Sie sind auf Seite 1von 10

SEPARATION SCIENCE AND ENGINEERING

Chinese Journal of Chemical Engineering, 18(3) 362371 (2010)


Modeling of Mass Transfer in Nonideal Multicomponent Mixture
with Maxwell-Stefan Approach
*

SONG Yiming ()
1,2
, SONG Jinrong ()
1,2
, GONG Ming ()
1,2
, CAO Bin ()
1,2
,
YANG Yanhong ()
1,2
and MA Xiaoxun ()
1,2,
**

1
School of Chemical Engineering, Northwest University, Xian 710069, China
2
Chemical Engineering Research Center of the Ministry of Education for Advanced Use Technology of Shanbei
Energy, Xian 710069, China
Abstract The Intalox metal tower packing was used to simulate an industrial relevant extractive distillation col-
umn for purifying azeotropic multicomponent mixture. In order to explain the inconsistencies in the modeling of
transfer process in nonideal multicomponent distillation column, a method was developed with equilibrium stage
models (EQ) and non-equilibrium model (NEQ) incorporated with Maxwell-Stefan diffusion equations in the
framework of AspenONE

simulator. Dortmund Modified UNIFAC (UNIFAC-DMD) thermodynamic model was


employed to estimate activity coefficients. In addition, to understand the reason for the diffusion against driving
force and the different results by EQ and NEQ models, explicit investigations were made on diffusion coefficients,
component Murphree efficiency and mass transfer coefficients. The results provide valuable information for basic
design and applications associated with extractive distillation.
Keywords rate based model, equilibrium model, mass transfer coefficient, Murphree efficiency, extractive distil-
lation, simulation
1 INTRODUCTION
There is considerable industrial interest in design
and optimization of extractive distillation due to the
large number of industrial columns in operation and
the potential of developing improved separation
schemes so as to minimize energy consumption [1-3].
As a result, various extractive distillation systems
have been investigated, such as solvent selection
methods [4], development of new extractive distilla-
tion systems [5-7], and introduction of a salt to the
solvent to improve the separation [8]. However, there
are few publications [9-12] on modeling and simula-
tion of azeotropic multicomponent extractive distilla-
tion using non-equilibrium (NEQ) approach. In indus-
trial design, chemical engineers usually develop their
design procedures for separation equipment using
Ficks law of diffusion [13, 14], in which the flux J
i
is
linearly dependent on molar average mixture velocity
and composition gradient
i
x
( )
i i i i i i
J c v v c D x = (1)
This constitutive relation is strictly valid only under
the following conditions: (i) binary mixtures, (ii) dif-
fusion of dilute species in a multicomponent mixture,
and (iii) in the absence of electrostatic or centrifugal
force field. For practical purpose, we usually use
Murphree vapor efficiency (Ei
MV
) for a plate and the
height equivalent to a theoretical plate (HETP) for
packings. The concept works quite well for binary sepa-
ration [14, 15], based on the assumption that the effi-
ciencies of two components are equal and the physical
properties are constant along the column. However, it
is difficult to relate the concept to the construction and
performance of equipment, and to apply it to nonideal
multicomponent mixtures [16, 17]. In addition, some
bizarre behavior may appear, such as unbounded
component Murphree efficiencies [17], and the diffu-
sion in ternary mixture is much more complex than
that in binary mixture [18] because of coupling between
species concentration gradients. To avoid the am-
biguous component efficiencies and the limitation of
Ficks law for describing diffusion in multicomponent
mixture, a realistic model is needed. It is now gener-
ally accepted that the Maxwell-Stefan formulation
provides the most general and convenient approach
for describing transport process in multicomponent
mixture [11, 15, 19]. Lao and Taylor [20] first developed
a model based on the Maxwell-Stefan formulation for
multicomponent distillation for a single tray. In this
study, the NEQ approach incorporated with Maxwell-
Stefan diffusion theory is applied to simulate an in-
dustrial relevant extractive distillation column, and the
mass transfer process and associated influence factors,
such as mass transfer coefficients, Murphree effi-
ciency and diffusion coefficients, are investigated.
2 SIMULATION PROCESURE AND MODELING
2.1 Numerical set up and conditions
Figure 1 illustrates the schematic presentation of
an extractive distillation column. The column consists
of 27 stages, including the total condenser (stage 1)

Received 2009-12-24, accepted 2010-03-29.
* Supported by the National Natural Science Foundation of China (20776118), Science & Technology Bureau of Xian
[CXY09019 (1)], Innovation Foundation for Graduated Student of Northwest University (08YJC21), Shaanxi Research Center
of Engineering Technology for Clean Coal Conversion (2008ZDGC-13).
** To whom correspondence should be addressed. E-mail: maxym@nwu.edu.cn
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 363
and partial reboiler (stage 27). Table 1 lists the principal
configuration data. In both EQ and NEQ approaches
the vapor phase is assumed to be thermodynamically
ideal. In addition, the test for the presence of the sec-
ond liquid phase is considered in EQ approach. To
determine the bulk properties in NEQ approach, which
is relative to the inlet and outlet properties for each
phase on each stage, mixed and countercurrent flow
are considered as flow patterns for bubble cap tray and
random packing respectively. The 25 mm Intalox
metal tower packing (IMTP) applied has a HETP
value of 0.42 m in the EQ simulation.
2.2 Mathematic modeling incorporated with
Maxwell-Stefan theory
2.2.1 Film model and Maxwell-Stefan theory
Simulation of distillation process is often carried
out with the equilibrium stage model [21]. However,
the equipment and flow pattern, which present the
hydrodynamic characteristics and influence the ther-
modynamic property, are beyond the scope of EQ ap-
proach. In most equipment, the flow of both phases is
highly turbulent [22-24], so on not-too large scales the
concentrations and temperatures in the bulk fluids can
be considered as uniform. Near the phase interface
turbulence dies out and eddies do not pass across it.
Fig. 2 presents the simplest model, the film model, for
mass and heat transfer [17, 22, 24].

Figure 2 The simplest non-equilibrium model
In the film model, it is assumed that all the resis-
tance to mass transfer is concentrated in a thin film
adjacent to the interface [24-26], in which the transfer
occurs by steady state molecular diffusion. Outside this
film, in the bulk fluid, all composition gradients are
wiped out by turbulent eddy. A fully turbulent flow of
bulk phase is adjacent to the thin film in laminar flow
parallel to the interface. Mass transfer through this
film is in the direction normal to the interface, and any
constituent molecular diffusion or convection parallel
to the surface resulted from composition gradients along
the interface is negligible. To calculate the interphase
mass transfer fluxes in multicomponent mixtures, it is
now generally accepted that the Maxwell-Stefan diffu-
sion formulation is adopted for the fluid phases [27-32],
in which chemical potential gradients are used as the
driving forces for diffusion. A linear relation is postulated

Figure 1 Schematic presentation of extractive distillation
column
Table 1 Principal configuration data
Input data and specified parameters
Operating specifications
feed stage of entrainer above stage 6
feed stage of C
4
above stage 19
temperature of entrainer 50 C
pressure of entrainer 1 MPa
pressure of C
4
stream 0.51 MPa
reflux ratio 2
C
4
condition vapor
flow rate of C
4
10000 kgh
1

mass fraction of C
4

n-butane 50%
1-butene 3%
cis-2-butene 16%
trans-2-butene 31%
flow rate of entrainer 80000 kgh
1

entrainer condition liquid
mass fraction of water in entrainer 8.30%
Column parameters
number of stage 27
diameter of packing in the column 1.25 m
number of wash tray 5
diameter of wash tray 1.2 m
random packing IMTP, 25mm-metal
wash tray bubble cap tray
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 364
between the driving forces and the fluxes:
1 1 1
n n N
j i i j j i i j
ij i
j j j i ij i ij
j i j i
x j x j x N x N
c c

= = =


= =

(2)
where x
i
represents the mole fraction in the fluid phase,
N
i
is the molar flux,
ij
is the Maxwell-Stefan diffu-
sivity, and
i
is the chemical potential gradient,
which is from the frictional drag of one group of mole-
cules moving through the others. It does not matter
whether the frictional drag arises purely from inter-
molecular collisions as in the simple kinetic theory of
gases or additionally from intermolecular forces be-
tween two groups of molecules. The Maxwell-Stefan
diffusivity is given by
( )
2 1 2
d
d
i
RT
x v v
z

= (3)
With this definition, the Maxwell-Stefan diffusivity
(m
2
s
1
) is related to the drag coefficient, and is easier
to interpret and predict than Ficks diffusivity. For an
ideal binary gas mixture it is equal to Ficks diffusivity.
Following the approach of Krishna and Taylor [24]
Eq. (2) can be recast in terms of the mass transfer
coefficients k
ij

1 1
1, 2, , 1
n n
j i i j
ij j
j j i ij
j i
x N x N
x i n
c k

= =

= =

(4)
where x
j
represents the difference in composition
between the bulk fluid and interface, and
ij
represents
thermodynamic correction factor related to the non-
ideal behavior. For a highly nonideal mixture, the ther-
modynamic factor is usually a strong function of the
mixture composition and vanishes in the region of the
critical point [17, 21]. The driving force contains con-
tributions due to mole fraction and activity coefficient
of species i, and the contribution of mole fraction gra-
dient is similar to that of concentration in Ficks equa-
tion. However, Ficks equation does not consider the
effect of nonideality via the activity coefficient [33, 34].
Thus the formulation is useful in relating the practical
coefficients to the molecular collision processes and
the intermolecular interactions in the mixture [24]. It
takes proper account of diffusional coupling between
the species transfer, i.e. the flux of any species de-
pends on the driving forces of all the species present
in the mixture.
2.2.2 Model formulation
For the system considered in the simulation as
shown in Fig. 2, our immediate task is to develop the
balances for describing the transport processes.
Material and energy balances for the bulk liquid
and vapor:
L F L L
1 , 1
0
j ij j i j ij ij j ij
F x L x N r L x

+ + + = (5)
V FV V V V V
1 1
0
j j j j j j j j
F H V H Q q V H
+ +
+ + + = (6)
V F V V
1 , 1
0
j ij j i j ij ij j ij
F y V y N r V y
+ +
+ + + = (7)
L FL L L L L
1 1
0
j j j j j j j j
F H L H Q q L H

+ + + = (8)
Material and energy balances for liquid and vapor film:
I fL L
0
ij ij ij
N r N + = (9)
V fV I
0
ij ij ij
N r N + = (10)
I L
0
j j
q q = (11)
V I
0
j j
q q = (12)
Phase equilibrium:
I I
0
ij ij ij
y K x = (13)
Mole fraction summation for bulk liquid and vapor:
1
1 0
n
ij
i
x
=
=

(14)
1
1 0
n
ij
i
y
=
=

(15)
Mole fraction summation for interface:
I
1
1 0
n
ij
i
x
=
=

(16)
I
1
1 0
n
ij
i
y
=
=

(17)
Mass flux for bulk liquid and vapor:
( ) ( )
( )
E L I
L L L
0
j j j j j j
j j t j
x z
x x
R N N x

+

=

(18)
( ) ( )
I I V V V
0
j j j j j t j
y y R N N y
+ =

(19)
where is the matrix of thermodynamical factors:
L
L
L
, , ,
, ,
ln
j j
ij
i k j i k ij
kj
T P
x
x

= +

(20)
V
V
V
, , ,
, ,
ln
j j
ij
i k j i k ij
kj
T P
y
y

= +

(21)
The symbol means fixing the mole fractions of all
other components except for the n-th component while
evaluating the differentiation. The inverse matrixes of
mass transfer coefficients R are
L
, ,
L I L L I L
1 , , , ,
1, , 1
n
ij mj
i i j
m j j i n j j j i m j
m i
x x
R
a k a k
i n

=

= +

=

(22)
L
L L L L L L , ,
, , , ,
1 1
1, , 1,
i k j ij
j j i k j j j i n j
R x
a k a k
i n i k

=



= (23)
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 365
V
, ,
V I L V I V
1 , , , ,
1, , 1
n
ij mj
i i j
m j j i n j j j i m j
m i
y y
R
a k a k
i n

=

= +

=

(24)
V
V I V V I V , ,
, , , ,
1 1
1, , 1,
i k j ij
j j i k j j j i n j
R y
a k a k
i n i k

=



= (25)
The heat stream for the bulk liquid and vapor phase is
given by
( )
I L L L L I L
1
0
n
j j j ij ij j j
i
a h q N H T T
=
+ =

(26)
( )
I V V V V V I
1
0
n
j j j ij ij j j
i
a h q N H T T
=
+ =

(27)
To reduce the size of Jacobian elements, the mass
transfer coefficients are written as
o
, , , ,
j
i k j j i k j
k k D

= (28)
where
o
j
k is a function of flow, temperature, composi-
tion and other properties but independent of compo-
nents i and k.
2.3 Correlation methods
Accurate transfer efficiency and knowledge of the
maximum hydraulic capacity [35] and pressure drop [36]
of a packing are essential for design and operation of
packed columns. The transfer process in packed col-
umns is dependent on many factors, such as flow rates
of vapor and liquid, physical properties, and vapor-liquid
equilibrium [37-42], which change from location to
location along the column. Most of the existing mod-
els for prediction of height of a transfer unit (HTU)
are inappropriate if the factors stated above change
along the column significantly, especially for nonideal
and chemical reaction systems.
2.3.1 Correlations for interfacial area and mass trans-
fer coefficients
For bubble cap tray, the correlation of Gerster et
al. [43] is used to estimate the interfacial area and
mass transfer coefficients, which is as follows.
Binary mass transfer coefficient for the liquid
and vapor:
( ) ( )
0.5
8 L
s L ,
L
,
L I
0.21313 0.15 4.127 10
i k
i k
F Lt D
k
a
+
= (29)
V L
w s ,
w
0.5 V
V, , b
I
104.85
0.776 4.567 0.2377
i k
i k s
Q
h F k
l
Sc A v
a


+ + =


(30)
Interfacial area:
I 0.375 0.247 0.515
b V L w
0.27 a A Re Re h = (31)
Superficial F-factor:
( )
0.5
V V
s s t
F u

= (32)
Average residence time for liquid:
L L w L
0.9998 / t h Zl Q = (33)
Liquid height:
L w L w s
0.04191 0.19 2.4545 / 0.0135 h h Q l F = + + (34)
Average volumetric flow rate per pass for liquid:
L L p
/ Q Q N = (35)
For IMTP packing, the correlation of Onda et al.
[44] is used to predict the interfacial area and mass
transfer coefficients.
Binary mass transfer coefficient for the liquid
and vapor:
( ) ( )
0.333
L
0.4 0.667
L 0.5
p p , L, , L
L
0.0051
i k i k
g
a d k Sc Re


=



(36)
( )
2
V 0.7 0.333 V
p p , V V, , p ,
5.23
i k i k i k
a d k Re Sc a D

= (37)
Effective interfacial area for mass transfer:

I
w t p
a a A h = (38)
Wetted surface area per unit volume:
0.75
0.1 0.05 0.2
c
w p L L L
1 exp
1.45
a a Re Fr We





=




(39)
2.3.2 Correlations for heat transfer coefficients
The Chilton and Colburn method is based on the
relationship of Chilton-Colburn analogy and calcu-
lates heat transfer coefficients from the binary mass
transfer coefficients. It is probably the most successful
and widely used analogy and proved to be the most
accurate [45-50].
Heat transfer coefficients for vapor and liquid:
2/ 3
L
L L L L
L L L
P
P
k C
C D


=



(40)
2/ 3
V
V V V V
V V V
P
P
k C
C D


=



(41)
Average diffusivity and mass transfer coefficient:
( )( )
( )( )
c c
c c
1
1 1
1
1 1
n n
ij kj ikj
i k i
j
n n
ij kj
i k i
x x D
D
x x

= = +

= = +
+ +
=
+ +


(42)
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 366
( )( )
( )( )
c c
c c
1
1 1
1
1 1
n n
ij kj ikj
i k i
j
n n
ij kj
i k i
x x D
K
x x

= = +

= = +
+ +
=
+ +


(43)
where n
c
is the number of components. The Chilton-
Colburn averaging parameter has the default value
of 10
4
. Eqs. (5)-(43) describe the transport process
for the investigated system and the numerical solution
represents the behavior in the column. The set of equa-
tions is solved using Newtons method. Property moni-
tors are configured to verify the convergence. How-
ever, for some properties, such as diffusivities, partial
molar enthalpies and activity coefficients, the deriva-
tives are not available for the property monitors as they
involve derivatives of matrices or derivatives of deriva-
tives. The next section gives the results of the simulation.
2.4 Thermodynamic property
In the nonequilibrium model, we assume that
equilibrium exists at the interface. Thermodynamic
model is needed to describe the equilibrium for the
mixture. Due to the highly nonideal behavior in extrac-
tive distillation, it is necessary to apply relative com-
plex thermodynamic method for accurate prediction of
the corresponding phase equilibria. The choice of spe-
cific thermodynamic model has a great effect on the
results of simulation [8, 27, 46]. Two models are usually
used in multicomponent distillation: multicomponent
equations of state and Gibbs excess energy models.
The latter are useful for strongly nonideal mixtures at
not-too-high pressures [27, 33]. In this study it is found
that the predictions, especially the temperature profiles,
by NRTL, UNIQUAC and ASOG are quite different. The
estimated profiles by UNIFAC and UNIFAC-DMD
with parameters regressed from VLE data are similar.
To regress the UNIFAC-DMD parameters, the values
provided by Detherm V2.0 are used in this work. Ta-
bles 2 and 3 show the definition of UNIFAC group
parameter and UNIFAC group vector, respectively.
Fig. 3 shows that the prediction results are in good
agreement with experimental data selected from
Detherm V2.0.
Table 2 Definition of UNIFAC groups
Group ID Group number Molecular structure
G1005 1005

G1010 1010

G1015 1015

G1060 1060

G1065 1065

G1070 1070

G1300 1300 H
2
O
G3450 3450 C
5
H
9
NO
Table 3 UNIFAC-DMD group vector containing the
UNIFAC group number and the number of
occurrence of each group
No. a


1 1010 1010 1015 1015 1300 3450
2 2 1 2 2 1 1
3 1015 1015 1065 1060
4 2 1 1 1
5 1070
6 1
Components: a, n-butane; b, 1-butene; c, cis-2-butene; d,
trans-2-butene; e, water; f, N-methylpyrrolidone.

Figure 3 Temperature predicted by UNIFAC-DMD with
parameters regressed by vapor liquid equilibrium data
cis-2-butene & n-butane; cis-2-butene & NMP; cis-2-butene
& I-butene
3 SIMULATION RESULTS
We have developed equations to describe the
thermo- and hydrodynamic behavior in the region ad-
jacent to the interface with the appropriate boundary
conditions at the interface, and considered the trans-
port process across the interface. Both EQ and NEQ
simulations are carried out in the framework of As-
penONE

. Fig. 4 shows column composition profiles


predicted by EQ and NEQ models and experimental
values, in which only the profiles of n-butane and
trans-2-butene in vapor phase are plotted for a better

Figure 4 Composition profiles predicted by EQ and NEQ
models
n-butane (NEQ); n-butane (EQ); experiment (n-butane);
trans-2-butene (NEQ); trans-2-butane (EQ)
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 367
view of the profiles without the loss of generality. The
composition profiles predicted by EQ and NEQ mod-
els differ significantly. The composition profile of
n-butane predicted by NEQ model shows an excellent
agreement with experimental data, while that by EQ
model differs significantly from the experimental val-
ues. In order to understand the reason, Murphree effi-
ciency, diffusion coefficients and mass transfer coeffi-
cients are investigated.
3.1 Murphree vapor efficiency and diffusion co-
efficients
In this section, we present information on com-
ponent Murphree vapor efficiencies and binary diffu-
sion coefficients. The NEQ approach does not use the
efficiency, but the efficiencies of each component on
each tray can be calculated from its results. Fig. 5 pre-
sents component Murphree vapor efficiencies calcu-
lated from the results of NEQ simulation. The differ-
ences in Ei
MV
are from the differences in the diffusivities
of binary pair vapor in the mixture, and Maxwell-Stefan
diffusion takes proper account of diffusional coupling
between the species. Moreover, the efficiencies vary
greatly from stage to stage and some are abnormal. Some
of the components have extreme values of 2400%
and +800%. In particular, the efficiencies of n-butane
are negative on some stages. Such odd behavior as
negative efficiency was observed experimentally for
acetone, methanol and water mixture in a sieve tray
column [51]. The main reasons are as follows. On the
one hand, the entrainer acts as a selective semiperme-
able filter, which lets those more soluble components,
1-butene, cis-2-butene and trans-2-butene, pass pref-
erentially. These components in the vapor phase enter
the liquid region through the interface and release its
latent heat. n-butane in the liquid absorbs the energy
released at the interface and moves from its low con-
centration region into n-butane rich vapor phase,
leading reverse diffusion of n-butane. On the other
hand, because of the poor solubility the equilibrium of
n-butane does not exist at the interface, so that the
driving force y
i
of n-butane in vapor boundary layer
vanishes, i.e. y
i
0. Its flux is resulted from the
movement of other components in the mixture, mainly
by 1-butene, cis-2-butene and trans-2-butene, so the
component efficiencies are greater than 100% and
may be positive or negative. At the same time, com-
ponent efficiency changes along the column signifi-
cantly. As a result, the mole fraction of component
(Fig. 4) on any stage with the NEQ approach is dif-
ferent from that by the EQ approach. In Fig. 5 with
increasing vapor loading from top to bottom, the effi-
ciencies for trans-2-butene, 1-butene and cis-2-butene
increase slightly. The situation is changed dramatically
when vaporous C
4
mixture is introduced. The sudden
jump discontinuity in the efficiencies appears from
stage 19 (feed stage of vaporous C
4
mixture) to stages
18 and 17, with the largest positive values (220% for
1-butene and 150% for cis-2-butene). The negative
efficiencies of n-butane appear from stage 11 to stage
23, which drive n-butane from the low concentration
liquid region into the rich vapor region. Tables 4 and 5
show the binary diffusion coefficients in vapor phase
on stages 8 and 22 respectively. The binary diffusion
coefficients of NMP have the smallest values and
those of n-butane are the next, which indicates that the
flux of n-butane is resulted from the movement of
other components. Table 6 shows the corresponding
mole fraction on stages 8 and 22, on which the driving
force of n-butane is much greater than that of others
and the mole fractions of n-butane in vapor phase are
much greater than that in liquid phase. n-butane moves
into vapor phase from liquid phase against the driving

Figure 5 Component Murphree vapor efficiencies along column
n-butane; trans-2-butene; cis-2-butene; 1-butene; NMP; water

Table 4 Binary diffusion coefficients in vapor
phase on stage 8 (10
6
m
2
s
1
)
a

0 1.0284 1.0216 1.0226 2.3961 0.7522


b

1.0284 0 1.0640 1.0651 2.5320 0.7880


c

1.0216 1.0640 0 1.0570 2.4994 0.7841


d

1.0226 1.0651 1.0570 0 2.4846 0.7831


e

2.3961 2.5320 2.4994 2.4846 0 1.7884


f

0.7522 0.7880 0.7841 0.7831 1.7884 0


Components: a, n-butane; b, 1-butene; c, cis-2-butene; d,
trans-2-butene; e, water; f, N-methylpyrrolidone.
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 368
force. The component Murphree efficiency in Fig. 5
indicates that the values of HETP or HTU are different
from stage to stage, so the application of classical HTU-
NTU approach and constant Ei
MV
for packed-column
design will lead to poor results compared with the
more rigorous approach with the Maxwell-Stefan model.
Olano et al. [52] also obtained similar conclusion for
other process. In addition, EQ model offers no explicit
information about diffusion and transfer across the
interface, so that the variation of Murphree vapor effi-
ciency along column is not considered. As a result,
process design based on EQ concept involves uncer-
tainties of Murphree efficiency and limitation of
Ficks diffusion theory. On the contrary, NEQ ap-
proach gives us quite a detailed understanding of what
is going on in the equipment.
3.2 Mass transfer rate and mass transfer coeffi-
cients of component
Since NEQ approach presents quite detailed be-
havior in the column, it is possible to investigate the
mass transfer rate of component across the interface,
diffusion coefficients and mass transfer coefficients on
each stage. Fig. 6 illustrates the transfer rate of com-
ponent across the interface along the column. The flux
of n-butane across the interface is from liquid phase to
vapor phase, and the fluxes of other components such
as 1-butene, cis-2-butene, trans-2-butene, water and
NMP are from vapor phase to liquid region. On the
Table 5 Binary diffusion coefficients in vapor
phase on stage 22 (10
6
m
2
s
1
)
a

0 1.0842 1.0773 1.0793 2.5292 0.7930


b

1.0842 0 1.1199 1.1215 2.6564 0.8312


c

1.0773 1.1199 0 1.1131 2.6230 0.8270


d

1.0793 1.1215 1.1131 0 2.6142 0.8255


e

2.5292 2.6564 2.6230 2.6142 0 1.8792


f

0.7930 0.8312 0.8270 0.8255 1.8792 0


Components: a, n-butane; b, 1-butene; c, cis-2-butene; d,
trans-2-butene; e, water; f, N-methylpyrrolidone.

Table 6 Mole fraction in vapor and liquid phase on stage 8 and 22
Mole fraction/molmol
1

Stage Phase
a


8 vapor 0.800337 0.017297 0.047475 0.126164 0.008403 0.000324
liquid 0.142389 0.004061 0.012969 0.032672 0.269374 0.538535
22 vapor 0.396424 0.040453 0.175672 0.374683 0.012286 0.000482
liquid 0.062889 0.009248 0.047499 0.095076 0.261728 0.523560
Components: a, n-butane; b, 1-butene; c, cis-2-butene; d, trans-2-butene; e, water; f, N-methylpyrrolidone.

Figure 6 Mass transfer rates of component across the interface predicted by NEQ models
(Positive values refer to transfer from vapor to liquid)
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 369
feed stages (6 and 19), the fluxes for C
4
components
change dramatically, mainly because the introduction
of fresh feed stream increases the driving force for
transfer and intensity of turbulence. The mass transfer
rate reduces from stage 6 to stage 19 but increases
from stage 20 to the bottom with the increase of vapor
loading. To investigate the mass transfer, it is custom-
ary to determine mass transfer coefficients [38, 45, 53, 54].
Eqs. (42) and (43) are used to calculate the average
diffusion and mass transfer coefficients of stage from
binary pair diffusion and mass transfer coefficients,
which are estimated with Eqs. (29), (30), (36) and (37).
Figs. 7-9 illustrate the average diffusion and mass
transfer coefficients in liquid and vapor phases along
the column. Both diffusion coefficients and mass
transfer coefficients in vapor phase are much greater
than those in liquid phase, mainly due to the differ-
ences in viscosity and density of the phases. In NEQ
approach, the movement of molecule or a group of
molecules depends on the driving forces of all the
species present in the mixture, while in EQ approach
the driving forces are merely the forces between the
key component and other species. Moreover, the in-
terfacial area is reduced as the deformation of fluid
surface is hindered due to the increase of inner friction
in the fluid, leading a decrease in mass transfer rate.

Figure 7 Average diffusion and mass transfer coefficients in liquid phase along column
D
L
; k
L


Figure 8 Average mass transfer coefficients in vapor phase and vapor Reynolds number along column
Re
V
; k
V


Figure 9 Kinematic viscosity and average diffusion coefficients of vapor

V
; D
V

Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 370
In addition, the mass transfer coefficient in vapor
phase is sensitive to the vapor loading, so it changes
dramatically on the feed stage. The diffusion coeffi-
cients for both phases and mass transfer coefficients in
liquid phase increase from the top to the bottom of the
column. This can be explained as follows. On the one
hand, the selective semipermeable filter lets the more
soluble components cross the interface preferentially
and release their condensation heat freely, leading an
enrichment of n-butane in the vapor film adjacent to
the interface. On the other hand, liquid n-butane ab-
sorbs the energy released and vaporizes, which also
enriches n-butane in the vapor film adjacent to the
interface. Consequently, n-butane diffuses from the
interface to the bulk vapor phase and carries away
those components diffusing towards the interface. The
resistance for diffusion and mass transfer is hence re-
duced, increasing the mass transfer and diffusion co-
efficients and leading the movement of n-butane from
its low concentration liquid region to n-butane rich
vapor phase. For other components, the mechanism is
the same in the liquid phase. On stage 6, where fresh
entrainer is introduced into the column, the diffusion
and mass transfer coefficients in liquid phase take the
smallest values (Fig. 7), but the flux of n-butane from
liquid phase to vapor region has the highest value (Fig.
6) due to the largest driving forces for extraction. On
the contrary, the mass transfer coefficient in vapor
phase takes relative large value, though the diffusion
coefficient on this stage has the smallest value. This is
mainly due to the relative large kinematic viscosity,
which can be traced back to vapor density.
4 CONCLUSIONS
The limitations of Ficks law for describing dif-
fusion in nonideal multicomponent mixture are dis-
cussed. The Maxwell-Stefan diffusion theory is used
to describe transfer processes and associated influence
factors are investigated on the basis of EQ and NEQ
simulations. The major conclusions are as follows.
(1) The flux of n-butane, which has the smallest
solubility, is resulted from the movement of all the
species present in the C
4
mixture. Diffusion against
driving force is explained by NEQ approach.
(2) The Murphree efficiencies of components in
C
4
mixture differ from each other and vary greatly
from stage to stage. The unbounded and negative effi-
ciencies are the results of diffusional coupling.
(3) The differences in the results with NEQ and
EQ are from the difference in the Murphree efficiencies
of components, which can be traced back to the differ-
ence in diffusivities D
y,ij
of the binary pair vapor phase.
(4) The NEQ approach takes proper account of
diffusional coupling between the species transfer, which
helps understanding of what is going on in the extrac-
tive distillation column, and avoids the uncertainties
of tray efficiency or HETP concept. The effects of
design parameter and equipment geometry are in-
cluded. This is an important extension of the classical
analysis using equilibrium stages.
Thus for the simulation of extractive distillation
column to separate n-butane from 1-butene, cis-2-butene
and trans-2-butene mixture, the rigorous NEQ ap-
proach is more appropriate.
NOMENCLATURE
A
b
active bubbling area on the tray, m
2

A
t
cross-sectional area of the column, m
2

a
p
specific surface of packing, m
2
m
3

a
w
wetted area, m
2

c molar concentration, molm
3

D diffusions coefficient, m
2
s
1

Maxwell-Stefan diffusions coefficient, m
2
s
1

d diameter, m
d
h
hydraulic diameter, m
d
p
nominal packing size, m
F feed stream, kgh
1

g gravitational acceleration, ms
2

H enthalpy, kJ
h specific enthalpy, kJkmol
1

h
L
liquid height, m
h
p
height of the packed section, m
h
w
average weir length per liquid pass, m
K equilibrium coefficient
k mass transfer coefficients, ms
1

L mass stream of liquid, kgh
1

M mole mass, kgkmol
1

N
p
number of liquid flow passes
n
c
number of component
Q heat input to stage, Js
1

Q
L
, Q
V
volumetric flow rate for liquid and vapor, m
3
s
1

q heat transfer rate, Js
1

R gas constant, JK
1
mol
1

r reaction rate, kmols
1

T absolute temperature, K
t
L
average residence time for liquid, s
u inner energy, kJkmol
1

V vapor stream, kgh
1

v velocity, ms
1

x mole fraction in liquid , molmol
1

y mole fraction in vapor , molmol
1

z height of package, m
heat transfer coefficient, Wm
2
K
1

thermodynamic correction factor
film thickness, m
heat conductivity, Wm
1
K
1

dynamic viscosity, Pas
kinematic viscosity, m
2
s
1

density, kgm
3

surface tension, Nm
1

to
surface tension of liquid to the stage
fugacity, Pa
fugacity coefficient

E
driving force caused by electric potential
Superscripts
F feed
f film
I interface
L liquid phase
V vapor phase
Subscripts
i index of component
j index of stage
Chin. J. Chem. Eng., Vol. 18, No. 3, June 2010 371
k bulk of phase
L liquid
m component
n last component
p packing
s surface
t total
V vapor
w wetted area
REFERENCES
1 Billet, R., Industrielle Destillation, Chemie Verlag, Weinheim (1995).
(in German)
2 Dssel, R., Stichlmair, J., Separation of azeotropic mixtures by batch
distillation using an entrainer, Comput. Chem. Eng., 19, 113-118 (1995).
3 Castillo, F.J.L., Towler, G.P., Influence of mlulticomponent mass
transfer on homogeneous azeotropic distillation, Chem. Eng. Sci.,
53 (5), 963-976 (1999).
4 Langston, P., Hilal, N., Shingfield, S., Webb, S., Simulation and op-
timization of extractive distillation with water as solvent, Chem.
Eng. Process, 44, 345-351 (2005).
5 Jimnez, L., Wanhschafft, O.M., Julka, V., Analysis of residue curve
maps of reactive and extractive distillation units, Comput. Chem.
Eng., 25, 635-642 (2001).
6 Lei, Z.G., Zhou, R.Q., Duan, Z.T., Process improvement on sepa-
rating C
4
by extractive distillation, Chem. Eng. J., 85, 379-386 (2002).
7 Repke, J.U., Ausner, I., Paschke, S., Hoffmann, A., Wozny, G., On
the track to understanding three phases in one tower, Trans IChemE,
Part A, Chem. Eng. R. & D., 85 (A1), 50-58 (2007).
8 Thomsen, K., Iliuta, M.C., Rasmussen, P., Extended UNIQUAC
model for correlation and prediction of vapor-liquid-liquid-solid
equilibria in aqueous salt systems containing non-electrolytes. Part b:
alcohol (ethanol, propanols, butanols)-water-salt systems, Chem.
Eng. Sci., 59, 3631-3647 (2004).
9 Eckert, E., Vank, T., Some aspects of rate-based modeling and
simulation of three-phase distillation columns, Comput. Chem. Eng.,
25, 603-612 (2001).
10 Kewis, W.K., Seader, J.D., The rate-based approach for modeling
staged separations, Chem. Eng. Prog., 85, 41-49 (1989).
11 Higler, A., Chande, R., Taylor, R., Baur, R., Krishna, R., Nonequi-
librium modeling of three-phase distillation, Comput. Chem. Eng.,
28, 2021-2036 (2004).
12 Yang, X.J., Yin, X., Ouyang, P.K., Simulation of 1,3-butadiene
production process by dimethylfomamide extractive distillation,
Chin. J. Chem. Eng., 17 (1), 27-35 (2009).
13 Eckert, J.S., Trays and packings: Selecting the proper distillation
column packing, Chem. Eng. Prog., 66, 39-44 (1970).
14 Fullarton, D., Selektivitt und bertragungsleitstung, Chem. Eng.
Proc., 20, 255-263 (1986). (in German)
15 Higler, A., Krishna, R., Taylor, R., Nonequilibrium cell model for
multicomponent (reactive) separation processes, AIChE J., 45 (11),
2357-2370 (1999).
16 Kooijman, H.A., Taylor, R., Modeling mass transfer in multicom-
ponent distillation, Chem. Eng. J., 57, 177-188 (1995).
17 Krishna, R., Wesselingh, J.A., The Maxwell-Stefan approach to
mass transfer, Chem. Eng. Sci., 52 (6), 861-911 (1997).
18 Toor, H.L., Diffusion in three-component gas mixtures, AIChE J.,
3, 198-207 (1957).
19 Zimmerman, A., Joulia, X., Gourdon, C., Gorak, A., Maxwell-Stefan
approach in extractor design, Chem. Eng. J., 57, 229-236 (1995).
20 Lao, M., Taylor, R., Modeling mass transfer in three-phase distilla-
tion, Ind. Eng. Chem. Res., 33, 2637-2650 (1994).
21 Wesselingh, J.A., Non-equilibrium modeling of distillation, Trans
IChemE., 75 (A), 529-538 (1997).
22 Krishna, R., Standart, G.L., A multicomponent film model incorpo-
rating an exact matrix method of solution to the Maxwell-Stefan
equations, AIChE J., 22, 383-389 (1976).
23 Krishna, R., Martinez, H.F., Sreedhar, R., Standart, G.L., Murphree
point efficiencies in multicomponent mixtures, Trans. Inst. Chem.
Eng., 55, 178-186 (1977).
24 Krishna, R., Taylor, R., Multicomponent Mass Transfer, John Wiley
& Sons, New York (1993).
25 Walter, G.W., The two-film theory of gas absorption, Chemical
and Metallurgical Engineering, 29 (4), 146-148 (1923).
26 Krishnamurthy, R., Taylor, R., A nonequilibrium stage model of
multicomponent separation processes. Part I: model, AIChE J., 3,
449-461 (1985).
27 Baehr, H.D., Thermodynamik 11. Auflage, Springer-Verlag, Berlin
(2002). (in German)
28 Seader, J.D., The rate-based approach for modeling staged separa-
tions, Chem. Eng. Prog., 85, 41-49 (1989).
29 Springer, P.A.M., Krishna, R., Crossing of boundaries in ternary
azeotropic distillation: influence of interphase mass transfer, Int.
Comm. Heat & Mass Transfer, 28 (3), 347-356 (2001).
30 Springer, P.A.M., van der Molen, S., Krishna, R., The need for us-
ing rigorous rate-based models for simulations of ternary azeotropic
distillation, Comput. Chem. Eng., 26, 1265-1279 (2002).
31 Springer, P.A.M., van der Molen, S., Baur, R., Krishna, R., Ex-
perimental verification of the Maxwell-Stefan formulation in de-
scribing composition trajectories during azeotropic distillation,
Trans IChemE., 80 (A), 654-666 (2002).
32 Springer, P.A.M., Baur, R., Krishna, R., Composition trajectories
for heterogeneous azeotropic distillation in a bubble-cap tray col-
umn, Trans IChemE., 81(A), 413-426 (2003).
33 Mersmann, A., Thermische Verfahrenstechnik, Springer Verlag,
Berlin (1980). (in German)
34 Ojeda Nava, J.A., Krishna, R., Influence of unequal component ef-
ficiencies on trajectories during distillation of a homogeneous
azeotropic mixture, Chem. Eng. Process, 43, 305-316 (2004).
35 Oluji, ., Development of a complete simulation model for pre-
dicting the hydraulic and separation performance of distillation
columns equipped with structured packings, Chem. Biochem. Eng.,
11 (1), 31-46 (1997).
36 Stichlmair, J., Bravo, J.L., Fair, J.R., General model for prediction
of pressure drop and capacity of countercurrent gas: liquid packed
columns, Gas Sep. Purif., 3, 19-28 (1989).
37 Rehfeldt, S., Stichlmair, J., Measurement and calculation of multi-
component diffusion coefficients in liquids, Fluid Phase Equilibria,
256, 99-104 (2007).
38 Billet, R., Schultes, M., Prediction of mass transfer columns with dumped
and arranged packings, Trans IChemE., 77 (A), 498-504 (1999).
39 Baur, R., Taylor, R., Krishna, R., Copati, J.A., Influence of mass
transfer in distillation of mixtures with a distillation boundary,
Trans IChemE., 77 (A), 561-565 (1999).
40 Antonio Rocha, J., Bravo, J.L., Fair, J.R., Distillation columns con-
taining structured packings: a comprehensive model for their per-
formance (II) mass-transfer model, Ind. Eng. Chem. Res., 35 (5),
1660-1667 (1996).
41 Zeck, S., Einflu von thermophysikalischen soffdaten auf die
auslegung und den betrieb von destillationskolonnen, Chem. Ing.
Teck., 62 (9), 707-717 (1990).(in German)
42 Xu, Z.P., Afacan, A., Chuang, K.T., Predicting mass transfer in
packed columns containing structured packings, Trans IChemE., 78
(A), 91-98 (2000).
43 Aspen Technology, AspenONE

2006 Reference Manual, Aspen


Technology Inc., Cambridge, MA (2008).
44 Onda, K., Akeuchi, H.T., Okumoto, Y., Mass transfer coefficients
between gas and liquid phases in packed columns, J. Chem. Eng.
Jpn., 1 (1), 56-61 (1968).
45 Billet, R., Schultes, M., Predicting mass transfer in packed col-
umns, Chem. Eng. Technol., 16, 1-9 (1993).
46 Gorak, A., Simulation thermischer trennverfahren fluider
vielkomponentengemische, Prozesssimulation, Schuler, H., eds.,
Wiley-VCH, Germany, 349-408 (2002). (in German)
47 Lucia, A., Amale, A., Taylor, R., Distillation pinch points and more,
Comput. Chem. Eng., 32, 1342-1364 (2008).
48 Rao, D.P., Prem Kumar, R.S., Pandit, P., Das, T.C.T., Multicomponent
tray efficiencies accounting for entrainment, Chem. Eng. J., 57,
237-246 (1995).
49 Strigle, R. F. Jr., Rukovena, F., Packed distillation column design,
Chem. Eng. Progr., 75 (3), 86-91 (1979).
50 Strigle, R.F., Jr, Random Packing and Packed Towers, Design and
Applications, GPC, Houston (1988).
51 Vogelpohl, A., Murphree efficiencies in multicomponent systems,
Ins. Chem. Eng. Symp. Ser., 56, 25-31(1979).
52 Olano, S., Nagura, S., Kosuge. H., Asano, K., Mass transfer in bi-
nary and ternary distillation by a packed column with structured
packing, J. Chem. Eng. Jpn., 28, 750-757 (1995).
53 Gao, X.Q., Ma, Y.G., Zhu, C.Y., Yu, G.C., Towards the mechanism
of mass transfer of a single bubble, Chin. J. Chem. Eng., 14 (2),
158-163 (2006).
54 Yuan, X.G., Yu, G.C., Computational mass transfer method for chemi-
cal process simulation, Chin. J. Chem. Eng., 16 (4), 497-502 (2008).

Das könnte Ihnen auch gefallen