You are on page 1of 10

Chemical Engineering Science 66 (2011) 2713–2722

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Modeling a non-equilibrium distillation stage using irreversible


thermodynamics
D.F. Mendoza a, S. Kjelstrup b,
a
National University of Colombia, Cra. 30 Cll. 45, Bogotá, Colombia
b
Department of Chemistry, Norwegian University of Science and Technology, 7491 Trondheim, Norway

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

Article history: We compare transport equations derived from non-equilibrium thermodynamics to a classical rate
Received 26 April 2010 model developed over the last 20 years, in terms of their ability to calculate the heat and mass fluxes by
Received in revised form modeling a segment of a packed distillation column. We show, using water and ethanol separation as
14 October 2010
an example, that there is a significant coupling between heat and mass transfer. Neglect of this
Accepted 14 March 2011
transport coefficient leads to variations in the magnitude, even the sign of the calculated heat flux,
Available online 21 March 2011
while the mass flux is less affected.
Keywords: & 2011 Elsevier Ltd. All rights reserved.
Distillation
Mass transfer
Rate-based model
Heat transfer
Non-equilibrium thermodynamics
Dufour effect

1. Introduction then estimated by multiplying the vapor heat transfer coefficient


by a factor, often 1000 (Krishnamurthy and Taylor, 1985b).
Modeling multicomponent distillation systems, where simul- Churchill (1997), in a critical review of the analogies between
taneous heat and mass transfer occurs, is a challenging task, not heat, mass and momentum transfer, points out that the most
only because of the interaction effects among the different widely used heat–mass transfer analogy, the Chilton and Colburn
components in the mixture, but also because of the interaction (1934) analogy, lacks of any mechanistic or theoretical basis, it
between the processes of heat and mass transfer. A relative is in functional error with respect to Prandtl number (Pr) and
successful model applied for modeling multicomponent distilla- gives significant numerical errors for large values of Pr as well as
tion has been developed by Krishnamurthy and Taylor (1985a). In for Pr  1. This means that there is still much to do on both
this model the equations come from the mass and energy theoretical and experimental grounds for modeling the simulta-
balances for the vapor and liquid phases, the phase equilibrium neous heat and mass transfer in heterogeneous systems.
in the vapor–liquid interface, and the rate equations for vapor– In a series of efforts for getting better models for simultaneous
liquid heat and mass transfer. heat and mass transfer in distillation, Kjelstrup and De Koeijer (2003)
There are different models for heat and mass transfer. The most and Van der Ham et al. (2010) obtained transport equations using
common model for the vapor–liquid heat and mass transfer is non-equilibrium thermodynamics for the film. Non-equilibrium (or
based on the film theory, where the film thickness is calculated irreversible) thermodynamics is a systematic theory for description
using empirical correlations and analogies between heat and mass of transport phenomena, where the local entropy production is used
transfer. Generally only mass transfer correlations are available to define the force–flux relations. This theory takes into considera-
and the heat transfer coefficients are estimated applying heat– tion some features that are presently not explicit in the theoretical
mass transfer analogies for both films, or by applying a heat–mass description of the vapor–liquid heat and mass transfer, namely, the
transfer analogy to estimate the vapor-side heat transfer coeffi- direct coupling between heat and mass transfer and lack of equili-
cient for the vapor film. The liquid-side heat transfer coefficient is brium across the liquid–vapor interface (Bedeaux and Kjelstrup,
2004). The property has been taken care of in the empirical handling
of experimental data, however (Taylor and Krishna, 1993).
The models used in non-equilibrium thermodynamics (Kjelstrup
 Corresponding author. Tel.: þ 47 73594179; fax: þ 47 73550877. and De Koeijer, 2003; Van der Ham et al., 2010; Bedeaux
E-mail address: signe.kjelstrup@chem.ntnu.no (S. Kjelstrup). and Kjelstrup, 2004) and non-equilibrium molecular dynamics

0009-2509/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2011.03.023
2714 D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722

simulations (Kjelstrup and Hafskjold, 1996) have shown that direct Here a flux is considered positive when directed from the liquid
coupling between heat and mass transfer in non-ideal mixtures have phase to the vapor phase. In a steady-state operation there is no
an important effect on the predicted fluxes when phase transitions accumulation of mass in any part of the system, which means that
are involved and should not be neglected in the modeling of mass transfer flows are equal on each side of the vapor–liquid
distillation processes. transfer region:
In this work the transport equations developed by Van der
N Lj,n N Vj,n ¼ 0 ð4Þ
Ham et al. (2010) are, therefore, included in a rate-based model
and compared with the transport equations based on the Chilton– The total mass balances for the liquid, MnL , and vapor, MnV phases
Colburn analogy for heat and mass transfer, in terms of the are
predicted values of the fluxes.
This work is organized as follows: the balance equations of both MnL : FnL þ Ln1 Ln N T,n ¼ 0 ð5Þ
rate-based models are presented in Section 2. In Section 3, we give
the flux equations derived from non-equilibrium thermodynamics. MnV : FnV þ Vn þ 1 Vn þN T,n ¼ 0 ð6Þ
These equations are compatible with the entropy balance by
The last term in Eqs. (5) and (6) is the total mass flow, of the c
definition. A brief overview of heat and mass transfer calculations P
components in the mixture, N T,n ¼ cj ¼ 1 N j,n .
using the approach recommended by Taylor and Krishna (1993),
The energy balance for the liquid phase, ELn, reads
together with the expression to calculate the entropy production in
the liquid–vapor region is presented in Section 4. Section 5 is ELn : QnL þ FnL HFn
L L
þLn1 Hn1 Ln HnL E Ln ¼ 0 ð7Þ
devoted to the model solution. In order to obtain numerical insight
The energy balance for the vapor phase, EVn, is
we give results for the water/ethanol system in Section 6. The
reasons for the selection of this systems lie in its practical EVn : QnV þ FnV HFn
V
þ Vn þ 1 HnV þ 1 Vn HnV þE Vn ¼0 ð8Þ
importance, in knowledge of the physical-chemical properties
required by the model, and because the results can be directly where QVn and QLn represent the external heat transfer to the vapor
compared with a previous model obtained by Kjelstrup and De and liquid phases, respectively, and Hp (p : L,V) is the enthalpy of
Koeijer (2003). Some relevant conclusions are given in Section 8. the stream. The last terms in Eqs. (7) and (8) are the net energy
flows across the liquid and vapor phases, respectively:
Z An
2. Standard balance equations
E pj ¼ p dA ðp : L,VÞ
Je,n ð9Þ
0
A distillation stage or packed section, as pictured in Fig. 1, can
Where Je is the flux of energy in the phase:
be understood as composed of three regions, namely, the liquid
bulk, the vapor bulk and the vapor–liquid region which includes X
c
p
Jep ¼ Jq0p þ Jkp H k ðp : L,VÞ ð10Þ
the vapor and liquid films as well as the interface.
k¼1
The standard balance equations, common to all models, are as
follows. The mass balances for a component j in the liquid, MLj,n, The first term in the right hand side, Jq0 , is the measurable heat flux
and vapor, MVj,n, phases for a segment n counting from the top to (we use the term heat flux interchangeably), the second term is
the bottom, of a distillation column can be written as the energy transported by component k associated with the
partial molar enthalpy, H k .
L
Mj,n : xj,Fn FnL þ xj,n1 Ln1 xj,n Ln N Lj,n ¼ 0 ð1Þ The energy balance around the vapor–liquid transfer region,
EIn, in the steady state, shows that the net energy change in the
V
Mj,n : yj,Fn FnV þyj,n þ 1 Vn þ 1 yj,n Vn þ N Vj,n ¼ 0 ð2Þ
two phases is zero:
where FLn
and FVn
are the molar flows of the liquid and vapor feeds EIn : E Ln E Vn ¼ 0 ð11Þ
in the stage n; Ln  1 and Vn þ 1 are the internal liquid and vapor
flows entering the stage; Ln and Vn are the liquid and vapor flows Eqs. (1)–(11) are general for all steady-state distillation models.
leaving the stage; xj and yj are the mole fractions of the Differences among models come from the models used to calcu-
component j in the liquid and vapor phases, respectively. The late the rates of heat and mass transfer and how the interface is
last term in Eqs. (1) and (2) represents the net gain or loss of considered.
component j in the phase due to vapor–liquid mass transfer,
calculated as the integral of the local mass transfer rates Jj over
the interface area, A, in a tray or packed segment. 3. Vapor–liquid heat and mass transfer rates from
Z An the entropy production
N pj,n ¼ p dA
Jj,n ðp : L,VÞ ð3Þ
0 Van der Ham et al. (2010) derived an expression for simultaneous
heat and mass transfer at stationary state, using non-equilibrium
thermodynamics. They considered the entropy production of a
certain number of control volumes in the liquid–vapor region,
including that of the interface. Applying the relations between forces
and fluxes to each control volume, see Fig. 2, and assuming that the
enthalpy is temperature-independent in the temperature range of
the V–L region that is considered, they obtained expressions for the
total driving forces in terms of the heat and mass fluxes.
0 G1 0 G G
10 1
Xq rqq    rqc J 0V
B GC B G B C q C
G CB
B X1 C B r1q    r1c CB J1 C
B C CB C
B ^ C¼B ^ ^ ^ CB ^ C
ð12Þ
@ A B @ A@ A
Fig. 1. Scheme of a non-equilibrium stage considering the interface and the films G G
XcG rcq    rcc Jc
next to it as a separated thermodynamic system.
D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722 2715

The entropy production calculated from the force–flux pairs,


IT
dSirr =dt, is with mechanical equilibrium:
IT Xc
dSirr
¼ Jq0V AX Gq þ Jj AX Gj ð17Þ
dt j¼1

where J’s are the fluxes calculated by Eq. (12), XG’s are the forces
Fig. 2. Scheme of the vapor–liquid region with m control volumes. I is the control
calculated by Eq. (13) and A is the interface area. Each term is
volume of the interface, ai and bi are the borders of the control volume, di is the
control volume length, Jj is the flux of component j, Jq0L and Jq0V are the measurable positive definite according to the second law of thermodynamics.
EB
heat fluxes in the vapor and liquid phase, respectively. The entropy production given by the entropy balance, dSirr =dt, is
!
EB Xc
dSirr Jq0V Jq0L
¼  Aþ Jk Aðs Vk s Lk Þ ð18Þ
dt TV TL k¼1
Here Jq0V is the measurable heat flux on the vapor side of the vapor–
liquid region, Jc is the stationary state molar flux of component c with where s Vk and s Lk are the partial molar enthalpies of the component
respect to the interface frame of reference; XG’s are the total driving k evaluated at the conditions of the vapor and liquid phases.
forces for heat and mass transfer, see Eq. (13); and rG’s are the overall
resistivities given by Eq. (14).
4. Vapor–liquid heat and mass transfer rates and entropy
The total driving forces are calculated from the temperatures
production using Taylor and Krishna’s approach
of the liquid and vapor bulk phases, TL and TV; and from the
chemical potentials of the components, mj , as
In Taylor and Krishna’s approach, the film theory is used
   together with the Maxwell–Stefan equations (Krishnamurthy
1 1 mj,V ðTL Þmj,L
XqG ¼  ; XjG ¼  , ð13Þ and Taylor, 1985b). The general expression used to calculate the
TV TL TL
mass transfer rates in the vapor phase is (Krishna and Standart,
where mj,V ðTL Þ is the chemical potential at the vapor-side bound- 1979):
ary of the vapor–liquid region evaluated at the liquid-side
ðN V Þ ¼ ½kV ½GV ½XV AðyI y V Þ þ N Vt ðy V Þ ð19Þ
boundary temperature and mj,L is the chemical potential at the
liquid boundary of the vapor–liquid region. where N Vt
is the net mass transfer in the vapor phase, y is theV

The elements of the matrix of total resistivities are given by average phase composition and yI is the composition in the vapor
side of the interface; ½kV  is the matrix of Maxwell–Stefan mass
X
m
transfer coefficients in the vapor and ½XV  is the correction factor
G
rqq ¼ r iqq ð14aÞ
i¼1
that takes into account finite transfer rates, and ½GV  is the matrix
of thermodynamic factors that accounts for the non-idealities in
the vapor mixture,
X
m
G
rjq G
¼ rqj ¼ ðr iqj þ r iqq DiVb H j Þ ð14bÞ ^
@lnf
i¼1 GVij  dij þyi i
ði,j ¼ 1, . . . ,c1Þ ð20Þ
@yj
^ is the fugacity coefficient of
where dij is the Kroneker delta and f
X
m i
G
rkj G
¼ rjk ¼ ðr ijk þr ijq DiVb H k þr ikq DiVb H j þr iqq DiVb H k DiVb H j Þ ð14cÞ i in the vapor mixture. The mass transfer rates in the liquid phase
i¼1 are calculated using an expression similar to Eq. (19):

where DiVb H k
is the difference of the partial molar enthalpy of ðN L Þ ¼ ½kL ½GL ½XL Aðx L xI Þ þ N Lt ðxL Þ ð21Þ
component k between the vapor phase and the border of the where N Lt is the net mass transfer in the liquid phase, x L is the
control volume i, bi; i.e., DiVb H k ¼ H k,V H k,bi , see Fig. 2. Enthalpies average composition of the liquid phase and xI is the composition
are evaluated at the average temperature of the vapor–liquid at the liquid side of the interface; ½GL  is the matrix of thermo-
region. dynamic factors in the liquid phase given by Eq. (20) where the
The resistivity in the control volume i, r i , for the liquid and activity coefficient, gi , replaces the fugacity coefficient, f^ i , and the
vapor films, is the product of the local resistivity, ri, evaluated at mole fractions in the liquid phase, xi, xj, replaces the vapor mole
the average temperature and composition of the control volume, fractions, yi, yj; ½kL  is the matrix of Maxwell–Stefan mass transfer
and the control volume length, di, i.e., r i ¼ r i di . For the interface, r i , coefficients, ½XL  is the matrix of correction factors that takes into
can be calculated directly from kinetic theory. The expressions for account other finite mass transfer rates. Explicit formulae for
the local resistivities are given in Section 6.1. these matrices can be found elsewhere (Krishna, 1977).
The measurable heat flux at the liquid side, Jq0L , can be Compositions at the liquid and vapor sides of the interface are
calculated from the vapor-side measurable heat flux, Jq0V , using calculated by assuming phase equilibrium at the interface:
the energy balance around the vapor–liquid region:
yIj Kj xIj ¼ 0 ð22Þ
X
c
Jq0L ¼ Jq0V þ DVL H k Jk ð15Þ where Kj is the equilibrium ratio for component j in the multi-
k¼1 component system evaluated at a given temperature, pressure
and composition. The mole fractions at the liquid and vapor sides
The term DVL H k is the difference between the partial molar
of the interface must sum to unity,
enthalpies of component k in the vapor and liquid phases.
The entropy production calculated by the force–flux products X
c
yIj 1 ¼ 0 ð23Þ
in non-equilibrium thermodynamics, Eq. (17), and by the entropy
j¼1
balance, Eq. (18), must be the same,
EB IT
X
c
dSirr dSirr xIj 1 ¼ 0 ð24Þ
 ¼0 ð16Þ j¼1
dt dt
2716 D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722

The measurable heat transfer for the liquid and vapor phases, 4.1. Entropy production for heat and mass transfer
Jq0L and Jq0V , is calculated as
The entropy production given by the entropy balance equation
Jq0L ¼ hL XLH ðTL TI Þ ð25Þ around the liquid–vapor region is
!
EB Xc
dSirr Jq0V Jq0L
Jq0V ¼ hV XVH ðTI TV Þ ð26Þ ¼  Aþ N k ðs Vk s Lk Þ ð33Þ
dt TV TL k¼1
L V
where h , h are the heat transfer coefficients for the liquid and The measurable heat flux, Jq0 , and the partial molar entropies, s k ,
vapor phases, respectively. XLH , XVH are corrections due to the effect are evaluated at the boundaries of the liquid and vapor films,
of a finite mass transfer rate on the heat transfer coefficients. which are the vapor and liquid phase conditions.
Expressions for these correction factors, given by the film theory,
can be found elsewhere (Taylor and Krishna, 1993; Bird et al.,
1960). This means that the empirical relations for heat and mass 5. Solving the distillation stage
transfer include indirectly a possibility for interaction between the
fluxes. 5.1. Taylor–Krishna model
Due to the difficulty of evaluating heat transfer coefficients for
each phase, Krishnamurthy and Taylor (1985b) suggested to use The Taylor–Krishna model uses 6cþ 7 equations, viz., the mass
the Chilton–Colburn analogy to estimate the vapor heat transfer and energy balances for the liquid and vapor phases, Eqs. (1), (2) and
coefficients from mass transfer coefficients as (5)–(8); the interface mass and energy balances, Eqs. (4) and (11);
the phase equilibrium and composition summation at the interface,
hV ¼ kVav CpV ðLe2=3 ÞV ð27Þ Eqs. (22)–(24); and the energy and mass transfer rates, Eqs. (21), (25)
and (26). These equations are solved simultaneously to calculate the
where kVav is the average value of all the binary mass transfer flow, temperature and composition of the vapor and liquid streams
coefficients in the multicomponent mixture, CVp is the molar heat leaving the stage, the heat and mass transfer rates, as well as, the
capacity of the vapor mixture and Le is the Lewis number, temperature and composition of the vapor and liquid sides of the
l interface.
Le ¼ ð28Þ
DrC^ p
5.2. Non-equilibrium thermodynamics model
where l, r, C^ p and D are the thermal conductivity, density, the
mass heat capacity and the average Fick’s diffusion coefficient of The model equations derived from non-equilibrium thermo-
the mixture, respectively. dynamics have 3cþ7 unknowns (see Fig. 1), namely: 2c composi-
The heat and mass transfer coefficients are functions of the tions: xn, yn; two temperatures: TLn, TVn; two flows: Ln, Vn; c mass
fluid flow mechanism, the fluid properties and the geometry. fluxes: N j ; one measurable heat flux: Jq0L (or Jq0V ) and the liquid–
L V
Most experimental correlations used in the prediction of mass vapor film thickness ratio d =d .
transfer coefficients have been obtained for binary systems at low These equations are solved in two steps. The first step
solute concentrations and low net mass transfer fluxes (Bird et al., calculates the interface heat and mass fluxes, as well as, the film
2002), therefore they are low-flux mass transfer coefficients. thickness ratio, the second step uses the values of the first step to
The relationship between the experimental low-flux heat and calculate the unknown flows and compositions.
mass transfer coefficients with the film thicknesses is (Krishna, The equations to be solved in the first step are Eqs. (12), (15)
1977) and (16). A numerical procedure to solve the cþ3 unknowns, is
explained in detail by Van der Ham et al. (2010).
l The equation set used in the second step is given by Eqs. (1),
h¼ ð29Þ
dH (2), (4)–(8), this set is used to calculate the 2c þ4 unknowns
which are: flows, compositions and temperatures of the vapor
½B1 ½G and liquid streams that leave the stage. The solution is achieved
½k ¼ ð30Þ
dm when the equations of step 2 are solved.

where l is the thermal conductivity calculated at the film’s


5.3. Film thickness
average temperature and composition, dH and dm are the effective
film thicknesses for heat and mass transfer, respectively, ½G is the
Non-equilibrium thermodynamics prescribes direct coupling
matrix of thermodynamic factors given in Eq. (20), ½B is a matrix
between heat and mass transfer; therefore, the effective film
of the Maxwell–Stefan diffusion coefficients which elements are
thicknesses for heat and mass must be the same since the heat
calculated according to (Krishna, 1977)
flow triggers a mass flow and vice versa.
zi Xc
zk The model based on non-equilibrium thermodynamics
Bii ¼ þ ði ¼ 1, . . . ,c1Þ ð31Þ
Dic k ¼ 1 Dik requires one film thickness, either the vapor or liquid film
kai thickness. The film thickness used in the model is vapor-side
mass transfer thickness as predicted by Eq. (30). The reasons for
 
1 1 this election are based on the fact that the heat and mass transfer
Bij ¼ zi  ði ¼ 1, . . . ,c1; i ajÞ ð32Þ
Dij Dic thicknesses in the vapor phase tend to be similar and because the
mass transfer coefficient comes from experimental data, whereas
Eqs. (29) and (30) were obtained for a planar film neglecting the heat transfer coefficient does not.
coupling between heat and mass and assuming constant proper-
ties, evaluated at the average film composition and temperature. 6. Cases studied
The values of dH and dm given by Eqs. (29) and (30) tend to be
considerably different in the liquid phase and relatively similar in In order to compare the model of Taylor and Krishna with the
the vapor phase (Frank et al., 1995b; Tosun, 2002). expressions from non-equilibrium thermodynamics, a segment of
D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722 2717

a packed column is modeled for the well-known water/ethanol where s and sc are the surface tension of the liquid and the
system. The information required for the simulation is given in critical surface tension of the packing, respectively; ReL is the
Table 1. Reynolds number based on the specific surface area, Eq. (41), FrL
The inlet flows, compositions and temperatures come in both liquid-phase Froude number, Eq. (42) and WeL is the Weber
cases from a simulation of a distillation column in Aspen plusTM number, Eq. (43).
using the Radfrac model with rate-based non-equilibrium column,
rL uL
this model uses Onda’s correlation (Onda et al., 1968) to calculate ReL ¼ ð41Þ
m L ap
the mass transfer coefficients and the interface area, and applies
the Chilton–Colburn analogy to estimate the heat transfer coeffi-
ap u2L
cients in both phases. Fr L ¼ ð42Þ
g
Onda’s correlation allows to estimate the binary low-flow
mass transfer coefficients for the vapor, kVij and liquid phases, kLij, rL u2L
as well as the interface area, a0 WeL ¼ ð43Þ
ap s
0:333
kVij ¼ ARe0:7
V Sc V ðap dp Þ2 ðap DVij Þ1 ð34Þ
The model of Taylor and Krishna and the equations from non-
where dp and ap are the nominal packing size and the specific equilibrium thermodynamics were programmed in MatlabTM to
surface area of the packing, A is a constant, it is 2.0 when dp is assess their prediction of the heat and mass transfer rates.
less than 0.012 m and 5.23 when dp is greater than or equal to The procedures used to estimate transport and thermodynamic
0.012 m; DVij is the vapor-phase Fick’s diffusion coefficient of properties are summarized in Table 2. The heat of mixing in the
species i in j; ReV and ScV are the vapor-phase Reynolds and liquid phase is estimated from the partial derivatives of the activity
Schmit numbers coefficients, with respect to the temperature using the NRTL model.
Detailed explanations about these methods and specific informa-
rV uV tion for the water/ethanol system can be found in the references
ReV ¼ ð35Þ
mV ap given in Table 2. A summary of the expressions for the local
resistivities in the liquid and vapor films are given in Table 3.
mV Detailed derivations of these expressions can be found elsewhere
ScV ¼ ð36Þ
rV DVij (Kjelstrup and De Koeijer, 2003; Kjelstrup and Bedeaux, 2008).
where rV is the vapor mass density, uV is the superficial velocity
and mV is the vapor viscosity. The mass transfer coefficient is 6.1. Calculation of transport properties
estimated from the following equation:
 L 0:333 The local resistivities in the vapor and liquid phases, given in
0:4 r Table 3, are calculated using the Maxwell–Stefan diffusion coeffi-
kLij ¼ 0:0051Re00:667
L Sc 0:5
L ða p dp Þ ð37Þ
mL DL cients, DEW , the thermal conductivity, l, the heat of transfer, qnE,
The liquid-phase Schmit and Reynolds numbers are and the molar density of the mixture, C. Kinetic theory is used to
estimate the interface resistivities.
mL The heat of transfer for ethanol in the vapor and liquid films,
ScL ¼ ð38Þ
rL DLij qnE, is given by

rL uL qE ¼ zw RT 2 GEE STE ð44Þ


Re0L ¼ ð39Þ
m L a0 Where zW is the mole fraction of water, GEE is the thermodynamic
The interfacial area density, a0 is calculated from factor of ethanol, and STE is the Soret coefficient of ethanol in the

a0 ¼ ap ð1expf1:45ðsc =sÞ0:75 Re0:1 0:05


L Fr L We0:2
L gÞ ð40Þ Table 2
Methods used for the calculation of transport and thermodynamic properties.
Table 1
Property Method Source
Specifications of the distillation column segment.

Pure liquid heat capacity Correlation Poling et al. (2008)


Specification Units
Pure vapor heat capacity Correlation Poling et al. (2008)
Vapor pressure Antoine Poling et al. (2000)
Geometry
Activity coefficient NRTL Renon and Prausnitz
Diameter 0.152 m
(1969)
Segment height 0.10 m
Vapor density Ideal gas
Packing Liquid density Rackett Poling et al. (2000)
Nutter ring Random Vapor viscosity Lucas Poling et al. (2000)
Material Steel Liquid viscosity Corresponding Teja and Rice (1981)
Size 0.025 m states
Critical surface tension 0.075 N m1 Vapor thermal Molar average Poling et al. (2000)
Surface density 167 m2 m  3 conductivity
Liquid thermal Fillipov Poling et al. (2000)
Liquid inlet
conductivity
Flow 14.32 mol s  1
Vapor Maxwell–Stefan Fuller et al. Poling et al. (2000)
Temperature 352.12 K
diffusivity
Pressure 1 atm
Liquid Maxwell–Stefan Correlation Kjelstrup and De Koeijer
Ethanol mole fraction 0.64
diffusivity (2003)
Vapor inlet Surface tension Corresponding Poling et al. (2000)
Flow 25.42 mol s  1 states
Temperature 352.48 K Liquid Soret coefficient Experimental Kolodner et al. (1998)
Pressure 1 atm data
Ethanol mole fraction 0.69 Vapor Soret coefficient Kinetic theory Hirschfelder et al. (1964)
2718 D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722

Table 3 the vapor phase is calculated using the Chilton–Colburn analogy


Expressions for calculation of local resistivities in the films. and the heat transfer coefficient in the liquid film is 1000 times
the value of the heat transfer coefficient in the vapor film. In CC2
Units
the Chilton–Colburn analogy is applied to calculate the heat
rEE Rzw J m s mol  2 K  1 transfer coefficients in the vapor and liquid films.
þ rqq q2
E
zE CDEW The model labeled NET denotes the rate equations derived
rEW zE J m s mol  2 K  1
 rEE
zW
from non-equilibrium thermodynamics, which takes into account
rWW  2 J m s mol  2 K  1 coupling between heat and mass transfer and non-equilibrium in
zE
rEE the interface.
zW
rqq 1 m s J1 K1
lT 2
rqE rqq qE m s mol  1 K  1
rqW zE m s mol  1 K  1
7.1. Predicted fluxes
 rqE
zW
The values obtained by the models CC1 and CC2 (Table 4)
show that ethanol is transferred from the liquid to vapor, whereas
Table 4 water is transferred from vapor to liquid; the net mass transfer is
Summary of results obtained using Taylor and Krishna’s model (CC1 and CC2) and from the liquid phase to the vapor phase. Both models predict a
equations from non-equilibrium thermodynamics (NET).
heat flux from vapor to liquid.
CC1 CC2 NET The major difference between models CC1 and CC2 is the
magnitude of the calculated heat fluxes in the liquid phase. The
Vapor heat flux in CC1 is 43% bigger than the estimated by CC2 because
V (mol s  1) 25.43 25.43 25.42 the Chilton–Colburn analogy, used in CC1 to predict the liquid-
yE 0.69 0.69 0.69
yW 0.31 0.31 0.31
phase heat transfer coefficient, gives a greater heat-transfer
TV (K) 352.42 352.42 352.33 resistance than the model CC2; the difference can be appreciated
when the values of the heat and mass transfer thicknesses, dH =dm ,
Liquid
L (mol s  1) 14.31 14.31 14.31 are compared. A bigger measurable heat flux in CC1 is evidenced
xE 0.64 0.64 0.64 by a higher temperature of the liquid stream leaving the stage,
xW 0.36 0.36 0.36 0.02 K.
TL (K) 352.19 352.17 352.21 The energy balance in the vapor–liquid region, Eq. (11), shows
Fluxes that differences in the heat fluxes can modify the mass transfer
JE (mol s  1 m  2) 0.43 0.43 0.41 fluxes, in the studied cases the fluxes of ethanol and water are
JW (mol s  1 m  2)  0.40  0.40  0.41
the same in both models, because the change in the heat flux in
Jq0L (W m  2)  404  282  1076
the liquid film, about 121 W m  2, is minor compared with the
Jq0V (W m  2)  307  307 219
enthalpy fluxes of ethanol and water, e.g., the heat flux needed to
A (m2) 0.30 0.30 0.30
evaporate 1 mol s  1 m  2 of ethanol, is  65500 W m2 .
Film thickness
Mass and liquid-phase measurable heat fluxes calculated by
dVm  105 (m) 1.9 1.9 1.9
the model NET (Table 4) agree qualitatively with the models CC1
dVH =dVm 0.79 0.79 1.00
and CC2; i.e., the measurable heat flux is directed from the vapor
dLm =dVm 0.26 0.26 0.25
to liquid, ethanol evaporates and water condenses. However, the
dLH =dLm 2.60 3.61 1.00
heat flux in the vapor phase has a different direction, it goes
Entropy production against the temperature gradient, from the interface to the vapor,
dSirr/dt(W K  1) 0.44 0.44 0.43 and the vapor-side of the interface temperature has a lower
temperature than the liquid-side, see Fig. 3.
The differences in the values of fluxes of heat and mass
mixture defined as the ratio between the thermal diffusion obtained by the model NET respect to the models CC have two
coefficient, DTE, and Fick’s diffusion coefficient, DEW (DeGroot and sources, viz., the film thickness ratios and the coupling effects.
Mazur, 1984) Unlike the CC’s models, the heat–mass film thickness ratio used,
dH =dm , in the NET model is always 1, this is mandatory because of
DTE
STE  ð45Þ coupling. The heat transfer thicknesses in the vapor film is  30%
DEW
thicker, whereas the heat transfer thickness in the liquid film is
The expressions for the calculation of resistivities at the interface about three times thinner than in the models CC1 and CC2, this
for a binary mixture are given elsewhere (Kjelstrup and De explains, in part, the higher heat flux in the liquid film and a
Koeijer, 2003). lower heat flux in the vapor film when compered with the CC
models.
The influence of coupling on the predicted heat and mass
7. Results and discussion fluxes are calculated from the expressions for the measurable
heat and component fluxes given by irreversible thermody-
A summary of the results obtained by the models using the namics. The local thermal force, Eq. (46), and chemical forces,
Chilton–Colburn analogy and the equations from non-equilibrium Eqs. (47) and (48), are expressed in terms of local resistivities and
thermodynamics is given in Table 4. The results are evaluated in fluxes
terms of their consistence with the entropy balance and in terms  
@ 1
of the values of the calculated fluxes. Additionally the influence of ¼ rqq Jq0 þ rqE JE þrqW JW ð46Þ
@Z T
coupling on the fluxes of heat and mass is assessed for the model
based on non-equilibrium thermodynamics.  
The models named CC are the rate-based models using the 1 @mE
 ¼ rEq Jq0 þ rEE JE þ rEW JW ð47Þ
Chilton–Colburn analogy; in CC1 the heat transfer coefficient in T @Z T,p
D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722 2719

   
1 @mW 1 @mE rEW rEq 0
 ¼ rWq Jq0 þrWE JE þ rWW JW ð48Þ JE ¼   JW  J ð50Þ
T @Z T,p rEE T @Z T,p rEE rEE q

 
here Z is a spatial coordinate increasing from the left (liquid) to 1 @mW rEW rEq 0
JW ¼   JE  J ð51Þ
the right (vapor), Jq 0 , JE and JW are the local fluxes of heat and mass rWW T @Z T,p rWW rWW q
and r’s are the local resistivities defined in Table 3. The local heat
and mass transfer fluxes, for the vapor and liquid films, are The relative influence of each term in the value of the heat fluxes
calculated from Eqs. (46)–(48) (Fig. 4), shows that coupling has a significant effect on the
  calculated heat fluxes in the liquid and vapor films. The magni-
1 @ 1 rqE rqW
Jq0 ¼  JE  JW ð49Þ tudes of the contributions to the heat flux JqE and JqW, tend to be
rqq @Z T rqq rqq
similar in both, the liquid and vapor films; nevertheless, coupling
effects are more significant in the vapor phase than in the liquid
phase, because the magnitude of the conductive heat transfer, Jqq,
is six times lower than in the liquid phase.
On the other hand, the effect of heat transfer on the fluxes of
ethanol and water, JEq and JWq, is negligible (Fig. 4). The insignif-
icant influence of heat transfer on mass transfer in distillation
operations has been recognized by Taylor and Krishna (1993).
The interface plays a key role for the positive heat flux obtained
in the vapor phase because the contribution of the mass flux to
heat flux in the interface is the greatest in the vapor–liquid region
(Fig. 4). The values JqE and JqW in the interface are two and five
times larger than in the vapor or liquid phases, respectively. This
behavior agrees with the work of Kjelstrup and Hafskjold (1996)
who demonstrated, using non-equilibrium molecular dynamics
simulations, the importance of coupling in the interface.
The influence of coupling on the heat fluxes and temperature
profiles depend on the sign and magnitude of the Soret or Dufour
coefficients. Delancey and Chiang (1964) acknowledged the
importance of heat–mass coupling on the heat fluxes, especially
in the liquid phase. They demonstrated that the Dufour effect (the
reciprocal of Soret effect) can even cause a temperature inversion.
In this work Soret coefficients in the liquid and vapor phases are,
on average, 1.1  10  3 and 7.9  10  4 K  1, respectively, the
orders of magnitude of these coefficients agree with the values
considered normal (10  3–10  5 K  1) for both, liquid and gaseous
solutions (DeGroot and Mazur, 1984). Besides, it is known that
Fig. 3. Composition and temperature profiles in the liquid–vapor region calcu- kinetic theory underestimates the interface resistivities when
lated using the NET model. Control volume number 100 represents the interface. compared with molecular dynamics simulations.

Fig. 4. Contribution to total heat and mass fluxes. Jqq heat conduction, JqE, JqW mass transfer contributions to heat transfer. JEE, JWW mass fluxes due to a chemical potential
gradient, JEW ethanol mass flux due to water mass transfer, JWE water flux due to ethanol mass transfer, JEq, JWq ethanol and water mass transfer due to heat transfer.
2720 D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722

7.2. Entropy production and thermodynamic consistency of does not violate the second law of thermodynamics because a
transport equations positive entropy production is obtained throughout the liquid–
vapor region.
The entropy production profile of the vapor–liquid region The vapor and liquid films generate most of the entropy in the
given by NET model (Fig. 5) shows that the phenomenon of heat liquid–vapor (68% and 32%, respectively). Even so, the entropy
transfer directed toward the direction of increasing temperature production in the interface is comparable with the entropy
generated in a control volume in the vapor phase.
The entropy production in the models CC1 and CC2 gives a
greater entropy production, about 2%, than the NET model
(Table 4), the discrepancy in the values is given by the difference
in the ethanol mass flux predicted by the model NET, because the
influence of heat transfer on the entropy production is rather
little, see Table 5 and Kjelstrup and De Koeijer (2003), Mendoza
and Riascos (2009).

7.3. The assumption of equilibrium at the interface and


coupling effects

The cases named IT (Table 5) are intended to assess the


influence of the interface resistivities and coupling in the liquid
and vapor films on the values of the fluxes of heat and mass
calculated by the NET model. In the case IT1, the interface
Fig. 5. Entropy production profile of the liquid–vapor region using the NET model. resistivities are neglected, i.e., equilibrium at the interface is
assumed. In IT2, the heat–mass coupling in the liquid phase is
Table 5 neglected by setting the Soret coefficient in the liquid phase as
Influence of coupling on the entropy production and predicted fluxes for heat zero. In IT3, coupling in the vapor phase is neglected by setting the
and mass. Soret coefficient in the vapor equal to zero; the case IT4 is intended
to resemble the classical film theory (equilibrium at the interface
IT1 IT2 IT3 IT4
and no coupling in the films), by setting the interface resistivities
Fluxes and the Soret coefficients in the liquid and vapor as zero.
JE (mol s  1 m  2) 0.41 0.41 0.41 0.41 In all IT cases the variation in the mass fluxes and in the liquid-
JW (mol s  1 m  2)  0.41  0.41  0.41  0.41 phase heat fluxes is less than 2% when compared with the mass
Jq0L (W m  2)  1077  1060  1099  1085
fluxes obtained by the model NET. The most important coupling
Jq0V (W m  2) 222 231  43  29
effect is in the vapor phase, cases IT3 and IT4, where the heat
A (m2) 0.30 0.30 0.30 0.30
fluxes exhibit a sign change compared to the NET model. The
Film thickness composition profiles obtained in the IT models overlap closely to
dV  105 (m) 1.9 1.9 1.9 1.9
the composition profile obtained by the NET model (Fig. 3);
dL =dV 0.25 0.25 0.25 0.25
whereas the temperature profile is more sensitive to coupling,
dSirr/dt (W K  1) 0.43 0.43 0.43 0.43
see Fig. 6.

Fig. 6. Influence of coupling on the heat fluxes and temperature profiles.


D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722 2721

Neglecting coupling has a deep effect on the physical under- Je flux of total energy, W m  2
standing of the system, mainly on the heat fluxes. According to the JEE ethanol flux due to a chemical potential
NET model the interface has a heat source (the heat of condensation), gradient, mol m2 s  1
but the models where coupling is neglected (CC1, CC2, IT3 and IT4) JEW contribution of the water flux to the ethanol flux,
are unable to predict such behavior. In the present work coupling mol m2 s  1
affects primarily the vapor phase; nevertheless, Delancey and Chiang JEq contribution of the heat flux to the ethanol flux,
(1964) showed that coupling in the liquid phase is also important to mol m2 s  1
describe adequately the heat transfer phenomenon in the liquid Jj flux of component j, mol m2 s  1
phase. The results support the idea that coupling should not be Jq 0 flux of measurable heat transfer, W m2
neglected in distillation processes and agree with the theoretical Jqq conductive heat transfer, W m2
analysis of Bedeaux and Kjelstrup (2005) who point out that JqE contribution of ethanol mass transfer to heat
transport equations based on Fick’s and Fourier’s laws alone, are transfer, W m2
not enough for the description of the simultaneous heat and mass JqW contribution of the water mass transfer to heat
transfer processes in two-phase regions. transfer, W m2
We have, therefore, seen that the prediction of the magnitude JWE contribution of the ethanol flux to the water flux,
and direction of the measurable heat flux differ largely between mol m2 s  1
the models investigated. A good prediction of the heat flux is JWW water flux due to a gradient in the chemical
required, not only because of its influence on the mass fluxes, but potential, mol m2 s  1
mainly because it is needed for good design. It is of primary JWq contribution of the heat flux to the water flux,
importance for diabatic and heat integrated columns, and in mol m2 s  1
studies of optimizations of the energy efficiency of distillation Kj equilibrium ratio of j, dimensionless
columns (Røsjorde and Kjelstrup, 2005). The outcome of such kij low rate mass transfer coefficient, mol m2 s  1
studies depends on accurate calculations of the heat transfer kij low rate mass transfer coefficient in an ideal
profile in the column(s). In addition, one may expect that it plays solution, mol m2 s  1
a role in the modeling of distillation systems where the thermal L flow of liquid, mol s  1
effects are significant, e.g., reactive distillation. In such columns, Le Lewis number, dimensionless
the thermal effects can modify the mass fluxes by a factor of 30 Nj mass transfer rate of j, mol  1 s  1
(Frank et al., 1995a), so coupling effects should clearly be taking R ideal gas constant, 8.3145 J mol  1 K  1
into account. It is then an advantage to have a systematic Re Reynolds number, dimensionless
procedure to derive pertinent transport equations. r local resistivities, see Table 3
ST Soret coefficient, K  1
s entropy of the stream, J mol  1 K  1
8. Conclusions
Sc Schmit number, dimensionless
sj partial molar entropy of j, J mol  1 K  1
We have studied a segment of a distillation column for water/
u velocity, m s  1
ethanol, and learnt by solving various sets of transport equations,
V flow of vapor, mol s  1
that the coupling between heat and mass transfer (the Dufour
We Weber number, dimensionless
effect) has a marked influence on the measurable heat flux. It
x mole fraction in the liquid phase, dimensionless
should be included in existing rate-based models if any precision
y mole fraction in the vapor phase, dimensionless
is needed in the modeling of heat fluxes.
z mole fraction, dimensionless

Greek
Nomenclature
d film thickness, m
A interface area, m2 dij Kroneker delta, dimensionless
A parameter used in Onda’s correlation, f^ j fugacity coefficient of component j,
dimensionless dimensionless
a0 interface area density, m2 m  3 g activity coefficient, dimensionless
ap specific surface area of the packing, m2 m  3 ½G matrix of thermodynamic factors, dimensionless
C molar density of the mixture, mol m  3 l thermal conductivity, W m  1 K  1
Cj molar density of j, mol m  3 m viscosity in Eq. (37), Pa s
c number of components, dimensionless mj chemical potential of component j, J mol  1
C^ p mass heat capacity, J kg  1 K  1 r density, kg m  3
Cp molar heat capacity, J mol  1 K  1 s surface tension, N m  1
½B matrix of Maxwell–Stefan coefficients, m2 s  1 sc critical surface tension, N m  1
Dkj Fick diffusion coefficient, m2 s  1 ½X matrix of correction factors for finite mass
DTk thermal diffusion coefficient, m2 s  1 K  1 transfer, dimensionless
Dkj Maxwell–Stefan diffusion coefficient, m2 s  1 ½XH matrix of correction factors for finite heat
dp nominal packing size, m transfer, dimensionless
dsirr/dt entropy production, W K  1 ½k matrix of binary mass transfer coefficients,
E energy transfer rate, W m s1
F flow of feed, mol s  1
Subscripts and superscripts
Fr Froude number, dimensionless
H molar enthalpy of the stream, J mol  1 E ethanol
Hj partial molar enthalpy of j, J mol  1 Fn feed in the stage n
h heat transfer coefficient, W m  2 K  1 G global
2722 D.F. Mendoza, S. Kjelstrup / Chemical Engineering Science 66 (2011) 2713–2722

H heat maxwell–Stefan theory—I. Model development and isothermal study. Chem.


I interface Eng. Sci. 50, 1645–1659.
Van der Ham, L.V., Bock, R., Kjelstrup, S., 2010. Modelling the coupled transfer of
i component i mass and thermal energy in the vapour–liquid region of a nitrogen–oxygen
NET non-equilibrium thermodynamics mixture. Chem. Eng. Sci. 65, 2236–2248.
j component j Hirschfelder, J.O., Curtis, C.F., Bird, R.B., 1964. Molecular Theory of Gases and
Liquids. Structure of Matter Series. Wiley, New York, USA.
k component k
Kjelstrup, S., Bedeaux, D., 2008. Non-Equilibrium Thermodynamics of Heteroge-
L liquid phase neous Systems, Series on Advances in Statistical Mechanics. vol. 16, World
m mass Scientific, Singapore.
n tray or segment number Kjelstrup, S., De Koeijer, G.M., 2003. Transport equations for distillation of ethanol
and water from the entropy production rate. Chem. Eng. Sci. 58, 1147–1161.
V vapor phase Kjelstrup, S., Hafskjold, B., 1996. Nonequilibrium thermodynamics simulations of
W water steady-state heat and mass transport in distillation. Ind. Eng. Chem. Res. 35,
4203–4213.
Kolodner, P., Williams, H., Moe, C., 1998. Optical measurements of the Soret
coefficient of ethanol/water solutions. J. Chem. Phys. 88, 6512–6524.
Krishna, R., 1977. A generalized film model for mass transfer in non-ideal fluid
mixtures. Chem. Eng. Sci. 32, 659–667.
Acknowledgments Krishna, R., Standart, G.L., 1979. A multicomponent film model incorporating a
general matrix method of solution to the maxwell–Stefan equations. AIChE J.
The authors appreciate financial support from the Norwegian 22, 383–389.
Krishnamurthy, R., Taylor, R., 1985a. A nonequilibrium stage model of multi-
Research Council, Storforsk Project 167336/V30. Valuable discus- component separation processes. Part I: model description and method of
sions with Leen van der Ham are appreciated. solution. AIChE J. 31, 449–456.
Krishnamurthy, R., Taylor, R., 1985b. A nonequilibrium stage model of multi-
component separation processes. Part II: comparison with experiment. AIChE
References J. 31, 456–465.
Mendoza, D.F., Riascos, C.A.M., 2009. Entropy production analysis in extractive
Bedeaux, D., Kjelstrup, S., 2004. Irreversible thermodynamics—a tool to describe distillation column using non-equilibrium thermodynamics and a rate based
phase transitions far from global equilibrium. Chem. Eng. Sci. 59, 109–118. model. In: de Brito Alves, R.M., Oller do Nascimento, C.A., Biscaia, E.C. (Eds.),
Bedeaux, D., Kjelstrup, S., 2005. Heat, mass, and charge transport and chemical 10th International Symposium on Processes System Engineering. Computer
reactions at surfaces. Int. J. Thermodyn. 8, 25–41. Aided Chemical Engineering, vol. 27. Elsevier, Amsterdam, The Netherlands,
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 1960. Transport Phenomena. John Wiley, pp. 789–794.
New York, USA. Onda, K., Takeuchi, H., Okumoto, Y., 1968. Mass transfer coefficients between gas
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2002. Transport Phenomena, second ed. and liquid phases in packed columns. J. Chem. Eng. Jpn. 1, 56–61.
John Wiley, New York, USA. Poling, B.E., Prausnitz, J.M., O’Conell, J.P., 2000. The Properties of Gases and Liquids.
Chilton, T.H., Colburn, A.P., 1934. Mass transfer (absorption) coefficients prediction McGraw-Hill, New York, USA.
from data on heat transfer and fluid friction. Ind. Eng. Chem. 26, 1183–1187. Poling, B.E., Thomson, G.H., Friend, D.G., Rowley, R.L., Wilding, W.V., 2008. Perry’s
Churchill, S., 1997. Critique of the classical algebraic analogies between heat, mass, Chemical Engineers’ Handbook, eighth ed. McGraw-Hill, New York, USA.
and momentum transfer. Ind. Eng. Chem. Res. 36, 3866–3878. Renon, H., Prausnitz, J.M., 1969. Estimation of parameters for the NRTL equation
DeGroot, S.R., Mazur, P., 1984. Non-equilibrium Thermodynamics. Dover, New for excess gibbs energies of strongly nonideal liquid mixtures. AIChE J. 8,
York, USA. 413–419.
Delancey, G.B., Chiang, S.H., 1964. Dufour effect in liquid systems. AIChE J. 14, Røsjorde, A., Kjelstrup, S., 2005. The second law optimal state of a diabatic binary
664–665. tray distillation column. Chem. Eng. Sci. 60, 1199–1210.
Frank, M.J.W., Kuipers, J.A.M., Krishna, R., Van Swaaij, W.P.M., 1995a. Modelling of Taylor, R., Krishna, R., 1993. Multicomponent Mass Transfer. John Wiley, New
simultaneous mass and heat transfer with chemical reaction using the York, USA.
maxwell–Stefan theory—II. Non-isothermal study. Chem. Eng. Sci. 50, Teja, A.S., Rice, P., 1981. Generalized corresponding states method for the
1661–1671. viscosities of liquid mixtures. Ind. Eng. Chem. Fundam. 20, 77–81.
Frank, M.J.W., Kuipers, J.A.M., Versteeg, G.F., Van Swaaij, W.P.M., 1995b. Modelling Tosun, I., 2002. Modelling in Transport Phenomena a Conceptual Approach.
of simulations mass and heat transfer with chemical reaction using the Elsevier, Amsterdam, The Netherlands.

You might also like