Você está na página 1de 16

Pergamon

Chemical En~lineerin9 Science, Vol. 52, No. 4, pp. 611-626, 1997


Copyright c) 1997 Elsevier Science Ltd
Printed in Great Britain. All rights reserved
P I I : S0009-2509(96)00425-3
0009 2509,/'97 $17.00 + 0.00

Dynamic numerical simulation of


gas-liquid two-phase flows
Euler/Euler versus Euler/Lagrange
A. Sokolichin* and G. Eigenberger
Institut fiir Chemische Verfahrenstechnik, Universit~it Stuttgart, B6blingerstr. 72,
D-70199 Stuttgart, Germany
and

A. Lapin and A. Liibbert t


Institut fiir Technische Chemie, Universit~it Hannover, Callinstr. 3, D-30167 Hannover,
Germany
(Received 18 January 1996; accepted 3 July 1996)

Abstraet--A dynamical, two-phase flow model in two- and three-space coordinates is presented. The gas-liquid flow is modeled by a Navier-Stokes system of equations in an Eulerian
representation. The motion of gas is modeled by a separate continuity equation. The Eulerian
approach with U P W I N D or TVD discretization and the Lagrangian approach for solving the
gas-phase equation are compared with each other on two two-dimensional test problems: the
dynamical simulation of a locally aerated bubble column and of a uniformly aerated bubble
column. The comparison shows that the results obtained with the TVD-version of the
Euler/Euler method and the Euler/Lagrange technique agree quantitatively. On the other hand,
it has not been possible to obtain similar agreement even qualitatively with the U P W I N D
technique, due to the influence of the numerical diffusion effects, which are inherent in the case
of U P W I N D discretization. Copyright 1997 Elsevier Science Ltd
Keywords: Modeling;

simulation;

fluid-dynamics; gas-liquid-flow; Euler/Euler;

Euler/

Lagrange.
1. INTRODUCTION
Numerical simulation is being recognized as a primary tool for improving the performance of process
equipment. In particular, for scale-up of chemical
reactors a reliable fluid dynamic reactor model is of
great benefit. Dynamic numerical simulation is thus
on the agenda of most big chemical companies and
many scientific research laboratories.
While the computing power of workstations and
mainframe computers, necessary to perform adequate
numerical simulations, increased considerably over
the last years, the appropriate basic simulation software is currently lagging behind. This is particularly
true for numerical codes which can be used to simulate gas-liquid two-phase flows.
As demonstrated by Lapin and Liibbert (1994),
Sokolichin and Eigenberger (1994) and Devanathan

et al. (1995), it is necessary to consider the dynamics of

* Corresponding author.
*Present address: Institut fiir Bioverfahrenstechnik,
Martin-Luther-Universifftt Halle-Wittenberg, Weinbergweg
23, D-06120 Halle, Germany.

the two-phase flow and the corresponding transient


flow behavior in order to account for the reactor
properties as mixing and heat transfer, which are of
interest to chemical engineers.
In literature, essentially two basic approaches to
dynamic flow simulations of two-phase gas-liquid
flows have been discussed. The first is an approach
where both the liquid motion and the gas-phase
motion are considered in a homogeneous way. These
two-fluid approximations are presented in Eulerian
representation and thus referred to as Euler/Euler
simulations (Torvik and Svendsen, 1990; Sokolichin
and Eigenberger, 1994). The second approach treats
only the liquid-phase motion in an Eulerian representation and computes the motion of the dispersed
gas-phase fluid elements in a Lagrangian way by individually tracking them on their way through the reactor. This approach has been termed Euler/Lagrange
representation (Webb et al., 1992; Lapin and Liibbert,
1994). Several numerical solution schemes which are
by no means equivalent have been applied to solve the
corresponding differential equation systems.

611

Downloaded from http://www.elearnica.ir

612

A. Sokolichin et al.

Before the available codes can be used for reactor


development it is necessary to validate them. Primarily, validation should be based on experiments
where the flow structures are similar to those of industrial reactors which are the final target of process
development. There are considerable difficulties in
such a direct validation procedure since the measurement techniques necessary to provide comprehensive
data from the turbulent flows prevailing in real chemical reactors are not sufficiently developed. Most
available measurement devices provide local fluid velocity data only. Usually, only long-time averaged
data are published. Even for bubble columns, which
can be regarded as the most simple two-phase reactors, gas-liquid flow patterns are available as longterm averages (Torvik and Svendsen, 1990; Grienberger and Hofmann, 1992). Consequently, a direct validation of transient flow structures in bubble columns
is presently not possible.
The best one can do at the moment is to make sure
that the codes predict at least qualitatively all characteristic properties of the flow which are known from
experience. In this contribution, such a comparison
will be based upon measurements in flat bubble columns with a wafer-type geometry where an essentially
two-dimensional flow structure prevails (Tzeng et al.,
1993; Becker et al., 1994).
It is the aim of the contribution to compare the
results of different codes, based upon the same fluiddynamical model, for two examples of a locally and
uniformly aerated flat bubble column. An Euler/Lagrange code is compared with two versions of an
Euler/Euler code where for the gas flow either a firstorder U P W I N D discretization or a second-order discretization is used. The stability of the second-order
discretization is based upon the concept of total variation diminution (TVD). Therefore, this code will be
referred to as the TVD method.
2. F L U I D - D Y N A M I C A L

MODEL

As pointed out by Landau and Lifschitz (1971), the


Navier-Stokes equation system, which is of fundamental importance to all single-phase flows, can also
be applied to two-phase flows if the dispersed phase
elements are small and do not significantly change the
overall fluid density and if the momentum of the
particles or bubbles can be neglected. Then the density p must be chosen as the effective density of the
dispersion and, similarly, the usual viscosity/~ must be
replaced by an effective viscosity #eff. This leads to the
following model equations:
~p

a-7 + v . (pu) = 0

(1)

stress tensor:
f ~ui

~u;

- -

dt

+ V.(puu) = -- Vp + V.T + pg

(2)

where u is the velocity vector, g is the acceleration


due to gravity, p is the pressure and T is the

?~u,\

(3)

The continuity equation (1) is usually combined with


the viscous momentum equation (2) to form the
Navier-Stokes equation system. Provided a proper
separate model is available for the effective viscosity
#elf, the system (1), (2) consists of 4 scalar equations
and contains 5 unknown variables (p, ul, Uz, u3, p).
The effective density, p, of the gas-liquid mixture
can be taken as the corresponding local average
p = ~pg + (1 - ~)p~

(4)

where e is the volume fraction or the local holdup of


the gas phase. The system of equations can be closed
with an additional continuity equation for the gas
phase:
O(ePo)
t?----f- + V. (epouo) = D

(5)

where D accounts for dispersive effects in the gas


phase due to random fluctuations of the bubble
motions. The gas velocity uo can be expressed as the
sum of liquid velocity u~ and slip velocity Usnp. For the
slip velocity, us~ip,various expressions can be found in
the literature depending on the pressure gradient, the
drag force, the added mass force, the Basset force, the
Magnus force and the Saffman lift force (see e.g.
Johansen, 1990). For the gas-liquid flow in bubble
columns, we assume the last four effects to be negligible. Then we get a simplified expression for the slip
velocity:
Vp
Uslip - -

Cdrag

(6)

where Cd,ag is a drag force coefficient for which a large


number of correlations can be found in literature,
depending upon whether single bubbles or bubble
swarms in stagnant or moving liquids are considered.
Ca~g depends primarily on the bubble size. This dependency is rather weak for air bubbles of 1-10 mm
mean diameter in water. According to Schwarz and
Turner (1988),
Car~g = 50 cmg3 s

(7)

can be used, leading to a mean bubble slip velocity of


about 20 cm/s, which is in complete accordance with
experimental velocity data of air bubbles in tap water.
The density of the liquid is assumed to be constant
within the bubble column while the density of the gas
phase depends on the local pressure p:
Pt = const.

Opu

2.

p
Pg = RTo"

(8)

(9)

Together with the equations representing the relationship between the velocities of both phases and the

Dynamic numerical simulation of gas-liquid two-phase flows


gas-liquid mixture,
pu = epgug + (1 - - e)plut

(10)

we get a closed system of differential and algebraic


equations which describes the dynamical behavior of
the two-phase flow. It can be solved numerically, if
p~ff and D are specified.
3, EFFECTIVE VISCOSITY AND BUBBLE PATH
DISPERSION

In order to determine the effective viscosity #eff of


the gas-liquid mixture, the standard k-s model developed for single-phase flows has been used in the
majority of publications on numerical simulations of
two-phase flows. However, at present it is not clear
how far such turbulence models, which have been
developed for single-phase flow, can be applied to
two-phase flows. The dispersed phase - - here the
rising gas bubbles - - obviously influences the effective
viscosity of the gas-liquid dispersion. Previous simulations showed (Becker et al., 1994) that gas-liquid
bubble flow can often be described with good qualitative and reasonable quantitative accuracy using
two-phase flow models without specific assumptions
about turbulence. In cases of insufficient quantitative
agreements, a moderate increase of the liquid viscosity
led to a substantial improvement (Becker et al., 1994).
On the contrary, in the air-in-water bubble columns discussed here, the standard k-e model would
predict an effective viscosity four orders of magnitude
larger then the liquid viscosity. This would substantially change the flow structure since it would
completely dampen out the transient motions and in
particular it would eliminate all the vortices, which
are well known to be present and can easily be seen in
such flows. In the often cited paper of Schwarz and
Turner (1988) the standard k-s model was used for the
case of a locally aerated bubble column. The authors
found a good agreement with the measurements.
However, it seems dangerous to generalize their result, since in their experiments the gas bubbles were
confined to a small portion in the middle of the
reactor only, while the rest of the column was essentially gas free. A comparison by the same authors with
a constant effective liquid viscosity also led to reasonable agreement with the experiments.
Since simple single-phase flow turbulence models
like the k-t- model turned out to be unsuccessful
(Becker et al., 1994), we simply assume that the effective viscosity/Lef
t of the gas liquid mixture is equal to
the viscosity of water.
Another important problem is the bubble path dispersion. When bubbles start from a point source at
the bottom of a bubble column with a sufficiently high
frequency, they interact with each other and do not
rise straight upwards even if the mean liquid velocity
is zero. Bubble wake effects (e.g. Fan and Tsuchiya,
1990) are the main reasons. As is well known, small
bubbles are accelerated in the wake of larger ones
and others are pushed aside. Hence, there is a con-

613

siderable path dispersion on a small scale which


appears as a random motion on the larger scale considered in our model.
This path dispersion effect is not restricted to
bubble plumes but is also present in bubble columns
which are aerated across their entire bottom. The
most simple way to consider this random spatially
dispersive effect is to extend the continuity equation of
the gas phase by a diffusion-like term as has already
been done in eq. (5). The corresponding diffusion
coefficient has been related to the turbulent eddy
viscosity of the liquid phase by Grienberger and
Hofmann (1992) and Torvik and Svendsen (1990).
This approach assumes an isotropic dispersion. However, since bubbles rise relative to the liquid predominantly in a vertical direction, dispersion will not be an
isotropic quantity.
A more general representation would be a dispersion tensor. However, presently there is neither
enough knowledge available to model the tensor elements nor enough experimental data to measure the
tensor elements reliably. We thus assume that the
term D in eq. (5j can be expressed as

D=:~

dxi [

(11)

where D~ (i = 1,2, 3) are some constant generalized


diffusion coefficients estimated from experimental
data.
4. NUMERICAL SOLUTION PROCEDURE

First, we introduce some simplifications into our


model. Since the density of the gas phase is much
smaller than the density of the liquid phase, we can
assume without significant loss of accuracy that
p = (1 - s)p/

(12)

u = ut.

(13)

and

Further, for the rest of this paper we assume the gas


phase to be incompressible. Under this assumption,
the gas continuity equation (5) simplifies to an equation for the local gas holdup s:

where
Vp
Ug = u + u~lip = u - Cdra--g"

(15)

In our numerical simulations of eqs (1) and (2) the


finite-volume method has been used. In the threedimensional case, the solution domain is discretized
into six-sided, rectangular control volumes. We take
the staggered grid formulation first used by Harlow
and Welch (1965), which means that the scalar quantitites are attached to the centers of the control volumes, and the velocity components are calculated for
the centers of the surfaces of the control cells.

614

A. Sokolichin et al.

Equations (1) and (2) fully correspond with the mass


and m o m e n t u m balances for the single-phase flow.
This means that these equations can be solved in the
same way as in the single-phase case and well-established iteration procedures can be applied. We use
the S I M P L E R technique of Patankar (1980). The only
modification required is to update the local density
values of the gas-liquid mixture, p, in the space domain at the end of each iteration loop. For this purpose we solve the gas holdup equation (14) and
substitute its result e into the expression (12).
The accuracy of the solving procedure for the gas
holdup equation (14) plays a crucial role in the
modeling of gas-driven gas-liquid flows, because the
resulting flow pattern directly depends upon the gas
holdup distribution in the reactor. The two methods
most frequently used to solve this equation are the
finite-volume method and the method of characteristics. Depending on which of these two methods is
used, the fluid dynamical model is referred to as an
Euler/Euler model or as an Euler/Lagrange model. In
the next sections these two approaches will be described in detail.
5. S O L V I N G T H E G A S H O L D U P E Q U A T I O N

In this section we will concentrate on the numerical


algorithms for solving the gas holdup equation (14).
We assume the components of u 0 to be known at the
faces of the control volumes at a given time from
the solution of eqs (1), (2) and (15). Furthermore, the
diffusion coefficients D~ are considered to be known
and constant.
5.1. Eulerian approach
We start with the finite-volume method for the gas
holdup equation in one spatial direction. The ideas
presented here can be extended in a rather straightforward way to two and three dimensions.
In one dimension, eq. (14) simplifies to

~:, = -- (eU)x + Dexx

(16)

where the indices t and x denote derivatives along the


corresponding variables. In the following, we will
omit the subscripts ofug, a, D1 and Xl for the reason of
simplicity. We use a finite-volume method in which e7
represents an approximation to the cell average of e at
time t, over the ith cell Ci = [ X i - 1 / 2 , X i + a / 2 ] . The
finite-volume formulation for the gas holdup equation
can be obtained through the integration of eq. (16)
over Ci x It., t.+a] and takes the form
E~' + 1 _ e7

At

(1/At) ~ ~'~Dex (xi- 1/2, t) dt based on the data d at time


h, where I is equal to n or to n + I, depending on what
kind of time integration (explicit or implicit) is used.
Unless otherwise stated, it is understood that all data
are taken at time t, + 1 and the superscript I will be left
out.
The usual method to approximate the convective
fluxes is the first-order U P W I N D method where
(<Uei-a
Fup(e;i) = I Uei

if U / > 0
if U < 0.

(18)

U replaces the velocity u72al/2 at the left side of the ith


cell Ci.
The U P W I N D method leads to a numerical diffusion in the order of[ UIAx/2. Usually, with the numerical grid resolutions which can be handled in two- or
three-dimensional calculations the true solutions become strongly smoothed. Hence, the accuracy of the
solutions is rather low.
The second-order central-difference method

ud(e; i) = U eil _ _1_+_ i e


2

(19)

works well in cases where only small t-gradients and


very low cell Peclet numbers (i.e. the low values of
UAx/D) are to be expected but it has difficulties if
e has steep gradients since then it is very dispersive
and tends to generate artificial oscillations.
Much better results can be obtained using a hybrid
method that uses the second-order flux in smooth
regions but involves some sort of limiting based on the
gradient of the solution so that near discontinuities it
reduces to the monotone U P W I N D method. The
stability theory of such flux-limiter methods is based
on the concept of the total variation diminishing of
the solution (for details see e.g. LeVeque, 1990), so we
will use the abbreviation TVD for this type of the
convective flux approximation. Note that the centraldifference flux (19) can be decomposed into the
U P W I N D flux plus a correction term:

FCd(e;i) = FUP(e;i) + 1Ul(el - el-l).

(20)

This suggest the following flux-limiter method:


FTVD(e;i) : FUp(g;i) q- 1Ul(g~- ~ , - ~ ) ~

(21)

where qbi is the limiter which depends on the local


nature of the solution. Note that if ~i = 0, then we
have the U P W I N D method while if tb~ = 1 we have
central difference. The limiter we will use here has the
form

A x [F(d;i + 1) - F(fl; i)]


~i = ck(Oi),
1

+ ~xx [D(e;i + 1) - D(et;i)]

Oi =

C'I - - e l - 1

(22)

~i - - E i - 1

(17)
where

where F(d;i) is some approximation to the average


convective flux (1/At)S,.,*n+l (eu)(xi_ 1/2, t) dt and D(d;i)
is some approximation to the average diffusive flux

I =

i-1
+1

if U ~ > 0
if U < 0 .

(23)

615

Dynamic numerical simulation of gas-liquid two-phase flows


We see that 0~ is the ratio of the slope at the neighboring interface in the upwind direction to the slope of
the current interface. One standard limiter we use in
our calculations is the superbee limiter
0(0) = max(0,min(1,20),min(2,0)).

Ax
is satisfied for each i. If the implicit TVD method is
used, eq. (17) leads to a system of non-linear algebraic
equations, which has to be linearized and to be solved
iteratively. Violation of condition (25) may lead to
negative central coefficients in the resulting system of
linear equations which may cause severe convergence
problems. That is why we prefer to use the explicit
approximation to the convective fluxes if the TVD
method is applied. At high space resolution of the
solution domain, condition (25) requires very small
time steps, which leads to a considerable increase of
computation time. The computation time can be drastically reduced if one uses a finer time mesh only for
solving the gas holdup equation and keeps larger time
increments for solving the other model equations.
For the diffusion term we exclusively use the implicit second-order central-difference flux approximation
'~ii -- ~;i

D(<i) = D - Ax

nm

(24)

The implicit TVD method described above is unconditionally stable, while the explicit one is stable only if
the condition

UAt

After each integration time step the gas holdup


~: can be calculated for every control cell as the total
volume Vapea of the GPPs within the cell divided by
the cell volume

126)

based on the data e"+J at time t,+ 1. The explicit


approximation to the diffusive fluxes requires very
small time steps in order to maintain stability
(At < (Ax)2/2D). It is therefore very ineffective on fine
spatial meshes.
5.2. Lagrangian approach
The Lagrangian method of solving the gas holdup
equation does not require solving any explicit equation of the gas holdup. The gas phase is represented by
a finite number of dispersed gas-phase particles
(GPP), and the gas holdup ~: can be calculated from
size and position of the GPPs in the solution domain.
We intentionally speak about dispersed gas-phase
particles rather than about gas bubbles for reasons
which will become clear later. At the beginning of the
simulation we assume the gas holdup in the reactor to
be equal to zero, so initially no GPPs are present in
the solution domain. After starting the aeration, the
GPPs enter the solution domain through the gas
sparger area of the bubble column according to the
predefined superficial gas velocity. If no bubble path
dispersion takes place in the gas phase (D~ = 0, i = 1,
2, 3) the GPPs will move through the reactor with the
velocity u 0. Then the bubble trajectories can simply be
calculated by solving an ordinary differential equation system describing the buoyancy.

VGppd

= -

Pg
where n is the number and m the mass of the GPPs in
the cell volume and p is the density of the gas.
However, if the total number of GPPs is not very
high, this way of calculating the gas holdup can lead
to sharp discontinuities in the gas holdup profiles
which might lead to convergence problems during the
solution for the Navier Stokes equation. In order to
avoid these difficulties, the GPPs are represented by
spatial density distribution functions as described in
detail by Lapin and Liibbert (1994). The density functions may overlap each other and also extend across
the boundaries of the computational cells. In this way
each G P P may contribute to the gas holdup not only
within the actual control cell but also, in particular
while it is moving near the cell border, in adjacent cells.
If random bubble motion effects cannot be neglected, they can be described by random shifts of the
positions of the individual GPPs which are computed
in the following way:
x n ( t ) -~ x n ( t ) q-

~ix~tdi

(27)

Here xT(t) is the ith coordinate of the nth G P P at time


t, gi (i-- 1, 2, 3) are independent random numbers
uniformly distributed in the range [ - 1/2, 1/2], At is
the length of the next time step and di (i = 1, 2, 3) are
coefficients characterizing the strength of the disturbances, which still have to be calculated. Thus, essentially a random motion has been superimposed on the
bubble motion corresponding to a Brownian or generalized diffusive motion.
In order to calculate the coefficients di in such a way
that the diffusive motion can be compared with the
approach used in the Euler/Euler solutions, consider
a three-dimensional diffusion equation

?':- c ('D, Oe"~


Ct Cx~\ CxU

(28)

on the whole space domain with the initial condition


c,(x, t = 0) -- 6(x).

(29)

The analytical solution of this equation is the threedimensional Gaussian distribution function
1

e(x,t)=

I~

i=1,2.3

2 ~'>4D''.x/vitr~
~/:__e -

{30)

The standard deviation 0"2(t) of the function


r(x,t) = x/x 2 + x 2 + x 2

(31)

can be calculated as
t~2(t) =

fR r2(x,t)e,(x,t)dx = 2t(D1 + D 2 -b D31.


(32)

616

A. Sokolichin et al.

On the other hand, we can determine the numerical


solution of eq. (28) analogous to an ink-drop dispersion experiment by releasing M particles with associated volume Ax~ A x 2 A x 3 / M
at some initial time
t = 0 from the point x = 0 (the discrete analogon of
the 6-function) and let them disperse, i.e. changing
their coordinates in space and time according to eq.
(27). The numerical solution e(x, t) calculated from the
particle distribution at time t will then converge to the
probability distribution function of single gas-phase
particle position x(r) in the limit for Ax--, 0 and
M ~ ~ . The position of a single particle x(t) at time
t = n At is defined by
x , ( n A t ) = ( ~ + ~2 + ... + / , ? ) , ~ t t d , .

(33)

The distribution of x(t) converges for fixed t and


n ~ ~ , A t = tin --, 0 to the three-dimensional Gauss
distribution function by the central limit theorem of
the probability theory, with
t~2(t) = (r2(x,t)) = ~2(d 2 + d22 + d~).

(34)

From eqs (32) and (34) we are now able to express the
relationship between d~ and D~:

di

2~t'.

(35)

From the numerical point of view it is not necessary


or even not correct to associate a dispersed gas-phase
particle with a single gas bubble. The number of
GPPs and the number of gas bubbles in the reactor
may be different. If the volume of a single gas bubble is
much smaller than the volume of a control cell and
many of them are within this single cell, then the
number of the GPPs can be taken smaller than the
number of the gas bubbles. In this case, one G P P
represents a bubble cluster.
On the other hand, if the gas holdup is low and the
volume of a single gas bubble is larger than the control volume element, then it might be of advantage to
represent such big bubbles by a number of GPPs in
order to obtain a more continuous distribution of the
gas across the numerical grid.
The total number of the GPPs in the solution
domain is controlled by the particle generation rate,
which depends on the grid resolution and the bubble
size distribution, but in every case it must match the
predefined superficial gas velocity.
It should be mentioned that the Lagrangian approach using eq. (27) is not the optimal way to solve
continuous equations like eq. (14). In particular, problems with the number of GPPs arise through the task
of representing the gas diffusion terms in eq. (14) by
means of the Lagrangian approach. Physically, the
diffusion approach describes a gas transport from the
regions with high gas concentration to regions with
lower gas concentration. Equation (27), however, describes the random component of the movement of
the GPPs. This approach can lead to unphysical results if only a small number of GPPs per unit cell are
present, because it allows for a transition from a con-

trol cell with lower gas holdup into the control cell
with the higher one. As an example we can imagine
a random jump of one G P P from a control cell
containing a single G P P to the adjacent cell containing two GPPs. In the following, examples of twophase flow with many small bubbles are considered.
Then it is no problem to represent the diffusional
component in eq. (14) adequately with a Lagrangian
approach.
6. T E S T

CASE

In order to compare the simulation results produced with both methods we use the example of
a partially aerated flat bubble column. This test case is
described in detail in Becker et al. (1994), hence only
a brief description will be given here. The apparatus
has a rectangular cross-section with the following
dimensions: width 50cm, depth 8 cm and height
150 cm.
Glass walls on the front and the back allow observation and photographic documentation of the
multiphase flow. For gas dispersion a single frit, flushmounted on the bottom of the apparatus at the distance of 14.5 cm from the left side of the column has
been used. In this flat column, an essentially twodimensional flow structure develops, depending on
the gas flow rate used. At superficial gas velocities
below 1.5 mm/s, the flow depicts a transient character.
This is shown by results obtained with a gas throughput of 0.66 mm/s. Several liquid circulation cells can
be observed in the column. They continuously change

oq.:. ,e4-.~.l,, ..t '~..':..


L'

" " -~:.'r'

: 4,,'."

""

'

~"

.. "..,_ . , ~ ~ . ~ :
.:.
;' : "~"~"2 "

y ~ . . ' ~ , ,.

. ~Y,,'.::-

/g-i::

~.

!~.

~l,

Fig. l. Locally aerated bubble column: binary and inverted


photographs of the oscillating bubble swarm at two different
times (Becker et al., 1994).

Dynamic numerical simulation of gas liquid two-phase flows


their location and their size. The bubble swarm
motion is influenced by these vortices and therefore
rises in a meander-like way (Fig. 1).
7. S I M U L A T I O N RESULTS W I T H O U T BUBBLE PATH
DIFFUSION

All numerical simulations assume a two-dimensional rectangular geometry with height 150 cm and
width 50 cm. A regular numerical grid with 150 x 50
grid points was used. The simulation results obtained
with three different numerical algorithms were compared with each other: the Eulerian approach with
U P W I N D discretization of the gas holdup equation
(short: UPWIND), the Eulerian approach with TVD
discretization (short: TVD) and the Lagrangian approach for the gas equation (short: LAGRANGE).
Let us first neglect the path diffusion effects in the
gas phase. This means that the coefficients Dj in
eq.(14) and di in eq. (27) are assumed to be zero.
Figure 2 depicts the gas holdup pattern in the bubble
column, 5 s after the onset of the aeration. A tremendous influence of the numerical diffusion in the Eulerian solution obtained with the UPWIND discretization
technique can be recognized. This is not due to the
Eulerian approach as the results obtained with the
TVD method demonstrate. The results obtained with
the TVD method look much more similar to the gas
distribution which results from the Lagrangian simulation. We thus can conclude that the U P W I N D
technique leads to strong numerical diffusion effects.
The amount of numerical diffusion in vertical and
horizontal directions is proportional to the local vertical and horizontal gas velocity components. In the

UPWIND

TVD

617

first 10 s after the onset of the aeration the vertical


velocity component prevails over the horizontal velocity component in the region where the gas phase is
present, leading to a much higher numerical diffusion
in the vertical direction than in the horizontal one (see
Fig. 2, left). The evolution of the velocity field during
the first 48 s after the beginning of the aeration obtained with the Lagrangian approach (Fig. 3) shows,
however, a continuously changing velocity pattern,
leading to different local numerical diffusion effects at
each time step. This means that the effect of numerical
diffusion of the U P W I N D method is completely uncontrollable and its influence on the distribution of
the gas phase has an unpredictable character.
The comparison of the liquid velocity patterns 60 s
after the onset of the aeration (Fig. 4) shows a very
good agreement between the TVD and the LAGRANGE results. Also the UPWIND solution shows
a good qualitative agreement with the other two solutions. For the better quantitative comparison between
the simulation results obtained with all three
methods, the vertical liquid velocity profiles at height
100 cm are presented in Fig. 5. We see that the TVD
and the LAGRANGE solutions are close to each
other, whereas the velocity variation in the U P W I N D
solution is about a factor 2 smaller.
Figures 6 and 7 show the comparison of the liquid
velocity patterns at t = 120 and t = 180 s. Even 180 s
after the onset of the aeration, the TVD and the
LAGRANGE solutions are in good qualitative agreement, whereas the UPWIND solution already leads to
different results at 120 s. The comparison of the evolution of the vertical liquid velocity component in time,

LAGRANGE

[]
0.28%

[]

0.42%

0.56%
0.69%

0.97%

[]

1.25%

N
1.39%
[]
Fig. 2. Locally aerated bubble column. Distribution of the gas holdup 5 s after the beginning of the
aeration calculated with three models. Diffusion term is assumed to be zero.

A. Sokolichin et al.

618
Time: 9.4 s

Time: 11.2 s

Time: 13.0 s

..-,

Time: 15.0 s

Time: 19.0 s

~1~;i, :. :-:-:,~,l

td'
'r

',~'.'l

I'

~ , :
' ','t

'"'""~ ~

~'"'-~/

< :.:,:,,,,

:: :'I I

I ~'/illl~ ','.'~t
,:I
,,t,,:-;i

'-~II~iittb
"i:','

h;.. ",'.<.>:,:,:,
:,:..:t~:,.,:,:,..:,:+ !~"."71','<,

i!!~i}

.~,:;: ...:.:.:,:.:
,,,h~ ,..:.:.:.:,:, l~:.::::',:::

:::::::::::::::::::::::!::'-

,'>,]~l{r6-?,h'--'")ilNi
,,',~t~-::,S;J, u ~ ,i

i::: :!:: f~~::',::

,,.,, ,.._.;,,,,>:,:,:, I'~/,;',';,'",':;

:.i,?) i:::':.;:;':;.'.

!!?i i iiiiiiiii!iiii '"'" "-"'"'"""

tu.
',~llt,{--_<f/>,,,
i~..,,rll.>_-'..;,',,,,,,,

Time: 27.0 s

Time: 35.0 s

~..',, },,.-:---:..,?

'..>:. >>>>>>: i.

) :,:': ~..:.; ..:...

t)

~-'~l I
t~)t,.:,
,

t! 7#/D}~,,'.",
l,j~tll',: : : '

ff~fif;
I1
[!:J:71

".-

Time: 42.0 s

!t,.

f ";'7.~>>1

'?i

Time: 48.0 s

['r//,

~.~'~ i~

,'t','."

,,,'-~?! ~--',
~_

,:,;.,.

It,'.i~t

~!I,!%I1[1" ;':-,-.,#1
.'--'J1

I/b-:,II.-:--:-.'.',,,,,
P,:::.: ::::: :.: :;:,.::

"."

Time: 22.0 s

~'~ S,:: ;:'I


;!
~'tL ::t'

iglIlI

~////~4,': :'

~/~,~,..-.,

lit,':i~ [~f!2"::',

i!ti! '>':::
N\\':::-:';

71 "><

fthttltih]i:
I 'J

~I,':'

;Ii

,':i
i

tI!::Nt!L::
dlltlfftl,',:::,'i
iltfJl/t!(,:.::+

#-','

7zi,,]ttlt4

:,~tIlIl[?:::::!Nltl
t,\\\"~.-_"!,Jlttt
iI',t;i,,g

-:,,',

7,~.",: : 7;17tmI

Fig. 3. Locally aerated bubble column. Evolution of the liquid velocity field during the first 48 s after the
beginning of the aeration. Lagrangian approach without diffusion. (Here and subsequently velocity vectors
are shown only at each 8th grid point; the vertical vector in the left bottom corner of each plot corresponds
to the velocity of 10 cm/s.)

plotted for some fixed position A in the reactor (Fig.


8), also shows quantitative agreement between the
T V D and the L A G R A N G E solutions. This is a rather
striking result since different numerical solutions of an
intrinsically unstable dynamical system tend to deviate more and more from each other as time proceeds.
As a matter of fact, such a deviation can also be
observed for longer simulation times. Figure 9 shows
the long-time behavior of the liquid velocity component at point A. After about 4 rain the agreement between the TVD and the L A G R A N G E solution
vanishes. Later on, however, a quasiperiodic solution
is established which is again in close accordance for

the two methods whereas the U P W I N D solution


shows a completely different single-periodic behavior.
Let us now look at the void fraction distribution
calculated with the three methods for t = 60s
(Fig. 10). If we compare the TVD and the L A G R A N G E
solution with the photographs in Fig. 1, we must state
that the radial dispersion in the gas phase cannot be
reproduced by both methods, whereas the U P W I N D
solution seems to perform much better. Even under
strongly fluctuating flow conditions (see Fig. 3) the
spread of the bubble plume calculated with T V D or
L A G R A N G E method is much smaller than observed
experimentally. This shows that the different medium

619

Dynamic numerical simulation of gas-liquid two-phase flows


UPWIND

TVD

LAGRANGE

Fig, 4. Instantaneous liquid velocity field at 60 s after beginning of the aeration calculated with different
methods (no diffusion is considered).

The apparently good performance of the U P W I N D


solution (Fig. 10, left) is of course a consequence of the
numerical diffusion which in our case happens to have
about the right order of magnitude. If the space grid is
further refined however, the numerical diffusion of the
U P W I N D solution decreases and the spread of the
bubble plume would also decrease.

20
""N

10

~'/ ...."........................
~,,"*~.....
/
/

,.-'
.-"

.....
\

....,....

/,."

-'II....

-10

x\

...............i !
LAGRANGE
UPWIND

-20

-30
0

....

S. S I M U L A T I O N

10
20
30
40
DISTANCE FROM THE LEFT WALL [cm]

50

Fig. 5. Vertical liquid velocity profiles at height 100 cm.


Time = 60 s after beginning of the aeration calculated with
different methods (no diffusion is considered).
size vortices of the computed flow field do not disperse the bubble flow sufficiently. Instead, the dispersion is caused by numerous small vortices and flow
variations caused by the liquid flow around individual
bubbles or bubble clusters. Since our model does not
resolve these small-scale phenomena, some appropriate corrections become necessary. In the following,
diffusion term D in the gas holdup equation (5) will be
used as a first approximation.

RESULTS

WITH DIFFUSION

The preceding example showed that a gas-phase


dispersion model is necessary to obtain a physically
reasonable distribution of the gas phase. We will
therefore use the procedure described in Section 5.2 to
obtain compatible dispersion parameter values for the
TVD and LAGRANGE solutions. Before doing so,
a validation of the algorithm, eq. (27), as well as of the
equivalence relation, eq. (35), is necessary since they
have been obtained on more or less intuitive arguments. For the validation only the gas holdup equation (14) for a constant liquid velocity field will be
solved with TVD and LAGRANGE methods and the
results compared. Note that the Lagrangian technique is used here in order to solve the continuous gas
dispersion term in eq. (14). Computationally this is
not the most efficient way of using this technique,
however, this allows for a direct comparison of the
results obtained with different techniques.

620

A. Sokolichin et al.
UPWIND

TVD

LAGRANGE

Fig. 6. Instantaneous liquid velocity field at 120 s after beginning of the aeration calculated with different
methods (no diffusion is considered).

UPWIND

TVD

LAGRANGE

Fig. 7. Instantaneous liquid velocity field at 180 s after beginning of the aeration calculated with different
methods (no diffusion is considered).

Dynamic numerical simulation of gas-liquid two-phase flows

621

20
~ .

']'wrjD

..
t

10

......... UPWIND

i
i"

i~

..,.
,

1500

.....

[tl

/01

90o
35]1_

-10

-20

-30

3O

60

90

120

150

180

TIME [s]

Fig. 8. Vertical liquid velocity at position A calculated with different methods (no diffusion is considered).

15
0

-15
-30
>[..,

15

>

-15

-30
15
0
-15

-30
0

500

1000

1500
TIME [s]

2000

2500

3000

Fig. 9. Long-time vertical liquid velocity fluctuation at position A (see Fig. 8) calculated with different
methods (no diffusion is considered).

8.1. T e s t p r o b l e m
Let us consider the gas holdup equation (14) in two
dimensions. We assume the components of ug and the
diffusion coefficients Di to be known and constant,
and consider uniform convection only in the vertical
direction and diffusion only in the horizontal direction. U n d e r these assumptions eq. (14) simplifies to
e t = - - (~U)x + Deyy

(36)

where x and y denote the vertical and horizontal


coordinate directions, u the vertical velocity and D the

horizontal diffusion coefficient. We assume the following values for u and D: u = 20 cm/s, D = 4.1(6) cm2/s;
the latter corresponds to the disturbance coefficient
d = 10 cm/s t/2. This means that two G P P s with the
same coordinates at time t = to can be a maximum
distance of 10 cm apart at time t = to + 1, if time step
At = 1 s is used.
We solve eq. (36) in the same calculation domain
as described in Section 7 and with the initial condition e(t = 0, x, y ) = 0, and the boundary condition
e(t, x, y = 0) = 0.05 for x ~ [ 1 3 em, 16 cm]. After 7.5 s

622

A. Sokolichin et al.
UPWIND

LAGRANGE

TVD

0.13%
0.25%

0.38%

0.63%
0.76%

1.01%

I
1.14%
[]

1.27%

[]

Fig. 10. Locally aerated bubble column. Distribution of the gas hold-up 60 s after the beginning of the
aeration calculated with three models. Diffusion term is assumed to be zero.

~.44
0.89

1.33

1.77
2.66
3.10

4.43

[]
a

Fig. 11. Simulation results for convection-diffnsion test problem with constant vertical velocity and
constant horizontal diffusion:(a) stationary TVD solution; (b) instantaneous positions of 45,630 GPPs in
Lagrangian method; (c)Lagrangian solution with 45,630GPPs; (d) Lagrangian solution with
180,566 GPPs.

of real time, the gas front reaches the top of the


calculation domain and the solution of eq. (36) becomes stationary. The corresponding numerical
stationary solution calculated with TVD method is
shown in Fig. 11 (a). In the frame of the LAGRANGE
method no stationary solution can be reached, because the positions of the GPPs [Fig. ll(b)] are
continuously changing in time. However, if the generation rate of GPPs is high enough, the time variation of the calculated gas holdup distribution varies

only slightly in time, so we can speak of a quasistationary solution of the Lagrangian approach. One
such quasistationary distribution of the gas holdup is
shown in Fig. 1l(c). At that time the total number of
GPPs in the calculation domain equals 45,630. Although a good qualitative agreement can be found
between Figs 11 (a) and (c), the Lagrangian solution is
not very smooth. Only after increase of the generation
rate of GPPs by a factor 4 can we get the smooth
solution presented in the Fig. 11 (d) (corresponding to

623

Dynamic numerical simulation of gas-liquid two-phase flows


20
A,

t.0

[-~.-~, LAGRANGE(180566GPP~)~

LAGRANGE
UPWIND

630 GPPs )l

.:.,,
:.

I m ~ . ,rv D

~TVD

0.8

t0

0.6

-10

0.4

>
0.2

0,0

-20

-313
0

10
20
30
40
DISTANCE FROM THE ~
WALL [crn]

50

Fig. 12. Calculated gas holdup profiles at height 100 cm for


convection~liffusiontest problem with constant vertical velocity and constant horizontal diffusion: stationary TVD
solution, Lagrangian solutions with different numbers of
GPPs.

t
t

30

60

90
TIME lsl

120

150

180

Fig. 13. Vertical liquid velocity at position A (see Fig, 8)


calculated with different methods (d2 = 5 cm/sa'Z).

15
0

-15
TVD
15
0
-15
LAGRANGE ]

15
0
-15
0

500

1000
TIME [s]

1500

2000

Fig. 14. Long-time vertical liquid velocity fluctuation at position A (see Fig. 8) calculated with different
methods (d2 = 5 cm/st'2).

the total n u m b e r of G P P s = 180,566). For better


quantitative comparison between the solutions, the
gas holdup profiles at the height 100 cm calculated
with different models are presented in Fig. 12.
As a result it can be stated that gas-phase dispersion can be modeled with equal accuracy by both the
Eulerian-TVD as well as the Lagrangian approach
provided that a rather high n u m b e r of G P P s is used in
the latter case. The required big n u m b e r of G P P s may
present computational problems if industrial-scale reactors have to be simulated, since the computational
time is roughly proportional to the G P P number,

8.2. Two-phase .flow with diffusion


Let us now compare the simulation results for the
whole two-phase system in the presence of diffusion in
the gas phase. Since a comparison of Fig. 1 l(b) with
the photographs in Fig. 1 shows a somewhat larger
spread in the calculated flow, only half of the disturbance coefficient of Fig. 11 (b) is used in the horizontal
direction: d2 = 5 c m / s 1/2 (resp. D 2 = 1.041(6)cruZ/s).
The diffusion in the vertical direction is assumed to be
negligible (dl = 0, D1 = 0). Figures 13 and 14 now
correspond to Figs 8 and 9 (without diffusion). We can
see that the TVD and the L A G R A N G E solutions

624

A. Sokolichin et al.
again show very similar long-time behavior and the
U P W I N D method gives a totally different solution.
Figure 15 shows the distribution of the GPPs in
L A G R A N G E solution at two different times, and we
can now observe a great similarity with the photographs in Fig. 1.

Fig. 15. Simulation results for a locally aerated bubble column with Lagrangian approach (dz = 5 cm/sl/2). Instantaneous positions of GPPs at two different times.

9. UNIFORM AERATION
The next test example is the dynamical simulation
of a bubble column which is aerated uniformly over
its entire bottom. Visual observation shows that at
low superficial gas velocity a so-called homogeneous
flow structure prevails, where the bubbles rise uniformly through an essentially stagnant liquid. As the
superficial gas velocity is increased, an instationary
flow structure develops, where vortices are created
close to the gas distributor and move upwards and
sideways in a rather irregular way. Long-term
measurements of the gas holdup distribution and of
the liquid velocities show the well-known picture of
an increased gas holdup in the middle of the column,
leading to an overall liquid circulation with upflow in
the center and downflow near the walls (Grienberger
and Hofmann, 1992).
The simulation results for the same flat column as
specified in section 6 with uniform aeration over the
entire bottom obtained with the Lagrangian approach
(d2 = 5 cm/s 1/2) are given in Fig. 16. The simulations
show that above a minimum value of the superficial
gas velocity of about 2 cm/s an unsteady flow structure

Fig. 16. Uniformly aerated bubble column: instantaneous (left, middle) and long-time-averaged (right)
simulation results of liquid velocity field. Lagrangian approach with diffusion (d2 = 5 cm/sl/2). Superficial
gas velocity equals 2 cm/s.

Dynamic numerical simulation of gas-liquid two-phase flows

625

20
0
-20
-40
20

0
-20
-40

2 1
0

-20

--40

0
10
20
30
40
50
DISTANCEFROMTHE LEFTWALL[cm]

Fig, 17. Uniformly aerated bubble column: long time averaged vertical liquid velocity profilesat three differentheights
calculated with different methods.

develops. If the calculated local velocities are averaged over a longer time period, as is done in the
usual bubble column measurements, a regular flow
structure with one overall circulation cell results
(Fig. 16, right).
It is not possible to make a direct quantitative
comparison between instantaneous flow pattern resuits obtained with all three methods, because of the
chaotic character of the solution. However, we can
make an indirect comparison through the calculation
of the long-time-averaged velocity patterns. The corresponding vertical liquid velocity profiles at three
different heights calculated with LAGRANGE, TVD
and U P W I N D methods are shown in Fig. 17. As in
the case of a locally aerated bubble column we can
state a very good quantitative agreement between the
TVD and the LAGRANGE solutions. The U P W I N D
method leads again to quantitatively different results.
The influence of the numerical diffusion in the case of
a uniformly aerated bubble column is, however, not so
high as in the case of a locally aerated bubble column,
due to the smoother distribution of the gas phase.
For the transport phenomena in the two-phase
flow, the instantaneous and not the long-time-averaged velocities are of decisive importance. If we now
take a look at the representative instantaneous liquid
velocity patterns calculated with the LAGRANGE
(Fig. 16, middle), the TVD (Fig. 18, left) and the
U P W I N D (Fig. 18, right) methods, we can see that the
U P W I N D solution depicts a qualitatively different
behavior with a much lower number of vortices as in
the other two solutions. So we can state that also in
the case of a uniformly aerated bubble column, the
U P W I N D method leads to a qualitatively different
solution compared to the LAGRANGE and the TVD
approaches.

Fig. 18. Uniformly aerated bubble column: instantaneous


simulation results of liquid velocity field calculated with
TVD (left) and UPWIND (right) methods.

10. CONCLUSIONS
From the chemical reaction engineering point of
view, fluid dynamical models are required for a proper
description of fluid mixing and contacting patterns,
i.e. they model the way by which materials flow
through the reactor and contact each other in order to
react chemically (e.g. Levenspiel, 1989). Hence, the
local transport properties are of primary importance.
In this light it is essential that numerical diffusion
effects which corrupt the numerical simulation results
are kept under control. Such numerical diffusion effects are of particular importance for the bubble column reactors considered in this paper, since flow in
bubble columns is essentially buoyancy driven. Strong
diffusional transports, however, may degradate the
density gradients. Numerical diffusion will, thus, lead
to incorrect driving forces in the simulations.
Simple numerical solution techniques such as the
commonly applied U P W I N D technique may lead to
unacceptable numerical diffusion effects. This artificial diffusion exceeds the naturally appearing diffusion considerably, often by orders of magnitude. In
principle it would be possible to compensate for this
deficiency by using finer numerical grids. Practically,
this counter measure is limited by the available computing power. Even with the finest grids which can be
handled with today's computers, the numerical diffusion effects appearing in the U P W I N D solutions are
much larger than the real ones.
Consequently, more sophisticated numerical integration schemes must be applied which are much
more immune to numerical diffusion. Here we
discussed the TVD as a reasonable alternative. It is

626

A. Sokolichin et al.

shown that it provides results which are in the same


order of accuracy as the solutions obtained with the
Euler/Lagrange method which is not affected by numerical diffusion.
A comparison of the numerical solutions of the
model equations obtained with the TVD-technique
which can be used in Euler/Euler representations and
the LAGRANGE technique showed that the resulting
flow patterns agree quantitatively over a surprisingly
long period of simulation time. This is particularly
interesting since the model equations are capable of
instable chaotic solutions where two solutions with
slightly different initial conditions will not lead to the
same long-time results.
Consequently, the results obtained with the TVD
and the LAGRANGE technique can be regarded to
be equivalent. The results presented can also be considered as a kind of validation of both numerical
codes since both solution procedures are much different. On the other hand, it has not been possible to
obtain similar agreement even qualitatively with the
U P W I N D technique, which is the common
approach to handle the gas-phase motion in the
Euler/Euler approach.
The implementations of the TVD and the
LAGRANGE techniques differ considerably from
each other with respect to the computing times required. In systems which are not sensitive to diffusion,
the LAGRANGE technique is considerably faster
than the TVD technique, since for reasons of stability,
the TVD method requires much smaller time steps for
the integration of the gas holdup equation than the
LAGRANGE method. In cases where diffusion effects
cannot be neglected, the LAGRANGE method becomes less effective if the dispersion effects are to be
modeled by a diffusion-type continuous equation.
When large bubble numbers are to be considered,
the LAGRANGE method might become slower because of the big number of GPPs to be handled.
However, it proved to be possible to follow the trajectories of individual bubble clusters instead of single
bubbles.
In cases where the gas holdup is too large (> 10%)
none of the presented techniques can provide reliable
results since the bubble-bubble interactions must then
be taken into account. Presently, there is no reasonable physical model available for such situations.
Acknowledgement

Support of this work through Deutsche Forschungsgemeinschaft is gratefully acknowledged.

Cdrag

d
D
g
P
t
T
U

NOTATION
drag force coefficient, g/(cm 3 s)
disturbance coefficient, cm/s 1/2
diffusion coefficient, cm2/s
acceleration due to gravity, 981 cm/s 2
pressure, dyn/cm 2
time, s
stress tensor, dyn/cm 2
velocity vector, cm/s

Greek letters
6(x)
three-dimensional Dirac's delta function
e
gas holdup, dimensionless
/~
viscosity, g/(cm s)
p
density, g/cm 3
V
gradient operator, cm-1
Subscripts
eft
g
l

effective
gas phase
liquid phase
REFERENCES

Becker, S., Sokolichin, A. and Eigenberger, G. (1994)


Gas-liquid flow in bubble columns and loop reactors. Part II. Comparison of detailed experiments
and flow simulations. Chem. Engng Sci. 49,
5747-5762.
Devanathan, N., Dudukovic, M. P., Lapin, A. and
Ltibbert, A. (1995) Chaotic flow in bubble column
reactors. Chem. Engng Sci. 50, 2661-2667.
Fan, L S., Tsuchiya, K. (1990) Bubble Wake Dynamics
in Liquids and Liquid-Solid Suspensions. Butterworth-Heinemann, Boston, U.S.A.
Grienberger, J. and Hofmann, H. (1992) Investigations and modelling of bubble columns. Chem.
Engng Sci. 47, 2215-2220.
Harlow, F. H. and Welch, J. E. (1965) Numerical
calculation of time-dependent viscous incompressible flows of fluid with free surface. Phys. Fluids 8,
2182-2189.
Johansen, S. T. (1990) On the modelling of disperse
two-phase flows. Ph.D. Thesis, The Norwegian Institute of Technology, Trondheim.
Landau, L. D. and Lifschitz, E. M. (1971) Lehrbuch
der Theoretischen Physik, Vol. VI, Hydrodynamik,
2. Aufl. Akademie, Berlin.
Lapin, A. and Ltibbert, A. (1994) Numerical simulation of the dynamics of two-phase gas-liquid flow
in bubble columns. Chem. Engng Sci. 49, 3661 3674.
Levenspiel, O. (1989) The Chemical Reactor Omnibook. OSU Book Stores, Corvallis.
LeVeque, R. J. (1990) Numerical Methods for Conservation Laws. Birkh~iuser, Ziirich.
Patankar, Suhas V. (1980) Numerical Heat Transfer
and Fluid Flow. McGraw-Hill, New York, U.S.A.
Schwarz, M. P. and Turner, W. J. (1988) Applicability
of the standard k-e turbulence model to gas-stirred
baths. Appl. Math. Modelling 12, 273-279.
Sokolichin, A. and Eigenberger, G. (1994) Gas liquid
flow in bubble columns and loop reactors. Part I.
Detailed modelling and numerical simulation.
Chem. Engng Sci. 49, 5735-5746.
Torvik, R. and Svendsen, H. F. (1990) Modelling of
slurry reactors: a fundamental approach. Chem.
Engng Sci. 45, 2325-2332.
Tzeng, J.-W., Chen, R. C. and Fan, L.-S. (1993) Visualization of flow characteristics in a 2-D bubble column and three-phased fluidized bed. A.I.Ch.E.J.
39, 733 744.
Webb, C., Que, F. and Senior, P. R. (1992) Dynamic
simulation of gas-liquid dispersion behaviour in
a 2-D bubble column using a graphics mini-supercomputer. Chem. Engng Sci. 47, 3305-3312.

Você também pode gostar