Você está na página 1de 12

Chemical Engineering Science 132 (2015) 259–270

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Porous media Eulerian computational fluid dynamics (CFD) model


of amine absorber with structured-packing for CO2 removal
Dung A. Pham a, Young-Il Lim a,n, Hyunwoo Jee b, Euisub Ahn b, Yongwon Jung b
a
CoSPE, Department of Chemical Engineering, Hankyong National University, Jungang-ro 327, Anseong-si 456-749, Gyeonggi-do, South Korea
b
GS E&C, Gran Seoul, 33 Jonro Jongno-gu, Seoul 110-121, South Korea

H I G H L I G H T S

 A porous media CFD model was developed for amine absorber with structured-packing.
 Porous resistance, gas–liquid momentum exchange, and liquid dispersion were included.
 Mass transfer between two phases and chemical reaction in liquid phase were involved.
 Pressure drop, liquid holdup and CO2 concentration were obtained from the CFD model.

art ic l e i nf o a b s t r a c t

Article history: A gas–liquid Eulerian porous media computational fluid dynamics (CFD) model was developed for an
Received 3 December 2014 absorber to remove CO2 from natural gas by mono-ethanol-amine (MEA). The three dimensional
Received in revised form geometry of the amine absorber packed with nine elements of Mellapak 500.X (M500X) was constructed
1 April 2015
for the CFD domain. The momentum conservation equation included the porous resistance, gas–liquid
Accepted 3 April 2015
momentum exchange and liquid dispersion to replace the structured-packing by the porous media
Available online 20 April 2015
model. The mass conservation equation involved the mass transfer of CO2 gas into the MEA solution and
Keywords: one chemical reaction. The mesh independent test was performed on coarse, medium and fine meshes
Natural gas and the medium mesh (37,400 cells) was selected for further investigation. Parameters of the CFD model
CO2 removal
were adjusted to fit experimental data (wet pressure drop, liquid holdup, and CO2 mole fraction along
Amine absorber
the column height) measured in a CO2-MEA system on M500X. This study demonstrated that the porous
Structured-packing
Computational fluid dynamics (CFD) media CFD model can treat both hydrodynamics and CO2 removal in the amine absorber with
Porous media model structured-packing.
& 2015 Elsevier Ltd. All rights reserved.

1. Introduction monoethanolamine (MEA) and diethanolamine (DEA), are often


used for the absorbent because of low volatility, thermal stability,
Carbon dioxide (CO2) removal plays a crucial role in many high reactivity, and low cost (Hikita et al., 1977). The performance
industrial processing activities such as oil refining, the synthesis of of the amine absorber is generally determined by mass-transfer
ammonia, and natural gas purification (Aroonwilas et al., 2003). In efficiency of column packings providing the gas and liquid con-
these processes, CO2 and hydrogen sulfide (H2S) are considered to tacting area (Aroonwilas et al., 2001).
be impurities, and the acid gases must be reduced to prevent Structured packings are well recognized as column internal
corrosion to materials and harmfulness to both the environment devices offering excellent mass transfer efficiency and low pressure
and human beings (Rahimpour et al., 2013). drop (Aroonwilas et al., 2001; Beugre et al., 2011; Owens et al., 2013;
CO2 is separated or captured by absorption, adsorption, mem- Sun et al., 2013). The structured-packing has been widely accepted in
brane, and cryogenic technologies (Pires et al., 2011). The amine many industrial applications to distillation, dehydration, extraction,
absorption technology is well-established with numerous indus- and absorption due to a large effective area per unit volume ranging
trial installations (Zhuang et al., 2011) and ready for large scale use from 125 to 750 m2/m3 and a large void fraction (Aroonwilas et al.,
(Kwak et al., 2012). Aqueous solutions of ethanolamine, especially 2001; Haroun et al., 2014, 2012; Shilkin et al., 2006).
In recent years, many experimental and modeling approaches
have been reported for amine absorbers with the structured-
n
Corresponding author. Tel.: þ 82 31 670 5207; fax: þ 82 31 670 5209. packings. The CO2 reaction kinetics with MEA and DEA were studied
E-mail address: limyi@hknu.ac.kr (Y.-I. Lim). at various temperatures (Hikita et al., 1977). The vapor–liquid

http://dx.doi.org/10.1016/j.ces.2015.04.009
0009-2509/& 2015 Elsevier Ltd. All rights reserved.
260 D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270

equilibrium of amine-CO2 systems was presented as a function of tuned with the aid of a CO2 concentration profile experimentally
temperature and amine concentration (Penttilä et al., 2011). measured by Aroonwilas et al. (2003). This study shows that the
The mass transfer efficiency of structured-packings was investi- porous media CFD model can represent both hydrodynamics and
gated in a MEA-CO2 system according to operating and design mass transfer in the amine absorber with structured-packing for
parameters (Aroonwilas et al., 2001). A mechanistic model that was CO2 removal.
able to predict hydrodynamics and mass transfer performance was In Section 2, the CFD calculation domain is described for an
proposed, and its results were compared to experimental data for amine absorber packed with nine elements of M500X. Section 3
several structured-packings and operating conditions (Aroonwilas presents the porous media CFD model including porous resistance,
et al., 2003). A hydrodynamic analogy-based model (HAM) with a momentum exchange, liquid dispersion, mass transfer and chemi-
simplified internal packing geometry was presented to predict cal reaction. Section 4 covers the mesh independent test, the
temperature and composition profiles in distillation columns model parameters adjustment, and the effect of mass transfer.
(Shilkin et al., 2006). A two-dimensional tubular channel model Finally the conclusion is followed in Section 5.
was solved by the volume of fluid (VOF) method, and the liquid
holdup and mass transfer were studied as a function of the liquid
flow rate and structured packing geometry (Haroun et al., 2012). 2. Amine absorber for CO2 removal
The computational fluid dynamics (CFD) was a powerful tool to
investigate effective interfacial area, liquid holdup, and fluid flow The dimension of an amine absorber was obtained from the
behaviors in structured packings (Haroun et al., 2014). The fluid literatures (Aroonwilas et al., 2003, 2001). The geometric char-
dynamic performance of packing internals such as gas and liquid acteristics (i.e., specific surface area, voidage, corrugation angle,
distributors, and liquid collectors was analyzed by using three- and element height) of three Mellapak structured-packings are
dimensional (3D) CFD simulation, compared to experimental listed in Table 1. The corrugated structured-packing height was
measurements of the gas velocity at the top of a packed column 2.205 m corresponding to the nine packings of M500X (Mellapak
(Mohamed Ali et al., 2003). A multi-scale CFD simulation sequen- 500.X). The voidage or porosity (ε) is defined as
tially approaching from small to large scales of a structured- VG þVL
packing column was presented to examine the liquid holdup and ε¼ ¼ αG þ αL ð1Þ
V
pressure drop (Raynal and Royon-Lebeaud, 2007). Another multi-
where VG and VL are the gas and liquid volumes, respectively, and V is
scale strategy combining a small scale gas–liquid VOF model with
the total volume. The sum of the gas (αG ¼ V G =V) and liquid
a unit network model was proposed to calculate the liquid holdup
(αL ¼ V L =V) volume fractions is the voidage (ε). Fig. 1 illustrates that
and liquid distribution (Sun et al., 2013). The representative
the representative elementary volume (V) includes VG, VL, as well as
element unit (REU) or periodic element unit (PEU) corresponding
the solid volume of corrugated sheets. The transport equations will
to a few centimeters of packing was used to calculate the pressure
be expressed inside the elementary volume (see Section 3).
drop and mass transfer by 3D CFD simulation (Lautenschleger
et al., 2015; Raynal and Royon-Lebeaud, 2007; Sun et al., 2013).
Recently, Owens et al. (2013) solved gas phase hydrodynamics 2.1. Geometry of amine absorber and liquid distributor
by a 3D turbulent CFD model, considering a real geometry of a
corrugated and perforated structured-packing (about 50 cm The three dimensional (3D) geometry of the amine absorber is
height) with over 50 million meshes. The dry pressure drop was depicted in Fig. 2. The amine absorber consists of nine elements
obtained directly from the CFD simulation and validated with (element height of 24.5 cm), a liquid distributor in the top, and a
experimental data. Fourati et al. (2013) addressed a gas–liquid bottom side for the uniform gas inlet. The diameter of the column is
porous media CFD model incorporated with the porous resistance, equal to that of the packing element (¼0.1 m). The liquid distributor
momentum transfer between the two phases, and liquid disper- with twelve holes of 1.2 cm diameter is located at 0.6 cm just above
sion. The liquid spreading coefficient was estimated from experi- the packing. The front and top views are shown in Fig. 2b and c,
mental measurements. Since the porous media model simplified respectively. The mono-ethanol-amine (MEA) solution is injected
the complex structured-packing geometry to a homogeneous through the liquid distributor, and natural gas containing 15 mol%
porous material, the mesh number was reduced much. The study CO2 enters from the bottom of the amine absorber.
showed that the porous media CFD model can reflect hydrody- Since the structured-packing was assumed as a porous medium
namics and gas–liquid interactions of structured-packings. How- in this study, the geometric characteristics of the packing were not
ever, there remains a need for a complete porous media CFD fully considered. However, the porous media zone mimicked key
model that can treat mass transfer and chemical reaction in the features of the structured-packing by means of the porous resis-
amine-CO2 system. tance and liquid dispersion. The porous media model will be
In this study, the gas–liquid Eulerian porous media CFD model presented in Section 3.
is extended to consider mass transfer and chemical reaction in the
amine absorber with structured-packing. Model parameters of the 2.2. Meshing of CFD calculation domain
porous resistance and momentum exchange are adjusted to fit
experimental data (pressure drop and liquid holdup) measured on The mesh structure of the amine absorber is shown in Fig. 3.
Mellapak 500.X (M500X). The mass transfer rate coefficient is The top of the column has a fine hexahedral mesh structure to
represent the twelve holes in the liquid distributor, while a
relatively coarse mesh structure is used in the porous zone just
Table 1 below the distributor (see Fig. 3c and d). The non-conformal
Geometric characteristics of Mellapak structured-packings.
meshing strategy is applied to reduce the number of meshes at
Packing M250X M500X M500Y the interface between the liquid distributor and the porous zone.
Keeping the same number of meshes in the non-porous zone
Specific surface area, as (m2/m3) 250 500 500 (top and bottom of the column), coarse, medium, and fine meshes
Voidage (or porosity), ε (–) 0.98 0.91 0.91 (about 12,000, 37,000, and 103,000, respectively) are implemented
Corrugation angle, θ (o) 60 60 45
Element height, he (m) 0.253 0.245 0.205
to the porous zone in order to investigate the effect of the number
of meshes (see Section 4.1).
D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270 261

Fig. 1. Gas, liquid and total volumes within structured-packing; (a) structured packing, (b) representative elementary volume.

Fig. 2. Geometry of amine absorber for CFD calculation domain.

3. Porous media Eulerian CFD model Eulerian approach that treats the two phases flow as the inter-
penetrating continua. The porous resistance and liquid dispersion
A 3D porous media CFD model was developed for the amine were considered to imitate hydrodynamics of the structured-
absorber. The fluid flow was assumed to be incompressible and packing. The momentum and mass transfers between the two
isothermal. The gas and liquid phases were modeled by the phases were taken into account and one chemical reaction
262 D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270

Fig. 3. Non-conformal mesh structure in CFD calculation domain.

 
occurred in the liquid phase. The momentum and mass equations
!
∂ αL u L   , , ,  ! 
! , ! ! ,
ρL ¼  ρL u L U ∇ αL u L  ∇ P þ μL ∇ U ∇ αL u L þ ρL αL g  S L
deal with quantities that are averaged over a representative ∂t
elementary volume (V) (Fourati et al., 2013), as shown in Fig. 1b. ð3Þ
!
where P is the pressure, g is the gravitational acceleration,
,
and m
3.1. Continuity equation is the viscosity. The momentum source term (S ) includes the
! !
porous resistance ( F porous ), the momentum exchange ( F exch;GL ),
!
The continuity equation for the gas and liquid phases is and the liquid dispersion ( F disp ):
expressed as follows: , αG ! ! !
,   SG ¼ F porous;G þ ε F exch;GL þ F disp;G
∂  ! ε
αG ρG ¼ ∇ U αG ρG u G  r GL
∂t , αL ! ! !
∂  ,   S L ¼ F porous;L  ε F exch;GL þ F disp;L ð4Þ
! ε
αL ρL ¼  ∇ U αL ρL u L þ r GL ð2Þ
∂t
here, αL =ε ¼ V L =ðV L þ V G Þ means the liquid volume fraction to the
!
where ρ is the density (kg/m3), u is the interstitial volume- empty volume of packing (Fourati et al., 2013).
averaged velocity vector, and r GL (kg/m3/s) is the total mass
transfer rate per unit volume from gas to liquid. The subscripts G 3.2.1. Porous resistance force
and L stand for gas and liquid. The gas and liquid volume fractions The porous resistances (or solid–fluid drag forces) were derived
(αG and αL ) change with time (t) and space (x, y, and z) that are the from a generalized Ergun correlation in packed-beds (Fourati et al.,
,
independent variables. The Nabla operator (∇ ) is a spatial deriva- 2013; Iliuta et al., 2004):
tive vector. The continuity equation means the overall mass  
balance in each phase. The first term of Eq. (2) is the accumulation , a UE1 a2s b U E2 as αG ρG !  !
F porous;G ¼ ð1  f e Þ U μG þ U  u G u G
term of the total mass within the unit volume. The second term 36 ε 6 ε
!
b UE2 εas !  !

denotes the convection term of the total mass. , a U E1 εas 2
F porous;L ¼ f e U 2 μL þ U ρL  u L  u L ð5Þ
36 αL 6 αL
3.2. Momentum conservation equation
where E1 and E2 are the Ergun coefficients, as [m2/m3] is the
The following assumptions and limitations should be recog- specific surface area of the packing, and fe is the fraction of wetting
nized in the porous media momentum model (ANSYS, 2013): area. a and b are the modification factors for the Ergun coefficient.
(1) the effect of turbulence is ignored in the porous medium, The wetting-area fraction (fe) is the ratio of the effective contacting
(2) the porosity (ε) of packing is isotropic, (3) the interstitial area (ae) between the gas and liquid phases to the specific surface
velocity is available with the porosity, and (4) the porous media area (as), which plays a crucial role in connecting hydrodynamics
momentum resistance is calculated separately in each phase. and mass transfer. Since the effective interfacial area (ae) cannot be
The porous,media model was constructed by adding momen- directly captured because of an interpenetrating behavior of
tum sources (S ) into the conventional incompressible Newtonian multiphase flow in the Eulerian approach, a function inspired
fluid momentum equation (or Navier–Stokes equation). from Tsai et al. (2011) is proposed:
 
"  4=3 #0:116
!
∂ αG u G   , , ,   ae ρ Q
! , ! ! ! , f e ¼ ¼ 1:34 L g 1=3 Q ¼ 5:1189AhL
2:854
ð6Þ
ρG ¼  ρG u G U ∇ αG u G  ∇ P þ μG ∇ U ∇ αG u G þ ρG αG g  S G as σ Lp
∂t
D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270 263

,
where σ [N/m] is the surface tension of MEA solution, Q [m3/s] is (9) and (10). u D is the drift velocity given by:
the liquid flow rate, Lp [m] means the wetted perimeter, A [m2] is , 
 
the cross-sectional area of column, and hL is the liquid holdup. The ,
u G =αG ,
second equation of Eq. (6) was regressed from air–water experi- u D;G ¼  f spread ∇ αG
αG
mental data (Suess and Spiegel, 1992) for liquid holdup (hL) versus , 
 
liquid load (qL) in M500X (see Section 4.3 for the experimental ,
u L ,
u D;L ¼  f spread ∇ αL ð12Þ
data). When the physical properties (ρL, σ, and Lp) are assumed to αL
be constant, fe is expressed as a function of hL and has an average
where fspread is the spread factor whose dimension is the length.
value of 0.75 (or ae ¼375 m2/m3) at a mean liquid holdup of , ,
Here, it is evident that ∇ αG ¼  ∇ αL from Eq. (1). Note that the
hL ¼0.094 (see Section 4.2). The effective surface area (ae) for mass !
magnitude of the gas dispersion force ( F disp;G ) is much smaller
transfer changes dynamically within the porous media domain !
than that of the liquid dispersion force ( F disp;L ). The spread factor
according to the local liquid holdup (hL) calculated from the
was obtained from the liquid flow distribution interpreted with a
Eulerian CFD model. It is expected that Eq. (6) is valid in the range
convection-diffusion equation of qL (liquid load or liquid volu-
of 0.05 rhL r 0.15.
metric flux, m3/m2/h) (Fourati et al., 2013):
The first and second terms on the right hand side in Eq. (5)
 
represent the viscous and inertial loss of momentum, respectively. ∂qL 1 ∂ ∂qr
¼ f spread r ð13Þ
In general, the solid–fluid drag forces are described as the follow- ∂z r ∂r ∂r
ing forms (ANSYS, 2013):
where qr is the radial liquid load. A good spreading factor was
 
, , 1 , , found to be fspread ¼7.4 mm for M250X (Fourati et al., 2013). In the
F porous ¼ μRvis Uu þ ρu Rin U u ð7Þ
2 present simulation, this value was also used for M500X because of
the same corrugated angle (601).
The viscous and inertial momentum losses in porous media
Thus, the viscous and inertial loss resistances (Rvis and Rin) are ! !
( F porous ), momentum exchange drag force ( F exch;GL ), and mechan-
given in the analogy of Eqs. (5) and (7): !
ical dispersion force ( F disp ) were supplied as the user defined
a U E1 a2s b U E 2 as α G function (UDF).
Rvis;G ¼ ð1  f e Þ U ; Rin;G ¼ ð1  f e Þ U
36 ε 3 ε
a UE1 εa2s b U E2 εas 3.3. Species transport equation
Rvis;L ¼ f e U ; Rin;L ¼ f e U ð8Þ
36 α2L 3 αL
It was assumed that there are two species such as CH4 and CO2
The gas–solid drag (KGS) and liquid–solid drag (KLS) coefficients in the gas phase. It was also supposed that MEA and other species
are rearranged from Eq. (5) for further usage. in the liquid phase do not vaporize into the gas phase. Even
though the reactions during the absorption of CO2 into the
a UE1 a2s bU E2 as αG ρG !  aqueous amine solution are remarkably complex, just one chemi-
K GS ¼ U μG þ U  u G
36 ε 6 ε cal reaction is considered in this study for the CO2-MEA system.
a UE1 εa2s b U E2 εas ! 
 Assuming that mass transfer between the gas and liquid phases
K LS ¼ U 2 μL þ U ρL  u L  ð9Þ occurs for only CO2, the mass conservation equation is expressed
36 αL 6 αL
in the two phases.
 
∂ αG ρG;i ! ,   
¼  u G U ∇ αG ρG;i þ DG;i ∇2 αG ρG;i  r GL;i ; i ¼ CO2
∂t
3.2.2. Gas–liquid momentum exchange  
! ∂ αL ρL;k ! , 
The gas–liquid momentum exchange ( F exch;GL ) may be written ¼  u L U ∇ αL ρL;k þ DL;k ∇2 ðαL ρL;k Þ þ r GL;i þ αL Rk ;
∂t
as follows (Fourati et al., 2013)
k ¼ MEA; CO2 ; HORNH3þ ; HORNHCOO  ð14Þ
 
! c U E1 a2s d UE2 as ! ! ! !
F exch;GL ¼ f e U μG þ U ρG  u G  u L  U ð u G  u L Þ where Di is the diffusion coefficient of species i, rGL is the mass
36 εαG 6 ε
transfer rate between gas and liquid applied to only CO2, and Rk is
! !
¼ K IG ð u G  u L Þ ð10Þ the homogeneous chemical reaction rate which will be described
in detail in Section 3.3.2. Owing to the continuity equation, the
where c and d are the modification factors for the Ergun coeffi-
mass conservation equation is given for (N  1) species where N is
cients. The two parameters should be adjusted according to the
the number of species.
type of structured-packings. KIG is the drag coefficient of momen-
tum exchange.
3.3.1. Mass transfer rate (rGL)
In general, the overall mass transfer coefficient derived from
3.2.3. Liquid dispersion force
the two film theory is used (Gabrielsen et al., 2006; McCabe et al.,
The radial dispersion force model for structured-packing was
2005):
originated from the mechanical dispersion in a trickle-bed reactor
(Fourati et al., 2013; Lappalainen et al., 2009). In this work, the 1 1 1
¼ þ ð15Þ
original model by Lappalainen et al. (2009) was used to consider K x Ekx Hky
the liquid dispersion phenomena in the amine absorption column.
! where kx and ky are the mass transfer coefficients in the liquid and
The gas and liquid dispersion forces ( F disp ) are defined as:
gas films, respectively, Kx is the overall liquid mass transfer
, , , , coefficient, E stands for the enhancement factor promoted by
F disp;G ¼ αG K GS u D;G þ εK IG ðu D;G  u D;L Þ
chemical reaction, and H is the Henry’s constant. Most of the
, , , ,
F disp;L ¼ αL K LS u D;L  εK IG ðu D;G  u D;L Þ ð11Þ overall mass transfer resistance (1/Kx) comes from the liquid-side
resistance (1/kx) in the CO2-amine system (Hamborg and Vers-
where the drag coefficients (KGS, KLS, and KIG) are found in Eqs. teeg, 2012; Tunnat et al., 2014). Thus, the overall mass transfer
264 D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270

coefficient may be assumed to: which is expressed as an Arrhenius equation with the pre-
K x ffi Ekx ð16Þ exponential factor (k0) and activation energy (Ea). Five species
exist in the liquid phase: MEA, H2O, HORNH3þ , HORNHCOO  , and
CO2(l). Since the molar concentration (C i ¼ ρL;i =M w ) is preferred
The enhancement factor (E) was expressed as a function of the rather than the mass concentration (ρL;i ), the chemical reaction
MEA concentration (Gabrielsen et al., 2006; Tobiesen et al., 2007). rate in Eq. (20) is converted into:
If the MEA concentration does not change much throughout the
dC CO2
reaction zone, the factor can be treated as a constant. In fact, the RCO2 ¼ ¼ kc C CO2 C MEA ð21Þ
dt
MEA concentration varied from 2.6 to 3.0 kmol/m3 in this amine
absorber with the nine packs. A simple mass transfer rate for CO2
with only the liquid film resistance is applied in this study: The mass transfer rate (r GL ) and chemical reaction rate (RCO2 )
0 were also incorporated into the species transport equation by an
r GL ¼ kx ae ðρnL;CO2  ρL;CO2 Þ ð17Þ
UDF. After the MEA-CO2 reaction, the physical properties of liquid
where ρnL;CO2 is the equilibrium liquid mass concentration of CO2 at (diffusivity, viscosity and surface tension) will change, and they
the gas–liquid interface. k0 x is a lumped value including the liquid may affect the effective interfacial area and the mass transfer
film resistance and the enhancement by the chemical reaction, coefficient. However, the effects were ignored in this simple
which depends on the system and can be estimated from theore- reaction model.
tical or empirical correlations (Aroonwilas et al., 2003). The ρnL;CO2 is
determined by Henry’s law: 3.4. Solution strategy and model parameters
P CO2
ρnCO2 ¼ M w;CO2 ð18Þ The continuity equation, momentum and mass conservation
H CO2 MEA
equations were solved by using ANSYS Fluent (ANSYS Inc., USA).
where M w;CO2 is the molecular weight of CO2, P CO2 [Pa] is the partial The phase coupled SIMPLE method was chosen for pressure–
pressure of CO2 in the gas phase, and H CO2 MEA is the Henry’s velocity coupling. The second-order upwind and QUICK schemes
constant of CO2 in the aqueous MEA solution. The Henry’s constant were used for the spatial discretization of the momentum and
for the MEA-CO2 aqueous system was obtained from the vapor– mass equations, and the volume fraction, respectively.
liquid equilibrium in N2O–CO2 analogy (Clarke, 1964; Penttilä Table 2 shows parameter values of the porous media model
et al., 2011). which were used for the additional source term in the momentum
H CO2 water equation. The Ergun coefficients (E1 and E2) reported for M250X by
H CO2 MEA ¼ H N2 OMEA ð19Þ Iliuta and Larachi (2004) were taken without modification. How-
H N2 Owater
ever, as M500X was a packing material in this study, the four
modification factors of the Ergun coefficients (a, b, c, and d) were
The Henry’s constants of CO2 and N2O in pure water (H CO2 water applied in Eqs. (8) and (10) and adjusted for M500X (see Section
and H N2 Owater ) as a function of temperature was found elsewhere 4.3). The liquid spread factor (fspread) proposed by Fourati et al.
(Penttilä et al., 2011). To obtain H N2 OMEA at a liquid mole fraction (2013) was taken. The wetted perimeter (Lp) was 4.05 m
of MEA (xMEA) of 0.06 (or 31.7 wt%), a regression equation for M500X.
estimated from experimental data of Penttilä et al. (2011) is shown The model parameters related to mass transfer and chemical
in Fig. 4. The Henry’s constant used in this study was reaction are reported in Table 3. The liquid mass transfer coeffi-
H CO2 MEA ¼3690 Pa m3/mol at 304 K. cient (kx) was initially guessed from Aroonwilas et al. (2003) and
tuned with experimental data (see Section 4.3). The amine
3.3.2. Chemical reaction rate (R) absorber was operated at T¼304 K, and P ¼1 atm. The reaction
For the CO2-MEA system, the reaction rate may be expressed by rate coefficients (k0 and Ea) were obtained from Hikita et al. (1977).
the 2nd-order kinetics (Hikita et al., 1977): Table 4 presents the boundary conditions and the simulation
parameters. The three mesh numbers of the porous media zone
CO2 ðlÞ þ 2HORNH2 2HORNH3þ þ HORNHCOO 
were tested to investigate the effect of the mesh number on the
d
RCO2 ¼ ρL;CO2 ¼  kc ρL;CO2 ρL:MEA CFD solution. The multiphase Eulerian model was calculated by
dt
the unsteady-state, sequential (or pressure-based segregated), and
kc ¼ k0 e  Ea =ðRTÞ ð20Þ implicit solver. Gas and liquid inlet boundary conditions (inlet BC)
where kc is the chemical reaction rate constant depending on T, were taken from experimental conditions of Aroonwilas et al.
(2003). The outlet boundary condition (outlet BC) of gas was set to
15000 the mass flow outlet, while that of liquid was the pressure outlet.
The maximum iteration number was 100 at each time step and the
13000 convergence tolerance was 1  10  3.
At the beginning of the simulation, gas and liquid were injected
HN2O - MEA (Pa-m 3 /mol)

11000 into the packed column which was initially empty. After a wetting

9000
Table 2
Porous media model parameters.
7000

HN2O- MEA = - 0.3122T 2 + 315.3T - 61950 Parameters Value


5000 R² = 0.997
Ergun coefficients (–) E1 ¼ 160, E2 ¼ 0.16
3000 Modification factors of Ergun coefficients (–) a¼ 1.9, b¼ 1.37, c ¼0.17, d ¼1.16
273 293 313 333 353 373 393 413 Liquid spread factor (mm) fspread ¼ 7.4
Temperature (T, K) Wetted perimeter (m) Lp ¼ 4.05
Surface tension of MEA solution (N/m) σ¼ 0.062
Fig. 4. Henry’s law constant of N2O (H N2 O  MEA ) in 6 mol% MEA solution with Viscosities of gas and liquid (kg/m/s) μG ¼ 1.72  10  5, μL ¼ 1.0  10  3
respect to temperature (T).
D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270 265

period of about 35 s, the mass transfer and chemical reaction that the mesh number barely affect the gauge pressure (Pgage) and
started and it took about 15 s to reach a stable state. Thus, the CO2 mole fraction (yCO2 ).
operating time about 50 s was required to obtain a stable solution. However, some discrepancy of the liquid holdup (hL) is
observed in Fig. 5d according to the mesh resolution. The liquid
holdup of the coarse mesh shows a substantial difference from
4. CFD simulation results those of the medium and fine meshes. Since the accuracy of the
CFD solution comes in general at the cost of computational time
The gas–liquid Eulerian porous media CFD model was first of all on the fine mesh, the medium mesh was selected for further
solved on the three different mesh numbers. Then, liquid disper- researches.
sion was examined along the height of the porous medium
column. The CFD model parameters were adjusted with experi- 4.2. Liquid dispersion and liquid holdup
mental data such as the wet pressure drop, liquid holdup, and CO2
concentration profile along the column height. Finally, the effect of Fig. 6 displays the liquid dispersion with time at the top of the
mass transfer and chemical reaction was discussed. column. Due to the dispersion force in Eqs. (11) and (12) with
fspread ¼7.4 mm, the amine solution spreads out radially and fills up
4.1. Mesh independent test the column.
The liquid holdup of the whole column is shown in Fig. 7. The
A high-resolution mesh having 36,480 hexahedral cells was radially-averaged liquid holdup is a little smaller at both the top
built in the non-porous zone (top and bottom of the column) and the bottom than in the middle (see Fig. 7b), because a certain
because of the inlet and outlet of gas and liquid. The mesh length from the distributor is necessary for even liquid spreading,
structure of the non-porous zone was fixed. The mesh indepen- and the gas upstream at the bottom slightly reduces the liquid
dent test was performed for only the porous zone. holdup.
The coarse (11,880 hexahedral cells), medium (37,440 hexahe- It is worth noting that the present porous media model cannot
dral cells) and fine (103,455 hexahedral cells) meshes were built in capture gas and liquid turbulences, boundary surfaces between
the porous zone. The effect of the mesh number was examined on the gas and liquid phases, and directional hydrodynamics appear-
the total dry pressure drop, and gauge pressure, CO2 mole fraction ing in the real structured packing. Thus, this model is suitable to
and liquid holdup along the column height, as shown in Fig. 5. The identify an overall column performance rather than rigorous
liquid holdup (hL) is a liquid volume fraction (αL) averaged over an hydrodynamics inside the structured packing that are available
interesting volume: in a detailed CFD model with the real geometry of packing.
P
αL;i ΔV i
4.3. Parameter adjustment of porous media CFD model.
hL ¼ i P ð22Þ
ΔV i
i The parameters of the porous media CFD model were adjusted
where ΔV is the cell volume in the computational domain. Since with the experimental data measured in an amine absorber
the structured-packing was regarded as a homogeneous porous packed with M500X. Fig. 8 shows the wet pressure drop per unit
material with viscous and inertial momentum loss, the effect of length (ΔPwet/L, Pa/m) versus the F-factor (gas load factor) defined
the mesh number was not significant on the total dry pressure as:
drop, gauge pressure and CO2 concentration profile. The ratio of pffiffiffiffiffi
F  f actor ¼ uGS ρG ð23Þ
the total dry pressure drop (ΔPdry ¼ 10.19 Pa) on the coarse grid to
the maximum pressure drop ((ΔPdry)max ¼10.43 Pa) on the fine where uGS is the superficial gas velocity. The three experimental
grid is only 0.986, as seen in Fig. 5a. In Fig. 5b and c, it is more clear data measured in M500Y, M500X and M250Y at qL ¼ 24.4 m3/m2/h
were obtained from elsewhere (Sulzer, 2013; Tsai et al., 2011). The
Table 3
modification factors (a and b) of Ergun coefficient in Eq. (8) were
Model parameters of mass-transfer and chemical reaction. slightly changed to fit the packing, M500X (see Table 2). The CFD
simulation results with the model parameters tuned for M500X
Parameters Value agree well with its experimental data.
0 In Fig. 9, the liquid holdup obtained from the CFD simulation is
Liquid mass-transfer coefficient (m/s) kx ¼ 0.00104
Operating pressure (atm) P¼ 1.0 compared to experimental data of M500Y from Suess and Spiegel
Operating temperature (K) T ¼ 304 (isothermal) (1992). It was indicated that M500X and M500Y had the same
Pre-exponential factor (m3/kmol/s) k0 ¼ 5.928  104 liquid holdup (Spiegel and Duss, 2014). The water liquid load (qL)
Activation energy (cal/gmol) Ea ¼ 4.274  103 varies from 5 to 75 m3/m2/h and the F-factor is fixed at 0.21 Pa0.5
Diffusivities of gas and liquid (m2/s) DG ¼ 2.8  10  7, DL ¼ 2.88  10  9
(or qG ¼38.5 kmol/m2/h). The liquid holdup was averaged over the

Table 4
Boundary condition (BC) and simulation parameters.

Parameters Value

Mesh number (Ncell) About 12,000, 37,000, and 103,000 cells


Eulerian multiphase model Un-steady state, sequential, and implicit solver
Bottom-side BC Incompressible ideal gas inlet (back flow), liquid (pressure-outlet), gas load (natural gas) ¼38.5 kmol/m2/h, and gas mole composition:
CH4 ¼ 0.85, CO2 ¼0.15.
Top-side BC Gas (mass flow outlet)
Liquid distributor BC Liquid (mass flow inlet), liquid load ¼ 22.9 m3/m2/h, and liquid mass composition: MEA¼0.317, H2O¼0.683.
Maximum number of 100
iterations
Convergence tolerance 1  10  3
Total operation time (s) 50
266 D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270

1.000 2.5
103,000 cells

Height from column bottom (h, m)


37,000 cells
2
12,000 cells
0.995
ΔPdry /(ΔPdry )max

1.5
0.990
1

0.985
0.5

0.980 0
0 40 80 120 0 2 4 6 8 10 12
Number of cells (Ncell, x103 cells) Gauge pressure (Pgage, Pa)

2.5 2.5
103,000 cells 103,000 cells

Height from column bottom (h, m)


37,000 cells
Height from column bottom (h, m)

37,000 cells 12,000 cells


2 2
12,000 cells

1.5 1.5

1 1

0.5 0.5

0 0
0 5 10 15 20 0.090 0.092 0.094 0.096 0.098 0.100
Gas-phase CO 2 mole fraction (yCO2, %) Liquid holdup (hL)

Fig. 5. Mesh independent test on coarse, medium and fine meshes; (a) ratio of total dry pressure drop to maximum pressure drop (ΔPdry/(ΔPdry)max), (b) gauge pressure
(Pgage) along column height, (c) CO2 mole fraction (yCO2 ) along column height, (d) liquid holdup (hL) along column height.

Fig. 6. Liquid dispersion with time near liquid distributor.

whole packing zone. The liquid holdup increases with the increase and d). The other parameters such as fspread, fe (or ae), k0, Ea, and
of the liquid load. The modification factors (c and d) of the Ergun H CO2 MEA were taken from the existing references without mod-
coefficient in Eq. (10) were adjusted to fit the experimental data ification or with a simple regression. The mass transfer (r GL ) and
(Suess and Spiegel, 1992). chemical reaction (RCO2 ) rates in Eqs. (17) and (20) were applied to
The validation of the CFD model for the M500X packing itself calculate the overall CO2 removal efficiency on M500X. In Fig. 10a,
was performed, adjusting the Ergun modification factors (a, b, c, the contour of the CO2 mole fraction (yCO2 ) is shown for the whole
D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270 267

Fig. 7. Liquid holdup (hL) distribution (a) of whole column, and (b) along column height.

0.18
M500Y (Tsai et al. 2011)
CFD result (this study)
80 M500X (Sulzer, 2013)
M250Y (Tsai et al. 2011) 0.16 Experiment (Suess & Spiegel, 1992)
M500X, CFD result (this study)

0.14
60
Liquid hold-up (h L)

0.12
Δ Pwet /L (Pa/m)

0.10
40

0.08

20 0.06

0.04
0 20 40 60 80
0
0.4 0.6 0.8 1 3
Liquid load (q L, m /m /h)2

F-factor (Pa0.5)
Fig. 9. Water liquid holdup (hL) versus liquid load (qL) at F-factor¼ 0.21 Pa0.5.
Fig. 8. Wet pressure drop per unit length (ΔP wet =L) with respect to F-factor at
qL ¼24.4 m3/m2/h.

4.4. Effect of mass transfer and chemical reaction


column. The radially-averaged mole fraction along the height is
0
compared to experimental data in Fig. 10b. Only the liquid mass The effect of the mass transfer coefficient (kx ) on the CO2
0
transfer coefficient (kx ) was tuned to fit the experimental data removal efficiency is shown in Fig. 11. The mass transfer coefficient
carried out at qG ¼38.5 kmol/m2/h and qL ¼22.9 m3/m2/h was perturbed at 0.001047 0.0002 m/s. The higher mass transfer
0
(Aroonwilas et al., 2003). The adjusted kx was 0.00104 m/s. coefficient leads to the higher CO2 removal, as expected.
268 D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270

Fig. 10. Gas phase CO2 mole fraction (yCO2 , %) along column height (h, m). (a) CFD results in whole column, (b) comparison between CFD results and experimental data.

A mass transfer and chemical reaction model reduced from Eq. 2.5
(14) without hydrodynamics was taken into account for the Exp. (Aroonwilas et al., 2003)
comparison with the CFD model. Let the reduced model under
CFD result (k'x=0.00084 m/s)
the assumption of a homogeneous contact between the gas and
liquid phases (i.e., ae ¼390 m2/m3 or fe ¼ 0.78): CFD result (k'x=0.00104 m/s)
2
Height from column bottom (h, m)

CFD result (k'x=0.00124 m/s)


∂ 
ρ ¼  r GL;i ; i ¼ CO2
∂t G;i
∂  
ρ ¼ þ r GL;i þ Rk ; k ¼ MEA; CO2 ð24Þ 1.5
∂t L;k

The three ordinary differential equations were solved by a time


integrator for 11.3 s which is required to pass through M500X
1
(h ¼2.205 m) at the same gas flow rate as that of the CFD
simulation, qG ¼38.5 kmol/m2/h.
0
When the mass transfer coefficient (kx ) and chemical reaction
rate constant (kc) are the same as those of the CFD simulation, the
result of the reduced model is compared with the experimental
0.5
data and the CFD result of this study in Fig. 12. The height (h) of
the experimental data was converted to the time (t) by the gas
0
velocity (uG ¼0.194 m/s). The reduced model with kx ¼0.00104 m/s
2 3
and ae ¼390 m /m shows a lower CO2 removal efficiency than 0
that of the CFD result. The difference results from ae and the liquid 0 5 10 15
velocity (uL). The ae varies along the column and uL is about Gas-phase CO 2 mole fraction (yCO2, %)
0.068 m/s in the CFD model, while uL is regarded as the same 0
Fig. 11. Sensitivity of mass transfer coefficient (kx ) on CO2 mole fraction (yCO2 )
velocity as uG in the reduced model. Since the slow liquid flow along column height.
increases the absorption residence time, the CO2 removal effi-
ciency of the CFD model is higher.
Owens et al. (2013) who fully considered the detailed structure holdup without parameter tuning necessary in the porous media
of a structured-packing (Mellapak N250Y, inner diameter of CFD model (see Table 2) at the expense of computational time.
14.6 cm and height of 47.5 cm) used 50 million meshes to calculate Furthermore, the effective specific area (ae) would be computed
dry pressure drops of nitrogen gas by a turbulent CFD model. The directly by the rigorous CFD model. The porous media CFD model
rigorous CFD model may predict the wet pressure drop and liquid using about 3000 times smaller meshes than the rigorous one
D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270 269

12 be applied for the development and design of the liquid distributor


Exp. (Aroonwilas et al., 2003) and collector. This paper provided a potential to predict the CO2
CFD result (k x =0.00104m/s)
10 removal efficiency by using appropriate parameters of the porous
Reduced model (kx =0.00104m/s, a e=390m2/m 3, u G=0.194m/s)
media model for a given structured-packing.
8
time (t, s)

6 Nomenclatures

Latin letters
4

A cross-sectional area of column (m2)


2
a modification factor of Ergun equation
ae effective interfacial area (m2/m3)
0 as specific surface area of structured-packing (m2/m3)
0 2 4 6 8 10 12 14 16
b modification factor of Ergun equation
Gas-phase CO 2 mole fraction (y CO , %)
c modification factor of Ergun equation
Fig. 12. Comparison of CO2 mole fraction (yCO2 ) versus time (t) for experimental Ci molar concentration of species i in liquid phase (kmol/
data, CFD and reduced model results. m3 )
d modification factor of Ergun equation
should be improved to obtain more reliable and predicable D diffusivity coefficient (m2/s)
hydrodynamic characteristics of structured-packing. The CO2 E1, E2 Ergun coefficients
removal efficiency of the amine absorber was computed efficiently Ea activation energy (cal/gmol)
!
by the porous media model on the much reduced mesh. F disp liquid dispersion force (N/m3)
fe fraction of wetting area
!
F momentum exchange force (N/m3)
!exch
5. Conclusions F porous porous resistance force (N/m3)
fspread spread factor (m)
The three dimensional geometry of an amine absorber packed g gravitational acceleration (m/s2)
with nine elements of Mellapak 500.X (M500X) was constructed he element height (m)
for the porous media CFD simulation. The amine absorber was hL liquid holdup (m3/m3)
operated to remove 15 mol% CO2 contained in natural gas by using H Henry’s constant (Pa m3/mol)
MEA (mono-ethanol-amine) as an absorbent. To represent the k0 pre-exponential factor (m3/kmol/s)
structured-packing by the porous media CFD model, the porous kc chemical reaction rate coefficient (m3/kmol/s)
resistance force, the gas–liquid momentum exchange, and the KGS gas–solid drag coefficient (kg/m3/s)
liquid dispersion force were added into a typical momentum KIG momentum exchange coefficient at the gas–liquid inter-
conservation equation. The mass conservation equation included face (kg/m3/s)
the mass transfer between the gas and liquid phases, and one KLS liquid–solid drag coefficient (kg/m3/s)
chemical reaction in the liquid phase. The two phases were kx liquid-side mass transfer coefficient (m/s)
0
modeled as a continuum by using the Eulerian approach. The kx lumped mass transfer coefficient (m/s)
three additional momentum forces, mass transfer and chemical Kx overall liquid mass transfer coefficient (m/s)
reaction were supplied by user-defined functions. The gas–liquid ky gas-side mass transfer coefficient (m/s)
Eulerian porous media CFD model was calculated by a finite Lp wetted perimeter (m)
volume-based solver, ANSYS Fluent. Mw molecular weight (kg/kmol)
The coarse (11,800 cells), medium (37,400 cells) and fine P pressure (atm)
(103,400 cells) meshes were built for the mesh independent test. Q liquid flow rate (m3/s)
The number of meshes affected little the pressure drop and CO2 qG gas load (kmol/m2/h)
removal efficiency. However, the liquid holdup of the coarse mesh qL liquid load (m3/m2/h)
showed a significant difference from those of the medium and fine r radial component in a cylindrical coordinate system (m)
meshes. The medium mesh was selected among the three meshes rGL mass transfer rate between gas and liquid (kg/m3/s)
as a compromise of accuracy and computational cost. The porous Ri chemical reaction rate of species i (kg/m3/s)
media momentum equation including the liquid dispersion force Rin inertial momentum loss (1/m)
reflected the characteristics of radial liquid dispersion of Rvis viscous momentum loss (1/m2)
structured-packing well. This study demonstrated that the porous S momentum source term (N/m3)
media CFD model can predict the wet pressure drop, liquid t time (s)
holdup, and CO2 removal efficiency in the amine absorber, adjust- T temperature (K)
!
ing the modification factors of the Ergun coefficients (a, b, c, and u interstitial volume-average velocity (m/s)
0 uGS superficial gas velocity (m/s)
d), and mass transfer coefficient (kx ).
In this study, the isotropic coefficients ignoring the directional V volume (m3)
property of structured-packing were used for the porous resis- xi liquid mole fraction of species i
tance, momentum exchange and liquid dispersion. An anisotropic yi gas mole fraction of species i
approach could reflect the complex structure of corrugated and
perforated sheets. It would be better to use a mass transfer Greek letters
coefficient dependent on the chemical reaction. More rigorous
chemical reactions in the liquid phase must be considered in the αG volume fraction of gas phase
future. The porous media CFD model presented in this study may αL volume fraction of liquid phase
270 D.A. Pham et al. / Chemical Engineering Science 132 (2015) 259–270

ε packing void fraction (porosity) Hikita, H., Asai, S., Ishikawa, H., Honda, M., 1977. The kinetics of reactions of carbon
θ corrugation angle (deg.) dioxide with monoethanolamine, diethanolamine and triethanolamine by a
rapid mixing method. Chem. Eng. J. 13, 7–12.
σ surface tension of MEA solution (N/m) Iliuta, I., Petre, C.F., Larachi, F., 2004. Hydrodynamic continuum model for two-
μ viscosity (kg/m/s) phase flow structured-packing-containing columns. Chem. Eng. Sci. 59,
ρ mass concentration or density (kg/m3) 879–888.
Kwak, N.-S., Lee, J.H., Lee, I.Y., Jang, K.R., Shim, J.-G., 2012. A study of the CO2 capture
pilot plant by amine absorption. Energy 47, 41–46.
Subscripts Lappalainen, K., Manninen, M., Alopaeus, V., 2009. CFD modeling of radial spread-
ing of flow in trickle-bed reactors due to mechanical and capillary dispersion.
Chem. Eng. Sci. 64, 207–218.
D drift Lautenschleger, A., Olenberg, A., Kenig, E.Y., 2015. A systematic CFD-based method
G gas phase to investigate and optimise novel structured packings. Chem. Eng. Sci. 122,
452–464.
L liquid phase
McCabe, W.L., Smith, J., Harriott, P., 2005. Unit Operations of Chemical Engineering,
seventh ed. McGraw-Hill, NY.
Mohamed Ali, A., Jansens, P.J., Olujić, Ž., 2003. Experimental characterization and
computational fluid dynamics simulation of gas distribution performance of
liquid (re)distributors and collectors in packed columns. Chem. Eng. Res. Des.
Acknowledgements 81, 108–115.
Owens, S.A., Perkins, M.R., Eldridge, R.B., Schulz, K.W., Ketcham, R.A., 2013.
Computational fluid dynamics simulation of structured packing. Ind. Eng.
This research was supported by a grant from the Program of Chem. Res. 52, 2032–2045.
KAIA funded by the Ministry of Land, Infrastructure and Transport Penttilä, A., Dell’Era, C., Uusi-Kyyny, P., Alopaeus, V., 2011. The Henry’s law constant
of the Korean government. (08 GASPLANT F02). The authors of N2O and CO2 in aqueous binary and ternary amine solutions (MEA, DEA,
DIPA, MDEA, and AMP). Fluid Phase Equilib. 311, 59–66.
appreciate the editing contribution of Patrick Bresnahan. Pires, J.C.M., Martins, F.G., Alvim-Ferraz, M.C.M., Simões, M., 2011. Recent develop-
ments on carbon capture and storage: an overview. Chem. Eng. Res. Des. 89,
1446–1460.
References Rahimpour, M.R., Saidi, M., Baniadam, M., Parhoudeh, M., 2013. Investigation of
natural gas sweetening process in corrugated packed bed column using
computational fluid dynamics (CFD) model. J. Nat. Gas Sci. Eng. 15, 127–137.
ANSYS, 2013. ANSYS Fluent User’s Guide. ANSYS Inc, Release 15.0. Raynal, L., Royon-Lebeaud, A., 2007. A multi-scale approach for CFD calculations of
Aroonwilas, A., Chakma, A., Tontiwachwuthikul, P., Veawab, A., 2003. Mathematical gas–liquid flow within large size column equipped with structured packing.
modelling of mass-transfer and hydrodynamics in CO2 absorbers packed with Chem. Eng. Sci. 62, 7196–7204.
structured packings. Chem. Eng. Sci. 58, 4037–4053. Shilkin, A., Kenig, E.Y., Olujic, Z., 2006. Hydrodynamic-analogy-based model for
Aroonwilas, A., Tontiwachwuthikul, P., Chakma, A., 2001. Effects of operating and efficiency of structured packing columns. AlChE J. 52, 3055–3066.
design parameters on CO2 absorption in columns with structured packings. Sep. Spiegel, L., Duss, M., 2014. Chapter 4—structured packings. In: Olujić, A.G. (Ed.),
Purif. Technol. 24, 403–411. Distillation. Academic Press, Boston, pp. 145–181.
Beugre, D., Calvo, S., Crine, M., Toye, D., Marchot, P., 2011. Gas flow simulations in a Suess, P., Spiegel, L., 1992. Hold-up of mellapak structured packings. Chem. Eng.
structured packing by lattice Boltzmann method. Chem. Eng. Sci. 66, Process. 31, 119–124.
3742–3752. Sulzer, 2013. Structured Packings for Distillation, Absorption and Reactive Distilla-
Clarke, J.K.A., 1964. Kinetics of absorption of cardon dioxide in monoethanolamine tion. Sulzer Chemtech Ltd.
solutions at short contact times. Ind. Eng. Chem. Fundam. 3, 239–245. Sun, B., He, L., Liu, B.T., Gu, F., Liu, C.J., 2013. A new multi-scale model based on CFD
Fourati, M., Roig, V., Raynal, L., 2013. Liquid dispersion in packed columns: and macroscopic calculation for corrugated structured packing column. AlChE J.
experiments and numerical modeling. Chem. Eng. Sci. 100, 266–278. 59, 3119–3130.
Gabrielsen, J., Michelsen, M.L., Stenby, E.H., Kontogeorgis, G.M., 2006. Modeling of Tobiesen, F.A., Svendsen, H.F., Juliussen, O., 2007. Experimental validation of a
CO2 absorber using an AMP solution. AlChE J. 52, 3443–3451. rigorous absorber model for CO2 postcombustion capture. AlChE J. 53, 846–865.
Hamborg, E.S., Versteeg, G.F., 2012. Absorption and desorption mass transfer rates Tsai, R.E., Seibert, A.F., Eldridge, R.B., Rochelle, G.T., 2011. A dimensionless model for
in chemically enhanced reactive systems. Part II: Reverse kinetic rate para- predicting the mass-transfer area of structured packing. AlChE J. 57, 1173–1184.
meters. Chem. Eng. J. 198–199, 561–570. Tunnat, A., Behr, P., Görner, K., 2014. Desorption kinetics of CO2 from water and
Haroun, Y., Raynal, L., Alix, P., 2014. Prediction of effective area and liquid hold-up in aqueous amine solutions. Energy Procedia 51, 197–206.
structured packings by CFD. Chem. Eng. Res. Des. 92, 2247–2254. Zhuang, Q., Pomalis, R., Zheng, L., Clements, B., 2011. Ammonia-based carbon
Haroun, Y., Raynal, L., Legendre, D., 2012. Mass transfer and liquid hold-up dioxide capture technology: issues and solutions. Energy Procedia 4,
determination in structured packing by CFD. Chem. Eng. Sci. 75, 342–348. 1459–1470.

Você também pode gostar