You are on page 1of 9

Chemical Engineering Science 76 (2012) 49–57

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Science


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

Numerical simulation of the liquid distribution in a trickle-bed reactor


M. Martı́nez a, J. Pallares a,n, J. López a, A. López b, F. Albertos b, M.A. Garcı́a b, I. Cuesta a, F.X. Grau a
a
University Rovira i Virgili, Department of Mechanical Engineering, Av. Paı̈sos Catalans 26, 43007 Tarragona, Spain
b
Repsol Technological Center, Carretera de Extremadura PK 18, 28931 Móstoles, Spain

a r t i c l e i n f o abstract

Article history: Numerical simulations of the two-phase flow distribution in a trickle-bed reactor used for fuel
Received 13 January 2012 hydrodesulfurization are reported. As a first step, the heat and mass transfer, as well as the chemical
Received in revised form reactions, are not considered. The reactor has four packed-beds and a distribution tray above the
1 April 2012
catalytic beds equipped with cylindrical chimneys. The flow distribution at the outlet of the circular
Accepted 11 April 2012
chimney predicted by the simulations is not axisymmetric because of the spatial distributions of
Available online 19 April 2012
the liquid and gas inlets in the chimney. This causes that the liquid entering the packed beds is
Keywords: distributed in three main streams. For the simulation of the two-phase flow in the packed beds, an
Multiphase flow Eulerian three-phase model that considers the particles of catalyst as a granular static phase has been
Multiphase reactors
used following the Holub single slit model for particle–fluid interaction to compute the liquid–solid and
Computation
gas–solid drag coefficients. Numerical simulations of the dispersion of water–air flow in a column of
Hydrodynamics
Liquid distribution glass beads using this model were initially carried out and results were found to be in reasonable
Packed bed agreement with numerical and experimental data available in the literature. The simulations consider
the flow dispersion in the central region of the reactor bed as well as in the region close to the
cylindrical lateral wall of the reactor. In both cases most of the liquid spreading takes place in the top
part of the bed. The distributions of the liquid volume fraction do not change significatively as the
depth of the bed is increased except in the third bed and at the interface of two beds with different
porosity.
& 2012 Elsevier Ltd. All rights reserved.

1. Introduction chimney and the two-phase flow distribution at the outlet of the
chimney is strongly affected by the trajectories of the jets.
Trickle-bed reactors are commonly used in the petrochemical These trajectories depend on the spatial localization and velo-
industry for fuel hydrodesulfurization processes. In these type of cities of the liquid inlets and the drag force produced by the gas
reactors gas and liquid flow co-currently downward through a on the liquid jets. The fluid dynamics of liquid jets and the effect
packed bed of catalytic solid particles. A uniform distribution of of the gas cross flow on the jet trajectory and breakup have
the two-phase flow is important in these reactors because flow received considerable attention because of its implications in, for
non-uniformities in the bed produce the reduction of the effec- example, fuel injection and chemical mixing (Gopalan et al., 2004;
tiveness of the bed and may lead to the formation of hot spots. Wu et al., 1997). Harter et al. (2001) and Raynal and Harter (2001)
An extensive review of multiphase reactors can be found in studied numerically and experimentally the liquid and gas dis-
Dudukovic et al. (1999). tribution in a chimney of a distribution tray similar to that
Distribution trays are used in trickle-bed reactors to improve considered in this study. These authors compared the overall
liquid and gas mixing before entering the beds. Some types of liquid flow topology in the chimney predicted numerically using
distribution trays, like the one used in this study, are equipped the Volume of Fluid (VOF) model (Hirt and Nichols, 1981) with
with elements which help to improve phase mixing, as for experimental flow visualizations. They also observed that the
example cylindrical chimneys. The chimneys installed on the tray trajectories of the jets inside the chimney were strongly dependent
usually have several lateral inlets of liquid while the gas enters on the physical properties of the liquid.
through the top of the chimney. The flow topology inside the The interest to obtain a uniform well mixed two-phase flow in
chimney is important because it determines the flow distribution catalytic beds has motivated most of the studies carried out to
at the inlet of the packed bed. Liquid jets are formed inside the determine the liquid spreading in packed beds. Baker et al. (1935)
reported experimental studies in columns of different sizes and
filled with particles of different shapes and sizes to evaluate the
n
Corresponding author. Tel.: þ34 977 559 682; fax: þ 34 977 559 691. radial liquid spreading. These authors found that initial uniform
E-mail address: jordi.pallares@urv.cat (J. Pallares). distribution is essential, since the flow from a single stream needs

0009-2509/$ - see front matter & 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ces.2012.04.017
50 M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57

a length equivalent to four or five column diameters to become


uniform. Cihla and Schmidt (1957) proposed a series of analytical
solutions to an advection–diffusion equation to predict the liquid
spreading in a column, for different types of inlet distributors.
The mathematical model assumes a radial diffusion term with an
adjustable diffusion coefficient and an advection term along the
axial direction. Tsochatzidis et al. (2002) studied experimentally
the effects of the flow maldistribution at the inlet of a packed bed
column on the radial liquid distribution and determined, as Baker
et al. (1935) had already hinted, that progressively the flow
becomes more uniform. They found that the length needed to
achieve a uniform radial liquid distribution is reduced with
increasing flow rates. Boyer and Fanget (2002) measured the
gas–liquid flow distribution using gamma-ray tomography in a
packed bed for different inlet flow distributions. The visualiza-
tions of the liquid fraction showed a quantitative agreement with
a hydrodynamic model based on Kozeny–Carman formalism (see
for example Boyer and Fanget, 2002). Jiang et al. (2002a,b)
simulated the macroscale multiphase flow in packed beds and
compared the results with available experimental data.
The CFDLIB package (Gaffney and Kashiwa, 2003), based on the Fig. 1. Sketch of the reactor.
Eulerian k-fluid model, was used in their simulations. The Holub
et al. (1992) model for particle–fluid interaction and the Attou
and Ferschneider (2000) model for gas–liquid interaction were
implemented in the model. The comparison showed that the
k-fluid model predicted reasonably well the pressure gradient and
the global liquid saturation in the trickle flow regime for liquid
upflow in a cylindrical packed bed. Boyer et al. (2005) studied
liquid spreading from a point source in a column through gamma-
ray tomography and CFD simulation. CFDLIB was also used in
their simulations. The comparison showed that the prediction of
the radial liquid spreading was strongly affected by the capillary
pressure term. Several models for this term were tested and a
modification of this capillary pressure term was proposed to fit
experimental data better.
This paper analyzes both the flow distribution in a chimney
installed in the distribution tray and the isothermal two-phase
dispersion in the packed beds. The significance of the study of the
flow through a chimney is twofold. Firstly, the numerical simula-
tion allows the detailed description of the two-phase flow
topology inside the chimney and, secondly, the flow distribution
at the outlet of the chimney obtained in the simulations is
important because it determines realistic inlet conditions to the
packed beds of the reactor. The knowledge of these inlet condi-
tions is needed for the correct prediction of the two-phase
dispersion in the beds. The numerical studies that consider the
macroscopic full three-dimensional dispersion phenomena in
packed beds are very scarce in the open literature. Simulations
were first conducted to predict the two-phase flow dispersion in
laboratory conditions for which numerical and experimental data
are available. Then, numerical simulations of the flow dispersion
in the packed beds of an industrial size reactor are analyzed using
inlet boundary conditions obtained in the simulation of the flow
through a chimney. To the authors’ knowledge, this level of
modelization has not been reported previously.

2. Physical model
Fig. 2. Geometry and boundary conditions of one chimney system. Data in Table 2.
2.1. Flow in a chimney

The reactor considered in this study has a distribution tray Fig. 2 shows the physical model of the chimney considered.
equipped with cylindrical chimneys. The tray is located above the The chimney has six lateral circular orifices and a top gas inlet.
catalyst bed to produce an improved two-phase flow distribution The geometry of the chimney is symmetric with respect to a
at the inlet of the catalytic fixed bed. A sketch of the reactor is vertical plane and consequently the computational domain con-
shown in Fig. 1. sists of a section of 1801, as shown in Fig. 2. The exterior zone
M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57 51

surrounding the chimney contains the horizontal gas–liquid Table 2


interphase. The diameter of this zone corresponds to a half of Computational domain dimensions.
the distance between the centers of two neighbor chimneys on
Element Size (m)
the distribution tray. It should be noted that a simulation with a
larger diameter of the exterior zone produced very similar results D1 0.050
to those presented here. A semicylindrical lower region has been D2 0.150
attached to the bottom outlet of the chimney to determine the D3 0.200
H1 0.330
liquid and gas distribution at the inlet of the first bed. The height H2 0.283
of this region corresponds to the distance between the tray that
supports the chimneys and the first packed bed.
The injection of the liquid through this boundary of the 2.2. Validation of two-phase flow dispersion models
domain, as shown in Fig. 2, avoids the predetermination of
the liquid height outside the chimney and consequently on the Before attempting the simulation of the flow dispersion in the
semicylindrical exterior boundary of the domain. It should bed of the reactor, two models for the simulation of two-phase
be noted that if a predetermined liquid height is set in this flow in porous media have been tested to predict the experi-
exterior boundary and that does not correspond to the exact mental and numerical results of Boyer et al. (2005).
steady state liquid height an unrealistic curvature of the liquid–gas Boyer et al. (2005) studied numerically and experimentally the
interface would be obtained. two phase flow dispersion in a vertical cylindrical column of
As a first step toward the complete modelization and simula- spherical glass particles of 1.99 mm of diameter. The column had
tion of all the relevant physical and chemical phenomena occur- an inner diameter of 0.4 m and a bed height of 1.80 m. Although
ring in a hydrodesulfurization reactor, this study is focused on the the axial porosity distribution is reported, an average porosity of
isothermal incompressible flow distribution, so the chemical 0.365 has been considered for the present simulation. The fluids
reactions and the associated heat and mass transfer processes used were water and air at room temperature. The liquid was
have not been considered. The physical properties of the gas and introduced through a tube of 8 mm of inner diameter located in
the liquid are assumed to be constant and they are shown in the center of the top of the bed while the air was injected all
Table 1. around the liquid tube over the whole cross-section of the bed.
The liquid and gas flow rates were 0.128 m3/h and 45 m3/h,
respectively.
2.1.1. Boundary conditions
The mass flow rates of liquid and gas through the inlet 2.3. Two-phase flow dispersion in the packed beds
boundaries of the computational domain of the individual chimney
considered correspond to the total mass flow rates at the inlet of the The hydrodesulfurization reactor considered, which is currently
reactor divided by the number of chimneys installed on the tray. in operation, has four different beds of particles of different shapes
Since the main objective of this study is the prediction of the liquid and sizes. The height and porosity of each individual bed are, from
height, the internal jet formation and the flow distribution at the top to bottom, 150 mm and 0.33, 150 mm and 0.53, 150 mm and
outlet of the chimney, the boundary conditions adopted are 0.45 and 15 560 mm and 0.4, respectively. The physical properties of
the liquid and the gas phases are evaluated at the operating
 Gas inlet: the gas enters through the upper surface of the conditions of the reactor. Two different cases have been studied:
computational domain with a constant and uniform velocity the first one considers the flow dispersion in the bed underneath a
distribution of 4:45  102 m=s (see Fig. 2). group of chimneys installed in a central region of the reactor, far
 Liquid inlet: the liquid enters the domain with a uniform from the cylindrical lateral wall of the reactor. The second case
velocity of 7:27  102 m=s through the bottom annular area considers the dispersion near the walls of the reactor.
corresponding to the distribution tray (see Fig. 2). The computational domain for the central region case consists
in a prism of rhombic cross section. The dimensions of this cross
section, which is centered in the projection of one chimney, were
It should be noted that the simulation is performed with a
selected according to the symmetry elements of the distribution
time marching scheme until a steady state is reached. Initially the
of chimneys in the tray (see sketch at the left in Fig. 3).
domain is considered to be full of gas. The liquid is introduced
The contours of the liquid volume fraction shown in Fig. 3 are
from the bottom annular section and the gas through the top
discussed in Section 4.3.1. The velocities and volume fraction
semicircular boundary. These boundary conditions, although not
distributions imposed on the inlet top surface of the computa-
being exactly coincident with the real process, especially for the
tional domain were taken from the simulation of the flow in a
liquid, do not influence the final steady state of the flow inside the
chimney. The sketch at the left in Fig. 3 shows the computational
chimney or the liquid height surrounding the chimney because
domain for this case. The projection of the chimney is indicated in
they only affect the transient process. The gas–liquid mixture
the sketch at the left in Fig. 3 on the top of the bed. The four
exits the domain at a constant pressure through the surfaces of
different beds are also indicated in this figure. Periodic boundary
the semicylindrical lower region attached to the bottom outlet of
conditions are imposed on the lateral faces of the domain to
the chimney. Symmetry conditions are imposed at the vertical
model the effect of the surrounding chimneys. The height of the
symmetry plane of the domain while the non-slip condition is set
fourth bed has been considered to be 450 mm, because prelimin-
on the tray and on the walls of the chimney.
ary simulations carried out predict fully developed flow condi-
tions for bed depths larger than 70 mm under the present flow
Table 1
Physical properties. conditions.
The computational domain for the near-wall region, shown in
Phase Density (kg/m3) Viscosity (kg/m s) the sketch at the left in Fig. 4, consists in a prism of rectangular
cross section. The contours of the liquid volume fraction shown in
Liquid 611 1.7  10  4
Gas 20.8 2.52  10  5
Fig. 4 are discussed in Section 4.3.2. The projections of the
chimneys contained in the rectangle representing the top of the
52 M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57

bed are indicated in the sketch at the left in Fig. 4. As shown in standard k–E turbulence model (Launder and Spalding, 1972) has
Fig. 4, the top surface of the computational domain contains the been used in the simulation of the flow in a chimney.
circular projection of one chimney located near the wall and three The simulation has been carried out with a time marching
halves of the projection of one chimney distributed along the scheme until the steady state is reached. The time steps used for
perimeter of the top surface. Specific distributions of volume time integration range from 0.001 to 0.01 s. About 10 iterations
fractions and inlet velocities are imposed on the top of the bed as per time step are needed to satisfy the convergence criteria.
inlet conditions, according to numerical simulations of the flow at Pressure–velocity coupling is solved using the SIMPLE
the exit of one chimney. The no slip condition is imposed at the scheme (Versteeg and Malalasekera, 1995). The discretization
wall and symmetry boundary conditions are imposed at the scheme used for pressure is PRESTO! (Patankar, 1980), while
lateral faces of the domain, which are perpendicular to the the momentum discretization is a second order upwind scheme.
y-direction (see sketch at the left in Fig. 4). At the boundaries The Geo-Reconstruct scheme (Fluent V6.1, 2003) is used as
opposed to the wall, the distributions of velocities and volume discretization for the volume fraction.
fractions imposed were taken from the simulation of the flow in a The total number of grid nodes in the domain is 150 000. The
central region of the bed. grid used combines tetrahedral and hexahedral elements. Hex-
ahedrals have been used where possible, because they provide
greater precision and numerical stability. Near the holes of the
3. Mathematical model chimney a fine tetrahedral mesh has been used to keep the
accuracy of the hexahedral mesh. The grid inside the chimney is
The momentum governing equations, as well as the correspond- stretched near the walls to resolve the momentum boundary
ing boundary conditions, have been numerically solved using the layer. Standard wall functions have been used for the simulation
commercial code Fluent (Fluent V6.1, 2003). in this near-wall region. Simulations with a finer mesh of 340 000
grid nodes were also carried out. In those simulations, results
were very similar to those obtained with a coarser mesh,
3.1. Flow in a chimney especially regarding liquid height and flow distribution.

The two-phase flow has been simulated using the VOF model
(Hirt and Nichols, 1981) available in the code. This model is 3.2. Validation of two-phase flow dispersion models
suitable for two-phase flow simulations where phases are segre-
gated since it predicts a well defined gas–liquid interphase. The The models considered are the porous media model and the
granular phase model. In the porous media model a pressure drop
term, based on the Ergun equation, is added to the momentum
equations. This term consists of a viscous term, proportional to
the viscosity (m) of the fluid phase, and an inertial term, propor-
tional to the square of the velocity (U), as shown in Eq. (1)
m r
DP ¼ U þC 9U 9U ð1Þ
a i 22 i i
In Eq. (1) the viscous and inertial resistance coefficients, a and
C2, can be computed using Eqs. (2) and (3), respectively

1 D2p E3
-a ¼ ð2Þ
a 150 ð1EÞ

3:5 ð1EÞ
C2 ¼ ð3Þ
Dp E 3

In the granular phase model particles are defined as a static


Fig. 3. Computational domain for the central region simulation (left) and contours
granular phase with the corresponding volume fraction. The phase
of liquid volume fraction in horizontal slices of the four different beds (right). interaction is modeled through drag coefficients according to the
The position of the chimney is indicated at the top of the right figure. single-slit model proposed by Holub et al. (1992).

Fig. 4. Computational domain for the near-wall region simulation (left) and contours of liquid volume fraction in horizontal slices of the different beds (right). The position of the
chimneys is indicated at the top of the right figure. The wall of the reactor is located at x¼0.
M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57 53

The two-dimensional axisymmetric computational domain has through user defined functions, has been chosen for the simulation
been divided into 5050 rectangular finite volumes using a uniform of the packed beds of the reactor. As a first step, this model has been
grid distribution similar to that used by Boyer et al. (2005). As used to simulate the isothermal three-dimensional two-phase flow
reported by Jiang et al. (2000) and also according to preliminary dispersion in the reactor bed without considering the chemical
simulations using the k–E turbulence model, the Reynolds stress reactions and the heat transfer processes that occur in a real reactor.
term does not influence the behavior of the two phase flow in the The implementation of the chemical reactions and the heat transfer
trickle bed for the flow conditions considered (see Table 3). is left for future studies.
Consequently, the simulations have been carried out without any The simulation of the two-phase flow dispersion in the packed
turbulence model. The simulation of the two-phase flow dispersion beds has been carried out for the steady state. Pressure and
in the packed column has been performed for the steady state. The velocity have been coupled using the phase coupled SIMPLE
phase coupled SIMPLE scheme has been used to couple pressure and scheme, while the discretizations used have been a second order
velocity, while a second order upwind discretization for momentum upwind for momentum and QUICK for volume fraction.
and QUICK discretization for volume fraction has been used. A hexahedral regular mesh of 1 346 700 cells has been used in
As suggested by Jiang et al. (2002a,b) the inclusion of a the numerical simulation corresponding to the central region
capillary pressure term (Eq. (4)) in the momentum equations case, while a hexahedral regular mesh of 1 228 500 cells has been
increases the flow dispersion used in the simulation of the near-wall region.
ryl ðpl pg Þ ð4Þ

Two cases with different capillary pressure correlations have 4. Results and discussion
been simulated, one with the correlation given by Attou and
Ferschneider (2000) (Eq. (5)) and another with the correlation 4.1. Flow in a distribution chimney
proposed by Grosser et al. (1988) (Eq. (6))
    Fig. 5 shows the iso-surface of liquid volume fraction of 0.3 to
ys 1=3 1 1 rg
pl ¼ pg 2ss ð1f Þ þ F ð5Þ portray the three-dimensional topology of the liquid–gas interphase.
1yg dp dmin rl It can be seen the shape of the jets inside the chimney and the
pffiffiffiffiffiffiffiffiffi    horizontal liquid–gas interphase outside the chimney. It should be
180ys 1ys yl
pl ¼ pg ð1f Þss 0:48 þ 0:036 ln ð6Þ noted that in the regions inside the iso-surface the volumetric
ð1ys Þde yl fraction of liquid ranges from 0.3 to 1, while in the outer regions
In Eq. (5) the pressure factor, F, can be computed as the liquid is dispersed or the flow is composed mostly by gas. This
rg rg can be clearly seen in Fig. 6, which shows the liquid volume fraction
F ¼ 1 þ 88:1 ð7Þ distribution inside the chimney and its surroundings.
rl rl
It can be seen in Figs. 5 and 6 that for the flow conditions
and the wetting efficiency, f, is given in Eq. (8) according to El- considered the liquid enters the chimney through the lower four
Hisnawi (1981) holes of the chimney, which has a total of six holes distributed on
  the wall as shown in Fig. 2. The hydrostatic pressure generated by
yl 0:224
f ¼ 1:021 ð8Þ the liquid height outside the chimney produces the liquid jets
1ys
inside the chimney. As expected, and as shown in Fig. 5, the
It should be noted that the computation of capillary pressure trajectory of the two bottom liquid jets is more horizontal than
term has been added to the simulation code using a user-defined that of the two top liquid jets according to the liquid height above
function. the holes. Fig. 5 shows that the opposed liquid jets merge near the
In the granular phase model, the fluid–particle interaction has axis of the chimney and they are deflected perpendicularly
been modeled through drag coefficients according to the single toward the walls of the chimney. As it can be seen in Fig. 5 the
slit model proposed by Holub et al. (1992). In this model the drag liquid descends attached to the walls toward the exit of the
force, that has to be included in the corresponding momentum chimney distributed in two main currents.
equations for each phase, can be expressed as The values of the discharge coefficients of the orifices, calcu-
lated from the flow rate and the liquid height above the orifice
F Dks ¼ yk ys X ks ðuk us Þ ð9Þ
predicted by the simulation, range between 0.58 and 0.62, in
where the exchange or drag coefficient Xks is defined as agreement with the typical value of 0.6 found in the literature
1 (Churchill, 1988).
X ks ¼ ðAks mk U k þ Bks rk U 2k Þ ð10Þ For the flow conditions considered the drag effect of the gas on
ð1EÞ9uk 9
the trajectory of the liquid jets is not important. Simulations
and carried out keeping the liquid flow rate conditions but decreasing
ð1EÞ2 25% the gas flow rate showed no significant differences in the
Aks ¼ 180 ð11Þ trajectories of the jets but differences were observed in the flow
y3 dp2
distribution at the exit of the chimney.
ð1EÞ Fig. 7 shows contours of liquid axial velocity in the vertical
Bks ¼ 1:8 ð12Þ symmetry plane of the chimney. Because of the coordinate
y3 dp
system adopted, the negative velocity values correspond to the
The particular forms of the drag forces used for the simulation liquid flowing downwards. As it can be seen, the velocity
and given in Eqs. (9)–(12) have been included in the simulation distribution is not axisymmetric with respect to the vertical axis
code through user-defined functions. of the chimney. According to Fig. 7, the gas, which enters by the
upper left corner of the chimney, drives the liquid toward the
3.3. Two-phase flow dispersion in the packed beds right. The liquid stream that leaves the chimney by the lower
right side (as shown in Fig. 7) descends with higher velocity, up to
The Eulerian granular phase model, with the fluid–particle 1.8 m/s, than the liquid stream at the left side, about 1.3 m/s at
interaction model of Holub et al. (1992) implemented in the code maximum. In addition, the liquid stream shown at the right in
54 M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57

Fig. 6. Contours of liquid volume fraction in the vertical plane which crosses the
centers of the orifices.

Fig. 5. Iso-surface of liquid volume fraction 0.3.

Fig. 7 is also displaced to the outer region at the exit of the


chimney, while the left current descends following a more
vertical path.
Fig. 8 shows the liquid distribution at the outlet of the
computational domain which corresponds to the inlet of the first
packed bed. The arrow depicted in Fig. 8 indicates the inlet
direction of the gas into the chimney. As it can be seen in Fig. 8,
the liquid is mainly distributed in three regions along the circular
boundary of the chimney projection according to the topology of
the liquid falling from the exit of the chimney (see Fig. 5). It can be
seen that the drag effect of the gas is evident in the distribution of
the liquid. The distribution of liquid at the exit of the computa- Fig. 7. Contours of liquid axial velocity (m/s) in the symmetry plane of the
tional domain considered is used as a boundary inlet condition for chimney.
the simulations of the two-phase flow dispersion in a packed bed,
which are presented and discussed later in this paper. Fig. 9a, the liquid spreading for the conditions considered is clearly
underpredicted by the porous media model. As shown in Fig. 9a, the
4.2. Validation of two-phase flow dispersion models porous media model is not capable to reproduce the data reported
by Boyer et al. (2005).
4.2.1. Porous media model The effect of the addition of the capillary pressure term using
Fig. 9a shows the comparison of the experimental and numerical Eq. (5) on the liquid distribution is shown in Fig. 9b. The results
results of Boyer et al. (2005) and the present numerical results obtained using Eq. (6) are not significantly different in comparison
obtained with the porous media model. The lines plotted in this with those shown in Fig. 9b and consequently have been omitted.
figure correspond to a value of liquid volume fraction of 0.12. It As it can be seen, the model with the capillary term predicts
should be noted that the liquid is introduced at the center approximately the same liquid spreading in comparison with the
(r o4 mm) of the top of the bed (z ¼ 1:8 m). As it can be seen in default porous media model and it does not provide a noticeable
M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57 55

improvement of the radial spreading. A simulation carried out 1.8


with a porosity of 0.339, corresponding to the minimum value of
the distribution reported by Boyer et al. (2005), shows no
significant differences with the one using an average value of
E ¼ 0:365.
1.6
4.2.2. Granular phase model
Fig. 10 shows contours of liquid volume fraction for the
granular phase model and their comparison with the experimen-
tal and numerical simulation results of Boyer et al. (2005). As it
can be seen in Fig. 10, the granular phase model overpredicts 1.4
liquid spreading at the top of the bed, due to the large gradient in
volume fraction between the liquid and gas phases at the inlet.

z (m)
After that sudden flow dispersion in the first 15 cm of the packed
bed, the liquid plume flows downwards with little additional
spreading. At a depth of 1 m into the bed, the flow dispersion
1.2
predicted by the granular phase model is as accurate as the
numerical prediction from Boyer et al. (2005). The difference
between both results at that depth is that the Boyer et al. (2005)
simulations underpredict slightly liquid spreading while the

0.8
0 0.05 0.1 0.15 0.2
r (m)
Experimental data by Boyer et al. (2005)
Numerical data by Boyer et al. (2005)
Fluent results

Fig. 10. Contours of liquid volume fraction of 0.12 in a r–z plane of the cylindrical
bed. Comparison between present simulations and experimental and numerical
Fig. 8. Liquid volume fraction contours at the inlet of the first bed. simulation results by Boyer et al. (2005).

1.8
1.8

1.6
1.6

1.4 1.4
z (m)
z (m)

1.2 1.2

1 1

0.8 0.8
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
r (m) r (m)
Experimental data by Boyer et al. (2005)
Experimental data by Boyer et al. (2005) Numerical data by Boyer et al. (2005)
Numerical data by Boyer et al. (2005) Fluent results with average porosity
Fluent results Fluent results with minimum porosity

Fig. 9. (a) Contour of liquid volume fraction of 0.12 in a r–z plane of the cylindrical bed and (b) contour of liquid volume fraction of 0.12 in a r–z plane of the cylindrical
bed. Effect of the addition of the capillary pressure term according to the model of Attou and Ferschneider (2000).
56 M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57

Fluent simulations overpredict flow dispersion in the same


magnitude. Thus, it can be concluded that the granular phase
model, unlike the porous media model, fits reasonably well the
experimental data. Since the overprediction of the liquid spread-
ing is due to the large gradients at the inlet of the bed, and the
flow conditions in the reactor are smoother than the ones of this
validation case, the granular phase model can be used to predict
the flow dispersion in the packed beds of the reactor.

4.3. Two-phase flow dispersion in the packed beds

4.3.1. Central region


The liquid distribution in the central region of the trickle-bed
reactor can be seen in Fig. 3, where liquid volume fraction
contours are shown at four different slices, one for each bed at
a depth of 75 mm into the bed. Fig. 3a shows liquid spreading
radially toward the outer part of the domain as compared to the
initial liquid volume fraction distribution shown in Fig. 8. No
significant change is seen in the liquid distribution from the first
bed (Fig. 3a) to the second one (Fig. 3b), due to the higher porosity Fig. 11. (a) Comparison between the central (left) and near-wall (right) regions
of the latter, 0.53, as compared to the 0.33 value of the first bed. through contours of liquid volume fraction. Horizontal slices at a depth of 70 cm.
The liquid spreads radially in the third bed (Fig. 3c), where the (b) Comparison between the central (left) and near-wall (right) regions through
contours of liquid axial velocity. Horizontal slices at a depth of 70 cm. (For
porosity is 0.45, lower than in the second bed. The same happens
interpretation of the references to color in this figure caption, the reader is
in the fourth bed (Fig. 3d), where porosity (E ¼ 0:40) is lower than referred to the web version of this article.)
in the third bed. However, that change is not as significant as the
one happening in the third bed, due to the smaller difference in
porosity between the third and fourth beds as compared to the Table 3
difference in porosity between the second and third beds. These Range of maximum values for the liquid velocities.
results show that one of the main factors of the liquid spreading is
Case Velocity (m/s)
the change in porosity. When the flow moves from a more porous
bed to a less porous bed, it spreads radially toward the outer Flow in the chimney 1.8
region of the bed, as it has less available space to flow down- Flow in the column 0.7
Flow in the reactor 0.2
wards. When the opposite happens, that is, the flow moves from a
less porous bed to a more porous bed, the resistance to downward
flow is smaller and the simulation predicts no further radial
spreading. In the cases where the difference in porosity is large, as the colorbar are negative due to the z-axis orientation. As it is
in the transition from the first bed (E ¼ 0:33) to the second one shown, the liquid descends with higher velocity in the near-wall
(E ¼ 0:53), the liquid flow tends to occupy the new available space region than it does in the central region. When looking at both
in the inner region of the plume. Fig. 11a and b, one can see that the zones with a larger volume
fraction of liquid correspond to those where liquid descends
faster. A closer look at Fig. 11a explains the difference in
4.3.2. Near-wall region velocities. In the half chimney projection placed at the left
Fig. 4 shows liquid volume fraction contours at the four beds, boundary of the near-wall region contours (right), it can be seen
75 mm deep into each one of them. As previously seen in the case how the liquid volume fraction is higher than in the rest of the
of the flow underneath a chimney located in the central region of domain. When looking at the velocity contours of Fig. 11b, the
the reactor, the liquid spreads slightly in the first bed (Fig. 4a) to same phenomenon is observed: the highest liquid velocity is
the outer part of the computational domain. There is no signifi- found in the half chimney projection placed at the left boundary
cant change in the liquid distribution in the second bed (Fig. 4b), in the near-wall region contours (right). This is due to the
as it happened in the central region, due to the higher porosity of boundary conditions used in the simulations for the near-wall
the second bed. The largest change in liquid volume fraction region. In that left boundary where the projection of half chimney
distribution happens in the third bed (Fig. 4c), where the liquid is placed, the results of the central region simulation for the same
spreads radially toward the outer part of the domain. This plane (center of one chimney’s projection) were imposed as
tendency continues, although in a slower manner, in the fourth boundary condition, which makes the liquid to descend through
bed (Fig. 4d). The same reasoning for the flow behavior in the the beds with a higher velocity than it should and as a conse-
central region can be also applied here to the near-wall region. quence, have a larger volume fraction of liquid.

4.3.3. Wall influence


Fig. 11a shows the comparison of liquid volume fractions 5. Conclusions
contours for the central (left) and near-wall (right) regions at a
depth of 70 cm. If one compares the liquid volume fraction Numerical simulations of the two-phase flow both in a
contours of the nearest-wall chimney projection with those of distribution chimney and the packed beds of a hydrodesulfuriza-
the central region, it can be seen that the liquid is more tion reactor have been reported. The opposed liquid jets entering
concentrated in the near-wall region than it is in the central the chimney merge near the axis of the chimney where the jets
region. In order to explain this different behavior, liquid axial are diverted perpendicularly toward the walls of the chimney,
velocity contours are shown in Fig. 11b, both for the central (left) producing two descending liquid currents attached to the inner
and near-wall (right) regions at a depth of 70 cm. The values in wall of the chimney. These currents are not symmetrically
M. Martı́nez et al. / Chemical Engineering Science 76 (2012) 49–57 57

distributed at the outlet because of the drag effect caused by the Subscripts
gas entering through the upper part of the chimney. The liquid
distribution at the outlet of the computational domain, which g gas phase
corresponds to the inlet of the first bed of the reactor, shows that k phase index
the liquid is distributed in three main regions along the circular l liquid phase
boundary of the chimney projection. This liquid distribution is s solid phase
used as a boundary condition in the simulations of the two-phase
flow dispersion in the packed beds.
For the flow conditions considered in the bed, the porous
media model is not adequate to simulate the flow dispersion in a Acknowledgments
catalytic bed, even the inclusion in the model of the capillary
pressure effect. A granular phase model has been used for the two This study was financially supported by Repsol, the Spanish
simulations carried out to predict the flow dispersion in the Ministry of Science and Technology and FEDER under project
trickle bed. One simulation considers the flow underneath the DPI2010-17212.
central region of the reactor, far from the walls, and the other
considers the flow in the bed near the cylindrical lateral wall of
References
the reactor. In both cases, most of the liquid spreading takes place
in the first and third beds. The distributions of the liquid volume
Attou, A., Ferschneider, G., 2000. A two-fluid hydrodynamic model for the
fraction do not change significatively as the depth of the bed is transition between trickle and pulse flow in a cocurrent gas–liquid packed
increased except in the third bed and at the interface of two beds bed reactor. Chem. Eng. Sci. 55, 491–511.
with different porosity. The simulations, with the boundary Baker, T., Chilton, T.H., Vernon, H.C., 1935. The course of liquor flow in packed
towers. In: Wilmington, Delaware, Meeting.
conditions chosen, show that the liquid phase descends through Boyer, C., Fanget, B., 2002. Measurement of liquid flow distribution in trickle bed
the beds with higher velocity in the near-wall region than in the reactor of large diameter with a new gamma-ray tomographic system. Chem.
central region, which translates into a larger volume fraction of Eng. Sci. 57, 1079–1089.
Boyer, C., Koudil, A., Chen, P., Dudukovic, M.P., 2005. Study of liquid spreading
liquid in the near wall region than in the central region. The
from a point source in a trickle bed via gamma-ray tomography and CFD
simulations indicate that, for the flow conditions considered, a simulation. Chem. Eng. Sci. 60, 6279–6288.
completely uniform liquid distribution is not achieved. The values Cihla, Z., Schmidt, O., 1957. A study of the flow of liquid when freely trickling over the
packing in a cylindrical tower. Collect. Czech. Chem. Commun. 22, 896–907.
of local liquid volume fractions at the bottom of the computa-
Churchill, S.W., 1988. Viscous flows: the practical use of theory. In: Butterworths
tional domain considered range between 0 and 0.3. Series in Chemical Engineering.
The results presented in this paper show that a reasonable Dudukovic, M.P., Larachi, F., Mills, P.L., 1999. Multiphase reactors—revisited.
prediction of the liquid distribution in packed beds can be Chem. Eng. Sci. 54, 1975–1995.
El-Hisnawi, A.A., 1981. Tracer and Reaction Studies in Trickle-Bed Reactors. D.Sc.
obtained. Thus, numerical simulations can be used to analyze Thesis. Washington University, St. Louis, USA.
current systems or to help in the design of new ones. For a Fluent 6.1 Documentation, 2003. Fluent Inc., Lebanon, NH, USA.
complete and more accurate prediction the inclusion of chemical Gaffney, E.S., Kashiwa, B.A., 2003. User’s Manual for CFDLib Version 02.1. LA-UR-
03-0657.
reactions and heat and mass transfer processes can be considered Gopalan, S., Abraham, B.M., Katz, J., 2004. The structure of a jet in cross flow at low
as a next step. velocity ratios. Phys. Fluids 16 (6), 2067–2087.
Grosser, K.A., Carbonell, R.G., Sundaresan, S., 1988. Onset of pulsing in two-phase
cocurrent downflow through a packed bed. AIChE J. 34, 1850.
Nomenclature Harter, I., Boyer, C., Raynal, L., Ferschneider, G., Gauthier, T., 2001. Flow distribu-
tion studies applied to deep hydro-desulfurization. Ind. Eng. Chem. Res. 40,
5262–5267.
C2 inertial resistance coefficient (m  1) Hirt, C.W., Nichols, B.D., 1981. Volume of fluid (VOF) method for the dynamics of
de equivalent diameter (m) free boundaries. J. Comput. Phys. 39, 201–225.
Holub, R.A., Dudukovic, M.P., Ramachadran, P.A., 1992. A phenomenological model
dmin minimum diameter (m) for pressure drop liquid holdup and flow regime transition in gas–liquid
dp particle diameter (m) trickle flow. Chem. Eng. Sci. 47, 2343–2348.
f wetting efficiency (dimensionless) Jiang, Y., Khadilkar, M.R., Al-Dahhan, M.H., Dudukovic, M.P., 2000. Single phase
flow modeling in packed beds: discrete cell approach revisited. Chem. Eng. Sci.
F pressure factor (dimensionless) 55, 1829–1844.
FD drag force per volume unit (N m  3) Jiang, Y., Khadilkar, M.R., Al-Dahhan, M.H., Dudukovic, M.P., 2002a. CFD of multiphase
p capillary pressure (Pa) flow in packed-bed reactors. I. k-Fluid modeling issues. AIChE J. 48, 701–715.
Jiang, Y., Khadilkar, M.R., Al-Dahhan, M.H., Dudukovic, M.P., 2002b. CFD of
DP pressure drop (Pa)
multiphase flow in packed-bed reactors. II. Results and applications. AIChE J.
U velocity (m/s) 48, 716–730.
X drag coefficient (kg/m3/s) Launder, B.E., Spalding, D.B., 1972. Lectures in Mathematical Models of Turbu-
lence. Academic Press, London, England.
Greek letters Patankar, S.V., 1980. Numerical Heat Transfer and Fluid Flow. Hemisphere,
Washington, DC.
a viscous resistance coefficient (m2) Raynal, L., Harter, I., 2001. Studies of gas–liquid flow through reactors internals
using VOF simulations. Chem. Eng. Sci. 56, 6385–6391.
E porosity (dimensionless) Tsochatzidis, N.A., Karabelas, A.J., Giakoumakis, G.A., Huff, G.A., 2002. An investi-
y volume fraction (dimensionless) gation of liquid maldistribution in trickle beds. Chem. Eng. Sci. 57, 3543–3555.
m viscosity (Pa s) Versteeg, H.K., Malalasekera, W., 1995. An Introduction to Computational Fluid
Dynamics. The Finite Volume Method. Pearson Education Limited.
r density (kg/m3) Wu, P.K., Kirkendall, K.A., Fuller, R.P., 1997. Breakup processes of liquid jets in
s surface tension (N m  1) subsonic crossflows. J. Propul. Power 13 (1), 64–73.

You might also like