Sie sind auf Seite 1von 10

Chemical Engineering Science 62 (2007) 5870 – 5879

www.elsevier.com/locate/ces

Prediction of pressure drop and liquid holdup in trickle bed reactor using
relative permeability concept in CFD
Arnab Atta, Shantanu Roy, K.D.P. Nigam ∗
Department of Chemical Engineering, Indian Institute of Technology Delhi, Hauz Khas, New Delhi 110 016, India

Received 6 January 2007; received in revised form 2 June 2007; accepted 11 June 2007
Available online 15 June 2007

Abstract
A Computational Fluid Dynamics (CFD) model based on porous media concept is presented to model the hydrodynamics of two-phase flow
in trickle-bed reactors (TBRs). The aim of this study is to develop a comprehensive CFD based model for predicting hydrodynamic parameters
in trickle-bed reactors under cold-flow conditions. The two-phase Eulerian model describing the flow domain as a porous region has been used
to simulate the macroscale multiphase flow in trickle beds operating under trickle flow regime using FLUENT 6.2 software. The closure terms
for phase interactions have been addressed by adopting the relative permeability concept [Sàez, A.E., Carbonell, R.G., 1985. Hydrodynamic
parameters for gas–liquid cocurrent flow in packed beds. A.I.Ch.E. Journal 31, 52–62]. The model has been evaluated by comparing predictions
with the data (collected under a varied set of laboratory conditions) available in the open literature. It is shown that while being relatively
simple in structure, this CFD model is flexible and predictive for a large body of experimental data presented in the open literature.
䉷 2007 Elsevier Ltd. All rights reserved.

Keywords: Trickle-bed; Porous media; Relative permeability; CFD

1. Introduction Numerous authors have studied and reported experimen-


tal data on pressure drop and liquid saturation in trickle-bed
Trickle bed reactors (TBRs) are multiphase reactors in which reactors (e.g. Specchia and Baldi, 1977; Rao et al., 1983;
gas and liquid phases flow co-currently downward over a solid Szady and Sundaresan, 1991; Al-Dahhan et al., 1997). An
catalyst packing. TBRs find widespread use in petroleum refin- exhaustive state of art reviews on hydrodynamic parameters of
ing, chemical and process industries, pollution abatement and TBR can be found in Saroha and Nigam (1996). The previous
biochemical industries. Design and scale up of TBRs contin- attempts for describing trickle bed hydrodynamics can be cate-
ues to be a major challenge for chemical engineers. The design gorized mainly into two different classes of work. The classical
and scale-up of trickle bed reactors depend on key hydrody- approach is empirical wherein correlations are developed to
namic variables such as liquid volume fraction (liquid satura- ‘fit’ the experimental data (e.g. Larkins et al., 1961; Ellman
tion), particle scale wetting and overall gas–liquid distribution. et al., 1988, 1990; Larachi et al., 1991). Another approach is
These variables are difficult to determine experimentally and to describe hydrodynamics in phenomenological manner, i.e.,
interactions between these are as yet poorly understood. Even assuming a simple picture of the microscale flow pattern, and
though numerous experimental studies have been reported in then integrate that depiction to address the entire bed (e.g.
measurement of these variables, predicting them from first prin- Holub et al., 1992; Iliuta et al., 2000).
ciple hydrodynamic simulations is difficult as yet and no co- Increasing computational power and development of Com-
herent and conclusive methodology for doing so has yet been putational Fluid Dynamics (CFD) has allowed promising appli-
espoused. cations of numerical simulations to the modeling of multiphase
flow in TBR (e.g. Attou and Ferschneider, 1999; Propp et al.,
2000; Souadnia and Latifi, 2001; Jiang et al., 2002a,b; Gunjal
∗ Corresponding author. Tel./fax: +91 11 26591020. et al., 2003, 2005). Table 1 summarizes these reported attempts
E-mail address: nigamkdp@gmail.com (K.D.P. Nigam). and highlights the key features in each of these efforts.
0009-2509/$ - see front matter 䉷 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2007.06.008
A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879 5871

Table 1
Summary of previous CFD models

S. No. Authors Key features of model Remarks

1. Attou and Ferschneider (1999) Model developed on the basis of area-averaged mass and As the model is one-dimensional, it cannot accommo-
momentum balance equations of each fluid date the variation of radial bed porosity distribution

Capillary pressure gradient was deduced from a momentum


balance analysis at the gas–liquid interface

The liquid–solid and gas–liquid interaction forces are for-


mulated on the basis of the Kozeny–Carman equation

2. Propp et al. (2000) Flow is assumed to be governed by equations of flow in The code was tested with several test problems, no
porous media explicit validation with experimental results for different
hydrodynamic parameters were presented
Use of high-resolution finite-difference methods to dis-
cretize governing equations.

Examined the effects of Ergun equation, capillary pressure


and variable porosity

3. Souadnia and Latifi (2001) One-dimensional computational model is used with the Development in 2D model required for better prediction
finite volume technique combined with Godunov’s method. of parameters

Drag forces are accounted for through the equations devel-


oped by Sàez and Carbonell (1985)

Porosity is assumed uniform and constant

4. Jiang et al. (2002a,b) Two fluid approach using CFDLIB (Los Alamos National Able to capture some of the key features of the hydro-
Laboratory) dynamics

Drag-exchange coefficients are calculated by the model of Bed structure implementation is involved
Attou et al. (1999)

Capillary pressure is incorporated via J-function

Bed structure implementation is resolved through statistical


implementation of sectional porosities

5. Gunjal et al. (2003) Two fluid approach using the closure of Attou et al. (1999) The first effort to simulate the reactor in 3D
have been used in 2D as well as 3D geometry

Liquid flow distribution and RTD were studied incorporat- Though the qualitative prediction of hysteresis was car-
ing the effect of capillary pressure and porosity variation ried out but the development in quantitative comparison
is required.

6. Gunjal et al. (2005) Extension of their own previous model (Gunjal et al., 2003) Though the simulation of periodic flow can be used
for simulating the spray flow regime and hysteresis on to understand some key features, still development is
pressure drop essential in this field for better prediction

Attempted to simulate the periodic flow

While the CFD models (Table 1) have tried to resolve the have advanced to a level (except perhaps the MRI imaging
existing complexities of the TBR to a great extent, consider- studies of Prof. Gladden’s group at Cambridge (Sederman and
able debate still persists on the exact nature of equations to be Gladden, 2001, 2005)) wherein the flow features predicted at
solved for TBR simulations. Depending on the problem for- the particle and liquid-rivulet scale in the TBR can be antic-
mulation and requirement of specific results, some of the open ipated with fidelity. Even in the MRI studies (Sederman and
issues include the number of phases to be solved, the nature Gladden, 2001, 2005; Lim et al., 2004), which can be indeed
of the model (steady state or transient), the formulation of the intriguingly sophisticated, the studies only aid in high resolu-
equations (Euler–Euler or Euler–Lagrangian), the level of de- tion flow visualization and a one-on-one comparison with the
tails in describing the packed bed (i.e., the particle scale mod- CFD computation is not possible. This is because while in the
els) and, the appropriate boundary and initial conditions to be experiments only a single realization of the flow is observed,
used (Ranade, 2002). Most of the literature available (Jiang the ‘averaged equations’ (which are in principle written for
et al., 2002a,b; Gunjal et al., 2003, 2005) dealing with packed the ensemble moments of the flow variables) represent an ‘av-
bed flow simulations use a three-phase Eulerian model in which erage’ of all possible realizations. Thus in the final analysis,
the solid velocity is identically set to zero. Such a calculation is one is left with comparing the average flow variables (such as
nevertheless computationally demanding and yet experiments overall pressure drop, holdup, etc.) and from a limited number
5872 A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879

of experiments, the profiles (radial, axial) of these variables. complex phenomena and it is difficult to develop a purely pre-
Therefore, whether getting into greater levels of detail with vol- dictive approach for the flow hydrodynamics. Therefore, the
ume and ensemble averaged equations (consequently making need of some fitting parameters is unavoidable.
the analysis more involved) is a justifiable direction of work, The closures used in this model are the relative perme-
is questionable. ability model developed by Sàez and Carbonell (1985). The
That being the present scenario, and the fact that even the fluid–fluid interaction model (developed theoretically, Attou
two-fluid CFD models have a lot of questionable assumptions and Ferschneider, 1999) has a notable feature of not having
(even though deceptively at a micro-scale), the wisdom of using any adjustable parameter in the closure but is found to predict
a highly computationally intensive model for predicting global results accurately when incorporated into the CFD framework.
profiles in a TBR may be questioned. In light of that, in this This aspect leads to the drawback of complicated incorpora-
work, a less computationally intensive, yet first-principle based tion of different particle size and shape effect in that model.
CFD model has been presented using the porous media concept. Hence, the relative permeability drag force model is rather log-
The independent experimental data sets, reported by Specchia ical and a promising alternative. This model has a remarkable
and Baldi (1977), Szady and Sundaresan (1991), Rao et al. feature of being flexible for different particle sizes which can
(1983) and the ANN model developed by Iliuta et al. (1999) be incorporated without much complexity.
were selected in the present work to validate the predictions. In
addition, we compare our results with the numerical simulation 2.1. Governing equations of the flow
of Gunjal et al. (2005), which is based on a three phase Eulerian
concept. The idea was to test our model against the state of the The volume-averaged equations for each flowing phase can
art in TBR modeling and offer it as a viable modeling approach. be written as

2. Modeling • Continuity equation:

The existing hydrodynamic models can be broadly classi- j( s )


fied into two different categories on the basis of empirical ap- + ∇( u ) = 0,  = g, l. (1)
jt
proach and theoretical or semi-empirical approach (presented in
Table 2). The model presented here to describe the multiphase • Momentum balance equation:
flow is based on the Eulerian framework, which consists of
 
continuity and momentum equations of each fluid phase with ju
 + u ∇u = − (∇p −  g)
appropriate closures. To relate the drag forces with the flow ve- jt
locities and volume fractions of each phase, and to the physical + ∇( + R ) + F , (2)
properties of the gas, liquid and solid phases, various closure
terms have been reported. These have been obtained either from where F is the total drag force per unit of bed volume ex-
first principles (e.g. fluid–fluid interfacial force model (Attou erted by the phase ;  and R are, respectively, the volume
et al., 1999)) or by semi-empirical approaches (e.g. relative per- averaged viscous stress tensor and the turbulence stress ten-
meability model (Sàez and Carbonell, 1985); slit model (Holub sor of phase . Note that in general, p is phase-specific and
et al., 1992, 1993; Al-Dahhan and Dudukovic, 1994; Al-Dahhan the same pressure is not shared by both phases.
et al., 1998; Iliuta and Larachi, 1999)). Recently, Larachi et al.
(2000) have compared these models for liquid holdup and pres- In order to solve these equations the assumptions taken in
sure drop to experimental data in the trickling regime. this model are
Carbonell (2000) has theoretically reviewed (with the help
1. There is no inter-phase mass transfer.
of mean absolute relative errors calculated by Larachi et al.,
2. Both the flowing fluids are incompressible.
2000) all the approaches to obtain drag expressions for two-
3. The porous medium is taken to be isotropic, i.e., permeabil-
phase flow model and concluded that the relative perme-
ities are independent of direction.
ability model (Sàez and Carbonell, 1985) and the fluid–fluid
4. The porosity is uniform and constant.
interaction model (Attou and Ferschneider, 1999) are based on
5. Trickle flow regime is the operating flow regime, i.e.,
solid hydrodynamic principles and are able to predict the hy-
gas–liquid interaction is low so capillary pressure force can
drodynamic parameters with acceptable accuracy. Once again
be neglected. This means that we assume same pressure for
it is noteworthy that the only closure model used frequently in
both phases at any point in time and space.
CFD calculations for simulating the flow through packed bed is
6. The contribution of the turbulent stress terms to overall mo-
fluid–fluid interaction model (Attou et al., 1999) in most of the
mentum balance equation (2) is not significant. This as-
previous simulations (e.g. Jiang et al., 2002a,b; Gunjal et al.,
sumption has also been used by other authors (e.g. Jiang
2003, 2005).
et al., 2002a).
In line with Carbonell (2000), the approach to be used in
this work is based on the assumption that flow domain (fixed Inter-phase coupling terms accounted by Eq. (2) are based on
bed with catalyst particles) can be described as porous media. relative permeability concept developed by Sàez and Carbonell
The two-phase flow phenomena through porous media are very (1985), as discussed below.
A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879 5873

Table 2
Various approach to model hydrodynamics of TBR

Approach References Remarks

Empirical Larkins et al. (1961), Sai and Varma (1987), Based on dimensional analysis to produce ex-
Ellman et al. (1988, 1990), Larachi et al. plicit correlations for pressure drop and liquid
(1991), Xiao et al. (2000, 2001) and Pina holdup using flow variables and packing charac-
et al. (2001) teristics or using Lockhart–Martinelli parameter
which was proposed for open horizontal tubes

Theoretical or semi- Relative permeability model Sàez and Carbonell (1985) Ergun equation has been modified to account
empirical for the existence of a second flowing phase by
incorporation of relative permeability which has
been correlated as a function of liquid saturation
of each phase depending on the experimental
results

Slit model Holub et al. (1992, 1993), Al-Dahhan and It is also a modified form of the Ergun equa-
Dudukovic (1994), Al-Dahhan et al. (1998), tion. The flow through complex geometry of the
Iliuta and Larachi (1999) actual void space in the catalyst bed at the pore
level has been modeled by the much simpler
flow inside a rectangular slit assuming the width
of the slit is a function of bed porosity the angle
of inclination of the slit to the vertical axis is
related to a tortuosity factor for the packed bed.
The surface area per unit volume of solid in this
rectangular slit was made equal to the surface
area per unit volume of solid in the reactor

Model based on fundamental Tung and Dhir (1988), Narasimhan et al. The liquid–gas interfacial drag has been taken
force balance (2002) into account by first developing an expression
for the drag on a single bubble/slug and then
multiplying it by the number of bubbles/slugs
per unit volume of the porous layer

Fluid–fluid interfacial force Attou et al. (1999) To avoid the empiricism of the relative perme-
model ability and slit models, this model was developed
in which the drag force on each phase has con-
tributions from the particle-fluid interactions as
well as from fluid–fluid interactions. The model
was derived from a momentum balance on the
fluid flow around fully wetted particles

2.2. Relative permeability model cause when two fluids are simultaneously present in a porous
medium, one fluid’s ability to flow will be guided by the mi-
The concept of relative permeability is very commonly used croscopic configuration of the other fluid. In general, the ex-
to the various problems of multiphase flow through porous me- pression for drag force for single-phase flow is modified using
dia. Essentially, it is a concept that stems from the classical certain parameter (named as relative permeability, k , of that
Darcy’s Law, a macroscopic equation based on average quan- phase) to accommodate the presence of second phase (Sàez and
tities for modeling pressure drop through a porous medium at Carbonell, 1985). Sàez and Carbonell (1985) have modified
fixed superficial velocity for one phase flow. If a fluid of viscos- the Ergun equation for the single-phase pressure drop to calcu-
ity  is passing through a porous medium of absolute perme- late the two-phase flow pressure drop which can be represented
ability k in a homogeneous gravitation field with the flow rate in dimensionless form with the help of Reynolds and Galileo
U, then the pressure gradient ∇p across the medium is given numbers:
by Darcy’s equation  
F 1 Re Re2
k = A +B  g. (4)
U = − (∇p − g), (3)  k Ga  Ga 

where g denotes the acceleration due to the gravitational forces This model accounts for the reported non-linearity in the pres-
and  is the density of that single phase fluid. For fluid flow in sure gradient as a function of velocity (MacDonald et al., 1979),
horizontal direction g can be neglected. using the concept of relative permeability (Sàez and Carbonell,
While describing two-phase flow in porous media, it be- 1985). The constants A and B in Eq. (4) are the Ergun equation
comes very essential to modify the above stated equation be- coefficients for single-phase flow in the packed bed (Ergun,
5874 A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879

1952). The Reynolds and Galileo numbers are defined as While deriving these equations, Sàez and Carbonell (1985) as-
sumed that flow is one-dimensional and the liquid holdup does
  u de
Re = , not change along the bed length, which implies that pressure
 (1 − ) gradients in the liquid and gas phases are equal. Therefore, the
2 gd 3e 3 capillary pressure terms are neglected. Again by subtracting
Ga  = ,
2 (1 − )3 the momentum equation of gas and liquid phases from each
6Vp other, and assuming that the liquid density is much greater than
where de = . (5) gas density, the equation for calculating liquid holdup can be
Ap
found as
In order to consider the microscopic/local configuration of the    
1 Rel Re2l 1 Reg Re2g g
second fluid and to define the ability to flow of one fluid in A +B − A +B = 1. (10)
presence of other fluid, the term relative permeability (k ) was kl Ga l Ga l kg Ga g Ga g l
introduced. From the previous discussions as well as from Eq.
(4), it can be observed that the relative permeability corrects More detailed derivation of these equations can be found else-
the drag force expression for single-phase flow conditions to where (Sàez and Carbonell, 1985). Regarding the dependency
account for the flow of two phases. With this perception it can as well as sensitivity of relative permeabilities on different pos-
be stated that it is a different approach to account fluid–fluid sible parameters, very recently, Nemec and Levec (2005) have
interfacial drag forces in order to achieve the same goal i.e., shown through their extensive experimentation and analysis for
pressure drop and holdup. a wide rage of operating conditions and with typical shapes
Again since the relative permeability parameter has been and sizes of particles encountered in commercial trickle-bed
incorporated to accommodate the presence of a second phase, reactors, that relative permeabilities are solely functions of the
essentially it will be a function of phase saturation or holdup of corresponding phase saturation. Before making this conclusion
that corresponding phase. To determine the dependence of the these authors have carefully explored the effects of uncertain-
relative permeability on the saturation for each phase Sàez and ties associated with the phase relative permeabilities and also
Carbonell (1985) analyzed different data sets for liquid holdup have carried out the detailed study on the phenomenological
and pressure drop over a wide range of Reynolds and Galileo insights of the suitable correlations, e.g. the effect of particle
numbers in packed beds available in the literature till that time. shape and size, effect of flow rate and reactor pressure. Inter-
They made the hypothesis that liquid relative permeabilities are estingly, they have opposed the observation by Lakota et al.
only a function of reduced saturation (l ), which is represented (2002) on the particle shape dependency of the gas phase rel-
by the ratio of effective volume of flow of the liquid phase to ative permeability. They have argued that the effect of shape
the available volume of flow considering that the static liquid factor is accounted by Ergun constants however the relative per-
holdup (0l ) represents a portion of the void fraction occupied meability being the ratio between single and two-phase pres-
by stagnant liquid. Thus sure drop, this shape effect has been already taken care in that
respect.
l − 0l
l = . (6)
 − 0l 3. Boundary conditions and numerical solution

The gas phase relative permeability was correlated as a function Considering a two-dimensional axisymmetric domain
of the gas phase saturation. The empirical correlations were (Fig. 1), the above set of model equations were solved using
reported to be (Sàez and Carbonell, 1985): commercial software FLUENT 6.2 (of ANSYS. Inc., USA)
defining the solution domain as porous. The gas phase was
kl = 2.43
l ,
treated as primary phase and liquid phase was considered as
kg = sg4.80 , (7) secondary phase.
At the inlet, flat velocity profile for gas and liquid phases was
0
where sg = 1 − l . assumed and implemented. No slip boundary condition was set
The static liquid holdup (0l ) can be calculated by the fol- for all the impermeable reactor walls. At the bottom of the col-
lowing correlation given by Sàez and Carbonell (1985) umn, an outlet boundary condition was specified. The reference
pressure equal to atmospheric pressure was fixed at the outlet.
1 l gd 2p 2 As a patch for the initial conditions, the overall volume fraction
0l = where Eo∗ = . (8)
(20 + 0.9Eo∗ ) l (1 − )2 of the liquid phase was estimated using the correlation given
by Eq. (10). Unsteady state simulations were carried out with
After simplifying these expressions for a given particle diameter the time step of 0.005 s. Some preliminary numerical experi-
and the flow rates of gas and liquid flows, the following equation ments were carried out to identify the required number of com-
of motion can be used to compute the pressure drop: putational cells to obtain grid independent results. It was also
    ensured with the preliminary numerical experiments to have
p Fg 1 Reg Re2g discretization scheme independent results. These simulations
= = A +B g g. (9)
l g kg Ga g Ga g confirmed that the grid size taken was satisfactory, as further
A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879 5875

Air-water inlet

1.6 m

500 cells
for
1m
Axis B.C.

1.2 m
b
1.8

A B
1.6
1.4

height of the bed (m)


25 cells radially 1.2
Inlet

1 This work_outlet
Anderson and Sapre (1991)_outlet
Outlet 0.8

Fig. 1. Computational domain. 0.6

0.4
Table 3 0.2
Simulation parameters
G' C E F D G
0
Discretization scheme First order UPWIND -0.9 -0.6 -0.3 0 0.3 0.6 0.9
Pressure–velocity coupling SIMPLE algorithm Distance from the centre of the bed (m)
Relaxation parameters
Pressure 0.3
Fig. 2. (a) Computational domain used by Anderson and Sapre (1991). (b)
Momentum 0.7
Quantification of liquid spreading for constricted inlet at the top.
Volume fraction 0.2
Body forces 1.0
Convergence criterion 10−5
the geometry considered by Anderson and Sapre (1991) (where
the ratio between bed diameter to the particle diameter is 400)
(Fig. 2a). Following the correlations proposed by Cohen and
increase in number of grids did not appreciably affect the pre- Metzner (1981) for incorporation of porosity distribution, we
dicted results. The simulation parameters are summarized in have observed that the porosity variation/fluctuation ceases to
Table 3. exist after almost eight particle diameters from the bed wall.
Therefore, it can be concluded that incorporation of porosity
4. Results and discussion variation in this particular case will be ineffective as there is
a minimum possibility of liquid spreading near the wall for
Initially some numerical experiments were carried out to constricted inlet at the top. Taking the similar flow condition of
endorse the fact that in the real flow, the flow characteristic Anderson and Sapre (1991), we have completed our case study
is 2D or 3D even if the initial liquid inlet is uniform (1D). and it (Fig. 2b) shows that for constricted inlet at the top (line
This was to establish the versatility of our model and to make AB) there is some degree of spreading of liquid and the width
sure that we can indeed use the model in 2D and 3D scenarios has been quantified by this present CFD model (line CD). EF
without compromising its robustness, even though most of the represents the result of Anderson and Sapre (1991) through
experimental validation we present later in this work relate to which 80% of the flow is taking place where GG depicts the
1D effects. width of the bed. This prediction of liquid spreading would
One of the case studies involves liquid flow introduced over not have been possible without taking 2D model equations.
only a small part of the top of the bed which is identical to However, it may be noted that these simulations were carried
5876 A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879

Table 4
Details of operation conditions used for simulations

S. No. Source Bed diam- Bed l/D ratio Particle di- D/dp ratio Bed poros- Gas veloc- Liquid
eter, D (m) length, l ameter, dp ity ity (m s−1 ) velocity
(m) (m) (m s−1 )

1. Szady and Sundaresan (1991) 0.165 1.49 9.03 0.003 55 0.37 0.22 0.002–0.008
2. Specchia and Baldi (1977) 0.08 1.05 13.13 0.0027 29.63 0.38 0.2–0.8 0.0028
3. Rao et al. (1983) 0.0924 1.835 19.86 0.00627 14.77 0.373 0.13–0.95 0.004128
4. ANN Model 0.165 1.49 9.03 0.003 55 0.37 0.22 0.0004–0.006
5. Gunjal et al. (2005) 0.114 1 8.77 0.006 19 0.37 0.22 0.0017–0.0092

18 0.6
This work This work
Szady and Sundaresan (1991) Szady and Sundaresan (1991)
15 ANN
Gunjal et al. (2005) simulation
0.5 ANN
Gunjal et al. (2005) simulation
Pressure drop (kPa.m-1)

12

Liquid saturation (-)


0.4

9
0.3

6
0.2
3
0.1
0
0 0.002 0.004 0.006 0.008 0.01
Liquid Velocity (m.s-1) 0
0 0.002 0.004 0.006 0.008 0.01
Fig. 3. Comparative study of pressure drop with literature data Liquid Velocity (m.s-1)
(Vg = 0.22 m s−1 ).
Fig. 4. Comparative study of liquid saturation with literature data
(Vg = 0.22 m s−1 ).

out without taking porosity distribution in the bed. With this


of variation of the pressure drop and liquid saturation with dif-
understanding and perception, further numerical simulations
ferent liquid superficial velocity at a constant gas superficial
were carried out for different geometries and flow conditions on
velocity of 0.22 m s−1 . Figs. 5 and 6 show the model predic-
Sun Fire V880 server accounting uniform and constant porosity
tion for pressure drop and liquid saturation for experimental
of the bed for simplicity of the computation/model.
data of Specchia and Baldi (1977) for different gas superficial
The simulated results were first validated against vari-
velocities at a constant liquid velocity of 0.0028 m s−1 . In all
ous published experimental data of pressure drop and liquid
cases the results were also verified for accuracy with the nu-
holdup. The details of the experimental data set are given in
merical prediction based on the three-phase Eulerian simula-
Table 4.
tion carried out by Gunjal et al. (2005). It is noteworthy that
Two-phase pressure drop per unit length and dynamic liquid
while carrying out the simulations Gunjal et al. (2005) varied
holdup at different liquid and gas flow rates were obtained by a
their choice of Ergun’s constants for different data sets (e.g.,
series of numerical simulations to compare with the above men-
A = 180 and B = 1.8 for the data set of Szady and Sundaresan
tioned data sets. The values of Ergun’s constants (A and B) used
(1991); A = 500 and B = 3 for the data set of Specchia and
in the closure model are 180 and 1.8 for all the numerical sim-
Baldi (1977)). Again, they seem to have used the definition of
ulations. Ad hoc fitting of these parameters was not required.
Eötvos no. (Eo∗ ) while calculating static liquid holdup as
Figs. 3 and 4 show the comparison of simulated results with
the experimental data of Szady and Sundaresan (1991) and nu- l gd 2p 2
merical prediction of Gunjal et al. (2005) for observed pressure Eo∗ = (11)
l (1 − 2 )2
drop per unit length and total liquid saturation. While adopting
the experimental data set from Szady and Sundaresan (1991), in which 2 has been used instead of  in the denominator.
only the upper branch of pressure drop curve was taken, as that The simulations in this work are then validated with the
was the case for prewetted bed where capillary pressure can be experimental results of Rao et al. (1983) for pressure drop
neglected. Both the model results show reasonable agreement with varying gas flow rates (Fig. 7). As is clear, the same
A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879 5877

45 5
4.5
40
4

Pressure drop (kPa.m-1)


35
Pressure drop (kPa.m-1)

3.5
30 3

25 2.5
2
20
1.5
Gunjal et al. (2005) experiment
15 1
Gunjal et al. (2005) simulation
10 Specchia and Baldi (1977) 0.5 This work
This work 0
5 0 0.002 0.004 0.006 0.008 0.01 0.012
Gunjal et al. (2005) simulation
0 Liquid velocity (m.s-1)
0 0.2 0.4 0.6 0.8 1
Gas velocity (m.s-1) Fig. 8. Comparative study of pressure drop with 3-phase simulation
(Vg = 0.22 m s−1 ).
Fig. 5. Comparative study of pressure drop with literature data
(Vl = 2.8 kg m−1 s−1 ).

model seems to predict experimental observations from various


sources collected over a wide variety of conditions.
Finally, we compare our model prediction against the ANN
0.45 models of Iliuta et al. (1999). Iliuta et al. (1999) developed
0.4 state-of-art explicit correlations for hydrodynamic parameters
0.35 of TBR based on a large database of experimental observations
Liquid saturation (-)

0.3 upon the combination of dimensionless analysis and artificial


0.25
neural networks. This large data base was then used to train a
hybrid artificial neural network (ANN) model so that the ‘best’
0.2
predictions of pressure drop, holdup and wetting efficiency to a
0.15 given TBR could be predicted. This ANN model is also able to
Specchia and Baldi (1977)
0.1 predict the trickle-pulse transition boundary. The comparison
This work
0.05 of our simulated results with this model (Figs. 3 and 4) shows
Gunjal et al. (2005) simulation
0
that the predictions are satisfactory. In the present study, the
0 0.2 0.4 0.6 0.8 1 assumption of neglecting the capillary pressure (i.e., the liquid
Gas velocity (m.s-1) holdup is practically independent of the axial position in the
bed) leads to the fact that this model is strictly valid for low
Fig. 6. Comparative study of liquid saturation with literature data interaction regime i.e., trickling flow regime. This is evident
(Vl = 2.8 kg m−1 s−1 ). from the validation of predicted results with the results of ANN
model (Figs. 3 and 4). It shows that at low liquid velocity the
prediction is more accurate and as the liquid velocity increases
the regime moves towards the transition region where the model
18 equations cease to hold. It is noteworthy that through this com-
16 parison, we are effectively comparing our model against the
Pressure drop (kPa.m-1)

14 extensive experimental database of Iliuta et al. (1999).


12
Thus, even without changing the Ergun’s constants, the
10
present model can be used to predict pressure drop and liquid
saturation in low interaction regime with enough confidence.
8
Figs. 8 and 9 support this statement, which show the compari-
6
This work son of predicted pressure drop and liquid holdup for different
4
liquid velocity with the CFD results of Gunjal et al. (2005) and
2 Rao et al. (1983) their in-house experimental data for spherical particle of 6 mm
0
0 0.2 0.4 0.6 0.8 1 1.2
diameter with A=500 and B =2.4. In these cases (Figs. 8 and 9)
the gas superficial velocity was kept constant at 0.22 m s−1
Gas velocity (m.s-1)
and diameter of the bed was 0.114 m. Our model functions as
Fig. 7. Comparative study of pressure drop with literature data arguably well as their model in predicting their experimental
(Vl = 4.128 kg m−1 s−1 ). data, in fact, the pressure drop comparisons are better.
5878 A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879

0.29
Notation
Gunjal et al. (2005) experiment
Gunjal et al. (2005) simulation
A constant in the viscous term of the Ergun type
0.24 This work equation
particle surface area, m2
Liquid holdup

Ap
0.19 B constant in the inertial term of the Ergun type
equation
0.14 de equivalent particle diameter, 6Vp /Ap
dp particle diameter, m
Eo∗ modified Eötvos number, l gd 2p 2 /l (1 − )2
0.09
F drag force on the  phase per unit volume,
kg m−2 s−2
0.04 g gravitational acceleration, m s−2
0 0.002 0.004 0.006 0.008 0.01 0.012
Ga  Galileo number of the  phase, 2 gd 3e 3 /2 (1 − )3
Liquid velocity (m.s-1)
k relative permeability of  phase
Fig. 9. Comparative study of liquid holdup with 3-phase simulation l length of the reactor
(Vg = 0.22 m s−1 ). p pressure, Pa
Re Reynolds number of the  phase,  u de / (1 − )
s saturation of the  phase
u superficial velocity of the  phase, m s−1
5. Conclusions
U flow velocity, m s−1
In this study, we have developed a two-phase Eulerian CFD Vp particle volume, m3
model based on porous media concept to simulate gas–liquid Greek letters
flow through packed beds. The porous media model is advan-
tageous to handle gas–liquid interaction terms due to its ability l reduced saturation of liquid phase, l − 0l / − 0l
to lump the adjustable parameters as compared to the conven- 0l static liquid holdup
tional k-fluid CFD treatment of the problem.  bed voidage
The closures used in this model are the relative permeability  holdup of  phase
model developed by Sàez and Carbonell (1985). The predicted  viscosity, Pa s
results are verified for different sets of independent experimen-  density of  the phase, kg m−3
tal data (Szady and Sundaresan, 1991; Specchia and Baldi,  surface tension, N m−1
1977; Rao et al., 1983) and results obtained by ANN model Subscripts
(Iliuta et al., 1999) which incorporates a myriad variety of ex-
perimental information. Predicted values showed good agree-  gas/liquid phase
ment with the experimental data. The predicted results are also g gas phase
compared with the numerical results of Gunjal et al. (2005), l liquid phase
which are based on the three-phase Eulerian simulation.
As discussed earlier, we can propose this model for future Acknowledgment
studies on prediction of hydrodynamic parameters under high-
pressure operation provided the suitable correlations are avail- A. Atta is indebted to the All India Council for Technical Ed-
able (e.g. namely the new correlations developed for relative ucation (AICTE), India, for providing National Doctoral Fel-
permeabilities by Nemec and Levec, 2005), which can be in- lowship.
corporated in this present CFD model as a modification of the
References
closure.
One must however approach the problem of flow modeling Al-Dahhan, M.H., Dudukovic, M.P., 1994. Pressure drop and liquid holdup
in TBRs with caution. Many decades have been spent in devel- in high-pressure trickle bed reactors. Chemical Engineering Science 49,
opment of TBR technology, prior to the development of CFD, 5681–5698.
and the practice in the industry for designing TBRs is well es- Al-Dahhan, M.H., Larachi, F., Dudukovic, M.P., Laurent, A., 1997. High-
pressure trickle-bed reactors: a review. Industrial & Engineering Chemistry
tablished (even if based on heuristics in many cases and not
Research 36, 3292–3314.
totally scientifically based). Any further developments that we Al-Dahhan, M.H., Khadilkar, M.R., Wu, Y., Dudukovic, M.P., 1998. Prediction
may want to propose with CFD must be consistent with that of pressure drop and liquid holdup in high-pressure trickle-bed reactors.
prior knowledge. As such, the present contribution is the first Industrial & Engineering Chemistry Research 37, 793–798.
of a series of papers wherein we demonstrate the use of porous Anderson, D.H., Sapre, A.V., 1991. Trickle-bed reactor flow simulation.
A.I.Ch.E. Journal 37, 377–382.
media approach as a viable CFD method that is both consistent Attou, A., Ferschneider, G., 1999. A two-fluid model for flow regime transition
with prior know-how as well as reveals new information from in gas–liquid trickle-bed reactors. Chemical Engineering Science 54,
a design and scale-up perspective. 5031–5037.
A. Atta et al. / Chemical Engineering Science 62 (2007) 5870 – 5879 5879

Attou, A., Boyer, C., Ferschneider, G., 1999. Modeling of the hydrodynamics Larkins, R.P., White, R.R., Jeffery, D.W., 1961. Two-phase concurrent flow
of the cocurrent gas–liquid trickle flow through a trickle-bed reactor. in packed beds. A.I.Ch.E. Journal 7, 231–239.
Chemical Engineering Science 54, 785–802. Lim, M.H.M., Sederman, A.J., Gladden, L.F., Stitt, E.H., 2004. New insights
Carbonell, R.G., 2000. Multiphase flow models in packed beds. Oil & Gas to trickle and pulse flow hydrodynamics in trickle-bed reactors using MRI.
Science and Technology 55, 417–425. Chemical Engineering Science 59, 5403–5410.
Cohen, Y., Metzner, A.B., 1981. Wall effects in laminar flow of fluids through MacDonald, I.F., El-Sayed, M.S., Mow, K., Dullien, F.A.L., 1979. Flow
packed beds. A.I.Ch.E. Journal 27, 705–715. through porous media—The Ergun equation revisited. Industrial &
Ellman, M.J., Midoux, N., Laurent, A., Charpentier, J.C., 1988. A new Engineering Chemistry Fundamentals 18, 199–208.
improved pressure drop correlation for trickle-bed reactors. Chemical Narasimhan, C.S.L., Verma, R.P., Kundu, A., Nigam, K.D.P., 2002. Modeling
Engineering Science 43, 2201–2206. hydrodynamics of trickle-bed reactors at high pressure. A.I.Ch.E. Journal
Ellman, M.J., Midoux, N., Wild, G., Laurent, A., Charpentier, J.C., 1990. A 48, 2459–2474.
new improved liquid holdup correlation for trickle bed reactors. Chemical Nemec, D., Levec, J., 2005. Flow through packed bed reactors: 2. Two-phase
Engineering Science 45, 1677–1684. concurrent downflow. Chemical Engineering Science 60, 6958–6970.
Ergun, S., 1952. Fluid flow through packed columns. Chemical Engineering Pina, D., Tronconi, E., Tagliabue, L., 2001. High interaction regime Lockhart-
Progress 48, 89–94. Martinelli model for pressure drop in trickle-bed reactors. A.I.Ch.E. Journal
Gunjal, P.R., Ranade, V.V., Chaudhari, R.V., 2003. Liquid distribution and 47, 19–30.
RTD in trickle bed reactors: experiments and CFD simulations. Canadian Propp, R.M., Colella, P., Crutchfield, W.Y., Day, M.S., 2000. A numerical
Journal of Chemical Engineering 81, 821–830. model for trickle bed reactors. Journal of Computational Physics 165,
Gunjal, P.R., Kashid, M.N., Ranade, V.V., Chaudhari, R.V., 2005. 311–333.
Hydrodynamics of trickle-bed reactors: experiments and CFD modeling. Ranade, V.V., 2002. Computational Flow Modeling for Chemical Reactor
Industrial & Engineering Chemistry Research 44, 6278–6294. Engineering. Academic Press, London.
Holub, R.A., Dudukovic, M.P., Ramachandran, P.A., 1992. A Rao, V.G., Ananth, M.S., Varma, Y.B.G., 1983. Hydrodynamics of two-phase
phenomenological model for pressure drop, liquid holdup, and flow regime cocurrent downflow in packed beds. A.I.Ch.E. Journal 29, 467–483.
transition in gas–liquid trickle flow. Chemical Engineering Science 47, Sàez, A.E., Carbonell, R.G., 1985. Hydrodynamic parameters for gas–liquid
2343–2348. cocurrent flow in packed beds. A.I.Ch.E. Journal 31, 52–62.
Holub, R.A., Dudukovic, M.P., Ramachandran, P.A., 1993. Pressure drop, Sai, P.S.T., Varma, Y.B.G., 1987. Pressure drop in gas–liquid downward flow
liquid holdup, and flow regime transition in trickle flow. A.I.Ch.E. Journal through packed beds. A.I.Ch.E. Journal 33, 2027–2036.
39, 302–321. Saroha, A.K., Nigam, K.D.P., 1996. Trickle bed reactors. Reviews in Chemical
Iliuta, I., Larachi, F., 1999. The generalized slit model: pressure gradient, Engineering 12, 207–347.
liquid holdup and wetting efficiency in gas–liquid trickle flow. Chemical Sederman, A.J., Gladden, L.F., 2001. Magnetic resonance imaging as a
Engineering Science 54, 5039–5045. quantitative probe of gas–liquid distribution and wetting efficiency in
Iliuta, I., Larachi, F., Grandjean, B.P.A., Wild, G., 1999. Gas–liquid interfacial trickle-bed reactors. Chemical Engineering Science 56, 2615–2628.
mass transfer in trickle-bed reactors: state of-art correlations. Chemical Sederman, A.J., Gladden, L.F., 2005. Transition to pulsing flow in trickle-bed
Engineering Science 54, 5633–5645. reactors studied using MRI. A.I.Ch.E. Journal 51, 615–621.
Iliuta, I., Larachi, F., Al-Dahhan, M.H., 2000. Double-slit model for partially Souadnia, A., Latifi, M.A., 2001. Analysis of two-phase flow distribution in
wetted trickle flow hydrodynamics. A.I.Ch.E. Journal 46, 597–609. trickle-bed reactors. Chemical Engineering Science 56, 5977–5985.
Jiang, Y., Khadilkar, M.R., Al-Dahhan, M.H., Dudukovic, M.P., 2002a. CFD Specchia, V., Baldi, G., 1977. Pressure drop and liquid holdup for two
of multiphase flow in packed-bed reactors: I. k-fluid modeling issues. phase concurrent flow in packed beds. Chemical Engineering Science 32,
A.I.Ch.E. Journal 48, 701–715. 515–523.
Jiang, Y., Khadilkar, M.R., Al-Dahhan, M.H., Dudukovic, M.P., 2002b. CFD Szady, M.J., Sundaresan, S., 1991. Effect of boundaries on trickle-bed
of multiphase flow in packed-bed reactors: II. Results and applications. hydrodynamics. A.I.Ch.E. Journal 37, 1237–1241.
A.I.Ch.E. Journal 48, 716–730. Tung, V.X., Dhir, V.K., 1988. A hydrodynamic model for two phase flow
Lakota, A., Levec, J., Carbonell, R.G., 2002. Hydrodynamics of trickling through porous media. International Journal of Multiphase Flow 14,
flow in packed beds: relative permeability concept. A.I.Ch.E. Journal 48, 47–65.
731–738. Xiao, Q., Anter, A.M., Cheng, Z.M., Yuan, W.K., 2000. Correlations for
Larachi, F., Laurent, A., Midoux, N., Wild, G., 1991. Experimental study of dynamic liquid holdup under pulsing flow in a trickle-bed reactor. Chemical
a trickle bed reactor operating at high pressure: two-phase pressure drop Engineering Journal 78, 125–129.
and liquid saturation. Chemical Engineering Science 46, 1233–1246. Xiao, Q., Cheng, Z.M., Jiang, X., Anter, A.M., Yuan, W.K., 2001.
Larachi, F., Iliuta, I., Al-Dahhan, M.A., Dudukovic, M.P., 2000. Hydrodynamics behavior of a trickle bed reactor under forced pulsing
Discriminating trickle-flow hydrodynamic models: some recommendations. flow. Chemical Engineering Science 56, 1189–1195.
Industrial & Engineering Chemistry Research 39, 554–556.

Das könnte Ihnen auch gefallen