You are on page 1of 33

Accepted Manuscript

Title: CFD simulation of the hydrodynamics in an industrial


scale cyclohexane oxidation airlift loop reactor

Author: <ce:author id="aut0005"


author-id="S0263876217300102-
333dc3e2801188916b99eef8b8df1a25"> Limeng
Chen<ce:author id="aut0010"
author-id="S0263876217300102-
9f9522bd93144f0dc2123fb73c799738"> Zhishan
Bai

PII: S0263-8762(17)30010-2
DOI: http://dx.doi.org/doi:10.1016/j.cherd.2017.01.008
Reference: CHERD 2537

To appear in:

Received date: 27-6-2016


Revised date: 16-11-2016
Accepted date: 4-1-2017

Please cite this article as: Chen, Limeng, Bai, Zhishan, CFD
simulation of the hydrodynamics in an industrial scale cyclohexane
oxidation airlift loop reactor.Chemical Engineering Research and Design
http://dx.doi.org/10.1016/j.cherd.2017.01.008

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
CFD simulation of the hydrodynamics in an industrial scale cyclohexane

oxidation airlift loop reactor

Limeng Chen, Zhishan Bai*


State Environmental Protection Key Laboratory of Environmental Risk Assessment and Control on
Chemical Process, East China University of Science and Technology, Shanghai 200237, China

Graphical abstract

Highlights
 Circulation flow is influenced by the liquid circulation velocity and flux together
 Gas phase entrainment in the draft-tube is determined by the liquid circulation flux
 Gas holdup and liquid circulation flux increase with the increasing draft-tube area
 The horn-mouth can effectively enhance the gas-liquid separation at the reactor top
 Axial position of the draft-tube can influence the circulation impetus within limits
ABSTRACT

The two-fluid model was applied to simulate the hydrodynamics based on the process
conditions in a cyclohexane oxidation airlift loop reactor (ALR). A lab-scale ALR was first
investigated to validate the simulation methods and the simulated gas holdup and liquid
velocity agree well with the experimental data. Following the validation, computational
fluid dynamics (CFD) was then extended to study the industrial scale cyclohexane
oxidation ALR. The CFD results indicate that the key hydrodynamic parameters for the
circulation flow are the liquid circulation velocity and liquid circulation flux. The
increasing liquid circulation flux in the downcomer enhances the downcomer gas holdup,
while the increasing liquid circulation velocity in the riser reduces the riser gas holdup. The
factor influencing the gas phase entrainment in the draft-tube is the liquid circulation flux
rather than the liquid circulation velocity. With a constant reactor diameter, the gas holdup
and liquid circulation flux increases with the increasing cross-sectional area ratio of
downcomer to riser. The draft-tube with a horn-mouth can efficiently enhances the
gas-liquid separation at the reactor top. Within a certain range, the axial position height of
the draft-tube can influence the circulation driving force and resistance loss in the reactor.
The results of this study are significant for guiding production, the optimization and
scale-up study of this kind of reactors.

Keywords: Airlift loop reactor (ALR); Gas holdup; Liquid circulation velocity; Liquid
circulation flux; Draft-tube; Computational Fluid Dynamics (CFD)

Corresponding author. Tel.: +86 21 64253731.


E-mail address: baizs@ecust.edu.cn (Z. Bai).
Nomenclature
AD cross-sectional area of the downcomer (m2)
AR cross-sectional area of the riser (m2)
drag coefficient (dimensionless)
lift coefficient (dimensionless)
turbulence dispersion force coefficient (dimensionless)
constants for stand k-ε model (dimensionless)
, constants for stand k-ε model (dimensionless)
diameter of the orifices (m)
bubble diameter (m)
DE diameter of the draft-tube (m)
DF diameter of the horn-mouth (m)
DT diameter of the reactor (m)
⃗ interface forces (N)
drag force (N)
lift force (N)
turbulence dispersion force (N)
gravitational constant (m/s2)
HA height of the top clearance (m)
HB height between the draft-tube bottom and the baffle (m)
HE height of the draft-tub (m)
HF height of the horn-mouth (m)
̇ interfacial mass transfer (dimensionless)
turbulent kinetic energy per unit mass (m2/s3)
pressure (Pa)
QG gas inflow (m3/s)
QL liquid circulation flux (m3/s)
Reynolds number (dimensionless)
time (s)
⃗ velocity vector (m/s)
UG,R superficial gas velocity in the riser (m/s)
, liquid velocity in the downcomer (m/s)
VL liquid circulation velocity (m/s)
Greek letters
volume fraction (dimensionless)
density (kg/m3)
∇ gradient operator (dimensionless)
turbulent energy dissipation rate per unit mass (m2/s3)
molecular viscosity (Pa s)
kinematic viscosity (m2/s)
surface tension (N/m)
, turbulence model constants for k-ε equation (dimensionless)
⃗ shear stress (Pa)
Subscripts
b Bubble
D Downcomer
DF drag (force)
g (G) gas phase
k Phase
LF liquid phase
l (L) liquid phase
o Orifice
O Overall
R Riser
T Turbulence
TDF turbulence dispersion (force)
1. introduction

Cyclohexane oxidation reaction using air as oxidant is an important process in producing


cyclohexanone, hexanolactam and other chemicals. ALRs are adopted widely as the production
devices and they have the advantages of both stirred-tank reactors and bubble column reactors. This
kind of reactors have been widely applied in chemical industry, bioengineering and environmental
engineering due to their simple construction, good sealing and mixing, low energy consumption and
high capacity of liquid circulation (Behin, 2012; Bendjaballah et al., 1999; Chisti, 1989; Fu et al.,
2007; Hwang and Cheng, 1997; Jin et al., 2006; Luo and Al-Dahhan, 2008).
Generally, ALRs are divided into two types: internal airlift loop reactor (IALR) and external
airlift loop reactor (EALR). The IALR is split in two parts by a concentric draft-tube or a baffle
(Siegel and Robinson, 1992; Van Baten et al., 2003). The EALR consists of two separated columns
(a riser and a downcomer column) which are connected by horizontal connections at the top and
bottom of the reactor (Chisti et al., 1995; Mohanty et al., 2006). The reactor is characterized by four
distinct regions: the riser, the downcomer, the bottom clearance region and the top clearance region
(Luo et al., 2013) (that is the top gas-liquid separation region). A gas in the form of bubbles is
injected into the riser. Most of the bubbles are separated out from the top gas-liquid separation region
while only a few bubbles are entrained into the downcomer by the recirculation liquid phase. When
the liquid phase reaches the top of the reactor, a part of them flows out through the top overflow weir
while the other part gets into the draft-tube to participate circulation, thereby realizing the recycle
mixing between two-phase and gas-liquid reaction. Therefore, the draft-tube is one of the key
structures influencing the gas-liquid two-phase flow in the reactor.
The critical hydrodynamic parameters of cyclohexane oxidation ALR are the gas holdup, liquid
circulation velocity and liquid circulation flux. With a given air input, a higher liquid circulation
velocity and liquid circulation flux will produce a better substance flow and exchange, as well as a
more sufficient gas-liquid reaction and less side reaction (Behin, 2010). If the reactive material can
be updated and exchanged timely, there will be a higher selectivity for the cyclohexane oxidation
reaction, therefore, it will require a better circulation flow. In spite of simple structure, the factors
affecting the performance of reactors are enormous, including geometric parameters, physical
property parameters, industrial operating parameters and etc. In practical application, operating
conditions are the only remaining adjustable parameters once initial geometric structure parameters
are set and designed. Hence, in the design and optimization of the reactor, the effects of geometric
parameters, especially the parameters of internals, on hydrodynamics will usually be investigated
primarily.
In the last few decades, numerous experimental studies focusing on ALRs were conducted to
measure the key hydrodynamic parameters (Roy et al., 2006), and several empirical and
semi-empirical correlations were summarized. For instance, Chisti et al. (1988) proposed a
theoretical correlation based on the conservation of energy in 1988, which can predict the liquid
circulation velocity well only in a certain liquid volume (0.06-1.06m3). Besides, these studies are
mostly restricted to air-water system at room temperature and atmosphere pressure conditions, and
the correlations were obtained in a relatively specific reactor structure, fluid media and operational
conditions. However there exist obvious differences in reactor scale and operating parameters
between the lab-scale reactor and the reactor applied in the industry. Therefore, it is very risky and
unreasonable to extrapolate these results from the lab-scale reactor because of the distinctly limited
experiment conditions. Koide et al. (1983) studied the effects of the diameter ratio of the draft-tube
to the reactor on the hydrodynamic parameters with air-water system in a central air-inlet reactor, and
found that, in the range of DE/DT=0.5-0.75,an increase in the diameter ratio would lead to a decrease
in the mass transfer coefficient between gas and liquid but the overall gas holdup is essentially
unchanged. In the meantime, related correlations were proposed based on the research, but these
correlations can only be used when the diameter of the reactor is in the range of DT=0.1m-0.3m.
Hwang and Fan (1986) also reported the effects of the diameter ratio DE/DT on the overall gas holdup,
but a different result, namely the overall gas holdup decreased with the increase of DE/DT, was
obtained. The influences of fluid media on the gas-liquid flow have been extensively studied
(Al-Masry, 2004; Buchholz et al., 1978; Godbole et al., 1982; Schumpe and Deckwer, 1982).
According to the experiments, Wachi et al. (1991) observed that the physical property parameters (i.e.
viscosity, density and surface tension) of fluid media had significant effects on gas holdup and fluid
flow.
But, such experimental researches require much capital and time to build laboratory equipment,
especially for the study of the amelioration of reactor internals. Besides, it’s impossible to carry out
the experiments in an industrial scale reactor. And, it’s also difficult to gain the details of the flow
fields in view of the complex flow in the reactor and the limitations of experimental measures.
Consequently, as an effective tool, CFD has been widely applied in analysing and optimizing
simulation results.
Using CFD, Šimčík et al. (2011) discussed the effects of the draft-tube diameter and height on
gas holdup and liquid circulation velocity with a 40L IALR, and compared the simulation results
with the experiment data. It was observed that CFD simulation can obtain satisfying results in a
given range of superficial gas velocity. Ebrahimifakhar et al. (2011) found that the distance between
the draft-tube and the column had an important influence on the hydrodynamics in the reactor. When
the distance increased, the gas holdup and liquid circulation velocity in the riser increased.
Analogously, the optimal distance between the draft-tube and the reactor wall was detected by
Hekmat et al. (2010) using air-water system, which was most favourable for the mixing and mass
transfer between the two reacting phases. Kiattichai et al. (2016) studied the effect of downcomer
diameter (AR/AD) on the gas holdup and liquid velocity with a constant AR in a central air-inlet IALR
and showed that the gas holdup in both the riser and downcomer decreases with increasing
downcomer diameter which also leads the driving force to increase between the riser and
downcomer.
However, despite the fact that much research work has been performed, the scale-up study on
such a reactor (cyclohexane oxidation ILAR) system is barely reported in the literature and it’s not
also fully understood that how the reactor internals influence the details of two-phase flow.
Especially for the study of the draft-tube, most of the documents merely analyzed the effects of
draft-tube diameter on hydrodynamics in a lab-scale ALR and different researchers drew different or
even contrary conclusions. In this work, the effects of the geometric parameters of the draft-tube
including the cross-sectional area ratio of downcomer to riser (AD/AR), the diameter ratio of
horn-mouth to draft-tube(DF/DE), the height ratio of bottom circulation region to top gas-liquid
separate region(HB/HA), on hydrodynamic behavior were investigated in an industrial large-scale
cyclohexane oxidation ALR. Due to the lack of the relevant industrialized experiment and research
of the reacting dynamics, the CFD simulation was conducted based on the actual
technologic condition and industrial scale reactor. The simulation results can provide useful guiding
suggestions for the structure modification and optimization of the reactors. To validate
the accuracy and reliability of the simulation results, a lab-scale ALR was first investigated and the
simulated gas holdup and liquid velocity were compared with experimental data when different
interfacial momentum force models were considered, thus the relatively accurate and validated
models and methods were established.

2. Numerical models

As an effectively numerical computation tool, CFD has been universally employed because of
the advantages in saving time and expense compared with the traditional experimental researches. In
spite of the possible differences between the predicted results and real results due to some
simplifications and assumptions, they can still match well to some extent if appropriate numerical
models are chosen.
The numerical simulation of gas-liquid flow in an ALR is established based on two-fluid model.
The Euler-Euler two-fluid model is applied extensively in the research on gas-liquid two-phase flow,
especially in the practical engineering. In Euler-Euler approach, two phases are regarded as
interpenetrating and coexisting continua of fluids, the motion of each phase is governed by the
respective conservation equations. The Euler-Euler method on a two-fluid model was applied in the
present study.
In the ALR, a large number of bubbles collide with each other, which will make the breakage
and coalescence of the bubbles happen easily, especially in a high superficial gas velocity. A
population balance approach has been widely used by many researchers (Lehr et al., 2002; Wang et
al., 2005, 2003). Although it was reported that a more accurate flow field will be obtained by taking
the breakage and coalescence into consideration, the time consumed is considerably high and it’s
also hard to get convergent results, which limits its usage (Chen et al., 2005; Sanyal et al., 2005).
Thus, an averaged bubble size was used in the simulation. This simplification could be acceptable as
long as the effect of different bubble sizes due to different designs and operating conditions can be
reasonably considered. In consequence, the two-fluid model which is on the basis of the Euler–Euler
method without considering the breakage and coalescence model is extensively adopted, and it’s
adequately accurate to investigate the hydrodynamics in ALR (Luo and Al-Dahhan, 2011).
The reliability and accuracy of the simulated results in two-fluid model is mainly adjusted by
the interfacial momentum force models (e.g. drag force model, lift force model and so on). These
closure sub-models have a significant influence on gas-liquid two-phase flow fields. It’s
one of the most controversial issues on how to use these interfacial momentum force models,
therefore the comparison between simulated results and experimental data is very significant to
determine the appropriate models. If the appropriate models are chosen, the simulated results will
match well with the experimental data. Only in this way can the established models and methods be
applied and extrapolated to conduct the scale-up study of the reactor, and this is also the foundation
of this study.

2.1 Two-fluid model and basic governing equations

In two-fluid model, each phase must solve its own continuity equations and momentum
equations, and the coupling between two phases is conducted by pressure and interphase exchange
coefficients.
The two-fluid model established in this paper is based on following assumptions and
simplifications:
1) Both phases are regarded as interpenetrating and coexisting continua, and phase holdup is
expressed by

+ =1 (1)

where and are the volume fraction of gas and liquid respectively.

2) Gas-liquid two-phase flow in the reactor is treated as viscous incompressible flow and the mass
exchange source term, the heat exchange and reaction between two phases are neglected.
3) The breakage and coalescence of bubbles are neglected and the gas phase is treated as spherical
bubbles with uniform size, thus the average diameter is calculated approximately by the
following equation (Huang et al., 2010).
/
= 2.9

(2)
where is the diameter of the orifices of gas distributor.
4) The effect of the interphase forces including drag force, lift force and turbulent dispersion force
on hydrodynamics will be discussed, while the virtual mass force is not discussed. Sokolichin et
al. (2004) reported that the virtual mass force was not the main interphase forces. It had no
significance on simulation results and a negative effect on convergence was detected.
Based on the above assumptions, the governing equations can be described as follows:
Continuity equation:

( )+∇∙( ⃗ )= ̇ (3)

Momentum equation:

( ⃗ )+∇∙( ⃗ ⃗ )=− ∇ + ⃗+∇∙ ⃗ + ⃗ (4)

Where subscript k = l, g, ̇ stands for interfacial mass transfer, according to the previous

assumptions, ̇ =0. ⃗ denotes the momentum transfer between continuous phase and dispersion

phase, which is embodied by the interphase forces.


The two-equation turbulence model was adopted by almost all of researches for the numerical
simulation in the ALR and good results were obtained. The stand k-ε turbulent model is:
,
( ) + ∇( ⃗)=∇ ∇ + − (5)

,
( ) + ∇( ⃗)=∇ ∇ + ( − ) (6)

In the above two equations, the default constants are = 0.09, = 1.00 , = 1.3, =

1.44, = 1.92.

2.2 Interfacial momentum force models

Apart from the pressure term, it is generally believed that the contribution to interphase
momentum transfer in gas-liquid two-phase flow comes mainly from the interaction forces between
two phases, such as drag force, lift force, virtual mass force, turbulent dispersion force, etc. Here, the
virtual mass force is neglected. Similar approach (i.e. only the drag force and lift force were
considered in simulation) was used by other authors (Krishna et al., 1999; Mudde and Van Den
Akker, 2001) and the numerical results and experimental data have been well matched. Oey et al.
(2003) showed that it was accurate enough to obtain the global hydrodynamic parameters with the
drag force considered alone, it is usually the only one involved in gas-liquid flow simulations, e.g.,
Van Baten et al. (2003) and Huang et al. (2007). It was also reported that the lift force can influence
the local gas holdup (Talvy et al., 2007). But, it was reported by Song et al. (2015) that there was
obvious difference when the two simulated results were compared, especially in the distribution of
hydrodynamic parameters, in which one took only the drag force into consideration, the other one
considered the drag force as well as the lift force, turbulent dispersion force, etc. The latter matched
well with experiment data. In this paper, the effects of these interphase forces on gas-liquid
two-phase flow in the reactor were discussed.

Drag force model

Drag force is produced by the liquid around the gas and will exert resistance on rising bubbles
along the opposite direction. In gas-liquid two-phase flow, the drag force is considered as the most
important factor among all interfacial forces. It is calculated by:
= ⃗ − ⃗ ⃗ − ⃗ (7)

here is the bubble diameter and its calculation formula has been described previously.
stands for the drag coefficient which depends on bubble diameter, fluid characters, etc. Although
there are many available correlations for calculating the drag coefficient (Cockx et al., 1997;
Karamanev and Nikolov, 1992; Mohajerani et al., 2012), they are hardly contradictory.
In the present work, the drag coefficient was calculated by Schiller and Naumann (1935) model:
. .
= ≫
(8)
.

the Reynolds number is given by :


⃗ ⃗
= (9)

Lift force model

Lift force describes the interaction between bubbles and liquid shearing field. It derives from
the unsymmetrical pressure distribution on bubble surface which leads to the radial motion of
bubbles. The lift force has an important influence on two-phase distribution, especially on gas-liquid
two-phase flow where large shearing force exits, while the lift force plays a minor role in
liquid-liquid or liquid-solid two-phase. The lift force is calculated by:

= ⃗ − ⃗ × (∇ × ⃗ ) (10)

in the above equation, is the lift coefficient which is a constant and different values were
suggested by different authors (Dhotre and Smith, 2007; Drew and Lahey, 1987; Grienberger and
Hofmann, 1992).The most common lift coefficient value used by several researchers is 0.5 (Buwa
and Ranade, 2002; Delnoij et al., 1997; Zhang et al., 2006).

Turbulent dispersion force model

Turbulent dispersion force considers the disperse effect of the disperse phase in the continuous
phase and it is a consequence of the turbulent fluctuation and interface drag force. The turbulent
dispersion force is calculated by:
, ∇ ∇
=− ( − ) (11)
,

where is the turbulent dispersion force coefficient and it’s usually a constant which is in the
range 0.1-0.5 (Luo and Al-Dahhan, 2011), is the drag force coefficient.
In this work, a commercially available CFD code ANSYS FLUENT 14.5 was used to solve the
equations of the models described above. In the simulation, the boundary conditions of the inlet and
outlet were velocity inlet and pressure outlet respectively. All the walls were the stand non-slip
boundary condition. The reactor was full of liquid phase before simulation.

2.3 Model validation

The selection of the models has a great influence on the simulation results. Therefore, it’s
essential to validate the accuracy and reliability of the simulation results by experiments before
conducting the scale-up study of the reactor. In this study, a lab-scale ALR is chosen as a cold model
experiment platform, and simple experiments were carried out to measure the gas holdup and liquid
velocity. The overall gas holdup and liquid velocity in the downcomer were measured by volume
expansion method and trace technique, respectively. Fig.1 displays the schematic diagram of the
laboratory ALR, which is composed of an external column with a height of 1900mm and a diameter
of 377mm constructed by Plexiglas, and a center draft tube with a height of 1200mm and a diameter
of 150mm constructed by stainless steel. The O-ring gas distributor in which there are fifty-five
2-mm-diameter orifices is located at the bottom of the riser. The airflow rate varies in the range of
QG=2.0-6.0m3/h, namely the superficial gas velocity in the riser is in the range of
UG,R=5.9×10-3-17.9×10-3 m/s. The continuous operation and intermittent operation are adopted for the
gas phase and liquid phase, respectively. To be consistent with the cold model experiment, air was
treated as the gas phase and tap water was treated as the liquid phase in the CFD simulation. All
experiments were carried out at room temperature and atmospheric pressure conditions.
Fig. 1 - Schematic diagram for the laboratory ALR.
Fig.2 illustrates the comparisons of the simulated gas holdup and liquid velocity with the
experimental data when different interfacial momentum force models including drag force model, lift
force model and turbulent dispersion force model are considered. It can be seen that the interfacial
momentum force models have a great effect on the gas holdup and liquid velocity. An obvious
deviation is observed without considering the lift model and turbulent dispersion force model. The
simulation results are most approximate to the experiment data when the three sub-models
mentioned above are involved together. Therefore it’s not adequate to obtain the reliable results with
the drag force considered alone in the present study. The simulated gas holdup and liquid velocity are
a little lower than the experimental data, especially for high superficial gas velocity. The reason
might be that the bubble size distribution is wilder due to the strong turbulence, and the high
coalescence and breakage rate in the reactor in a higher superficial gas velocity, thus the method of
average bubbles diameter seems not to be enough to obtain the completely accurate simulation
results. Also, it perhaps indicates that the interfacial momentum forces models need to be considered
more fully due to the more instability in a higher superficial gas velocity.
(a) 4.0
D
3.5
D+L
D+TDF
D+L+TDF
3.0
Exp

2.5

G,O (%) 2.0

1.5

1.0

0.5
5 6 7 8 9 10 11 12 13 14 15 16 17 18 19
3
UG,R*10 (m/s)

(b) 0.50

D
D+L
0.45
D+TDF
D+L+TDF
0.40 Exp
vL,D (m/s)

0.35

0.30

0.25

0.20
5 6 7 8 9 10 11 12 13 14 15 16 17 18 19
3
UG,R*10 (m/s)

Fig. 2 - Comparison between simulation results with different interfacial momentum force models
considered and experimental data for: (a) the overall gas holdup; (b) liquid velocity in the downcomer.
(D: Drag Force Model; L: Lift Force Model; TDF: Turbulent Dispersion Force Model)
Table 1 - Physical properties of air, water and cyclohexane used in the CFD simulation.
density, viscosity, surface tension,
(kg/m³) (Pa·s) (N/m)
air 1.225 1.789×10-5 —
water 998 1×10-3 0.073
cyclohexane 620 1.758×10-4 0.0074

3. Results and discussion

Fig.3 exhibits the geometry configuration of the cyclohexane oxidation ALR applied in the
industry. Air gets into the reactor from the O-ring gas distributor where there are 1057
2-mm-diameter orifices. The draft-tube with a horn-mouth is located in the center of reactor in order
to achieve liquid circulation flow. The horn-mouth can contribute to the gas-liquid separation at the
top clearance and reduce the gas phase entrainment. The height and diameter of the reactor are
HT=8.8m and DT=2.6m, respectively. The non-dimensional parameters which vary with the
draft-tube diameter DE, the horn-mouth diameter DF and the draft-tube position height HB (or HA) are
in the range of AD/AR=0.0385-0.2198 (AD and AR are the area of the downcomer and the riser,
respectively), DF/DE=1-2.5 and HB/HA =0.0385-1.7. The draft-tube height (HE=5.9m) and the
horn-mouth height (HF=1.1m) are constant.

Fig. 3 - Geometry configuration of the cyclohexane oxidation ALR applied in the industry.
There are obvious distinctions in operating conditions, operating modes, physical parameters of
fluid media and so on between the cyclohexane oxidation ALR applied in the industry and the
laboratory ALR. In the industrial reactor, the operating pressure is about 1900kpa and the operating
temperature is 165-175℃, and continuous operation is applied for both phases (air and cyclohexane).
In spite of the differences mentioned above, there are no intrinsic distinctions in phase interactions
and flow regularity for the gas-liquid two-phase flow in the reactor. Therefore the models and
simulation methods validated before can also be applied to the industrial reactor. The numerical
simulation has been carried out with the full-scale geometry structure, and one half of the geometry
model was chosen as the computational domain in consideration of the symmetry of the model. All
simulations were conducted at a constant airflow QG=1970Nm³/h, corresponding to a superficial gas
velocity UG,R=0.0155m/s in the riser. According to the lab-scale study, the relative error between
experiment data and simulation results is 9-10% for the gas holdup and 5.8-6% for the liquid velocity
when the superficial gas velocity is in the range of UG,R=0.015-0.018m/s. And the higher the
superficial gas velocity, the greater relative error. The accuracy and relative error of scale-up results
are approximately corresponding in a same superficial gas velocity (UG,R=0.0155m/s).
3.1 Sensitivity of the hydrodynamics behavior to bubble diameter

According to the equation (2), the average bubble diameter was approximately equal to 4.5mm.
In order to consider the sensitivity of the hydrodynamics behavior to bubble diameter in the reactor,
three sizes of the bubble diameter (3 mm,4mm, 5mm) were used here. The radial distribution of the
gas holdup and liquid circulation velocity were studied compared with different bubble diameter.
Fig.4 displays the study of the sensitivity to bubble diameter. It can be seen that the smaller bubble
diameter will lead the higher gas holdup and smaller liquid circulation velocity. Nevertheless, the
bubble diameter has a slight effect on the gas holdup and liquid circulation velocity in the range of
db=3-5mm, similar study was reported by Kiattichai et al. (2016). Therefore the bubble diameter of
4mm was used in present work.
(a) 0.050
db=3mm
0.045
db=4mm
0.040 db=5mm

0.035
Draft-tube
0.030 wall
G (-)

0.025

0.020

0.015

0.010

0.005
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radial (m)

(b) 0.75

0.60 db=3mm
db=4mm
0.45
db=5mm
0.30

0.15

0.00
VL (m/s)

-0.15
Draft-tube
-0.30
wall
-0.45

-0.60

-0.75

-0.90

-1.05
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)

Fig. 4 - Radial distribution of (a) gas holdup and (b) liquid circulation velocity.
3.2 Effects of the cross-sectional area ratio of downcomer to riser on the hydrodynamic
behavior in the reactor

As previously stated,with a constant reactor diameter DT, CFD simulations were carried out to
study the influences of the cross-sectional area ratio of downcomer to riser AD/AR on the gas holdup,
liquid circulation velocity and liquid circulation flux in the reactor. It can be seen clearly in Fig.5, the
gas holdup increases with the increasing cross-sectional area ratio AD/AR. But the riser gas holdup
and overall gas holdup increase slowly while the downcomer gas holdup increase rapidly. For
example, when AD/AR varies from 0.1051 to 0.2198, the riser gas holdup and the overall gas holdup
increase by 6% and 8%, respectively, while the downmcomer gas holdup increases by more than
67%. An increase in AD/AR means a larger circulation area in the downcomer and a smaller
circulation area in the riser. Therefore the downcomer liquid circulation velocity reduces and the riser
liquid circulation velocity improves as shown in Fig.6. Although the liquid circulation velocity in the
downcomer reduces, the downcomer liquid circulation flux still improves due to the increase in the
downcomer circulation area (see Fig.6).
4.5
G,R in the riser
4.0
G,O in the overall reactor
3.5 G,D in the downcomer

3.0

2.5
G (%)

2.0

1.5

1.0

0.5

0.0
0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24
AD/AR (-)

Fig. 5 - Effect of the cross-sectional area ratio AD/AR on the average gas holdup.
Some studies (Baten et al., 2003; Gouveia et al., 2003; Rujiruttanakul and Pavasant, 2011) have
presented that a larger liquid circulation velocity will cause a shorter gas resident time when the
draft-tube diameter varies, thus the gas holdup will decrease. In this paper, it can be seen that the
riser gas holdup doesn’t reduce but increases slightly in spite of the augment of the riser liquid
circulation velocity. This is because, with a given air inflow, the superficial gas velocity increases as
the cross section area declines in the riser. It’s also found that a decrease in downcomer liquid
circulation velocity doesn’t reduce the downcomer gas holdup, which is different from the analyses
stated in some literatures (Ghasemi and Hosseini, 2012; Tian et al., 2013; Kiattichai et al., 2016). It
could be due to that the augmentation of the liquid circulation flux in the draft-tube makes more gas
phase entrained into the downcomer. All these indicate that the circulation flow in the reactor
depends on not only the liquid circulation velocity but also the liquid circulation flux. The main
influential factor for gas phase entrainment isn’t the liquid circulation velocity but the liquid
circulation flux in the downcomer. In fact,it sometimes pays more attention to the liquid circulation
flux rather than the liquid circulation velocity in the industry production process.
1.6 0.60
VL,D in the downcomer
1.4 VL,R in the riser 0.55

QL,D in the downcomer


1.2 0.50

1.0 0.45

QL (m /s)
VL (m/s)

0.8 0.40

3
0.6 0.35

0.4 0.30

0.2 0.25

0.0 0.20
0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24

AD/AR (-)

Fig. 6 - Effect of the cross-sectional area ratio AD/AR on the average liquid circulation velocity and liquid
circulation flux.
4.2

3.9 Ⅰ Ⅱ
3.6

3.3

3.0
G,R (%)

2.7
AD/AR=0.0385
2.4 AD/AR=0.0670

2.1
AD/AR=0.1051
AD/AR=0.1511
1.8 AD/AR=0.2198
1.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0
Axial height (m)

Fig. 7 - Axial distribution of cross-sectional average gas holdup in the riser above the gas distributor.
In addition, the axial distribution of cross-sectional average gas holdup in the riser above the gas
distributor (see Fig.7) and the radial distribution of gas holdup at different axial height in the reactor
(see Fig.8) were also investigated, because they are also important parameters for good mixing and
mass transfer. Fig.7 indicates that the cross-sectional area ratio AD/AR has an obvious effect on the
axial distribution of cross-sectional average gas holdup. The comparisons between different AD/AR
reveal that a larger cross-sectional area in the downcomer will produce a more level curve of the
axial distribution of gas holdup, thus the gas phase is distributed more uniformly along the axial
direction. It is also discovered that, the gas holdup declines dramatically along the axial height in
regionⅠ(gas-distributor action region) and it varies from the maximum to the minimum. As the axial
position continues to rise, the gas holdup gradually increases and tends towards stability at last in
region Ⅱ. Above-mentioned phenomenon can be explained by Fig.8 (two different values for AD/AR
are taken as examples). The radial distribution of gas holdup has similar varying pattern for different
AD/AR compared Fig.8 (a) and Fig.8 (b). There is a higher gas holdup near the gas distributor (Y=1)
than anywhere else above the gas distributor. This is also why the cross-sectional average gas holdup
decreases in regionⅠ. However, the gas distribution along the radial direction will be more uniform
with the increase of the axial position height, thus the cross-sectional average gas holdup would
increase and tend to be stable gradually in region Ⅱ. The bias flow of the gas which is induced by
the liquid flow field at the bottom of the draft-tube can be found in the reactor. Much of the gas
gathers near the draft-tube wall, which causes a lower gas holdup near the reactor wall. Therefore,
the distribution of the gas along the radial direction is inhomogeneous, especially in the vicinity of
the gas distributor. The difference lies in that the distribution of the gas along the radial and axial
direction is relatively uniform in the draft-tube although the gas holdup is slightly low at a lower
axial position height due to the upward buoyant force exerted on the bubbles. But sometimes, a
higher gas holdup in the draft-tube will also have a negative effect on cyclohexane oxidation reaction
because of the existence of over oxidation and side reaction in the production process. For example,
in order to improve the selectivity of the reactor, it will need to decrease the residence time of the
reaction products and the quantity of the gas entrained into the downcomer.
(a) 0.07
Y=1m
0.06
Y=2m
Y=3m
Y=4m
0.05 Y=5m
Draft-tube
wall
0.04
G (-)

0.03

0.02

0.01

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(b) 0.07
Y=1m
0.06
Y=2m
Y=3m
Y=4m
0.05 Y=5m
Draft-tube wall
0.04
G (-)

0.03

0.02

0.01

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)

Fig. 8 - Radial distribution of gas holdup at different axial height in the reactor:
(a) AD/AR=0.0385; (b) AD/AR=0.1051. (Y=0 is in the bottom of the gas distributor)

3.3 Effects of the diameter ratio of horn-mouth to draft-tube on the hydrodynamic behavior in
the reactor

In order to reduce the quantity of the gas entrained into the downcomer, the draft-tube with a
horn-mouth which can enhance gas-liquid separation at the top clearance is necessary. With a given
draft-tube diameter ( DE=1.4m),the diameter ratio DF/DE varies with the horn-mouth diameter DF.

Fig. 9 - The contours of the air volume fraction at the top of the reactor for different DF/DE:
(a) DF/DE=1; (b) DF/DE=1.375; (c) DF/DE=1.75; (d) DF/DE=2.125; (e) DF/DE=2.5.
Fig.9 displays the contours of the air volume fraction at the top of the reactor for different
DF/DE. It’s obvious that the horn-mouth can enhance the gas-liquid separation effectively at the top
gas-liquid separation region and the separation capability improves with the increase of the diameter
ratio DF/DE. It also can be seen that there is a highest local gas holdup at the turning. The gas holdup
is higher in the draft-tube wall than in the reactor wall. The reason lies in the high turbulence
intensity here. Likewise, as shown in Fig.10 and Fig.11, the gas holdup also hits the peak at the gas
distributor because the turbulent kinetic energy is strong here, and the distribution of gas holdup
along the axial and radial direction is also heterogeneous especially near the gas distributor. As
shown in Fig.10, the axial distribution curve of cross-sectional average gas holdup is similar
compared with different DF/DE, namely the gas holdup profiles have the same curvature. Like Fig.7,
firstly the cross-sectional average gas holdup decreases sharply, and then increases gradually with the
increasing axial height (see Fig.10). Besides, as illustrated in Fig.11, it is also obvious that the
distribution of gas holdup along the radial direction in the riser becomes more homogeneous at a
higher position than at a lower position due to the effect of turbulent dispersion. While the gas
holdup in the draft-tube is lower at a lower position, which further demonstrates that the downcomer
gas is difficult to arrive at the bottom due to the upward buoyant force.
4.2

3.9

3.6

3.3

3.0
G,R (%)

2.7
DF/DE=1.000
2.4
DF/DE=1.375
2.1 DF/DE=1.750
DF/DE=2.215
1.8
DF/DE=2.500
1.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0
Axial height (m)

Fig. 10 - Axial distribution of cross-sectional average gas holdup in the riser above the gas distributor.
(a) 0.07
DF/DE=1.000
0.06
DF/DE=1.375
DF/DE=1.750
0.05 DF/DE=2.215
DF/DE=2.500
0.04

G (-)
0.03
Draft-tube
wall
0.02

0.01

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(b) 0.050

0.045

0.040

0.035

0.030
Draft-tube
G (-)

0.025 wall
0.020
DF/DE=1.000
0.015 DF/DE=1.375
DF/DE=1.750
0.010
DF/DE=2.215
0.005
DF/DE=2.500
0.000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(c) 0.045

0.040

0.035

0.030

0.025
Draft-tube wall
G (-)

0.020
DF/DE=1.000
0.015
DF/DE=1.375
0.010 DF/DE=1.750
DF/DE=2.215
0.005
DF/DE=2.500
0.000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)

Fig. 11 - Radial distribution of gas holdup for different DF/DE at different axial height above the gas
distributor: (a) Y=1m; (b) Y=3m; (c) Y=5m. (Y=0 is in the bottom of the gas distributor)
On the other hand, Fig.11 also shows that the diameter ratio affects the riser gas holdup and the
downcomer gas holdup significantly. The gas holdup inside and outside of the draft-tube tends to
decease with the increasing diameter ratio DF/DE, which will lessen the overall gas holdup in the
reactor without a doubt. However, the downcomer gas holdup declines faster than the gas holdup in
the riser. For instance,the gas holdup in the riser and downcomer decrease by 20% and 93%
respectively when the diameter ratio DF/DE varies from 1 to 2.5. As a consequence of this, the
driving force ( , − , ) required for the liquid circulation and generated by the density difference
of the mixture phase between the riser and the downcomer doesn’t reduce, but enhances. With other
conditions unchanged, the liquid circulation velocity and liquid circulation flow in the reactor will
increase as well (see Fig.12). All these fully prove that the draft-tube with a horn-mouth can play a
key role in promoting liquid circulation flow. But a high liquid circulation velocity, in turn, also has
an adverse effect on gas holdup. The higher liquid circulation velocity is, the shorter the gas
residence time becomes, therefore the gas holdup would also decrease. This, of course, explains why
the gas holdup decreases when the diameter ratio DF/DE increases as well. The decreased gas holdup
has a bad effect on the cyclohexane oxidation reaction if the conversion ratio of the cyclohexane is
low.
Fig.12 also shows that the liquid circulation velocity inside of the draft-tube is more
homogeneous than that outside of the draft-tube, especially near the gas distributor (Y=1m) where
the vortex and high liquid turbulence exists. This is also the reason why the gas distribution along the
radial direction is more uniform in the downcomer. Besides, it can be noticed that the liquid
circulation velocity close to the draft-tube wall reaches the maximum, and the velocity close to the
reactor wall reaches the minimum in the riser. Note that the minimum is a negative, which indicates
the existence of local liquid backflow and back mixing in the riser. This part of liquid couldn’t
accomplish the circulation flow. The local backflow and back mixing occur mainly at the bottom of
the reactor and it’s not obviously observed in the upper regions of the reactor. The backflow also
indicates the existence of large eddy in the riser which is disadvantageous to mixing and mass
transfer. Furthermore, because of the non-slip boundary condition, the velocity is zero on the wall.
(a) 0.9

0.6

0.3

0.0

VL (m/s)
-0.3
Draft-tube
-0.6 wall D F/D E=1.000
D F/D E=1.375
-0.9
D F/D E=1.750
D F/D E=2.215
-1.2
D F/D E=2.500
-1.5
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(b) 0.9

0.6

0.3

0.0
VL (m/s)

-0.3
Draft-tube
DF/DE=1.000
-0.6 wall
DF/DE=1.375
-0.9 DF/DE=1.750
DF/DE=2.215
-1.2
DF/DE=2.500
-1.5
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(c) 0.9

0.6

0.3

0.0
VL (m/s)

-0.3
Draft-tube
-0.6 wall DF/DE=1.000
DF/DE=1.375
-0.9 DF/DE=1.750
DF/DE=2.215
-1.2
DF/DE=2.500
-1.5
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)

Fig. 12 - Radial distribution of liquid circulation velocity for different DF/DE at different axial height above
the gas distributor: (a) Y=1m; (b) Y=3m; (c) Y=5m. (Y=0 is in the bottom of the gas distributor)
3.4 Effects of the height ratio of bottom circulation region to top gas-liquid separation region
on the hydrodynamic behavior in the reactor

Under a certain draft-tube structure and size, the height ratio of bottom circulation region to top
gas-liquid separation region HB/HA is varied by merely adjusting the axial position height of the
draft-tube. The results show that the height ratio HB/HA affects the hydrodynamic behavior in the
reactor significantly within limits. The effects of HB/HA on the average gas holdup, average liquid
circulation velocity and liquid circulation flux with different HB/HA are illustrated in Figs.13-14.
With the increasing HB/HA, the riser gas holdup reduces slightly at first and then tends to be stable,
while the gas holdup in the draft-tube increases dramatically at first and then tends towards stability.
But the downcomer gas holdup varies faster than that in the riser, thus the overall gas holdup in the
reactor is inclined to augment at first and then keeps unchanged as well. With a given air input, the
gas holdup is mainly influenced by the liquid flow field. The increase of the liquid circulation
velocity in the riser accelerates the ascending motion of the gas and reduces gas residence time, thus
the riser gas holdup decreases. In contrast, more gas is entrained into the draft-tube due to the
increasing liquid circulation flux in the downocmer. However, when the liquid velocity and liquid
circulation flux approximately tend to be stable, the gas holdup no longer changes as well.
3.6

3.3

3.0

2.7

2.4

2.1
G (%)

1.8

1.5

1.2

0.9 G,R in the riser


0.6 G,O in the overall reactor

0.3 G,D in the downcomer

0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
HB/HA (-)

Fig. 13 - Effect of the height ratio HB/HA on the average gas holdup.
1.0 0.8

0.9
0.7
0.8
0.6
0.7
0.5
0.6
VL (m/s)

QL (m3/s)
0.5 0.4
VL,D in the downcomer
0.4
0.3
VL,R in the riser
0.3
QL,D in the downcomer 0.2
0.2
0.1
0.1

0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
HB/HA (-)

Fig. 14 - Effect of the height ratio HB/HA on the average liquid circulation velocity and liquid circulation
flux.
As shown in Fig.14, the liquid circulation velocity and liquid circulation flux increase with the
increment of height ratio HB/HA in a certain range. When HB/HA varies from 0.42 to 1.08, the liquid
circulation velocity and liquid circulation flux have little change. At this point, a slight decline will
occur in the liquid circulation velocity if continuing to increase HB/HA. This is because as the axial
position height of the draft-tube increases, the space of flow enlarges at the bottom of draft-tube, the
liquid flow resistance reduces, thus the liquid circulation velocity and liquid circulation flux increase.
But, on the other hand, an increase in the height of bottom circulation region (HB) will also lead to
the decrease of the top gas-liquid separation region (HA), which can reduces the driving force of
liquid circulation flow at the top of the reactor. Therefore the total circulation impetus tends to
stabilize instead of increasing continuously in the range of HB/HA= 0.42-1.08. It is also possible that
the total circulation impetus may lower if the HB/HA continues to increase.
Figs.15 and 16 display the distribution of gas holdup and liquid circulation velocity. Compared
with different HB/HA, the gas holdup and liquid circulation velocity in the riser vary obviously just in
the vicinity of the gas distributor (Y=1m). It nearly has no influence on the riser gas holdup and
liquid circulation velocity in the upper region (Y=3m and Y=5m). However, the influence is always
observably in any region of the draft-tube (Y=1, 3, 5m).
(a) 0.060
HB/HA=0.0385
0.055
HB/HA=0.0800
0.050
HB/HA=0.2273
0.045 HB/HA=0.4211
0.040 HB/HA=0.6875
0.035

G (-) 0.030

0.025 Draft-tube
0.020
wall

0.015

0.010

0.005

0.000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(b) 0.050

0.045

0.040

0.035

0.030
G (-)

0.025
Draft-tube
0.020 wall
HB/HA=0.0385
0.015
HB/HA=0.0800
0.010 HB/HA=0.2273
HB/HA=0.4211
0.005
HB/HA=0.6875
0.000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(c) 0.045

0.040

0.035

0.030

0.025 Draft-tube
G (-)

wall
0.020
HB/HA=0.0385
0.015
HB/HA=0.0800
0.010 HB/HA=0.2273
HB/HA=0.4211
0.005
HB/HA=0.6875
0.000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)

Fig. 15 - Radial distribution of gas holdup for different HB/HA at different axial height above the gas
distributor: (a) Y=1m; (b) Y=3m; (c) Y=5m. (Y=0 is in the bottom of the gas distributor)
(a) 0.8

0.6

0.4

0.2

0.0

VL (m/s) -0.2

-0.4
HB/HA=0.0385
Draft-tube
-0.6 HB/HA=0.0800
wall
-0.8 HB/HA=0.2273
HB/HA=0.4211
-1.0
HB/HA=0.6875
-1.2
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(b) 0.8

0.6

0.4

0.2

0.0
VL (m/s)

-0.2

-0.4
Draft-tube HB/HA=0.0385
-0.6 wall HB/HA=0.0800
-0.8
HB/HA=0.2273
HB/HA=0.4211
-1.0
HB/HA=0.6875
-1.2
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)
(c) 0.8

0.6

0.4

0.2

0.0
VL (m/s)

-0.2

-0.4
HB/HA=0.0385
Draft-tube
-0.6 HB/HA=0.0800
wall
-0.8 HB/HA=0.2273
HB/HA=0.4211
-1.0
HB/HA=0.6875
-1.2
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
Radius (m)

Fig. 16 - Radial distribution of liquid circulation velocity for different HB/HA at different axial height
above the gas distributor: (a) Y=1m; (b) Y=3m; (c) Y=5m. (Y=0 is in the bottom of the gas distributor)
4. Conclusion

In the present work, the likely influences of the draft-tube structure on the hydrodynamic
parameters such as average gas holdup, average liquid circulation velocity and liquid circulation flux,
and the detailed distribution of gas holdup and liquid circulation velocity in an industrial scale
cyclohexane oxidation ALR were investigated theoretically by applying CFD. The accuracy and
feasibility of the simulation results was validated by the experiment study. The main findings are
summarized as follows:
(1) CFD numerical simulation can reliably predict the gas-liquid two-phase flow performances in
the cyclohexane oxidation ALR. The geometry parameters of the draft-tube with a horn-mouth
have important influences on the hydrodynamic behavior in the reactor. Both the liquid
circulation velocity and the liquid circulation flux have important effects on the circulation flow.
(2) With a constant reactor diameter DT, the gas holdup in the reactor increases with the increasing
cross-sectional area ratio AD/AR. Also, when AD/AR increases, the riser liquid circulation velocity
improves and the downomcer liquid circulation velocity reduces. However, the total liquid
circulation flux and the entrainment of gas phase in the downcomer are still increasing. The
factor influencing the gas phase entrainment is the liquid circulation flux in the downcomer
rather than the liquid circulation velocity.
(3) The horn-mouth can effectively enhance the gas-liquid separation at the top gas-liquid separation
region. With a given draft-tube diameter DE=1.4m, when the diameter ratio DF/DE increases,
the capacity of gas-liquid separation increases, the function of flow guiding of the horn-mouth
heightens, and the driving force of liquid circulation improves, thus the liquid circulation
velocity and liquid circulation flux increases. The increasing liquid circulation velocity in turn
can reduce the gas holdup. However, the downcomer gas holdup declines faster than that in the
riser, which enhances the total driving force of liquid circulation as well.
(4) The axial position height of the draft-tube has a great significance on the total circulation
impetus and gas holdup in the reactor within a certain range. With the increase of the height ratio
HB/HA, the flow resistance reduces at the draft-tube bottom, but the flow driving force also
reduces at the reactor top. Therefore, it’s possible that the total circulation impetus increases at
first, then keeps approximately stable and decreases slightly at last. The increasing liquid
circulation flux enhances the gas holdup in the downcomer, while the increasing liquid
circulation velocity in the riser reduces the riser gas holdup. When the liquid circulation velocity
and liquid circulation flux tends to be unchanged, the gas holdup no longer varies as well.
(5) There exists the mal-distribution of the gas holdup and liquid circulation velocity, especially
nearby the gas distributor. The gas holdup and liquid circulation velocity are higher in the
vicinity of the draft-tube wall than in the reactor wall. However, the mal-distribution is alleviated
with the increasing axial height. The local backflow and back mixing also are observed in the
riser, which have a negative effect on the mixing and mass transfer in the reactor.

Acknowledgments

We acknowledge the support by National Basic Research Program of China (2014CB748500) and
National Natural Science Foundation of China (51578239, 51322805).

References
Al-Masry, W.A., 2004. Influence of gas separator and scale-up on the hydrodynamics of external
loop circulating bubble columns. Chem. Eng. Res. Des. 82, 381–389.
Baten, J. Van, Ellenberger, J., Krishna, R., 2003. Using CFD to Describe the Hydrodynamics of
Internal Air‐lift Reactors. Can. J. Chem. Eng. 81, 660–668.
Behin, J., 2012. Deinking in bubble column and airlift reactors: Influence of wastewater of Merox
unit as pulping liquor. Chem. Eng. Res. Des. 90, 1045–1051.
Behin, J., 2010. Modeling of modified airlift loop reactor with a concentric double-draft tube. Chem.
Eng. Res. Des. 88, 919–927.
Bendjaballah, N., Dhaouadi, H., Poncin, S., Midoux, N., Hornut, J.-M., Wild, G., 1999.
Hydrodynamics and Flow Regimes in External Loop Airlift Reactors. Chem. Eng. Sci. 54,
5211–5221.
Buchholz, H., Buchholz, R., Lücke, J., Schügerl, K., 1978. Bubble swarm behaviour and gas
absorption in non-Newtonian fluids in sparged columns. Chem. Eng. Sci. 33, 1061–1070.
Buwa, V. V, Ranade, V. V, 2002. Dynamics of gas–liquid flow in a rectangular bubble
column:experiments and single/multi-group CFD simulations. Chem. Eng. Sci. 57, 4715–4736.
Chen, P., Sanyal, J., Dudukovićć, M.P., 2005. Numerical simulation of bubble columns flows: Effect
of different breakup and coalescence closures. Chem. Eng. Sci. 60, 1085–1101.
Chisti, M.Y., 1989. Airlift Bioreactor, Elsevier: London and New York.
Chisti, M.Y., Halard, B., Moo-Young, M., 1988. Liquid circulation in airlift reactors. Chem. Eng. Sci.
43, 451–457.
Chisti, Y., G, F.W., Murray, M.-Y., 1995. Relationship between riser and downcomer gas hold-up in
internal-loop airlift reactors without gas-liquid separators. Chem. Eng. J. 57, B7–B13.
Cockx, A., Liné, A., Roustan, M., Do-Quang, Z., Lazarova, V., 1997. Numerical simulation and
physical modeling of the hydrodynamics in an air-lift internal loop reactor. Chem. Eng. Sci. 52,
3787–3793.
Delnoij, E., Kuipers, J. a. M., Swaaij, W.P.M., 1997. Dynamic simulation of gas-liquid two-phase
flow: effect of column aspect ratio on the flow structure. Chem. Eng. Sci. 52, 3759–3772.
Dhotre, M.T., Smith, B.L., 2007. CFD simulation of large-scale bubble plumes: Comparisons against
experiments. Chem. Eng. Sci. 62, 6615–6630.
Drew, D.A., Lahey, R.T., 1987. The virtual mass and lift force on a sphere in a rotating and straining
flow. Int. J. Multiph. Flow 13, 113–121.
Ebrahimifakhar, M., Mohsenzadeh, E., Moradi, S., Moraveji, M., Salimi, M., 2011. CFD simulation
of the hydrodynamics in an internal air-lift reactor with two different configurations. Front.
Chem. Sci. Eng. 5, 455–462.
Fu, C.C., Fan, L.S., Wu, W.T., 2007. Flow regime transitions in an internal-loop airlift reactor. Chem.
Eng. Technol. 30, 1077–1082.
Ghasemi, H., Hosseini, S.H., 2012. Investigation of hydrodynamics and transition regime in an
internal loop airlift reactor using CFD. Brazilian J. Chem. Eng. 29, 821–833.
Godbole, S.P., Honath, M.F., Shah, Y.T., 1982. Holdup Structure in Highly Viscous Newtonian and
Non-Newtonian Liquids in Bubble Columns. Chem. Eng. Commun. 16, 119–134.
Gouveia, E.R., Hokka, C.O., Badino, A.C., 2003. The effects of geometry and operational conditions
on gas holdup, liquid circulation and mass transfer in an airlift reactor. Brazilian J. Chem. Eng.
20, 363–374.
Grienberger, J., Hofmann, H., 1992. Investigations and modelling of bubble columns. Chem. Eng.
Sci. 47, 2215–2220.
Hekmat, A., Amooghin, A.E., Moraveji, M.K., 2010. CFD simulation of gas-liquid flow behaviour
in an air-lift reactor: Determination of the optimum distance of the draft tube. Simul. Model.
Pract. Theory 18, 927–945.
Huang, Q., Yang, C., Yu, G., Mao, Z., 2010. CFD simulation of hydrodynamics and mass transfer in
an internal airlift loop reactor using a steady two-fluid model. Chem. Eng. Sci. 65, 5527–5536.
Huang, Q.S., Yang, C., Yu, G.Z., Mao, Z.S., 2007. 3-D Simulations of an Internal Airlift Loop
Reactor using a Steady Two-Fluid Model. Chem. Eng. Technol. 30, 870–879.
Hwang, S.., Cheng, Y.L., 1997. Gas holdup and liquid velocity in three-phase internal-loop airlift
reactors. Chem. Eng. Sci. 52, 3949–3960.
Hwang, S.J., Fan, L.S., 1986. Some design considerations of a draft tube gas-liquid-solid spouted
bed. Chem. Eng. J. 33, 49–56.
Jin, B., Yin, P.H., Lant, P., 2006. Hydrodynamics and mass transfer coefficient in three-phase air-lift
reactors containing activated sludge. Chem. Eng. Process. Process Intensif. 45, 608–617.
Karamanev, D.G., Nikolov, L.N., 1992. Free Rising Spheres Do Not Obey Newton’s Law for Free
Settling. Am. Inst. Chem. Eng. J. 38, 1843–1846.
Koide, K., Kurematsu, K., Iwamoto, S., Iwata, Y., Horibe, K., 1983. Gas holdup and volumetric
liquid-phase mass trandfer coefficient in bubble column with draught tube and with gas
dispersion into tube. J. Chem. Eng. Japan 16, 413–419.
Krishna, R., Urseanu, M.I., van Baten, J.M., Ellenberger, J., 1999. Influence of scale on the
hydrodynamics of bubble columns operating in the churn-turbulent regime: experiments vs.
Eulerian simulations. Chem. Eng. Sci. 54, 4903–4911.
Lehr, F., Millies, M., Mewes, D., 2002. Bubble-size distributions and flow fields in bubble columns.
AIChE J. 48, 1226–1243.
Luo, H.P., Al-Dahhan, M.H., 2011. Verification and validation of CFD simulations for local flow
dynamics in a draft tube airlift bioreactor. Chem. Eng. Sci. 66, 907–923.
Luo, H.P., Al-Dahhan, M.H., 2008. Local characteristics of hydrodynamics in draft tube airlift
bioreactor. Chem. Eng. Sci. 63, 3057–3068.
Luo, L.J., Yuan, J.Q., Xie, P., Sun, J.W., Guo, W., 2013. Hydrodynamics and mass transfer
characteristics in an internal loop airlift reactor with sieve plates. Chem. Eng. Res. Des. 91,
2377–2388.
Mohajerani, M., Mehrvar, M., Ein-Mozaffari, F., 2012. CFD analysis of two-phase turbulent flow in
internal airlift reactors. Can. J. Chem. Eng. 90, 1612–1631.
Mohanty, K., Das, D., Biswas, M.N., 2006. Hydrodynamic modeling of a novel multi-stage external
loop airlift reactor. Chem. Eng. Sci. 61, 4617–4624.
Mudde, R.F., Van Den Akker, H.E.A., 2001. 2D and 3D simulations of an internal airlift loop reactor
on the basis of a two-fluid model. Chem. Eng. Sci. 56, 6351–6358.
Oey, R.S., Mudde, R.F., Van den Akker, H.E.A., 2003. Sensitivity study on interfacial closure laws
in two-fluid bubbly flow simulations. AIChE J. 49, 1621–1636.
Oey, R.S., Mudde, R.F., Van Den Akker, H.E.A., 2003. Numerical Simulations of an Oscillating
Internal-loop Airlift Reactor. Can. J. Chem. Eng. 81, 684–691.
Roy, S., Dhotre, M.T., Joshi, J.B., 2006. CFD Simulation of Flow and Axial Dispersion in External
Loop Airlift Reactor. Chem. Eng. Res. Des. 84, 677–690.
Rujiruttanakul, Y., Pavasant, P., 2011. Influence of configuration on the performance of external
loop airlift contactors. Chem. Eng. Res. Des. 89, 2254–2261.
Sanyal, J., Marchisio, D.L., Fox, R.O., Dhanasekharan, K., 2005. On the Comparison between
Population Balance Models for CFD Simulation of Bubble Columns. Ind. Eng. Chem. Res. 44,
5063–5072.
Schiller, L., Naumann, Z., 1935. A drag coefficient correlation, Z.Ver.Deutsch.Ing. Elsevier Ltd.
Schumpe, A., Deckwer, W.D., 1982. Gas holdups, specific interfacial areas, and mass transfer
coefficients of aerated carboxymethyl cellulose solutions in a bubble column. Ind. Eng. Chem.
Process Des. Dev. 21, 706–711.
Siegel, M.H., Robinson, C.W., 1992. Application of airlift gas-liquid-solid reactors in biotechnology.
Chem. Eng. Sci. 47, 3215–3229.
Šimčík, M., Mota, A., Ruzicka, M.C., Vicente, A., Teixeira, J., 2011. CFD simulation and
experimental measurement of gas holdup and liquid interstitial velocity in internal loop airlift
reactor. Chem. Eng. Sci. 66, 3268–3279.
Sokolichin, A., Eigenberger, G., Lapin, A., 2004. Simulation of Buoyancy Driven Bubbly Flow:
Established Simplifications and Open Questions. AIChE J. 50, 24–45.
Song, T., Jiang, K.X., Zhou, J.W., Wang, D.Y., Xu, N., Feng, Y.Q., 2015. CFD modelling of
gas-liquid flow in an industrial scale gas-stirred leaching tank. Int. J. Miner. Process. 142, 63–
72.
Talvy, S., Cockx, A., Liné, A., 2007. Modeling hydrodynamics of gas-liquid airlift reactor. AIChE J.
53, 335–353.
Tian, X.F., Zhang, J.C., Liu, X.L., Fan, J.F., 2013. Application of CFD in airlift loop reactor. Mod.
Chem. Ind. 33, 121–124.
Van Baten, J.M., Ellenberger, J., Krishna, R., 2003. Hydrodynamics of internal air-lift reactors:
Experiments versus CFD simulations. Chem. Eng. Process. Process Intensif. 42, 733–742.
Wachi, S., Jones, a. G., Elson, T.P., 1991. Flow dynamics in a draft-tube bubble column using
various liquids. Chem. Eng. Sci. 46, 657–663.
Wadaugsorn, K., Limtrakul, S., Vatanatham, T., Palghat, A.R., 2016. Hydrodynamic behaviors and
mixing characteristics in an internal loop airlift reactor based on CFD simulation. Chem. Eng.
Res. Des. 113, 125–139.
Wang, T.F., Wang, J.F., Jin, Y., 2005. Population balance model for gas - Liquid flows: Influence of
bubble coalescence and breakup models. Ind. Eng. Chem. Res. 44, 7540–7549.
Wang, T.F., Wang, J.F., Jin, Y., 2003. A novel theoretical breakup kernel function for
bubbles/droplets in a turbulent flow. Chem. Eng. Sci. 58, 4629–4637.
Zhang, D., Deen, N.G., Kuipers, J.A.M., 2006. Numerical simulation of the dynamic flow behavior
in a bubble column: A study of closures for turbulence and interface forces. Chem. Eng. Sci. 61,
7593–7608.

You might also like