You are on page 1of 41

Accepted Manuscript

Title: Determination of Liquid Mass-transfer Coefficients for


the Absorption of CO2 in Alkaline Aqueous Solutions in
Structured Packing using Numerical Simulations

Authors: Bing Dong, X.G. Yuan, K.T. Yu

PII: S0263-8762(17)30346-5
DOI: http://dx.doi.org/doi:10.1016/j.cherd.2017.06.017
Reference: CHERD 2720

To appear in:

Received date: 19-12-2016


Revised date: 26-4-2017
Accepted date: 15-6-2017

Please cite this article as: Dong, Bing, Yuan, X.G., Yu, K.T., Determination of Liquid
Mass-transfer Coefficients for the Absorption of CO2 in Alkaline Aqueous Solutions in
Structured Packing using Numerical Simulations.Chemical Engineering Research and
Design http://dx.doi.org/10.1016/j.cherd.2017.06.017

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Determination of Liquid Mass-transfer Coefficients for the
Absorption of CO2 in Alkaline Aqueous Solutions in Structured
Packing using Numerical Simulations

Bing. Donga, X. G. Yuana*, K. T. Yub


State Key Laboratory for Chemical Engineering, Chemical Engineering Research

Center, and School of Chemical Engineering and Technology, Tianjin University,

Tianjin, 300072, China

Author Information

Corresponding Author

*Tel.: +86 022 27402786. Fax: +86 022 27402786.

E-mail: yuanxg@tju.edu.cn

The authors declare no competing financial interest.

To whom correspondence should be addressed.


Tel: +86 022 27402786. Email: yuanxg@tju.edu.cn

Graphical Abstract
Highlights

• Mass transfer process in structured packing is simulated with Volume of Fluid model

• A numerical method to locate the mass-transfer area is proposed

• An approach to determine the liquid mass transfer coefficient in structured packing is

established.

•The results in liquid mass transfer coefficients are validated with published models.

ABSTRACT

A 3D volume of fluid (VOF) model was used and a convection-diffusion equation was

implemented to simulate the reactive absorption process of CO2 into NaOH aqueous

solution. The simulations were performed based on a representative element unit (REU)

of Mellapak 500Y structured packing using direct numerical simulation (DNS) and

turbulent simulation. We proposed a method to locate the gas-liquid interface, based on

which the effective area of mass-transfer was determined, using the information

provided by the VOF model. Using the simulation results and material balance around

the liquid phase in the REU, the liquid mass transfer coefficient was calculated. The

obtained mass transfer coefficient was shown to agree well with existing models. We

also showed a two-equation turbulent model which consumes much less computation

efforts than the DNS, but it can also be applied for the same purpose. The procedure

proposed in this paper is expected to be a general approach to determining

systematically the mass-transfer coefficient in structured packing without relying on

empirical correlations or experiments.

Keywords: Structured packing; Effective transfer area; VOF method; CFD;

Liquid mass transfer coefficient;


Nomenclature

A=cross-sectional area of column, m2

aeff =effective area for mass transfer between the gas phaseand liquid phase (m2 m-3)

ap= specific surface area of packing, (m2 m-3)

B=packing channel base, m

CA= CO2 concentration(average), (mol m-3)

cin= inlet concentration(average), (mol m-3)

cout= outlet concentration(average), (mol m-3)

cs= saturation concentration, (mol m-3)

Dl = molecular diffusivity for mass transfer (m2 s-1)

Dt= turbulent diffusivity for mass transfer (m2 s-1)

E = enhancement factor

fi= the interfacial friction factor

g =acceleration due to gravity (m s-2)

H =Henry’s constant in the NaOH solution (kmolm-3 atm-1)

HA= physical absorption heat of mole CO2 absorbed (J kmol-1)

HR=chemical reaction heat of mole CO2 absorbed (J kmol-1)

(Hco2)w =Henry’s constant in pure water (kmol m-3 atm-1)

h= Crimp height (m)

Ic=ionic strengths of aqueous electrolyte solutions (kmol m-3)


jA=mass flux, (kg s-1 m-2)

Ks=salting-out parameter (L mol-1); for the CO2-NaOH system, Ks ) 0.142 L mol-1

k=surface curvature (m-1)

kL =liquid-phase mass transfer coefficient without chemical reaction (m s-1)

k2 =second-order reaction rate constant (m3 kmol-1 s-1)

L=mass flux, kg/m2.s

Lp= wetted perimeter, (m)

Pt= total pressure of gas phase (atm)

Q= volumetric flow rate, (m3 s-1)

Ri= the rate of production of species I by chemical reaction

S= Channel side, (m)

Si= the rate of creation by addition

T =liquid temperature (K)

Ueff= effective liquid phase velocity (m s-1)

uy= Y direction’s liquid phase velocity (m s-1)

X= bed length, (m)

X CO2, I =molar concentration of CO2 at interface (kmol m-3)

yCO2=volume fraction of CO2 in gas phase

Z= coordinate along film thickness (m)

Greek Symbols

α = volume fraction
 = corrugation angel of packing

τ=exposure time (s)

μ= the volume-fraction-average viscosity, respectively (kg m-1 s-1)

ρL= the volume-fraction-average density (kg m-3)

δ=film thick (m)

 =surface tension, N/m

 =angle of top vertex (o)

 =the contact angle of the liquid

Subscripts

e=effective

g=gas

i =interface

l =liquid

1 Introduction

Structured packing, thanks to its higher efficiency and capacity, and lower pressure

drop comparing to trays or random packings (Wu et al., 2006; Zhang et al., 2001;

Zhongli et al., 2004), has been very widely used in gas-liquid mass transfer processes

like distillation, absorption and others. An effective model for mass transfer coefficient

is essential for designing properly gas-liquid mass transfer columns in structural

packing to guarantee profit margins for the corresponding processes. Liquid mass-
transfer coefficient in structured packings is crucial in many important applications

including CO2 capturing using reactive absorption by alkaline aqueous solution. Such

a process can be known as mass transfer controlled process. This is because the

chemical reaction happens instantaneously at the interface and the rate of transport of

the product species from the interface to the liquid bulk determines the rate of the

process.

Numerous models of mass-transfer coefficient as well as the related mass-transfer

area for structured packing can be found in the literature (Brunazzi et al., 1995; De Brito

et al., 1994; Murrieta et al., 2004; Nawrocki et al., 1991; Olujić, 1999; Rocha et al.,

1993, 1996; Xu et al., 2000), and a summary can be found in Wang et al. (2005). Most

of the liquid mass transfer coefficient models were based on the penetration theory, and

some of them were corrected by empirical parameters. It can be found, however, that

most of the models, empirical or semi-empirical correlated, were developed from and

checked against different sources of pilot plant date. As a result, the models are

generally case dependent, and the development of a reliable model for the underline

chemical absorption process becomes an expensive and nontrivial job.

CFD (Computational Fluid Dynamics) based numerical simulation of mass transfer

process in structured packing has drawn increasingly attention in the last one and half

decades (Haroun et al., 2014; Haroun et al., 2012; Khosravi Nikou and Ehsani, 2008;

Khosravi Nikou et al., 2008; Sebastia-Saez et al., 2014, 2015). The main advantage of

numerical simulation is that the mass transfer process and its detailed phenomena in

structured packings can be investigated in a fast, straightforward and economical way.

However, the experiment and the traditional theoretical calculation cannot achieve this

(Sebastia-Saez et al., 2014; Spiegel et al., 1996; Szulczewska et al., 2003).

A study on hydrodynamic and mass transfer performance of a structured packing by


CFD-based simulation was effectively carried out by Lautenschleger et al. (2015),

where they investigated the energy dissipation mechanism and optimized the geometry

of the packing to reduce its resistance but maintain the separation efficiency. Attempts

have also been made in the prediction of concentration distribution in packed columns

by incorporating CFD with a turbulent diffusion model (Liu et al., 2006a; Liu et al.,

2006b), but the emphasis of these works was placed on the modeling of the eddy

diffusivity of species in a single or pseudo-single phase, and a coefficient for the mass

transfer across the interface needed in a source term in the governing equations must be

provided by the user. Haroun et al. (2010a) performed a one-dimensional simulation to

analyze the gas-liquid mass transfer behaviors in a reactive absorption. The same

authors (Haroun et al., 2010b) investigated the transport process in structured packing

by a two-dimensional simulation of a cycle of the corrugation of the packing. Although

the simulation were limited in one or two dimensions to avoid the computational burden,

some postulations, i.e. the Higbie (1935) theory, for mass transfer in the packing were

effectively validated. Petre et al. (2003) first proposed a mesoscale approach employing

a conceptual representative elementary units (REU) for CFD simulation of structured

packing. This was based on the fact that the uniform structure of the structured packing

can be known as the repetitions of representative local structure. REU was used by Petre

et al. (2003) to build a model, which consists of a network of these REUs to simulate a

whole packing. Sun et al. (2013) also simulated hydrodynamics of structure packing

using REU. Chen et al. (2009) made use of the REU in the 3D two-phase flow for

simulating the hydrodynamics and mass-transfer behavior of commercial structured

packings. Haroun et al. (2014) used REU in their investigations on the interfacial

effective area and liquid hold-up in REU of Mellapak 250X.

However, although the CFD-based simulation has been extensively used to


investigate mass transfer processes in structured packings as discussed above, little

attempt has been made in estimating mass transfer coefficient for general purpose by

numerical simulation. In the present paper, we demonstrate how the liquid mass transfer

coefficient in the structured packing for a reactive absorption process of CO2 into NaOH

aqueous solution can be estimated using numerical simulation.

One of the main difficulties for mass transfer coefficient identification by CFD

simulation is in accurate estimation of the effective mass-transfer area. In previous

works, the packing surface was sometimes used as the interfacial area in CFD

simulation. This may be an adequate assumption for the solid-fluid interface, like those

used in Atmakidis and Kenig (2012) for reaction processes, where the packing surface

functions is catalyzer. For gas-liquid systems, however, some studies showed that the

interfacial area is smaller than the packing surface except for conditions of high liquid

loading (Brunazzi et al., 1995; Olujić, 1999), while others showed that the interfacial

area can be significantly larger than the packing surface (De Brito et al., 1994). Tsai et

al. (2011) presented a model for predicting interface area which was measured in a

0.427m structured packing column via absorption of CO2 from air into NaOH. VOF

method has been used in interface research (Augier et al., 2010; Haroun et al., 2010a,

b; Petre et al., 2003). Olsson and Kreiss (2005) consider the volume fraction takes a

value of 0.5 to define the interface. Wang et al. (2005) showed in their review paper

numerous available model for structured packing area, but, as they noted, the problem

has not been satisfactorily solved.

To cope with the difficulties, we propose in the present work to trace the gas-liquid

interface by numerical simulation using a 3-dimensional VOF model provided by

ANSYS Fluent in a REU of the structured packing. Based on the simulation, we propose

a procedure to locate the gas-liquid interface that corresponds to the effective mass-
transfer area. Using the mass-transfer area, the material balance around the liquid phase

and the concentration distribution, the mass transfer coefficient is finally evaluated. We

also test the viability of a turbulent model for the estimation of the mass transfer

coefficient, which needs much less computation effort than the DNS. Finally, we

validate the results of our simulation with existing models.

2 Reactions in CO2 absorption by aqueous NaOH solution

When CO2 is absorbed and reacts with aqueous NaOH solutions, the following three

overall reactions take place (Liu et al., 2006a):

C O2 g, H
A
C O2 , L (1)

CO2,L  OH 


H R , k2
 HCO3 (2)

HCO3  OH 
 CO32  H2O (3)

Reaction 1 denotes the physical absorption of CO2 by water, accompanied by the heat

of solution HA . Reaction 2 can be known as the rate controlling step because reaction

3, a proton-transfer reaction, is about 1000 times faster than reaction 2 (Onda et al.,

1968). Therefore, the overall reaction can be represented by equation 4.

CO2,L  2OH 
 CO32  H2O (4)

3 Proposed numerical procedure

3.1 Definition of the model geometry.

The typical geometry of structured packing is show in Figure 1a and b. To mitigate

computational burden, a REU of the structured packing is adopted as the geometric

model in our simulation. A REU is the most elementary portion that possess all the

geometrical feature of the structured packing, assuming that the hydrodynamic

behaviors of the structure packing can be captured by the ensemble of REU arranged
regularly. The REU used in the present study is shown in Figure 1c, which is used as

the control volume domain for the computation.

(a) (b)

(c)
Figure 1 (a) Geometry of Mellapak 500Y structured packed, (b). Packing channel
dimensions (Tsai et al., 2011) and (c). REU of the Mellapak 500Y

The REU element is composed of two adjacent sheets, both have two repetitions of

corrugations in both the horizontal and vertical directions. The geometrical parameters

we used in this paper for the REU are compatible with the commercial Mellapak 500Y

packing. The corresponding geometric dimensions are detailed in Table 1. On the open

faces of the REU boundary, the periodic boundary conditions are used for z and x

directions in our computation.

It should be noted that the fine surface texture on the metal sheet of the structured

packing may affect the fluid dynamics. However, with the mesh resolution we used in

our computation, the texture can hardly be considered. The surface texture of structured

packing has been usually neglected in the numerical simulation studies in the literature.

In the present work, we also neglected the surface texture for the reason of

simplification.
3.2 Governing equations.

3.2.1 Hydraulics model.

Gas-liquid flows in the REU are simulated with 3D volume of fluid (VOF) method,

implemented in ANSYS Fluent ™ 14 software, in which finite volume method is used

to solve the Navier-Stokes equations. The VOF method (Hirt and Nichols, 1981) is a

numerical approach for tracking free surfaces of the liquid based on the Eulerian

method and includes a transport equation of the volume fraction of the two phases

involved. The fluids in both phases are assumed to be Newtonian and incompressible

with no phase change and no heat transfer.

The continuity equation is

  ui
 0 (5)
t xi

The momentum equation is

   ui     ui u j  P  2ui
    g  F (6)
t x j xi x j xi


where F accounts for the forces other than the pressure, viscosity and gravity acting on

the fluid, ui and u j are velocity in different direction, t is time, P is transient pressure.

Accordingly,  and  are the volume-fraction-average density and viscosity

respectively given as

   L L  G G (7)

   L L  G G (8)

where  L , G ,  L and G are respectively the densities and viscosities of liquid and gas

phases.

In each control volume, the summation of all the phase’s volume fractions is equal
to unity. That is,

 L  G  1 (9)

where  L and  G represent the volume fractions of liquid and gas respectively, and a

value of 1 or 0 for either of them corresponds to a cell filled with the corresponding

phase or the other. If the value of the phase volume fraction is between 0 and 1 in a cell,

then the interface between the phases may be located in the cell. The solution for

volume fraction distribution can be obtained by solving its conservation equation:


i   ui i  0 (10)
t

We use the continuum surface force (CSF) model developed by Brackbill et al. ( 1992)


to account for the effect of surface tension F in our simulation

 k 
F  (11)
0.5   L  G 

where  is the surface tension coefficient, and k is the free surface curvature defined as

1  n  
k  n     n    n  (12)
n  n  

In equation (12), n is the unit normal vector, and n   

The unit normal vector to any cell next to the wall is represented by the following

equations (Brackbill et al., 1992):

n  nw cos   t w sin  (13)

where  is the contact angle of the liquid on the solid wall, n w and t w are the unit

vectors respectively normal and tangential to the wall.

The influence of the drag force can be described by the friction pressure drop model

given by Sun et al. (2013):


FLG  ae fi G  uG  uL  uG  uL (14)

where ae is the effective interfacial area per unit volume and can be calculated by

ae  aG  aL (Xu et al., 2009). uG and uL are gas and liquid velocity, f i is the

interfacial friction factor. The interfacial friction in countercurrent flow can be

descripted by correlation of Stephan and Mayinger (1992)

fi  0.079 ReG0.25 1  115  N  (15)

ReG  u G G  D  2  G (16)

N  3.95 1.8  3 / D  (17)

where D and   are, respectively, the dimensionless ratios of tube diameter and film

thickness to the Laplace length defined as:

LLa     L  G  g 
0.5
(18)

The tube diameter D is replaced by the hydrodynamic diameter d of the packing unit,

and its value can be considered as the channel side of the packing unit (S).  is

calculated by Nusselt (1916) film thickness given by

  3Q   L g sin  


0.333
(19)

where Q is the volume flow rate per unit of perimeter length and  is the corrugation

angle of packing.

In this paper, the liquid is 0.1kmol/m3 NaOH aqueous solution with the surface

tension of 30 mN/m and the contact angle is 25 deg (Rocha et al., 1993, 1996). CFD

simulation using the equations presented with sufficient number of cells can be known

as DNS. The number of cells used in the present work is approximately 7.5×105, which

is more than the number of cells suggested by Zhang et al. (2005) for DNS. The

governing equations were discretized by the finite volume approach, using second order
Finite Difference Method (FDM) for discretization of the convection terms (Albert et

al., 2014; Mousazadeh et al., 2013) .

In the present paper simulations are also carried out by a turbulence model, using the

RNG (re-normalization group) k-ε model (Khosravi Nikou et al., 2008; Zhang et al.,

2013), which is commonly used in engineering problem. In this model, the Reynolds

averaged Navier-Stokes equations for turbulent flow, by adopting the Boussinesq

postulate, are closed by the k and ε equations, where k and ε represent, respectively, the

average of turbulent fluctuating kinetic energy of the fluid and its dissipation rate. Note

that in the turbulent simulation, the obtained velocity is time-averaged velocity, instead

of the instantaneous velocity presenting in equations (5), (6), (10) and (14) etc. The

number of cells for the turbulence model is approximately 2.1×105.

3.2.2 Species transport Equations

When we choose to solve conservation equations for chemical species i through the

solution of a convection-diffusion equation, this conservation equation can take the

following general form:


t
 
 Ci      uCi    J i (20)

Where J i is the diffusion flux of species i, which is due to concentration gradients and

can be written as

J i     Dm  Dt  Ci (21)

where Dm is the diffusivity of species i in the mixture, and Dt is the turbulent diffusion

coefficient being valid only for turbulent simulation.

For the turbulent flow simulation, the form given by equation (21) implies that

Boussinesq postulate is also adopted to model the covariance of the fluctuating

quantities of the velocity and concentration. By the Boussinesq postulate, one assumes
that such a covariance is proportional to the gradient of the time-average concentration,

and the proportional coefficient is defined as the turbulent diffusivity Dt. Based on this

postulate, Sun et al. (2011; 2007) implemented a two-equation model, called c2   c

model, for the simulation of mass transfer processes in turbulent flow. According to this

model, the turbulent diffusivity is given by


1

 kc2  2
Dt  Cc 0 k   (22)
  c 
 

Where c is the fluctuating concentration and  c is its dissipation rate, and modeled

respectively by the following two equations.

    c 2 
2
c2 U i c2 D  C 
   Dm  t
    2 Dt    2 c (23)
t xi xi   c 2  xi   xi 

    c 
2
 c U i c D  c  C  
   Dm  t


 xi 
  Cc1Dt 2    Cc 4  c (24)
t xi xi    c   c  xi  k

The constants (Sun et al., 2011; Sun et al., 2007) in Eqs. (22), (23), (24):

Cc 0  0.11 , Cc1  1.8 , Cc 2  2.2 , Cc3  0.8 ,  c  1.0 and   c  1.0 .


,2

3.2.3 Interfacial Mass-transfer

Even though the chemical absorption of the CO2-NaOH system is a liquid-phase-

controlled system, the gas-phase overall phase mass transfer coefficient is normally

used in the traditional design method. This is due to the fact that the liquid mass transfer

coefficient depends also on the chemical reaction mechanism, and it is difficult to be

extrapolated to different concentration ranges or operating conditions since changes in

the reaction mechanisms can cause the liquid mass transfer coefficient vary

unexpectedly. However, the overall mass transfer coefficient of gas cannot capture such

dependence, and so it is usually used more for comparing different packings than for
designing (Eckert et al. 1967). In the present work, we proposed to isolate the liquid

mass transfer coefficient from the impact of the chemical reaction and estimate its value

by numerical simulation.

It is often convenient to consider the effect of a chemical reaction in terms of the

enhancement factor E, which is defined as the ratio of the amount of gas absorbed in a

given time into a reacting liquid, to the amount that would be absorbed if there were no

reaction (Danckwerts and Lannus, 1970). In the case of chemical absorption in the

present work, we assume that CO2 reacts quickly with OH1- to form CO32- after entering

the liquid phase, and the film coefficient for mass transfer in the liquid is to be enhanced

with a factor E. Or equivalently, according to Santiago and Farina (1970), the

equilibrium concentration of dissolved CO2 at the interface is enriched with the factor

E as illustrated by Figure 2(a). In order to estimate the liquid mass transfer coefficient

from the simulation, where the concentration distribution of CO32- in the liquid is to be

computed, and the mass transfer coefficient of CO32- will be estimated in the present

work. The stoichiometric coefficients of equations (4) indicates that the upstream

concentration of the driving force of the mass transfer of CO32- can be given by EXCO2,I,

as illustrated in Figure 2(b). As indicated in Figure 2, the liquid mass transfer coefficient

for CO32- physical absorption is defined on the basis of the film thickness *.
(a) (b)
Figure 2 Concentration distributions in the liquid film for interfacial mass transfer,
lR: the thickness of the reaction layer; : mass transfer resistant film for physical
absorption; *: mass transfer resistant film for chemical absorption. (a) The
absorption is enhanced as if the interfacial equilibrium concentration of CO2 in the
2-
liquid is increased to EXCO2,I, as suggested by Santiago and Farina (1970); (b) CO3
profile in the liquid film with chemical reaction enhancement

For an irreversible second-order reaction such as the CO2-NaOH reaction, Wellek et

al. (1978) reported a correlation for the calculation of enhancement factor:

1

 DX 1 1.35  
1.35
 1.35

  
DCO2,L k2 X OH 1 k L 
E  1     1
 
OH

 
 2 DCO , L X CO ,i  tanh D k X  (25)

  CO2,L 2 OH 1 k L 
2 2
  

where XOH- denotes the molar concentration of OH- in liquid phase, DCO2,L represents

the diffusivity of CO2 in the NaOH aqueous solution, and k2 is the second-order reaction

rate constant for the CO2-NaOH reaction. The diffusivity of CO2 in the NaOH aqueous

solution, DCO2,L can be determined by modifying the diffusivity of CO2 in the pure water,

DCO2,W (Hikita et al., 1976)


DCO 2, L  DCO 2,W 1   X OH 1  (26)

where  is a constant depending on the aqueous solution. For the NaOH aqueous

solution,  = 0.129 m3 Kmol-1. The diffusivity of CO2 in the pure water, DCO2,W is a

function of liquid temperature correlated (Pohorecki and Moniuk, 1988)

712.5 2.591105
lg DCO2,w  8.1764   (27)
T T2

The diffusivity ratio of D/DCO2,L is 1.67 at 20℃ (Danckwerts and Lannus, 1970) and

is assumed to be independent of the temperature in the range of 20-40℃.

For CO2-NaOH reaction, the second-order reaction rate constant, k2, was correlated
by Pohorecki and Moniuk (1988) as a function of temperature and ionic strengths, Ic,

in aqueous electrolyte

2382
log k2  11.895   0.221IC  0.016 I c2 (28)
T

According to Pohorecki and Moniuk (1988), the molar concentration of CO2 at

interface XCO2,I can be expressed by Henry’s law:

X CO2 ,I  HPt yCO2 (29)

where H is the Henry’s constant of CO2 in the NaOH solution, Pt is the total pressure of

gas phase, and yco2 is the volume fraction of CO2 in the gas phase.

The Henry’s constant for CO2 in aqueous NaOH solution can be calculated from the

semi-empirical mode (Danckwerts and Lannus, 1970):

0.9869 H
log   Ks I c (30)
 H CO2  w

 
2
log H CO2  9.1229  0.059044T  7.8857 105 T (31)
w

where (Hco2)w is the Henry’s constant for CO2 in pure water and Ks is the salting

parameter, denoting the sum of the contributions of the ions and gas molecules in the

liquid. For the CO2-NaOH system, Ks=0.142. Ic is ionic strengths of aqueous electrolyte

solutions.

3.3 Assumptions.

Five assumptions are adopted in the present paper for the simulation of the chemical

absorption of CO2 by NaOH aqueous solution in structured packing:

1. The absorption process is in steady state.

2. The fluids are incompressible.

3. Only the CO2 component in the gas phase is absorbed by NaOH aqueous solution,

and the NaOH aqueous solution does not transfer to gas phase.
4. Thermal effects of the absorption and of the reaction have no influence on the

fluid properties.

5. The concentrations of CO2 and OH1- in liquid are neglected.

By assumption 1, we imply that the mass transfer coefficient we estimate by the

numerical simulation is for steady state operation. By using the dynamic governing

equations presented in the previous sections, the mass transfer coefficient will be

evaluated when the results of the simulation came to stable. Assumptions 2 and 3, which

have been usually adopted by previous works, are for simplification purpose. Even

though thermal effect of the absorption is inevitable especially in the dissolution of the

gas in the liquid, we adopt the assumption 4 since the effect is weak. Assumption 5 is

based on the fact that CO2 can react completely with excess NaOH in the liquid.

3.4 Mass-transfer area.

A reliable effective mass-transfer area of the packing is an essential parameter for

estimate mass transfer coefficient. Under most conditions, the CO2- alkali reaction is

fast enough such that the CO2 reacts within a short distance of the gas-liquid interface

in the liquid. In the computing, the free surfaces of the liquid, or the interface between

gas and liquid, tracked by the VOF method is located in the cells where the volume

fraction of liquid falls into the range of (0-1). As a result, the interface given by VOF

computation is not just a surface, but a layer with a thickness of several cells, within

which the volume fraction of liquid phase is higher than zero and smaller than 1, as

illustrated in Figure 3. As volume fraction of liquid within the layer varies in the range

of (0-1), we need to find a particular volume fraction value, say qi, within that range to

locate the interface. In the present work, we determine such a value by comparing the

simulation with the experimentally correlated model.


Figure 3. Illustration of the definitions of the interfacial layer produced by VOF
computing and the interface location to be found

Many correlations for the effective mass-transfer area for structured packings can

be found in the literature (Wang et al., 2005). In the present work, a model that correlates

the ratio of effective area to specific surface area (ae/ap) as a function of the liquid Weber

and Froude numbers, most recently proposed by Tsai et al. (2011) is employed.

0.116
  Q 
43
aeff  
 1.34   g  
 L  
L 13
(32)
ap      
 p

where Q is volumetric flow rate,  is surface tension, with the wetted perimeter Lp

calculated by

4S
Lp= A (33)
Bh

where A is cross-sectional area of column, B is packing channel base, h is packing crimp

height and S is packing channel side.

Note that the structure and operating parameters for the simulation in the present

work are all within the applicable range of the Tsai’s model, which is defined by: the

specific area: 125-500m2/m3, liquid load: 2.5-75m3/m2.h, the gas velocity: 0.6-2.3m/s),

surface tension: 30-72 mN/m , and liquid viscosity: 1-15 mPa•s.


1.0 1.0
2.1X105
5
4.6X105
0.9 7.5X10 0.9
1.3X105
12.7X105

Fractional area,ae/ap
0.8
Fractional area,ae/ap

0.8

0.7
0.7
0.6
0.6
0.5
0.5
0.4
0.4
0.1 0.2 0.3 0.4 0.5 0.1 0.2 0.3 0.4 0.5

volume fraction of liquid volume fraction of liquid


(a) (b)

Figure 4 Comparison of cells number profile of the ae/ap from simulation in


different volume fraction (a)DNS model (B) turbulent model

We tested the grid dependency of our simulations and the results are given in Figure

4 by the fractional area with different liquid fraction value at different mesh resolutions.

It can be seen that the difference of calculated fractional area with the two mesh

resolutions is rather small in DNS model. In the turbulence model, the differences of

calculated fractional area with 2.1×105 and 4.6×105 meshes are rather small. In the

following computations, the cell number of 7.5×105 in DNS and 2.1×105 in turbulence

simulation were used.

1.0

0.9
Fractional area,ae/ap

0.8

0.7
Tsai's Model
0.6 DNS model
CFD turbulence model

0.5

0.4
0.1 0.2 0.3 0.4 0.5

volume fraction of liquid

Figure 5 Comparison of volume fraction of liquid profile of the ae/ap from


simulation and Tsai’s model for liquid load of 30 (m3/m2•h)

Figure 5 shows the dependence of the fractional area ae/ap on the volume fraction of

liquid in the interfacial layer by both the DNS and turbulent simulations. It can be seen

that, for a lower value of the volume fraction of liquid in the interfacial layer, the

interface tends to move rightward in Figure 3, and the interfacial area tends to be larger

as shown in Figure 5 because of the general positive curvature of the surface of liquid

on the packing sheet surface. In Figure 5, the interfacial area value predicted by Tsai’s

model is also plotted. By comparing the simulated to the model predicted results we

can find that the qi value should be taken as 0.2, because the simulated ae/ap value at

this volume of the liquid fraction is the most closed to the Tsai’s model prediction.

0.95
Tsai's Model
CFD laminar model
0.90 CFD turbulence model
Fractional area,ae/ap

0.85

0.80

0.75
10 15 20 25 30 35 40 45 50 55 60 65 70

Liquid load (m /m h)


3 2

Figure 6 Comparison of liquid load profile of the ae/ap from simulation and Tsai’s
model for volume fraction of liquid of 0.2
1.00

0.95 Tsai's Model

Fractional area,ae/ap
CFD laminar model
0.90
CFD turbulence model

0.85

0.80

0.75

0.70
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

F,m/s(kg/m3)0.5

Figure 7 Comparison of F factor profile of the ae/ap from simulation and Tsai’s
model for volume fraction of liquid of 0.2 and liquid load of 30 (m3/m2•h)

Further comparison between our numerical results and the Tsai’s model in different

liquid load and F factor are shown respectively in Figures 6 and 7. As seen from these

Figures, the simulation and Tsai’s model prediction on the ae/ap in liquid phase is in

satisfactory agreement if the qi takes the value of 0.2.

As seen from Figure 6, the ae/ap values calculated by either simulation or Tsai’s model

increased with increase of liquid load. This can be explained by the increases of both

the wetting areas on the packing surface and the turbulent motion of gas-liquid interface

with the increase of liquid load.

Various gas velocities were run in our simulation. Although the ae/ap calculated by

simulation fluctuated with increase of F factor, the dependence of the interfacial area

on the gas flow is shown to be very weak as shown in Figure 7. Such a result is

consistent with the Tsai’s model and many other structured packing models in the

literature as well (Wang et al., 2005) .


4 Results and discussions

4.1 Hydrodynamics.

A liquid distribution in the REU obtained by DNS is shown in Figure 8(a). As

discussed by McGlamery (1988), Nicolaiewsky and Fair (1999) for the mechanism of

the rivulet flow formation on the sheet surface of a structured packing, the effective

area tends to be governed by the spreading ability of the liquid within the individual

channels. Considering the surface tension effect, the liquid load tends to be more

determined by the distribution of liquid to the channels than the degree of coverage

within a channel.

(a)

(b) (c)

Figure 8 Contours of liquid film and concentration in the surface of REU of


Mellapak 500.Y structured packing, liquid load of 30 (m3/m2•h), F factor of 0.8
m/s(kg/m3)0.5 (a) geometry of a packing channel, (b) contour of liquid volume
fraction, (c) contour of CO32- concentration in the liquid on the sheet

In Figure 8(a), the simulated concentration field of CO32- in the liquid is depicted. In
all our simulations, low concentrations are observed near the top of the REU and high

concentrations near the bottom. Due to turbulence effects and the exposure time lasting,

the concentrations of CO32- in NaOH aqueous was increased in the liquid flow direction

through the REU. It is interesting to note that, at the spots where corrugates’ ridges of

the two adjacent sheets are brought to contact, the concentration is higher. An

explanation can be that at these conjunction points, the liquid films flowing down from

the two sheets are remixed, thus the liquid side mass transfer can be enhanced, and more

solute tends to be solved.

As shown in Figures 5, 6 and 7, the turbulent simulation can give satisfactory

approximation to the DNS for the gas-liquid two-phase flow in structured packing in

an absorption process.

(a)

(b) (c)

Figure 9 Simulated gas velocity in the section in the REU channel of the packing
for F factor of 0.4 m/s(kg/m3)0.5and liquid load of 30 (m3/m2•h), (a) geometry of a
packing,(b) turbulence model, (c) DNS model

The velocity contours in the section along the Y axis computed by DNS model and

turbulence model with the same liquid load and F factor are shown respectively in
Figures 9(b) and (c). As seen from Figure 9, although the DNS can simulated gas

velocity distribution in the REU more accurate than turbulence model, the velocity

distribution patterns by both methods are quite similar.

1.8x10-4 0.07
CFD turbulence model
CFD turbulence model
DNS model
DNS model
0.06
average film thickness m

-4
1.6x10
0.05

Hold up
0.04
1.4x10-4

0.03

1.2x10-4 0.02
10 20 30 40 50 60 70 10 20 30 40 50 60 70

Liquid load m /m h
3 2
Liquid load m3/m2h

(a) (b)
1.6x10-4 4.8x10-2
CFD turbulence model 4.6x10-2
DNS model CFD turbulence model
1.5x10-4 4.4x10-2 DNS model
average film thickness m

4.2x10-2
Hold up

1.4x10-4 4.0x10-2

3.8x10-2

1.3x10-4 3.6x10-2

3.4x10-2

1.2x10-4 3.2x10-2
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
3 0.5
F factor ( m/s(kg/m ) ) F fator ( m/s(kg/m3)0.5)

(c) (d)

Figure 10 Hydrodynamic simulated of the packing (a) liquid film thickness under
F factor 0.8 m/s(kg/m3)0.5, (b) liquid holdup under F factor 0.8 m/s(kg/m3)0.5, (c)
liquid film thickness under liquid load of 30 m3/m2•h, (d) liquid holdup under
liquid load of 30 m3/m2•h

As a result, the liquid film thickness, liquid holdup, wetted area fraction and

pressure drop variations with respect to liquid load estimated from the turbulent

simulation agree quite well with those by the DNS as shown in Figure 10.

4.2 Mass transfer coefficients.


By the simulation, the concentration distribution of CO32- in the liquid can be

obtained. The contours of phase volume fraction and the CO32- concentration

distributions in the liquid on a cross section of a packing channel by DNS and

turbulence model are shown in Figure 11. In Figures 11(b) and (c), the gas-liquid

interfaces are shown in black dotted lines. Figures 11(b) and (c), show the contours of

CO32- concentration distributions in the liquid for DNS and turbulent simulation

respectively.

(a)

(b) (c)

(d) (e)

Figure 11 Comparison of contours of CO32- concentration distributions in the liquid


and liquid distribution on a horizontal cross section of a packing channel, for F factor
of 0.8 m/s(kg/m3)0.5 and liquid load of 30 m3/m2•h (a) geometry of a packing channel;
(b) contours of liquid volume fraction by DNS, the identified interface is marked by
the dotted line, (c) contours of liquid volume fraction by turbulence simulation, , the
identified interface is marked by the dotted line; (d) contours of concentration by DNS,
(e) contours of concentration by turbulence simulation

The contour of turbulent diffusivity coefficient distributions in the liquid on a cross

section of a packing channel by turbulence model is shown in Figure 12. In Figure 12(c),

it can be seen that the value of Dt is higher (approaching to the value in the bulk flow)

near edge of the cross section of the packing channel, where the mixing of the liquid

between two sheets takes place.

(a)

(b) (c)

Figure 12 the contours of turbulent diffusivity coefficient in the liquid and liquid
distribution on a horizontal cross section of a packing channel, for F factor of 0.8
m/s(kg/m3)0.5 and liquid load of 30 m3/m2•h (a) geometry of a packing channel, (b)
contours of liquid distribution by turbulence simulation, (c) contours of turbulent
diffusivity coefficient by turbulence simulation

By using the simulated spatial distribution of the concentration, the average

concentrations of CO32- in the liquid at the inlet and outlet of the REU, the average

logarithmic concentration difference  c ln , and then the mass transfer coefficient can

be estimated using the mass balance of species CO32- around the liquid phase in the

REU. As a result, the liquid mass transfer coefficient can be given by (Atmakidis and
Kenig 2012) :

L(cout  cin )
kL  (34)
aeff X (c)ln

where

cin  cout
(c)ln  (35)
 c 
ln  in 
 cout 

cin  X co2  cin (36)


3

cout  X co2  cout (37)


3

In equation 34, L is mass flux of the interfacial mass transfer, X is bed length of the

REU. X co2 is molar concentration at interface. To validate the simulated liquid-side


3

mass transfer coefficient given by equation (34), comparisons can be made with three

kinds of models: a theoretical model, a semi-theoretical model, and an empirical model.

The theoretical and semi-theoretical models are based on the penetration theory, with

the exposure time related to the residence time of flowing liquid across a single

corrugation. They are given as the following:

(1) Penetration theory model using by Sebastia-Saez et al. (2015). The mass transfer

coefficient by Higbie (1935) theory is given as

Dl
kL  2 (38)


where Dl is diffusion coefficient of CO32- in liquid. According to the literature there are

several ways to calculate the exposure timeτ. Sebastia-Saez et al. (2015) used the ratio
between the distance traveled by the fluid element from the inlet and the velocity at the

interface:

X
 (39)
U eff

where X is bed length with structured packed, the velocity at the interface can be

calculated as:

 L g 2 sin     z 2 
U eff  1     (40)
2 L     

where  is the Nusselt film thickness given by equation (19).

(2) Semi-theoretical model used by Haroun et al. (2012)

This model has the same form of equation (38), but the exposure time is estimated

differently and given by

X
 K (41)
 L g sin  2   2 2 L

where K is shown to depend on the film regime: K≈0.85 for uniform film and K≈0.7

for film with recirculation,  is the angle of top vertex of the relevant geometrical

characteristics of the structured packing, X is bed length with structured packed.

(3) Henriques de Brito’s model (1992)

The empirical correlation of Henriques de Brito is given by

kL500Y  0.000233U LS
0.302
(42)
12
DNS model
11 CFD turbulencemodel
Haroun model
10
Henriques de Brito model
9 Sebastia-saez model
kL105(m/s)
8

4
15 20 25 30 35 40 45 50 55 60 65
Liquid load (m3/m2h)

Figure 13 Comparison of liquid load profile of the kL from simulation and


Empirical model for F factor of 0.8 m/s(kg/m3)0.5
The comparison between our numerical results and the various model predictions

on mass transfer coefficient for different liquid and gas loads are shown in Figures 13

and 14. As can be seen from Figures 13 and 14, despite of the differences among the

numerical results and those predicted by the models, they are fairly close. The empirical

correlation overestimated the liquid mass transfer coefficient than the theoretical and

semi-theoretical models in our case, and the numerical results are in between from the

middle to higher load for both liquid and gas. As seen from Figure 13, the simulated

liquid mass transfer coefficient is the most consistent with the penetration theory

predictions.
12
DNS model
CFD turbulence model
10 Haroun model
Henriques de Brito model
Sebastia-saez model
kL105(m/s) 8

0.2 0.4 0.6 0.8 1.0 1.2


F,m/s(kg/m3)0.5

Figure 14 Comparison of F factor profile of the kL from simulation and Empirical


model for liquid load of 30 (m3/m2•h)

Figure 14 shows that the kL predicted by the three models are independent of the

gas velocity. In the theoretical models, as shown in equation (40) and (41), the effect of

the gas flow is not considered, as this is a usual assumption that gas is much less dense

than the liquid and its effect can be neglected. The empirical correlation, on the other

hand, was obtained from experiments with gas flow, but it is usually costly or difficult

to accurately correlate experimentally the liquid side mass transfer coefficient with the

gas velocity. The simulation results by DNS shows that the gas velocity can have effect

on the liquid mass transfer coefficient but the effect is weak, because the absorption of

carbon dioxide by NaOH aqueous process is a liquid film controlled mass transfer

process, and the impact of the gas velocity should be small.

It is interesting to note from Figures 13 and 14 that the turbulent simulation gave

very close results to the DNS for the liquid side mass transfer coefficient. This is
interesting because the turbulent model needs significantly less dense cells, and thus

lower computational effort than those needed by the DNS.

It must be pointed out that the hypotheses for the DNS and the turbulent simulation

for the mass transfer process in the REU are quite different. In the DNS, with the

sufficient number of meshes, the mass transfer behaviors in both the bulk liquid and

near interface can be captured. Whereas in the turbulent simulation, by definition, the

diffusion of species is determined by the turbulent diffusivity Dt, which is much higher

than the molecular diffusivity D used in the DNS. The similarities between the results

of the two approaches, as shown in Figures 13 and 14, can then be explained by the

adeptness of the turbulent model we used. As the two-equation model (the RNG k-

accompanied by the c2   c model) relates Dt with the kinetic energy and its

dissipation of the turbulent flow, and thus, Dt value gets lower where the turbulent effect

gets lower, near the interface for example. As a result, the flow and diffusion behaviors

near the interface as well as in the bulk flow can be satisfactorily captured. This may

also be true even for the cases of low resolution of the meshes in the turbulent

simulation. As seen from Figure 11, the concentration gradient near the interface is

larger than the bulk of liquid, so the mass transfer resistance is concentrated there in the

layer near the interface. A turbulent simulation result of the concentration distributions

of CO32- is shown in Figure 11(e). Comparing with Figure 11(d), the concentration

gradient near the interface by the turbulent simulation appears lower. However, the
diffusion flux of CO32- may keep the same with that of Figure 11(d) because the value

of Dt in Figure 11(e) must be higher (approaching to the value in the bulk flow) near

the interface. Consequently, the turbulent model can be used effectively in the

procedure to estimate the liquid mass-transfer coefficient in the structured packing.

5 Conclusions

We have developed in the previous sections 3D VOF model to simulate the

hydrodynamics and mass transfer process for CO2 absorption into NaOH aqueous

solution in structured packing. A method to locate the interfacial area has been proposed

so that the mass-transfer area can be identified. Using the simulated results on the

concentration distribution and the mass-transfer area, we have established a mass

balance for the interested species around the liquid phase and solved it for the liquid

mass-transfer coefficient. The results were validated by the existing models. The

computational burden has been limited by adopting a REU for the simulation, and we

showed that a RNG k-ε turbulent model can be effectively applied. The work presented

in the above sections can be known as a CFD-based numerical procedure to determine

the liquid mass-transfer coefficient for the structured packings. Comparing to the

various empirical or semi-empirical models, the proposed procedure can provide a fast,

straightforward and cheap alternative approach to determining the liquid mass-transfer

coefficient for structured packings.

It must be pointed out that a weakness of the present work is the dependence of
experiments in locating the gas-liquid interface. This issue can be solved by increase

the mesh resolution so that the error in the mass-transfer coefficient generated by the

numerical location of the interface is acceptably small. Thus, the mesh density for

computing should depend not only on the traditional grid independence check, but also

on the error of the estimated mass-transfer coefficient. It would be thus interesting in

the future to test the sensibility of the mesh density in terms of the error of the estimated

mass-transfer coefficient, so that the mesh density can be optimized.

Acknowledgment

The authors acknowledge the financial support of the National Natural Science

Foundation of China (Contract 21376163) and the assistance from the staff in the State

Key Laboratories for Chemical Engineering (Tianjin University).

Reference

Albert, C., Marschall, H., Bothe, D., 2014. Direct Numerical Simulation of interfacial
mass transfer into falling films. Int J Heat Mass Transfer 69, 343-357.
Atmakidis, T., Kenig, E.Y., 2012. Numerical analysis of mass transfer in packed-bed
reactors with irregular particle arrangements. Chem Eng Sci 81, 77-83.
Augier, F., Koudil, A., Royon-Lebeaud, A., Muszynski, L., Yanouri, Q., 2010.
Numerical approach to predict wetting and catalyst efficiencies inside trickle bed
reactors. Chem Eng Sci 65, 255-260.
Brackbill, J., Kothe, D.B., Zemach, C., 1992. A continuum method for modeling surface
tension. J Comput Phys 100, 335-354.
Brunazzi, E., Nardini, G., Paglianti, A., Petarca, L., 1995. Interfacial area of mellapak
packing: Absorption of 1, 1, 1‐trichloroethane by Genosorb 300. Chem Eng
Technol 18, 248-255.
Chen, J., Liu, C., Yuan, X., Yu, G., 2009. CFD simulation of flow and mass transfer in
structured packing distillation columns. Chin J Chem Eng 17, 381-388.
Danckwerts, P.V., Lannus, A., 1970. Gas‐liquid reactions.
De Brito, M.H., Von Stockar, U., Bangerter, A.M., Bomio, P., Laso, M., 1994. Effective
mass-transfer area in a pilot plant column equipped with structured packings and
with ceramic rings. Ind Eng Chem Res 33, 647-656.
De Brito, M.H., Von Stockar, U., Bomio, P., 1992. Predicting the liquid phase mass
transfer coefficient k L for the Sulzer structured packing Mellapak, institution of
chemical engineers symposium series. hemsphere publishing corporation, pp.
B137-B137.
de Santiago, M., Farina I.H., 1970. Mass transfer with second order reaction -Numerical
solution. Chem Eng Sci 25(4):744-747
Eckert, J S, Foote E H, Rollison L R, Walter L. F, 1967, Absorption processes utilizing
packed towers. Ind Eng Chem, 59 (2): 41–47
Haroun, Y., Legendre, D., Raynal, L., 2010a. Direct numerical simulation of reactive
absorption in gas–liquid flow on structured packing using interface capturing
method. Chem Eng Sci 65, 351-356.
Haroun, Y., Legendre, D., Raynal, L., 2010b. Volume of fluid method for interfacial
reactive mass transfer: application to stable liquid film. Chem Eng Sci 65, 2896-
2909.
Haroun, Y., Raynal, L., Alix, P., 2014. Prediction of effective area and liquid hold-up in
structured packings by CFD. Chem Eng Res Des.
Haroun, Y., Raynal, L., Legendre, D., 2012. Mass transfer and liquid hold-up
determination in structured packing by CFD. Chem Eng Sci 75, 342-348.
Higbie, R., 1935. The rate of absorption of a pure gas into still liquid during short
periods of exposure.
Hikita, H., Asai, S., Takatsuka, T., 1976. Absorption of carbon dioxide into aqueous
sodium hydroxide and sodium carbonate-bicarbonate solutions. The Chemical
Engineering Journal 11, 131-141.
Hirt, C.W., Nichols, B.D., 1981. Volume of fluid (VOF) method for the dynamics of
free boundaries. J Comput Phys 39, 201-225.
Khosravi Nikou, M., Ehsani, M., 2008. Turbulence models application on CFD
simulation of hydrodynamics, heat and mass transfer in a structured packing. Int
Commun Heat Mass 35, 1211-1219.
Khosravi Nikou, M., Ehsani, M., Davazdah Emami, M., 2008. CFD simulation of
hydrodynamics, heat and mass transfer simultaneously in structured packing. Int J
Chem React Eng 6.
Klöker, M., Kenig, E.Y., Piechota, R., Burghoff, S., Egorov, Y., 2005. CFD‐based
Study on Hydrodynamics and Mass Transfer in Fixed Catalyst Beds. Chem Eng
Technol 28, 31-36.
Lautenschleger, A., Olenberg, A., Kenig, E.Y., 2015. A systematic CFD-based method
to investigate and optimise novel structured packings. Chem Eng Sci 122, 452-
464.
Liu, G., Yu, K., Yuan, X., Liu, C., 2006a. New model for turbulent mass transfer and its
application to the simulations of a pilot-scale randomly packed column for CO2-
NaOH chemical absorption. Ind Eng Chem Res 45, 3220-3229.
Liu, G., Yu, K., Yuan, X., Liu, C., Guo, Q., 2006b. Simulations of chemical absorption
in pilot-scale and industrial-scale packed columns by computational mass transfer.
Chem Eng Sci 61, 6511-6529.
McGlamery, G.G., 1988. Liquid film transport characteristics of textured metal surfaces.
UMI.
Mousazadeh, F., van Den Akker, H., Mudde, R.F., 2013. Direct numerical simulation
of an exothermic gas-phase reaction in a packed bed with random particle
distribution. Chem Eng Sci 100, 259-265.
Murrieta, C.R., Seibert, A.F., Fair, J.R., Rocha-U, J.A., 2004. Liquid-side mass-transfer
resistance of structured packings. Ind Eng Chem Res 43, 7113-7120.
Nawrocki, P., Xu, Z., Chuang, K., 1991. Mass transfer in structured corrugated packing.
Can J Chen Eng 69, 1336-1343.
Nicolaiewsky, E.M., Fair, J.R., 1999. Liquid flow over textured surfaces. 1. Contact
angles. Ind Eng Chem Res 38, 284-291.
Nusselt, W., 1916. Die Oberflachenkondesation des Wasserdamffes the surface
condensation of water. Zetrschr. Ver. Deutch. Ing. 60, 541-546.
Olsson, E., Kreiss, G., 2005. A conservative level set method for two phase flow. J
Comput Phys 210, 225-246.
Olujić, Ž., 1999. Effect of column diameter on pressure drop of a corrugated sheet
structured packing. Chem Eng Res Des 77, 505-510.
ONDA, K., SADA, E., TAKEUCHI, H., 1968. Gas absorption with chemical reaction
in packed columns. J Chem Eng Jpn 1, 62-66.
Petre, C.F., Larachi, F., Iliuta, I., Grandjean, B., 2003. Pressure drop through structured
packings: Breakdown into the contributing mechanisms by CFD modeling. Chem
Eng Sci 58, 163-177.
Pohorecki, R., Moniuk, W.d.w., 1988. Kinetics of reaction between carbon dioxide and
hydroxyl ions in aqueous electrolyte solutions. Chem Eng Sci 43, 1677-1684.
Rocha, J.A., Bravo, J.L., Fair, J.R., 1993. Distillation columns containing structured
packings: a comprehensive model for their performance. 1. Hydraulic models. Ind
Eng Chem Res 32, 641-651.
Rocha, J.A., Bravo, J.L., Fair, J.R., 1996. Distillation columns containing structured
packings: a comprehensive model for their performance. 2. Mass-transfer model.
Ind Eng Chem Res 35, 1660-1667.
Sebastia-Saez, D., Gu, S., Ranganathan, P., Papadikis, K., 2014. Micro-scale CFD study
about the influence of operative parameters on physical mass transfer within
structured packing elements. Int J Green Gas Con 28, 180-188.
Sebastia-Saez, D., Gu, S., Ranganathan, P., Papadikis, K., 2015. Micro-scale CFD
modeling of reactive mass transfer in falling liquid films within structured packing
materials. Int J Green Gas Con 33, 40-50.
Spiegel, L., Bomio, P., Hunkeler, R., 1996. Direct heat and mass transfer in structured
packings. Chem Eng Process 35, 479-485.
Stephan, M., Mayinger, F., 1992. Experimental and analytical study of countercurrent
flow limitation in vertical gas/liquid flows. Chem Eng Technol 15, 51-62.
Sun, B., He, L., Liu, B., Gu, F., Liu, C., 2013. A new multi‐scale model based on CFD
and macroscopic calculation for corrugated structured packing column. AlChE J
59, 3119-3130.
Sun, Z., Liu, C., Yu, G., Yuan, X., 2011. Prediction of Distillation Column Performance
by Computational Mass Transfer Method. Chin J Chem Eng 19, 833-844.
Sun, Z., Yu, K., Yuan, X., Liu, C., 2007. A modified model of computational mass
transfer for distillation column. Chem Eng Sci 62, 1839-1850.
Szulczewska, B., Zbicinski, I., Gorak, A., 2003. Liquid flow on structured packing:
CFD simulation and experimental study. Chem Eng Technol 26, 580-584.
Tsai, R.E., Seibert, A.F., Eldridge, R.B., Rochelle, G.T., 2011. A dimensionless model
for predicting the mass‐transfer area of structured packing. AlChE J 57, 1173-
1184.
Wörner, M., 2012. Numerical modeling of multiphase flows in microfluidics and micro
process engineering: a review of methods and applications. Microfluidics and
nanofluidics 12, 841-886.
Wang, G., Yuan, X., Yu, K., 2005. Review of Mass-Transfer Correlations for Packed
Columns*. Ind Eng Chem Res 44, 8715-8729.
Wellek, R., Brunson, R., Law, F., 1978. Enhancement factors for gas‐absorption with
second‐order irreversible chemical reaction. Can J Chen Eng 56, 181-186.
Wu, X., Tang, Z., Wang, G., Yuan, X., Yu, G., 2006. Performance of high specific
surface area wire gauze structured packing at high pressure. J Chem In Eng 57,
2582.
Xu, Y.Y., Paschke, S., Repke, J.U., Yuan, J.Q., Wozny, G., 2009. Computational
Approach to Characterize the Mass Transfer between the Counter‐Current Gas‐
Liquid Flow. Chem Eng Technol 32, 1227-1235.
Xu, Z.P., Afacan, A., Chuang, K.T., 2000. Predicting Mass Transfer in Packed Columns
Containing Structured Packings. Chem Eng Res Des 78, 91-98.
Zhang, P., Liu, C., Tang, Z., Yu, G., 2001. Experimental Determination of Axial Mixing
in Two-phase Flow Through Structured Packings at Elevated Pressure: Axial
Mixing in Gas Phase. J Chem In Eng 52, 381-382.
Zhang, X., Yao, L., Qiu, L., Zhang, X., 2013. Three-dimensional Computational Fluid
Dynamics Modeling of Two-phase Flow in a Structured Packing Column. Chin J
Chem Eng 21, 959-966.
Zhang, Z., Cui, G., Xu, C., 2005. Theory and modeling of turbulence. Tsinghua
University Press, Beijing 80, 81.
Zhongli, T., Chunjiang, L., Xigang, Y., Guocong, Y., 2004. Progress in research on
structured packing in high pressure services. Chem Ind Eng Prog 23, 353-357.
Table 1 Geometrical parameter of the REU used in simulations

Specific Channel Crimp Channel


Packing area, ap Corrugation base, B height ,h side S
type (m2/m3) angle,  (deg) (mm) (mm) (mm)
Mellapak 500 45 9.6 6.53 8.1
500Y

You might also like