that different aeration
mR mass of Na2 SO3 remaining in the column (kg) methods have a variety of ˛-factor values, and for fine-pore
mS mass of Na2 SO3 reacting with the dissolved diffusers, the initial ˛ decreases over time in the operation
oxygen (kg) because of fouling or scaling. This ˛-factor is also a function
mT total mass of Na2 SO3 introduced initially (kg) of process conditions such as the mean cell retention time
M molecular mass (kg/mol) (MCRT), the air flow rate, the horizontal velocity of the liq-
NB number of bubbles generated uid and also the configurations of the aeration system (Rosso
QG gas flow rate (m3 /s) et al., 2005; Gillot et al., 2000; Gillot and Héduit, 2000). Fur-
SB total bubble surface (m2 ) thermore, several studies have been continued to understand
Se surface coverage ratio at equilibrium a variation of the ˛-factor that is caused by contaminants
taeration aeration time (s) such as organic substances, surfactants, raw wastewater, etc.
T temperature (◦ C) (Gillot et al., 2000; Stenstrom and Gilbert, 1981; Mancy and
UB bubble rising velocity (m/s) Okun, 1960; Hébrard et al., 2000; Rosso and Stenstrom, 2006).
UG gas velocity through the orifice (m/s) However, according to the mixture of the various contamina-
VB bubble volume (m3 ) tions present in the liquid phases (surfactant, suspended solid,
VC gas chamber volume between the control valve microorganism, salt, organic substances, etc.), the current
and the orifice (m3 ) problem is to correctly understand and predict the evaluation
VL liquid volume in reactor (m3 ) of this ˛-factor.
VTotal total volume in reactor (m3 ) In fact, the determination of the kL a coefficient in a com-
plex system is not an easy task. As in the case of a biological
Dimensionless numbers
reactor, the well-known gas method of classical mass oxygen
Re bubble Reynolds number defined by
balance is difficult to apply because the biological respira-
Re = UB ·L ·DB /L
tion cannot be correctly evaluated. The literature about mass
ReOR hole Reynolds number defined by
transfer parameters shows that there is a limited number of
Re = UG ·G ·DOR /G
qualitative data related to the influence of contaminants or
We Weber number defined by We = UG2 ·D
OR · G /L
surface tension on this kL a value. In addition, the kL a values
are often global and insufficient to grasp the oxygen mass
Greek symbols
transfer mechanisms that directly affect the variation of the
∞ surface concentration when it is saturated
˛-factor (Vázquez et al., 1997; Akosman et al., 2004). For this
(mol/m2 )
purpose, it becomes essential to separate the parameters,
˛ ˛-factor obtained with the experimental
especially the liquid-side mass transfer coefficient (kL ) and
method
the interfacial area (a) (Bouaifi et al., 2001; Zhao et al., 2003a,b;
˛a ˛-factor in terms of interfacial area
Vasconcelos et al., 2003; Vázquez et al., 2000; Cents et al., 2001).
˛kL ˛-factor in terms of liquid-side mass transfer
To fill this gap, the general aim of this work is to study the
coefficient
variation of ˛-factor due to different operating conditions dis-
˛MC ˛-factor of the mixture of various contaminants
covered in the biological processes. A local experimental mass
calculated by Eq. (8)
transfer determination method will be applied. The interfacial
˛SC ˛-factor of the single contaminants calculated
area and the liquid-side mass transfer coefficient will be dis-
L liquid viscosity (Pa s)
sociated and analyzed to provide a better understanding of the
L liquid density (kg/m3 )
parameters which influence the oxygen mass transfer. Then,
L liquid surface tension (N/m)
these two parameters experimentally obtained will be used to
 L,0 surface tension when the solvent is pure (N/m)
determine the ˛-factor associated with the interfacial area and
the liquid-side mass transfer coefficient, and thus to predict
the overall ˛-factor.
In this paper, it will firstly present the material and the
experimental methods that are used in this study. Then,
chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215 1209

BIOBLOCK 915PM247 (3). The average gas flow rate is mea-


sured by using a soap film meter (7), through a funnel (1.5 cm
diameter) put on the orifice. Nitrogen flow (used for oxygen
elimination in the liquid phase and oxygen elimination at the
top of the bubble column) is controlled by a pressure gauge
(11). The UNISENSE oxygen microsensor, that has a quick
response time (as low as 50 ms), is used to measure the change
in dissolved oxygen concentration (9). All chemical solutions
(8) are injected at the top of the column. The operating condi-
tions are as follows: liquid height HL = 25 cm and temperature
T = 20 ◦ C.
Fig. 1 – Schematic diagram of the experimental set-up. (1) In this work, bubbles are generated by a single puncture
Pressure gauge, (2) gas flow meter, (3) electronic located at the membrane center which has a hole diame-
manometer, (4) bubble generation vessel, (5) membrane ter (DOR ) varied between 0.29 and 0.48 mm. Complementary
sparger, (6) bubble column, (7) soap film meter, (8) Chemical information about the experimental set-up can be found
solution, (9) oxygen microsensor, (10) acquisition computer (Painmanakul et al., 2005). It can be noticed that the gas
and camera, (11) nitrogen pressure gauge, (12) agitation flow rate which is studied (QG = 0.3–3.45 ml/s) in this work can
system. correspond to the static and the dynamic bubbling regimes.
However, the jetting regime is not considered (ReOR = 80–1000;
the characterization of liquid phases under a test will be WeOR = 0.09–4).
described, as well as the local mass transfer parameter deter-
mination (the interfacial area provided by sparger, volumetric 2.2. Liquid phase characterization
mass transfer coefficient obtained with the local experimental
method and liquid-side mass transfer coefficient). Finally, the To understand the effect of various contaminants on the vari-
effects of different contaminants (single and mixture) on the ation of the ˛-factor, it is important to initially characterize
mass transfer parameters and the ˛-factor values will be dis- the liquid phases under the test: tap water and aqueous solu-
cussed. Thus, a simple correlation for estimating the ˛-factor tions with surfactants, NaCl, glucose and inert microorganism
will be proposed and applied in whatever operating condi- (solid suspension with liquid secretions). Moreover, the com-
tions. bination of this various liquid phases is also investigated. In
this study, the contaminants have been chosen with regards
2. Materials and methods to their nature and application (wastewater treatment). The
selected contaminants (glucose and inert cell) are based on
2.1. Experimental set-up the real operating conditions that are obtained within the bio-
logical fermentation process. The summary series of liquid
Fig. 1 presents the experimental set-up used in this study. phase characteristics are defined and numbered in Table 1.
These experiments are carried out in a glass bubble column Given that these liquids are dilute aqueous solutions, their
(6), 0.05 m in diameter, 0.40 m in height. This column is fixed density and viscosity are assumed to be equal to those of tap
into a glass parallelepiped vessel (4) that has the size of 0.40 m water (997 kg/m3 and 10−3 Pa s, respectively).
width, 0.40 m length and 0.30 m height. The flow of air is mon-
itored by a pressure gauge (1) and regulated by a gas flow 2.2.1. Static surface tension
meter (2). Pressure drop created by the membrane sparger For a given liquid phase, the liquid phase characterization will
is determined using an electronic manometer type so-called firstly consist of determining their static surface tensions by

Table 1 – Chemical characteristics of liquid phases


No. Solution type Chemical name Ma (kg/mol) [C]b (mg/l)  L (mN/m)
−3
1 Tap water – 18 × 10 – 71.8

2 20,000 75.8
NaCl Sodium chloride 58 × 10−3
3 4,000 73.8

4 250 70.2
Glucose 6-(Hydroxymethyl)Oxane-2,3,4,5-tetrol 180 × 10−3
5 50 70.6

6 Inert cell – – 250 57.9

7 1,900 39.7
8 −3
200 60.45
Anionic surfactant Sodium laurylsulfate 382 × 10
9 110 65
10 50 69.78

11 Lauryl dimethyl benzyl ammonium 2,000 27.6


Cationic surfactant 400 × 10−3
12 bromine 110 42.4

13 Fatty alcohol C12/18, 10 EO, n-butyl 3,500 30.4


Non-ionic surfactant 700 × 10−3
14 end-capped 86 45.10

a
M is the molecular weight.
b
[C] is the concentration used for experiments.
1210 chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215

Table 2 – CMC value and characteristics adsorption parameter of surfactants


Solution type Ma (kg/mol) [C]b (mg/l)  L (mN/m) CMC (mg/l)  ∞ (mol/m2 ) K (m3 /mol) Se
−3
Tap water 18 × 10 – 71.8 – – – 0

1900 39.7 1
−3
200 60.45 −5
0.8
Anionic surfactant 382 × 10 1900 3.52 × 10 6.25
110 65 0.6
50 69.78 0.4

2000 27.6 1
Cationic surfactant 400 × 10−3 920 3.49 × 10−5 90.9
110 42.4 0.92

3500 30.4 1.00


Non-ionic surfactant 700 × 10−3 400 2.56 × 10−6 357.00
86 45.10 0.97

a
M is the molecular weight.
b
[C] is the concentration used for experiments.

using the Digidrop GBX (pendant drop method) and Krüss K6 as it would affect the coalescing properties of liquid phase.
(Wilhelmy plate method) tensimeters. The static surface ten- Thus, an adequate amount of Na2 SO3 has to be found for keep-
sion and the chemical properties of the liquid under test are ing a zero O2 concentration during the limited aeration step.
reported in Table 1. The lowest surface tension is obtained for The crucial advantages of this method are:
the cationic surfactant solution with 2000 mg/l and the high-
est for NaCl solution. Overall the resulting trend is found as • An excessive utilization of Na2 SO3 is avoided since the de-
follows: oxygenation is made by injection of nitrogen.
• No models for the mixing in the liquid or the gas phase are
L cationic < L non-ionic < L anionic < L cell < L glucose required.


= L water < L NaCl
Further details about this method and its comparison with
the classical dynamic method can be found in Painmanakul et
It can be noted that the low  L value in the case of the solu-
al. (2005). In this method, the local volumetric mass transfer
tion with inert cell can be linked to the effect of surfactant
coefficient, kL a, is expressed as:
molecules produced by microorganisms in the biological fer-
mentation processes. 1/2(MO2 /MNa2 SO3 )mS
kL a = (1)
taeration VL C∗L
2.2.2. Surfactant solution analysis
For the solution with surfactants (anionic, cationic and non- where MO2 is the molecular mass of oxygen; MNa2 SO3 is the
ionic types), the critical micelle concentration (CMC) and molecular mass of sodium sulfite; mS is the mass of Na2 SO3
adsorption parameters have been determined. The values of reacting with the oxygen dissolved during the steady state
CMC are reported in Table 2. In order to characterize the regime; taeration is the aeration time (120 s); VL is the liquid
adsorption of solute molecules at a gas–liquid interface, the volume in the bubble column; C∗L is the saturation oxygen
method based on the Langmuir theory has been used as in concentration in the liquid.
Loubière and Hébrard (2004).
The CMC of each surfactant and the characteristic adsorp- 2.3.2. Local interfacial area (a)
tion parameters such as adsorption constant at equilibrium As in Albespy et al. (2003), the local interfacial area is defined as
(K), surface concentration when it is saturated ( ∞ ) and sur- the ratio between the bubble surfaces (SB ) and the total volume
face coverage ratio at equilibrium (Se ) are reported in Table 2. in the reactor (VTotal ):

2.3. Mass transfer parameter determination SB


a = NB × (2)
VTotal
In this study, the local experimental approach presented in
NB is number of bubbles that can be deduced from the bubble
Sardeing et al. (2006) is used to firstly determine the volumetric
rising velocities (UB ) and the bubble formation frequency (fB ).
mass transfer coefficient and the corresponding local interfa-
In this work, the interfacial area is, therefore, expressed as Eq.
cial area which is provided by a single orifice sparger. Thus, the
(3):
local liquid-side mass transfer coefficient can be calculated.
SB HL SB
2.3.1. Local volumetric mass transfer coefficient (kL a) a = NB × = fB × × (3)
VTotal UB A · HL + NB · VB
The kL a determination method used in this study is proposed
by Painmanakul et al. (2005). This method is based on a mass A and HL are the cross-sectional area of the bubble column
balance on sulfite sodium (Na2 SO3 ) concentration during the and the liquid height, respectively. Moreover, according to the
aeration time. Nitrogen is firstly injected into the liquid phase hydrodynamic parameter determination method proposed by
in order to remove the dissolved oxygen presents in the bubble Painmanakul et al. (2005), the bubble diameters, the bubble
column. When the concentration of dissolved oxygen reaches formation frequency and the terminal rising velocities can be
nearly zero, fresh oxygen from the air, is introduced and reacts correctly determined by an image analysis. This experimental
with the small quantity of Na2 SO3 . This step was limited to approach can be thus applied to determine the value of (a) in
120 s. Note that an excessive use of Na2 SO3 should be avoided this work.
chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215 1211

As the bubbles in this study being mainly ellipsoidal, the


Table 3 – Slopes of the curve relating kL a to QG and the
bubble surface area is calculated as: ˛-factor of different liquid phases

E
 E + 1  No. Solution type Slope of kL a = f(QG ) ˛-Factor
SB = l2 1 + ln (4)
2 2 E−1
1 Tap water 1.381 1.000

where E is the bubble eccentricity defined in this work as ratio 2 1.591 1.152
NaCl
between the length, l, and the height, h, of an ellipsoidal bub- 3 1.454 1.053
ble. According to Eq. (3), the interfacial area is a function of 4 1.277 0.925
Glucose
the bubble formation frequency, the terminal bubble rising 5 1.375 0.996
velocity and the generated bubble diameter. 6 Inert cell 1.643 1.190

7 0.671 0.486
2.3.3. Liquid-side mass transfer coefficient (kL )
8 0.831 0.602
The volumetric mass transfer coefficient (kL a) is the product 9
Anionic surfactant
0.921 0.667
of the liquid-side mass transfer coefficient (kL ) and the inter- 10 1.092 0.791
facial area (a). The local liquid-side mass transfer coefficient
11 0.550 0.398
is simply determined by: Cationic surfactant
12 0.552 0.400

kL a 13 0.612 0.443
kL = (5) 14
Non-ionic surfactant
0.615 0.445
a

In this study, these two values are experimentally obtained


in the local conditions, giving a good accuracy. The average
and the maximum experimental errors for determining the kL of the ˛-factor (Rosso and Stenstrom, 2006). Therefore, it can
value have been estimated at ±15% and ±30%, respectively. be concluded that the correction ˛-factors depend on the liq-
uid phase types in the conditions in response to the static and
3. Results and discussion the dynamic bubbling regimes.
To thoroughly understand these phenomena, both interfa-
In this part, the effect of the single contaminants on the ˛- cial area and liquid-side mass transfer coefficient need to be
factor will be firstly presented and then the second part will considered separately.
deal with the mixture of various contaminants: the simple cor-
relation will be proposed in order to predict these obtained 3.1.2. Interfacial area (a)
values of ˛-factor. In this work, the bubble diameters vary between 2.02 and
6.74 mm while gas flow rates can change between 0.3 and
3.1. The study of the single contaminant present in the 3.45 ml/s. The results agree with the bubble generation phe-
liquid phases nomena, associated with membrane spargers (Painmanakul et
al., 2004). Over this bubble diameter range, the terminal rising
Only single contaminant solutions results are reported in bubble velocities (obtained experimentally) are nearly con-
this part. To present the following experimental results, each stant. They vary between 15 and 25 cm s−1 and are not within
operating condition linked to a particular solution has been the range of the UB values of Grace and Wairegi (1986) corre-
numbered as defined in Table 1. sponding to the contaminated and pure systems. Moreover, it
can be noted that the presence of contaminants can affect the
3.1.1. Variation of the ˛-factor bubble hydrodynamic parameters, especially the contamina-
The correction ˛-factor is defined as the ratio between kL a in tion of surfactants. By using the experimental results of the
different liquid phases and kL a in tap water. According to the bubble diameter (DB ) and the bubble rising velocity (UB ), the
linear augmentation of kL a values with the gas flow rate, the bubble formation frequencies (fB ) related to the gas flow rates
values of the ˛-factor in this study are calculated by the ratio can be calculated. Then, the local interfacial area (a) can be
between slopes of the curve relating kL a to QG obtained with determined.
each liquid phase under a test and those obtained with the Figs. 2 and 3 present the variations of the interfacial area,
tap water. The values of the slope obtained with the curve together with the gas flow rate for the different liquid phases.
relating kL a to QG and the ˛-factor calculated for every liquid The summary series of liquid phase characteristics are defined
phase under a test are shown in Table 3. and numbered as previously presented in Table 1.
According to Table 3, the values of ˛-factor that are obtained As shown in Figs. 2 and 3, in any liquid phase, the inter-
in the aqueous solutions, together with NaCl, are greater than facial area roughly increases linearly with the gas flow rate.
1 (1.1–1.2) whereas those obtained with another solution are Their values vary between 1.8 and 16.6 m−1 whereas the gas
less than 1 (0.3–0.9). These results are in good agreement with flow rates change between 0.3 and 3.45 ml/s. The values of
the values of the ˛-factor that can be obtained in real operat- (a) are directly linked to the bubble diameter and thus the
ing conditions (Zlokarnik, 1979; Gillot et al., 2000; Gillot and static surface tension of liquid phases under test: low values
Héduit, 2000; Chern et al., 2001). Noticeably, the surfactant of  L are associated with high values of (a) (Painmanakul et
contaminations have more influence on the reduction of the al., 2005; Loubière and Hébrard, 2004; Sardeing et al., 2006).
values of the ˛-factor than another liquid phase under the As a result, the interfacial areas related to the solutions with
test. However, as the gas flow rate or Re number increases surfactants and inert cell are significantly larger than those
into the turbulent regime, the surface renewal rate can be of other solutions. However, with higher  L value obtained
sufficiently high to shear surfactant molecules or another con- with the addition of NaCl, the increase of the (a) values can
taminant off the bubble surface: these can increase the values be observed at the high gas flow rates because of the preven-
1212 chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215

3.1.4. Determination of correction ˛-factor in term of a


and kL values
In order to understand the differences of the ˛-factor previ-
ously presented in Table 3, it is essential to separately consider
their variation related to the interfacial area and the liquid-
side mass transfer coefficient. The values of the ˛-factor
associated with the interfacial area (˛a ) are calculated by the
ratio between slopes of the curve relating (a) to QG obtained
with all liquid phases under the test and those obtained with
the tap water. Whereas, the ratio within the average values of
kL coefficient obtained with the high gas flow rate zone is used
to determine the ˛-factor that is related to the liquid-side mass
transfer coefficient (˛kL ). In this work, the values of ˛a and ˛kL
are, therefore, expressed as Eqs. (6) and (7), respectively:

Slope of a = f (QG )
˛a = (6)
Slope of a = f (QG )Tap water

Average kL value at high QG


˛kL = (7)
(Average kL value at high QG )Tap water

Table 4 demonstrates the ˛-factor calculated in term of the


values of (a) and kL for whatever the liquid phases under the
test.
According to Table 4, in the presence of surfactants
molecules, the ˛a values are greater than 1. On the contrary,
Figs. 2 and 3 – Interfacial area versus gas flow rate for the associated ˛kL values show in the opposite way: both ˛a
different liquid phases. and ˛kL values compensate each other in this type of liquid
phases. However, more significant variation are observed with
the values of ˛a than those observed with the ˛kL value in
tion of bubble coalescence phenomena provided by the NaCl the case of NaCl. The diminutive effect is also found with the
molecules presence in the liquid phase (Deckwer, 1992). More- glucose solution. These results show the interest of the dis-
over, the substantial increase of the a values can be observed sociation of the overall ˛-factor in terms of interfacial area
in the case of the solutions with inert cell. It can be noted (˛a ) and kL coefficient (˛kL ) in order to understand the role
that, not only the static surface tension, but also the presence
of inert cell (solid) in liquid phase can possibly affect the bub-
ble hydrodynamic parameters and thus the interfacial area. In
this study, the following overall trend is found:

aNon-ionic ∼
= aInert cell > aCationic > aAnionic > aNaCl > aGlucose ∼
= aWater

3.1.3. Liquid-side mass transfer coefficient (kL )


In this part, the kL coefficient is calculated from the experi-
mental values of the volumetric mass transfer coefficient and
the experimental values of the interfacial area (Figs. 2 and 3)
by Eq. (5).
Figs. 4 and 5 show the variation of the liquid-side mass
transfer coefficient, kL , with the gas flow rate for the different
liquid phases.
According to Figs. 4 and 5, the values of kL obtained vary
between 1 × 10−4 and 4 × 10−4 m s−1 for gas flow rates varying
between 0.3 and 3.45 ml/s. In every liquid phase, the kL values
constantly remain for the gas flow rates greater than 0.5 ml/s.
Nevertheless, in case of the lower gas flow rates, an increase
in the kL values is observed. It can be noted that the results
in this study agree with the three zones of the kL coefficients
proposed by Sardeing et al. (2006), Calderbank and Moo-Young
(1961) and Painmanakul (2005). Moreover, it can be noted that
the presence of surfactants, even in little quantities, can have
more significant effects, not only on the values of kL a and (a),
but also on the calculated kL values than those of another con-
taminant. These results certainly affect the oxygen transfer Figs. 4 and 5 – Liquid-side mass transfer coefficient versus
mechanism. gas flow rate for different liquid phases.
chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215 1213

Table 4 – The ˛-factor in terms of a and kL values for different liquid phases
No. Solution type ˛-Factor Slope a = f(QG ) ˛a Avg kL (×10−4 m s−1 ) ˛kL ˛SC = ˛a × ˛kL  (%)

1 Tap water 1 3.58 1.000 3.84 1.000 1.000 0.00

2 1.152 4.22 1.179 3.87 1.008 1.188 3.12


NaCl
3 1.053 3.89 1.087 3.82 0.995 1.087 3.19

4 0.925 3.29 0.919 3.71 0.966 0.888 −4.01


Glucose
5 0.996 3.79 1.059 3.69 0.961 1.017 2.14

6 Inert cell 1.19 6.76 1.888 2.44 0.635 1.200 0.83

7 0.486 4.3 1.201 1.65 0.430 0.516 6.19


8 0.602 3.9 1.089 2.07 0.539 0.587 −2.45
Anionic surfactant
9 0.667 3.85 1.075 2.5 0.651 0.700 4.97
10 0.791 3.62 1.011 2.86 0.745 0.753 −4.79

11 0.398 4.37 1.221 1.31 0.341 0.416 4.63


Cationic surfactant
12 0.4 3.84 1.073 1.51 0.393 0.422 5.45

13 0.443 5.54 1.547 1.14 0.297 0.459 3.70


Non-ionic surfactant
14 0.445 4.98 1.391 1.28 0.333 0.464 4.20

of physico-chemical factors that influence the mass transfer contaminants (˛MC ). In this study, the following correlation is
mechanism. proposed:
In conclusion, the determination of these ˛a and ˛kL val-
L Mixture 
n
ues can provide a better understanding of the effect of each
contaminant on the oxygen transfer processes. ˛MC = (n − 1) × × (˛a,i · ˛kL ,i ) (8)
L Water
i=1

3.2. The study of the mixture of various contaminants where ˛a,i and ˛kL ,i are the values of ˛-factor associated
present in the liquid phases with the interfacial area and the liquid-side mass transfer
coefficient of component (i), respectively. These values are
3.2.1. Variation of the ˛-factor previously presented in Table 4. The values of  L Water and
In order to understand the effect of liquid phase properties in  L Mixture are the measured static surface tension obtained
the actual operating conditions that are observed in the bio- with tap water and the mixture of various contaminants,
logical process (fermentation or supernatant wastewater), the respectively. Additionally, n is the tap water phase plus the
combination of various contaminants is selected and investi- number of contaminant phases injected in the liquid phases
gated in this work. The different series of their mixtures are and it is thus equal to 6 in this study.
numbered as defined in Table 5. In Fig. 6, the values of ˛-factors calculated by Eq. (8) are
As shown in Table 5, Number 17 has the lowest  L mixture compared with those obtained with the experimental method
while Number 18 has the highest  L mixture . In comparison with for any liquid phase presented in Table 5.
the results presented in Table 1, the mixture of various con- This figure shows that a relatively good agreement between
taminants can affect the variation of the static surface tension. the experimental and the predicted ˛MC is obtained (aver-
Moreover, it can be noticed that the presence of surfactants, age difference about ±20%). However, more experimental data
even in small quantities, can have significant effects on the are necessary to accurately validate this correlation. In the
reduction of the  L values. Moreover, the ˛-factor obtained future, other mixtures of various contaminants and different
with the mixture of various contaminants are reported and gas spargers should be tested to extend the operating condi-
compared with those obtained with tap water in Table 5. tion ranges. Furthermore, the improvement in the accuracy
According to Table 5, the values of ˛-factor are very low
(0.3–0.5). This ˛-factor range corresponds to those obtained
with the high concentration of the single surfactant present
in the liquid phase (Table 4). In addition, these results confirm
the major influence of surfactants, even in small quantities,
on the ˛-factor reduction and thus oxygen transfer efficiency.
This, in the real operating condition of the aeration process,
is significantly important to have a great care in determin-
ing and applying an appropriate ˛ value related to aeration
systems design. A simple correlation for estimating ˛-factor
should also be studied and proposed.

3.2.2. Prediction of the ˛-factor in various contaminants


Regarding their importance, the static surface tension ( L ) and
the values of the ˛-factor in terms of the interfacial area (˛a )
and the liquid-side mass transfer coefficient (˛kL ) obtained
experimentally with the single contaminant are chosen to pre- Fig. 6 – Comparison of experimental and predicted the
dict the overall ˛-factor obtained with the mixture of various ˛-factor values.
1214 chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215

Table 5 – The ˛-factor obtained with four types of various contaminant mixtures
No. NaCl (mg/L) Glucose (mg/L) Surfactants (mg/L)  L Mixture (mN/m) Slope of kL a = f(QG ) ˛-Factor
(×10−3 ml−1 )

Anionic Cationic Non-ionic

1 – – – – – 71.2 1.381 1.000


16 4,000 50 110 110 86 42.5 0.548 0.397
17 4,000 250 1900 110 86 41 0.457 0.331
18 20,000 50 110 110 86 47 0.678 0.491
19 20,000 250 1900 110 86 45 0.562 0.407

of the experimental method presented here can be applied parameters for predicting the variation of the ˛MC values.
in the fermentation media and the industrial and/or urban - The simply proposed correlation has allowed a relative good
wastewaters. coincidence between the experimental and predicted ˛MC to
be obtained (average difference about ±20%).
4. Conclusion
In the future, it is essential to continue studying care-
The objective of this work is to study the variation of the ˛- fully in the contamination of the solid phase and the active
factor due to the different operating conditions discovered microorganism in order to grasp the effect of the physical mass
in the biological processes. For this purpose, the effect of transfer mechanism and the bacteria growth rate on the oxy-
the liquid phases with single contaminant and with mix- gen transfer process. In addition, it is obvious that the results
ture of various contaminants is characterized in terms of the observed in our little bubble column volume have to be vali-
mass transfer parameters and the ˛-factor values. The sim- dated into a tall bubble column and at higher gas flow rates.
ple correlation for estimating ˛-factor is proposed and applied In order to validate the ˛-factor predicting correlation pro-
whatever the operating conditions. posed in this study, more experimental data are necessary
For the study of the single contaminant presence in liquid with other mixtures of various contaminants and different
phase, the following results have been obtained: gas spargers to extend the operating condition ranges such as
the fermentation media and also the industrial and/or urban
- The presence of surfactants, even in small quantities, can wastewaters.
have more significant effects on the ˛-factor values than
those of another contaminant. These results, thus, affect References
the oxygen transfer mechanism.
- The modification of surfactant concentration (i.e. of surface Akosman, C., Orhan, R. and Dursun, G., 2004, Effects of liquid
tension value and of surface coverage ratio at equilibrium property on gas holdup and mass transfer in co-current
downflow contacting column. Chem Eng Process, 43: 503–509.
(Se )) shows a direct effect on gas–liquid mass transfer.
Albespy, S., Bonnet, S., Marie, C., Painmanakul, P., Loubière, K.,
- The contaminants under the test, especially surfactants,
Hébrard, G. and Mietton-Peuchot, M., 2003, Bubble generation
considerably affect the bubble generation process (a bub- with a ceramic gas sparger classically used in the wine
ble diameter and a bubble rising velocity) as well as the micro-oxygenation, In Proceedings of the 9ème Congrès de la
interfacial area (a). The (a) values increase with the gas flow Société Française de Génie des Procédés
rates whatever the liquid phases. The large values of (a) are Bischof, F. and Sommerfeld, M., 1991, Studies of the bubble
obtained with small bubble size; therefore, these values are formation process for optimisation of aeration systems, In
Proceeding of the International Conference on Multiphase Flows
logically related to the  L values of the liquid phases.
Bouaifi, M., Hébrard, G., Bastoul, D. and Roustan, M., 2001, A
- The presence of surfactants, even in little quantities, can comparative study of gas hold-up, bubble size, interfacial area
have more significant effects, not only on the values of (a), and mass transfer coefficients in stirred gas–liquid reactors
but also on the kL values, than those of another contami- and bubble columns. Chem Eng Process, 40: 97–111.
nant. Calderbank, P.H. and Moo-Young, M.B., 1961, The continuous
- The determination of ˛-factor in terms of interfacial area phase heat and mass-transfer properties of dispersions.
and kL coefficient (˛a and ˛kL , respectively) can be used to Chem Eng Sci, 16: 39–54.
Cents, A.H.G., Brilman, D.W.F. and Versteeg, G.F., 2001, Gas
understand the differences of the overall ˛-factor and thus
absorption in an agitated gas–liquid–liquid system. Chem Eng
the role of physico-chemical factors that influence the mass Sci, 56: 1075–1083.
transfer mechanism. Chern, J.M., Chou, S.R. and Shang, C.S., 2001, Effects of impurities
on oxygen transfer rates in diffused aeration systems. Water
Res, 35(13): 3041–3048.
The studies relating to the mixtures of various contami-
Côté, P., Bersillon, J.L. and Huyard, A., 1989, Bubble-free aeration
nants show that:
using membranes: mass transfer analysis. J Membr Sci, 47:
91–106.
- The ˛-factors obtained with the mixture of various contam- Couvert, A., Roustan, M. and Chatellier, P., 1999, Two-phase
inants (˛MC ) range between 0.3 and 0.5 and correspond to hydrodynamic study for a rectangular air-lift loop reactor with
an internal baffle. Chem Eng Sci, 54(21): 5245–5252.
those obtained with the high concentration of the single
Deckwer, W.D., (1992). Bubble Column Reactors. (John Wiley & Son
surfactant present in the liquid phase.
Ltd, Baffins Lane, Chichester, England).
- The liquid surface tension and the ˛-factor in terms of inter- Gillot, S. and Héduit, A., 2000, Effect of air flow rate on oxygen
facial area (˛a ) and kL coefficient (˛kL ) determined in the case transfer in an oxidation ditch equipped with fine bubble
of the single contaminant have proven to be the significant diffusers and slow speed mixers. Water Res, 34(5): 1756–1762.
chemical engineering research and design 8 6 ( 2 0 0 8 ) 1207–1215 1215

Gillot, S., Capela, S. and Heduit, A., 2000, Effect of horizontal flow Rice, R.G., Tupperainen, J.M.I. and Hedge, R., 1981, Dispersion and
on oxygen transfer in clean water and in clean water with hold up in bubble columns: comparison of rigid and flexible
surfactants. Water Res, 34(2): 678–683. sparger. Can J Chem Eng, 59: 677–687.
Grace, J.R. and Wairegi, T., (1986). Properties and characteristics of Rosso, D. and Stenstrom, M.K., 2006, Surfactant effects on
drops and bubbles. In Encyclopedia of Fluid Mechanics, ␣-factors in aeration systems. Water Res, 40(7): 1397–1404.
Cheremisinoff. (Gulf Publishing Corporation, Huston, TX), pp. Rosso, D., Iranpour, R. and Stenstrom, M.K., 2005, Fifteen years of
43–57 (Chapter 3) off-gas transfer efficiency measurements on fine-pore
Hébrard, G., Bastoul, D. and Roustan, M., 1996, Influence of the aerators: key role of sludge age and normalized air flux. Water
gas spargers on the hydrodynamic behaviour of bubble Environ Res, 77(3): 266–273.
columns. Trans IChemE, 74(A): 406–414. Rosso, D., Libra, J.A., Wiehe, W. and Stenstrom, M.K., 2008,
Hébrard, G., Destrac, P., Roustan, M., Huyard, A. and Audic, J.M., Membrane properties change in fine-pore aeration diffusers:
2000, Determination of the water quality correction factor full-scale variations of transfer efficiency and headloss. Water
using a tracer gas method. Water Res, 34(2): 684–689. Res, 42(10–11): 2640–2648.
Kessener, H.J. and Ribbius, F.J., 1935, Practical activated sludge Sardeing, R., Painmanakul, P. and Hébrard, G., 2006, Effect of
research. J Proc Inst Sew Purif, 6(3): 50–56. surfactants on liquid-side mass transfer coefficients in
Loubière, K. and Hébrard, G., 2003, Bubble formation from a gas–liquid systems: a first step to modelling. Chem Eng Sci,
flexible hole submerged in an inviscid liquid. Chem Eng Sci, 61(19): 6249–6260.
58: 135–148. Stenstrom, M.K. and Gilbert, R.G., 1981, Effects of alpha, beta and
Loubière, K. and Hébrard, G., 2004, Influence of liquid surface theta factors in design, specification and operations of
tension (surfactants) on bubble formation at rigid and flexible aeration systems. Water Res, 15: 643–654.
orifices. Chem Eng Process, 43: 1361–1369. Vasconcelos, J.M.T., Rodrigues, J.M.L., Orvalho, S.C.P., Alves, S.S.,
Mancy, K.H. and Okun, D.A., 1960, Effects of surface active agents Mendes, R.L. and Reis, A., 2003, Effect of contaminants on
on bubble aeration. J Water Pollut Control Fed, 32(4): 351–364. mass transfer coefficients in bubble column and airlift
P. Painmanakul, 2005, Analyse locale du transfert de matière contactors. Chem Eng Sci, 58: 1431–1440.
associé à la formation de bulles générées par différents types Vázquez, G., Cancela, M.A., Riverol, C., Alvarez, E. and Navaza,
d’orifices dans différentes phases liquides Newtoniennes: J.M., 2000, Application of the Danckwerts method in a bubble
étude expérimentale et modélisation, Ph.D. Thesis (INSA column: effects of surfactants on mass transfer coefficient
Toulouse, France). and interfacial area. Chem Eng J, 78: 13–19.
Painmanakul, P., Loubière, K., Hébrard, G. and Buffière, P., 2004, Vázquez, G., Cancela, M.A., Varela, R., Alvarez, E. and Navaza,
Study of different membrane spargers used in wastewater J.M., 1997, Influence of surfactants on absorption of CO2 in a
treatment: characterisation and performance. Chem Eng stirred tank with and without bubbling. Chem Eng J, 67:
Process, 43: 1347–1359. 131–137.
Painmanakul, P., Loubière, K., Hébrard, G., Mietton-Peuchot, M. Zhao, B., Wang, J., Yang, W. and Jin, Y., 2003, Gas–liquid mass
and Roustan, M., 2005, Effects of surfactants on liquid-side transfer in slurry bubble systems: I. Mathematical modeling
mass transfer coefficients. Chem Eng Sci, 60: 6480–6491. based on a single bubble mechanism. Chem Eng J, 96:
Reardon, D.J., 1995, Turning down the power. Civ Eng, 65(8): 54–56. 23–27.
Rice, R.G. and Howell, S.W., 1986, Elastic and flow mechanics for Zhao, B., Wang, J., Yang, W. and Jin, Y., 2003, Gas–liquid mass
membrane spargers. AIChE J, 32(8) transfer in slurry bubble systems: II. Mathematical modeling
Rice, R.G. and Lakhani, N.B., 1983, Bubble formation at a puncture based on a single bubble mechanism. Chem Eng J, 96: 29–35.
in a submerged rubber membrane. Chem Eng Commun, 24: Zlokarnik, M., 1979, Sorption characteristics of slot injectors.
215–234. Chem Eng Sci J, 34: 1265–1271.

You might also like