Sie sind auf Seite 1von 12

Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Investigation of hydrodynamic behavior in random


packing using CFD simulation

Jia-Lin Kang a , Ya-Cih Ciou b , Dong-Yang Lin b , David Shan-Hill Wong b,∗ ,
Shi-Shang Jang b,∗
a Department of Chemical and Material Engineering, TamKang University, New Taipei City, Taiwan, ROC
b Department of Chemical Engineering, National Tsing Hua University, Hsinchu, 30013, Taiwan, ROC

a r t i c l e i n f o a b s t r a c t

Article history: Computational fluid dynamics (CFD) simulations of countercurrent gas–liquid flow in ran-
Received 6 October 2018 dom packings of Raschig rings were carried out in this study. This model used gravity
Received in revised form 22 April simulation to construct the random packing structure, and a volume expansion-recovery
2019 method to improve the meshing quality. A simple feedback control scheme was applied to
Accepted 29 April 2019 control the gas inlet flow rate so that pressure drop can be estimated. The generated char-
Available online 7 May 2019 acteristics of the packing structure such as the number of packing elements, dry surface
area and porosity were found to be close to experiments. CFD predictions of hydrodynamics
Keywords: properties are also validated by the experimental data using a small column section. The
CFD simulation results indicate that our CFD model was able to capture the essential hydrodynamics behav-
Random packing ior of the gas–liquid countercurrent flow. Hence, the approach presented in this study can be
Multiphase used as a basis for studying the effect of detailed packing-geometry design on hydrodynamic
Automatic stacking and mass transfer characteristics.
Hydrodynamic behaviors © 2019 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction uid flow. The model can predict fairly well pressure drop, liquid hold-up
and wetted area of countercurrent gas–liquid flow in various structural
Packings are widely used in distillation, absorption, and stripping packings. Van Baten and Krishna (2002) presented using separated flow
columns. Effective mass transfer in packing reduces the height of a channels of the gas and liquid phases and interface between two chan-
column and leads to a reduction in capital cost. Alternatively, using a nels to study the mass transfer in KATAPAK-S structural packing. The
better packing can increase the number of transfer units, and hence, result showed that the predictions of Sherwood numbers of gas and
reduce the reflux in a distillation column and the absorption solvent liquid phases. The gas-phase Sherwood number agrees with empiri-
usage leading to reduced operating costs and increased throughput cal correlation, but the liquid phase Sherwood number is significantly
in revamping. Development of packings used to be empirical, relying lower. Valluri et al. (2005) adopted volume of fluid (VOF) to simulate
on experiments to collect data on the height of mass transfer units, the dynamic evolution of waves on laminar falling wavy films at low
liquid holdup, interfacial area, and pressure drop. Empirical correla- to moderate Reynolds numbers over corrugated surfaces in an REU of
tions of these data were developed. It is desirable that computational Mellapak-500.Y. Chen et al. (2009) reported that the VOF method devel-
fluid dynamics can be used to enable a more rational design of these oped a 3D gas–liquid countercurrent flow model for simulating the
packings. hydrodynamics and mass-transfer behavior in a typical REU of struc-
There have indeed been many studies using computational fluid tural packing. The gas inlet and the liquid outlet were on the same
dynamics (CFD) used to investigate the hydrodynamic behavior of side, whereas the gas outlet and the liquid inlet were on the same side.
structured packing. Structural packing can be easily simplified into a The result showed that the model can predict the liquid holdup, wet-
representative elementary unit (REU) to reduce computation load. Iliuta ted area, and separation efficiency. Haroun et al. (2012) investigated
and Larachi (2001) presented a one-dimensional two-fluid model of two mass transfer from the gas phase to the liquid along with a corru-
parallel inclined and interconnected slits to simulate the gas and liq- gated metal sheet with liquid and gas flowing in the same direction.


Corresponding authors.
E-mail addresses: dshwong@che.nthu.edu.tw (D.S.-H. Wong), ssjang@mx.nthu.edu.tw, ssjang@che.nthu.edu.tw (S.-S. Jang).
https://doi.org/10.1016/j.cherd.2019.04.037
0263-8762/© 2019 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
44 Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54

Nomenclature ˛l Volume fraction of the liquid phase


˛g Volume fraction of the gas phase
ae Effective specific surface areaea, m2 /m3 gl Surface tension coefficient, N/m
ap Total specific surface area, m2 /m3 k Effective turbulent Prandtl numbers for k
Atot, packing Surface area of total packing, m3 ε Effective turbulent Prandtl numbers for ε
Agl Gas–liquid contact area, m l Viscosity of liquid, cp
Ap,s Surface area of a single packing g Viscosity of gas, cp
Abot,j Bottom area of jth mesh mx Viscosity of mixture, cp
C1ε , C2ε , C Constants for turbulence model t Turbulence viscosity, cp
COH− Concentration of hydroxyl ion in the liquid l Density of liquid, kg/m3
phase, kmol/m3 g Density of gas, kg/m3
DCO2 Diffusivity in the liquid phase, m2 /s mx Density of mixture, kg/m3
e(t) Error between the set point of the gas inlet flow p,s Density of a single packing, kg/m3
rates and the actual value of the gas inlet flow ε Void fraction of packing column/turbulence
rates dissipation rate for turbulence model
F s Additional body force εl Liquid holdup
F gl Surface tension force, N εexp Porosity of experiment
G Gas volume flow rate, l/min w Wall contact angle
Gk Generation of turbulence kinetic energy due to
the mean velocity gradients, J/m3 /s
g Gravity, m/s2
Haroun et al. (2014) used the VOF model for predicting the gas and liq-
HCO2 Henry’s law constant of CO2 , m3 -Pa/kmol uid contact area of a Mellapak 250X REU with countercurrent flows with
hinit Height of initial state, m one velocity inlet and one pressure outlet boundaries. The inlet bound-
hS.S Height of steady state, m ary was given positive and negative values for liquid and gas phases,
hcol Height of the column, m respectively, to implement countercurrent flows. The simulated results
KD Coefficient for the derivative terms in PID showed that when liquid velocity increased, the gas and liquid contact
KI Coefficient for the integral terms in PID area increased, but liquid holdup decreased. The simulated results had
KP Coefficient for the proportional terms in PID good agreement with experimental data.
CFD simulations of truly random packing are complicated by the
k Turbulence kinetic energy, m2 /s2
need to generate an actual random packing stack. Moreover, a large
ks Free surface curvature
number of mesh cells are required to describe the flow field near contact
kOH− Second-order reaction rate constant,
points between packings. Hence, specialized random stack generation
m3 /kmol/s procedures are required to overcome the challenge of meshing quality.
kg ’ Liquid film mass transfer coefficient, Caulkin et al., (2007, 2009, 2006) presented a digital packing algorithm
kmol/m2 /Pa/s to generate the randomly packed beds of spheres and cylinders. The
mp,tot Total mass of packing in the absorber, kg results showed that the approach is capable of predicting the bulk and
mp,s Mass of a single packing, kg local voidage of beds of complex geometry with a reasonable degree
n Normal vectors of accuracy. Atmakidis and Kenig (2009, 2012) used a modified ballis-
n bot,j Normal vector of jth mesh tic approach to generate a random packing of non-contacting spheres.
nw Normal vectors of a packing wall The influence of the confining walls on pressure drop in packed beds
for moderate tube/particle diameter ratios and mass transfer phe-
Np,exp Number of packing of experiment
nomenon were studied. Lopes and Quinta-Ferreira (2009) provided an
P Pressure, Pa
Eulerian multiphase model to predict the liquid holdup and pressure
P̄ Bias of pressure
drop in the trickling flow regime through a regular non-contacting
Pt Top pressure, Pa spherical packing. The simulated result had a good agreement with
Pb Bottom pressure, Pa the experimental correlations of mass transfer coefficients. Bai et al.
R Ideal gas constant, m3 Pa/K/kmol (2009) used the discrete element method to generate a random non-
tw Tangential vectors contacting cylindrical packing structure. Particle shrinkage was used
T Temperature, K to achieve good meshing quality. The concept of “porosity correction
Vl Volume the of the liquid, m3 factor” was introduced.
Vtot Total volume of the flow field, m3 Dixon et al. (2013) studied four methods of contact point modifica-
tions to study the drag coefficient of air flow through a packed bed of
Vcol Volume of an absorber, m3
spheres. The result showed that global expansion or shrinkage meth-
U Velocity vector, m/s
ods (gaps or overlaps) which lead to non-negligible changes in bed
Uj Velocity vector of jth mesh
porosity could lead to substantial errors for drag coefficient and pres-
UT Transposed velocity vector, m/s sure drop. Less error can be obtained with local modifications (caps and
ug Gas velocity, m/s bridges) only near the contact points. Heat transfer study was also pre-
yCO2 ,in Inlet CO2 mole fraction, mol/mol sented. Boccardo et al. (2015) presented using open-source software,
yCO2 ,out Outlet CO2 mole fraction, mol/mol Blender, to fast generate random packing of spherical, cylindrical and
Z Height of packing column, m trilobes. The generated packing properties and hydrodynamics were
validated with air-flow experimental data. The packing structure was
exported as solid volumes, and the intergranular space was meshed.
Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54 45

This was similar to a bridging approach. Partopour and Dixon (2017) where l and g are the viscosities of the liquid and gas,
presented the shrink-wrap method with an automated package for respectively. The continuity equation of the mixture fluid is
packed bed generation for handling the meshing challenges. The result expressed as follows:
showed that the shrink-wrap method can speed up the process of
the geometry and mesh generation while it maintains the structural ∂  
(mx ) + ∇  =0
 mx U (6)
features of the bed. Boccardo et al. (2019) investigated particle trans- ∂t
port and deposition in different catalytic systems using the geometric
models representing different porous media. The geometric models The momentum equation was modified as follows:
were a number of random packings of spheres created by rigid body
simulations. The result showed that the calculated particle deposi- ∂  
 +∇

U
 mx U

 = −∇P

 + t ∇ 
U
mx U
tion efficiency using the CFD matched well with the classical filtration ∂t

theory. U
+∇  mx g + F s
 T + ∇ (7)
The aforementioned studies were usually applied to trickle bed
reactors with single phase flow or using two-phase flow in a co-current
where P is the pressure, t is the turbulence viscosity, g is the
manner. Hence the boundary condition at the inlet can be easily set
force of gravity, and F s is any additional body force.
by defining the mass flow or velocity of a given distribution geome-
try. An open boundary condition with given pressure can be ascribed The k − ε model was employed as the turbulence equa-
to the outlet. In this way, the pressure drop can be calculated. In a tion (Launder and Spalding, 1974). Turbulent kinetic energy,
structural packing, gas and liquid flow are relatively stratified. To sim- k, and turbulence dissipation rate, ε, were used to estimate
ulate countercurrent flow in a structural packing, Haroun et al. (2014) the Reynolds stress in the Reynolds-Averaged Navier–Stokes
used uniform velocity to inject liquid at the inlet surface and assumed equations (Kajishima and Taira, 2017). The equations are
a uniform negative gas velocity at the same inlet-surface. An open expressed as follows:
pressure-outlet surface was used. However, in countercurrent flow
through a random packing, we should not assume in a priori how the ∂
 t
 
(mx k) + ∇  =∇
 · (mx Uk) · mx + 
∇k + Gk − mx ε (8)
air will flow out of the liquid inlet side. ∂t k
The purpose of this study was to present a CFD approach that can
be used to simulate a representative random packing section that can ∂
 t
 
(mx ε) + ∇  =∇
 · (mx Uε) · mx + 
∇ε
study the effect of the detailed geometry of the packing elements. In ∂t ε
this approach, we have adopted a packing generation procedure that
can approximate the porosity of the actual packing but retain good ε2
+ C1ε Gk − C2ε mx (9)
meshing quality. A new method of setting boundary conditions for k
countercurrent gas–liquid flow was presented. Experimental data were
where k is the turbulent kinetic energy, ε is the turbulence
collected using a laboratory packed column to validate the simulation
results.
dissipation rate, C1ε and C2ε are numerical constants, and k
and ε are the effective turbulent Prandtl numbers for k and
2. CFD modelling approach ε, respectively. The term Gk is the generation of turbulence
kinetic energy owing to the mean velocity gradients and is
defined as follows:
2.1. Governing equations
 
Gk = t ∇  −∇
 ·U T
 ·U (10)
The VOF method in the multiphase flow model of Ansys
®
Fluent , a commercial CFD simulator was used. The standard
where U  means the velocity vector. The turbulence viscosity,
k − ε method was adopted in the turbulence model. The gov-
t , is a function of k and ε, as shown below:
erning equations are described as follows.
The continuity equation of the water phase is given by
k2
t = mx C (11)
∂   ε
(˛  ) + ∇  =0
 ˛l l U (1)
∂t l l The model constants C1ε , C2ε , C , k , ε in the above equa-
®
tions are the default values of Ansys Fluent , which were
where ˛l and l are the volume fraction and density of the
provided by Launder and Spalding (1974).
liquid phase, and U is the velocity of the multiphase fluid
The additional body force considered in this work includes
mixture. The volume fraction of the liquid phase obeys the
surface tension forces between the gas–liquid interface which
transport equation:
was calculated by continuum surface force (Ataki and Bart,
∂˛l   2006; Brackbill et al., 1992) model as follows:
+∇  l = ˛l ∇
 U˛ 
 ˙· U (2)
∂t

mx kn
F gl = gl 1
(12)
The volume fraction of the air phase is defined as the void 2 (g + l )
not occupied by water:
where gl is the surface tension coefficient between the gas
 is the unit normal of the gas–liquid interface given
liquid and n
˛g = 1 − ˛l (3)
by

The mixture density, mx , and viscosity, mx , are expressed  l


 = ∇˛
n (13)
as follows:
 ·n
k=∇  (14)
mx = ˛l l + (1 − ˛l ) g (4)
We also include the option that the interface is modified
mx = ˛l l + (1 − ˛l ) g (5) by the packing wall if the cell is next to the wall, i.e., the unit
46 Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54

normal is determined by the contact angle of packings, shown ®


Table 1 – Setting of the motion study in Solidworks .
as below:
Frame number (per second) 16
Geometric accuracy 10 (maximum value is 10)
n  w cosw + tw sinw
=n (15)
3D contact resolution 10 (maximum value is 10)
Gravity(m/s2 ) 9.8
where n  w and tw are the unit normal and tangential vectors Packings Movable and contactable
to the wall respectively; and w is the contact angle between Column and track Unmovable and contactable
the wall and the liquid. w was set as 70◦ (Ma et al., 2007) for
plastic packings in this study.
The coupled algorithm was adopted in this study because it Table 2 – Packing characteristics of automatic, and
experimental stacking.
has a higher tolerance of divergence with low-quality mesh or
with large time steps (Ahipo et al., 2011). Compared with the Case Number of Packing surface Porosity
packing m2
pressure-based segregated algorithm, the coupled algorithm area
m3
elements
can solve the momentum, continuity, volume of fraction,
and pressure gradient equations simultaneously and is the Automatic 23.8 ± 2.2 483 ± 45 0.79 ± 0.02
implicit coupling-equation solver. Experimental 24.3 495 0.79
The gas and liquid phases considered were air and water.
Both are assumed to be incompressible by small changes in
pressure. The randomly stacking is generated by the collision between
packings. The detailed setting is shown in Table 1.
2.2. Random packing generation As the simulation started, the packing rings fell into the
absorber by gravity and formed a stack as Fig. 1b shows. The
To simulate the automatic stacking of random packing by grav- height of the stacked packing structure was over 70 mm. The
®
ity, “motion study” which is an add-in function in Solidworks , stacked packing structure was cut to a height of 22 mm and is
was used. Fig. 1, shows an example of the automatic stacking used to create a flow field. 22 mm of column height was used to
procedure. In Fig. 1a, a column having an inner diameter of reduce computational cost and maintain random structures.
25 mm and a height of 100 mm was established as a small- The number of packing rings that can be stacked in the
scale absorber. A track was built at the top of the absorber, experimental bench absorber, Np,exp , can be estimated by
and the track was filled with over 80 packing elements. In this weighting:
study, Raschig rings with an inner diameter of 4 mm, outer
diameter of 6 mm, and height of 6 mm were used. All pack- m p,tot
Np,exp = (16)
ings are set as movable and contactable items for simulating mp,s
falling and collision effects while the column and track are
set as unmovable items. Friction force is not considered, and where mp,tot is the total mass of packing in the absorber, and
the gravity force is set as 9.8 m/s2 . The geometry accuracy mp,s is the average mass of a single packing ring. The porosity
and 3D contact resolution are set as high level. The high the of packing, εexp , is calculated as follows:
geometry accuracy makes the closer to actual geometry the
mesh becomes. The high 3D contact resolution to prevent the −1
mp,tot p,s
interpenetration of packings. The number of animation frame εexp = 1 − (17)
Vcol
per second is 16 to generate falling pathways of the packing.
where p,s is the solid density of the packing material, and Vcol
is the volume of the absorber. The packing surface area, ap,exp ,
is calculated by

Ap,s Np,exp
ap,exp = (18)
Vcol

where Ap,s is the surface area of a single packing.


Different geometric structures of natural stack packing
were constructed by adjusting the angle of the track and the
distance between the track and absorber. To ensure a random
stacking, five separated structures were generated for each run
condition with specific gas and liquid flow rates.
The averages and standard deviations of the number of
rings, surface area, and porosity were compared with experi-
mental values in Table 2. It can be seen that the results of the
automatic stacking procedure were very close to the experi-
mental values. However, the number of rings packed in the
column section was still slightly smaller than the experimen-
tal value, leading to smaller surface areas and higher porosity.
This was partly because of the need to avoid the direct contact
Fig. 1 – Real image of the automatic stacking procedure: (a) of packings that leads to poor meshing quality, as described
before stacking, and (b) after stacking. in the next section.
Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54 47

Fig. 2 – Schematic diagram of interval establishment. The dot pattern indicates the added volume.

2.3. Meshing quality marking in the blue of 5 mm diameter at the top. The 4 circu-
lar areas were evenly distributed around the range of 12 mm
If the packing rings are in contact, there will be direct contact diameter from the origin of the top area. The rest of the areas,
between the packing elements, a dead zone will be created, marking in the green, at the top were set as the outlet.
and a mesh grid cannot be generated by the ANSYS default The velocity at this inlet boundary was set to provide the
mesh generator. To overcome this bottleneck, the outer diam- specified liquid flow rate Lsp . Other parts of the top surface
eter and height of the packing were enlarged from 6 mm to were set as a pressure outlet with a fixed pressure of 0 Pa. The
6.3 mm, and the diameter of the absorber was reduced from entire bottom surface marking in the red was set as a pressure
25 mm to 24.7 mm before the automatic stacking procedure. outlet. For a pressure outlet, it is possible that the velocity is
After stacking, the packing and absorber were returned to negative, i.e. materials are coming in instead of going out. In
their original sizes to generate interval between packings. The such cases, we must designate what is the phase condition
schematic diagram of interval establishment is shown in Fig. 2. of the material flowing in. We assume that only gas can flow
We called this method “expansion and recovery”. into the section at both pressure outlets, i.e. the gas phase is
Dixon et al. (2013) reported that global enlargement or designated as the back flow phase. The net gas flow, G, into the
shrinkage methods would lead to large porosity changes, bottom pressure outlet can be calculated by the dot product of
resulting in substantial errors in the calculation of drag coef- the average velocity vector and the bottom surface-area vector
ficients. Local corrections at the contact points are preferred. with an outward normal vector:
However, in a random packing, contact points are random, cre-
ating caps or bridges at the contact point are cumbersome. In 
elements at the bottom outlet
G=− n  j ˛g,j
 bot,j Abot,j · U (19)
our expansion and recovery method, all packings are returned
to its original size. While there will be inevitable disturbance j=1

locally at the contact point, the error in overall porosity is


found to be small as shown in Table 2. where n  j tis the
 bot,j is the normal vector, Abot,j is the area, U
velocity and ˛g,j is the gas volume fraction of the jth mesh at
2.4. Boundary conditions for flow fields the bottom outlet.
Since we did not know the pressure drop required to sus-
The wall of the column and packing are set as a non-slip tain a given gas flow rate, a proportional-integral-derivative
wall. However, the setting of boundary conditions at the top (PID) controller was used to adjust the pressure so that the
and bottom surfaces need some explanations. In the study of calculated gas inlet was equal to the specified gas inlet flow
gas–liquid flow in a packing column, it is customary to carry rate Gsp . The equation of the adjusted pressure was shown as
out experiments at fixed liquid and gas flow rates. However, below:
only three types of boundary conditions, mass flow, velocity t
®
and pressure, can be defined in Ansys Fluent . In the setting of d
Padj = P̄ + KP e (t) + KI e () d + KD e(t) (20)
the mass flow and velocity boundary conditions, only the aver- =0
dt
age values of the boundary surface can be defined but in the
countercurrent flow packing section, the liquid and gas flow where P̄ is the average pressure of the steady state; KP , KI , KD
in different directions. Hence, a simple definition of the flow are the coefficients for the proportional, integral, and deriva-
or velocity cannot be used. For a random packing, we have tive terms. e(t) is the error between the set point of the gas
no way of telling a priori where the liquid goes in, and the inlet flow rates and the actual value of the gas inlet flow rates.
gas comes out at the top inlet. Neither do we know where the The parameters used in this study list in Table 4. Note that the
liquid goes out, and gas comes in at the bottom outlet. parameters only used to maintain a steady flow rate and have
To overcome this problem, the following method was pro- not been optimized tuning.
posed in this study. The schematic diagram of our approach The detailed boundary conditions are shown in Table 5.
is shown in Fig. 3a. Fig. 3b shows the specification of the flow Fig. 3c shows that there was an upward pressure drop, and
field is 25 mm diameter and 22 mm height. The effects of the Fig. 3d shows that there was a well-distributed upward gas
diameters and heights in the flow field are provided in Appen- flow into the bottom surface. The dynamic changes in gas flow
dices A and B. A liquid inlet was created with 4 circular areas rate and pressure are shown in Fig. 4a and b. The dynamic
48 Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54

Fig. 3 – (a) Schematic diagram of the approach presented by this study, (b) boundaries used by this work, (c) simulated
pressure distribution, and (d) simulated inlet gas streamlines.

Fig. 4 – Dynamic changes of (a) inlet gas flow rate, (b) bottom pressure, (c) the gas mass balance and (d) the liquid mass
balance at L/G = 0.5.

changes in the gas and liquid mass balance errors are shown ditions of the convergence were set as 10−3 of the absolute
in Fig. 4c and d. residual for all monitoring parameters including continuity,
velocity, volume fraction of phases, and k and ε for viscous
2.5. Convergence and termination model. The simulations started from an initial transient state
in which the packing was filled with a stationary gas phase,
A workstation (18 cores/CPU × 4 CPUs, 2.2 GHz of Intel Xeon) and liquid flowed from the top inlet until a steady state is
was used in this study. The time-step size set as 0.001 s while reached. Steady state is characterized by the fact that gas and
solving a maximum of 20 iterations per time step. The con- liquid mass balance errors become stable and close to zero;
Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54 49

Table 3 – Geometry and operating conditions of packed


bed.
Diameter (mm) 25.4
Height (mm) 300
Operating pressure (atm) 1
Area per unit volume (m2 /m3 ) 495
Porosity 0.79
Gas flow rate (L/min) 1
Liquid flow rate (L/min) 0.1–0.5
Packing type Plastic Raschig rings

Table 4 – Parameters of PID.


Parameters Value

P̄ (Pa) 0.28
KP (Pa/m3 s−1 ) 60,000
KI (Pa/m3 ) 500
KD (Pa/m3 s−2 ) 120

Table 5 – Boundary conditions used by this work.


Boundaries Type Materials Backflow Specification
phase

Liquid inlet Velocity Water 0.021–0.106 m/s


Top outlet Pressure Air 0 kPa-g Fig. 6 – Schematic illustration of packed-bed absorber
Bottom outlet Pressure Air Adjusted by PID
apparatus (Liu et al., 2015).
control to obtain
the correct gas
Agl in the flow field using an iso-surface. The fractional area,
inlet flow rate of
1 L/min ae /at , is treated as the gas–liquid interfacial area given as

ae Agl
= (21)
aP Atot, packing

where ae is the effective area, aP is the packing surface area,


and Atot, packing is the surface area of the total packing.
The volume of liquid in the flow field was treated as the
liquid holdup. The equation of liquid holdup fraction, εl , is
expressed as follows:

˛ ıV l,j
V j l,j
εl = l = (22)
Vtot Vtot

Where Vl and Vtot are the volume of the liquid and the total
volume of the flow field, respectively, and ˛l,j and ıV l,j are the
liquid volume fraction of the jth mesh and the volume of the
jth mesh, respectively.
dP
The pressure drop, dh , is estimated by

dP P − Pt
= b (23)
dh hcol

Where Pb and Pt are the pressures of the bottom and top outlet,
Fig. 5 – Iso-surface for air volume fraction = 0.5. respectively.

and stability in the gas–liquid interfacial area, liquid holdup, 3. Experiments


and pressure drop were also reached. We have carried out the
dynamic simulations for 4 s and found that the system usually 3.1. Apparatus
settles into a steady state after 2 s, as shown in Fig. 4.
A bench-scale packed-bed absorber (Liu et al., 2015) was used
2.6. Calculations of hydraulic properties to carry out the experiments for validating our CFD results. As
Fig. 6 shows, the equipment contained four parts of removable
The area of the interface where both the gas phase and the liq- glass packed sections. In each part of the packed section, the
uid volume fractions were at 0.5 as the gas–liquid contact area inner diameter was 25.4 mm, and the height was 300 mm. A
®
was calculated by using the iso-surface curve in Ansys Fluent 2-L sump was located at the bottom of the equipment. The
(Ataki and Bart, 2006). Fig. 5 shows the gas–liquid contact area diameter of the sump was 160 mm. In this study, only one
50 Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54

packed section was used for investigating the interfacial area


and liquid holdup. The plastic Raschig rings with an outer
diameter of 6 mm, an inner diameter of 4 mm, and a height
of 6 mm were poured into the section and shaken. The poros-
ity and packing surface area were determined to be 0.79 and
495 m2 /m3 , respectively, by weighing the column sections and
the packing rings, as described in Section 2.2. The inlet gas
flow was 1 L/min, and five liquid flow rates from 0.1 L/min to
0.5 L/min between the detailed operating conditions are sum-
marized in Table 3.

3.2. Interfacial area measurement

The interfacial area was measured by the absorption of 36%


CO2 with a 0.1-M NaOH solution (Wang, 2015). The effective
interfacial area, ae was calculated as follows: Fig. 7 – Comparison of the fractional area obtained by the
experiment, CFD simulation, and empirical correlations.
yCO ,in
2
ug ln( y )
CO2 ,out
ae ≈ (24)
Zkg ’RT
section as the pressure drop, the calculation of which was the
where uG is the linear velocity of the gas phase, yCO2 ,in and
same as Eq. (22).
yCO2 ,out are the CO2 mole fraction of the inlet and outlet,
respectively, Z is the height of the column, kg ’ is the reaction-
enhanced liquid-film mass-transfer coefficient with chemical 4. Results and discussions
reactions, R is the ideal gas constant, and T is the temper-
ature. It is well known that in the CO2 absorption by sodium 4.1. Gas–liquid interfacial area
hydroxide, the liquid mass transfer resistance is the dominant
resistance. kg is not dependent on the gas or liquid flow rates Fig. 7 shows the comparison of the fractional gas–liquid
and can be calculated as follows: interfacial area obtained by experiment, CFD simulations,
and empirical correlations found in the literature (Billet and
kOH− CIOH− DCO2 Schultes, 1993, 1999; Bravo and Fair, 1982; Onda et al., 1968).
k’g = (25)
HCO2 The CFD simulations were very close to the experimental
results. All three of the aforementioned empirical correla-
Where kOH− is the second-order reaction rate constant, COH− tions predicted larger values for the interfacial area, with
is the concentration of hydroxyl ions in the liquid phase, the results of Billet and Schultes (1993) being the lowest and
DCO2 is the diffusivity in the liquid phase, and HCO2 is Henry’s closest to the experimental values. The correlation by Onda
law constant of CO2 . The kOH− was set at 8400 m3 /kmol-s at et al. (1968) predicted an interfacial area that was signifi-
24 ◦ (Wang, 2015). The remaining parameters in Eq. (24) were cantly greater than the experimental values. Bear in mind
®
obtained from Aspen Plus . that empirical correlations in the literature are generalized
correlations developed for many types of packings by using
3.3. Liquid holdup measurement data obtained over much wider and gas–liquid-flow ranges
and with packing structure that is much larger in diame-
The liquid holdup was measured by the difference between ters. Hence, their absolute precisions should not be compared
the initial state and the steady state of the liquid level by with our CFD simulations results which were designed specif-
using pure water and air. To enhance the liquid-level differ- ically for our experiments. Nevertheless, the CFD simulations
ence, an additional glass-packed section filled with pure water showed the same trend and order as these generalized
was installed between the packed section and the sump as correlations.
an extension of the sump to monitor the liquid level change.
Because the cross-section area of the additional glass-packed
4.2. Liquid holdup
section was smaller than the cross-section area of the sump,
the liquid level difference was much easier to observe. The
Fig. 8 shows a comparison of liquid holdup between the sim-
liquid holdup, εl,exp , was calculated as follows:
ulation, experiment, and correlations. The correlations are
those of Stichlmair et al. (1989) and Billet and Schultes (1993).
(hinit − hS.S )
εl,exp = (26) As shown in the figure, the CFD model can fit the experimen-
hcol
tal liquid holdups. The correlations of both Billet and Schultes
where hinit is the height of the initial state, hS.S is the height of (1993) and Stichlmair et al. (1989) overestimated the liquid
the steady state, and hcol is the height of the packed section. holdups, with the correlation by Stichlmair et al. (1989) being
closer to experimental results.
3.4. Pressure drop measurement As in the case of interfacial area, liquid holdup correlations
are also generalized correlations developed for many types of
The pressure drop was measured by the differential Pressure packings over a much wider range of gas–liquid flows for larger
Transmitter (Yokogawa EJA-120A). The measurement range is columns. Similarly, CFD simulations showed the same trend
−1000 to 1000 Pa. The top and bottom pressures of the column and order as these generalized correlations.
Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54 51

L/G was increased, the structure of the packing supporter


caused that the pressure drop was increased rapidly simul-
taneously. The internal structure of the equipment was not
considered in the CFD simulation and the correlation. Thus,
when in the higher L/G, the pressure drop of the CFD simula-
tion was agreed with the pressure drop of the correlation by
Stichlmair et al. (1989), but lower than the pressure drop of the
experiment.

5. Conclusions

This study presented a novel approach to simulate the


hydrodynamics of countercurrent gas–liquid flow in random
packing. The approach employed a gravity simulation for
automatic packing structure generation, volume expansion-
Fig. 8 – Comparison of liquid holdup among the recovery method to avoid dead zones and poor meshing, and
experiment, CFD simulation, and empirical correlations. PID control of the gas flow rate by pressure adjustment as
an inlet-gas boundary condition. The results also showed
that the interfacial area, liquid holdup, and pressure drop
predicted values of the CFD model were quite close to the
experimental data. The results indicate that our CFD model
was able to capture the essential hydrodynamics behavior of
the gas–liquid countercurrent flow in a small random packing
section. Therefore, our approach presented in this study can
be used as a basis for studying packing-element design in the
future.
Whereas the columns used in both simulation and experi-
ments are relatively small. It should be pointed out that direct
simulation of large scale industrial columns with detailed
packing structure is not possible with the current computing
power. Our work provided a good representation of a local ran-
dom packing section so that simulation results can be used to
Fig. 9 – Comparison of pressure drop among the study the effect of packing geometry on hydrodynamics. The
experiment, CFD simulation, and empirical correlation. results can also be used to develop correlations that can be
used in multiscale simulations of industrially relevant large
4.3. Pressure drop columns.

The average and standard deviations of the steady state time- Acknowledgment
average pressure drop were compared with the pressure drop
of the experiment, and the correlation by Stichlmair et al. The authors acknowledged the financial support provided by
(1989) in Fig. 9. The result showed that the pressure drop of the Ministry of Science and Technology through the grant MOST
CFD simulations was lower than the experimental pressure 105-2221-E-007-141.
drop but was agreed with the pressure drop of the correlation.
At low L/G, the difference in the pressure drops of the CFD sim- Appendix A. Effect of length in the flow field
ulation and experiment was minor, but as the L/G increased to
greater than 0.3, the difference in the pressure drop between As Fig. A1 shows, the authors have carried out simulations of
the CFD simulation and experiment became larger. The exper- five heights: 10, 16, 22, 28, and 35 mm at a liquid flow rate of
imental pressure drop raised rapidly, while the CFD simulation 0.3 l/min, and a gas flow rate of 1 l/min. As figures shown, the
and correlation remained in low-pressure drops. The dif- interfacial area and liquid holdup can be reached stable after
ference between the CFD simulation and experiment was the height was higher than 22 mm. Thus, 22 mm height of the
caused by packing supporter in the small scale experimen- flow field was selected as the investigating length to reduce
tal equipment. The supporter was a porous disc. When the the computational cost.
52 Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54

Fig. A1 – Effect of (a) fractional area, (b) liquid holdup, and (c) pressure drop in various lengths of the flow field.

Appendix B. Effect of diameter in the flow field to prove that the approach presented in this study can be
used in larger scales. The results showed that the interfa-
As Fig. B1 shows, the authors ran simulations of three diame- cial area, liquid holdup, and pressure drop are not affected in
ters: 25.4, 50.8, and 101.6 mm at the liquid flow rate of 0.3 l/min, the flow fields with different diameters. Thus, 22 mm diame-
and the gas flow rate of 1 l/min to investigate the differ- ter of the flow field was selected to reduce the computational
ences of the interfacial area, liquid holdup, and pressure drop cost.
Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54 53

Fig. B1 – Effect of (a) fractional area, (b) liquid holdup, and (c) pressure drop in various diameters of the flow field.

References International Conference on Scale Space and Variational


Methods in Computer Vision., pp. 386–397, Springer.
Ataki, A., Bart, H.-J., 2006. Experimental and CFD simulation
Ahipo, Y., Auroux, D., Cohen, L.D., Masmoudi, M., 2011. A hybrid
study for the wetting of a structured packing element with
scheme for contour detection and completion based on
liquids. Chem. Eng. Technol. 29, 336–347.
topological gradient and fast marching
Atmakidis, T., Kenig, E.Y., 2009. CFD-based analysis of the wall
algorithms-application to inpainting and segmentation. In:
effect on the pressure drop in packed beds with moderate
54 Chemical Engineering Research and Design 1 4 7 ( 2 0 1 9 ) 43–54

tube/particle diameter ratios in the laminar flow regime. Haroun, Y., Raynal, L., Legendre, D., 2012. Mass transfer and
Chem. Eng. J. 155, 404–410. liquid hold-up determination in structured packing by CFD.
Atmakidis, T., Kenig, E.Y., 2012. Numerical analysis of mass Chem. Eng. Sci. 75, 342–348.
transfer in packed-bed reactors with irregular particle Haroun, Y., Raynal, L., Alix, P., 2014. Prediction of effective area
arrangements. Chem. Eng. Sci. 81, 77–83. and liquid hold-up in structured packings by CFD. Chem. Eng.
Bai, H., Theuerkauf, J., Gillis, P.A., Witt, P.M., 2009. A coupled DEM Res. Des. 92, 2247–2254.
and CFD simulation of flow field and pressure drop in fixed Iliuta, I., Larachi, F., 2001. Mechanistic model for
bed reactor with randomly packed catalyst particles. Ind. Eng. structured-packing-containing columns: irrigated pressure
Chem. Res. 48, 4060–4074. drop, liquid holdup, and packing fractional wetted area. Ind.
Billet, R., Schultes, M., 1993. Modelling of packed tower Eng. Chem. Res. 40, 5140–5146.
performance for rectification, absorption and desorption in Kajishima, T., Taira, K., 2017. Reynolds-Averaged Navier–Stokes
the total capacity range. The 3rd Korea-Japan Symposium. Equations, Computational Fluid Dynamics. Springer, pp.
Billet, R., Schultes, M., 1999. Prediction of mass transfer columns 237–268.
with dumped and arranged packings: updated summary of Launder, B.E., Spalding, D., 1974. The numerical computation of
the calculation method of billet and schultes. Chem. Eng. Res. turbulent flows. Comput. Methods Appl. Mech. Eng. 3,
Des. 77, 498–504. 269–289.
Boccardo, G., Augier, F., Haroun, Y., Ferre, D., Marchisio, D.L., 2015. Liu, J., Gao, H.-C., Peng, C.-C., Wong, D.S.-H., Jang, S.-S., Shen, J.-F.,
Validation of a novel open-source work-flow for the 2015. Aspen plus rate-based modeling for reconciling
simulation of packed-bed reactors. Chem. Eng. J. 279, laboratory scale and pilot scale CO2 absorption using aqueous
809–820. ammonia. Int. J. Greenhouse Gas Control 34, 117–128.
Boccardo, G., Sethi, R., Marchisio, D.L., 2019. Fine and ultrafine Lopes, R.J., Quinta-Ferreira, R.M., 2009. CFD modelling of
particle deposition in packed-bed catalytic reactors. Chem. multiphase flow distribution in trickle beds. Chem. Eng. J. 147,
Eng. Sci. 198, 290–304. 342–355.
Brackbill, J., Kothe, D.B., Zemach, C., 1992. A continuum method Ma, Y., Cao, X., Feng, X., Ma, Y., Zou, H., 2007. Fabrication of
for modeling surface tension. J. Comput. Phys. 100, 335–354. super-hydrophobic film from PMMA with intrinsic water
Bravo, J.L., Fair, J.R., 1982. Generalized correlation for mass contact angle below 90. Polymer 48, 7455–7460.
transfer in packed distillation columns. Ind. Eng. Chem. Onda, K., Takeuchi, H., Okumoto, Y., 1968. Mass transfer
Process Des. Dev. 21, 162–170. coefficients between gas and liquid phases in packed
Caulkin, R., Ahmad, A., Fairweather, M., Jia, X., Williams, R., 2007. columns. J. Chem. Eng. Jpn. 1, 56–62.
An investigation of sphere packed shell-side columns using a Partopour, B., Dixon, A.G., 2017. An integrated workflow for
digital packing algorithm. Comput. Chem. Eng. 31, 1715–1724. resolved-particle packed bed models with complex particle
Caulkin, R., Ahmad, A., Fairweather, M., Jia, X., Williams, R., 2009. shapes. Powder Technol. 322, 258–272.
Digital predictions of complex cylinder packed columns. Stichlmair, J., Bravo, J., Fair, J., 1989. General model for prediction
Comput. Chem. Eng. 33, 10–21. of pressure drop and capacity of countercurrent gas/liquid
Caulkin, R., Fairweather, M., Jia, X., Gopinathan, N., Williams, R., packed columns. Gas Sep. Purif. 3, 19–28.
2006. An investigation of packed columns using a digital Valluri, P., Matar, O.K., Hewitt, G.F., Mendes, M., 2005. Thin film
packing algorithm. Comput. Chem. Eng. 30, 1178–1188. flow over structured packings at moderate Reynolds numbers.
Chen, J., Liu, C., Yuan, X., Yu, G., 2009. CFD simulation of flow and Chem. Eng. Sci. 60, 1965–1975.
mass transfer in structured packing distillation columns. Van Baten, J., Krishna, R., 2002. Gas and liquid phase mass
Chin. J. Chem. Eng. 17, 381–388. transfer within KATAPAK-S structures studied using CFD
Dixon, A.G., Nijemeisland, M., Stitt, E.H., 2013. Systematic mesh simulations. Chem. Eng. Sci. 57, 1531–1536.
development for 3D CFD simulation of fixed beds: contact Wang, C., 2015. Mass Transfer Coefficients and Effective Area of
points study. Comput. Chem. Eng. 48, 135–153. Packing, M.S. CHE. University of Texas.

Das könnte Ihnen auch gefallen