Вы находитесь на странице: 1из 32

Activity Coefficient Calculation for

Binary Systems Using UNIQUAC

Project Work in the Course

Advanced Thermodynamics: With Application to


Phase and Reaction Equilibria
KP8108

Department of Chemical Engineering, NTNU

Ugochukwu E. Aronu

July, 2009
Table of Contents 2

1.0 Introduction 3

2.0 Thermodynamic Framework 4

2.1 Thermodynaic Equilibrium Condition 4


2.2 Partial Molar Gibbs Free Energy 5
2.3 Gibbs-Duhem Equation 6
2.4 Equilibrium in Heterogenous Closed System 7
2.5 Chemical Potential 8
2.6 Fugacity and Fugacity Coefficient 9
2.7 Activity and Activity Coefficient 10

2.7.1 Normalization of Activity Coefficient 11

2.8 Excess Functions 12

3.0 Vapour Liquid Equilibrium Calculation 13

3.1 Chemical Equilibria 14

3.1.1 Chemical Equilibrium Constant 15

3.2 Phase Equilibria 15


3.3 G E Model for Activity Coefficient 17

3.3.1 Local Composition Models 18

4.0 UNIQUAC 18

4.1 UNIQUAC Implementation 20

4.1.1 MEA-H20 System 21


4.1.2 Acetic Acid-H20 System 21
4.1.3 Other Systems 21

5.0 Results 25

Summary 25

Future Work 25

References 25

Derivation of Activity Coefficient Expression Based on UNIQUAC 27

2
1.0 Introduction
Design of industrial chemical process separation equipment such as absorption and distillation
columns as well as simulation of chemical plants require reasonably accurate correlation and
prediction of phase equilibria because it is an integral part of vapour liquid equilibrium (VLE)
modeling. It is thus necessary to develope thermodynamic models suitable for practical phase
equilibrium calculations.

In general, VLE models are based on fundamental equations for phase and chemical
equilibria. The basic quantities required for VLE calculations are; chemical reaction
equilibrium constants, Henry’s law constant, fugacity coefficients and activity coefficients
(Poplsteinova, 2004), phase equilibria calculation requires Henry’s law constant, fugacity
coefficients and activity coefficients. The differences in VLE models are mainly in the way
the phase non-idealities are treated. In the vapor phase non-ideality is either neglected or
represented by a fugacity coefficient calculated from one of the well-established equations of
state such as Soave-Redlich-Kwong equation of state, the models will then differ on the type
of equation of state that were applied. These differences, however are of minor importance
since the calculation of vapor phase fugacities are not crucial for the model performance. The
treatment of the liquid phase, on the other hand, is very important. Accurate prediction of the
liquid phase composition plays a key role in the VLE modeling. Since the evaluation of
equilibrium constants and fugacity coefficients are reasonably well established, the activity
coefficients were identified as the most important variables of the VLE model. Thus proper
representation of activity coefficients is desirable (Poplsteinova, 2004).

The system of interest in my research group is mainly CO2-H20-Alkanolamine. An activity


coefficient model for such system must be able to represent the non-ideality of an electrolyte
solution. Some activity coefficient models for non-electrolyte systems include, Wilson,
NRTL, UNIFAC, UNIQUAC. Some have been modified for electrolyte systems. Review of
electrolyte activity coefficient models were present by Maurer, 1983; Renon, 1986 and
Anderko et al., 2002. For the purpose of this course the universal quasi-chemical
(UNIQUAC) equations proposed by Abrams and Prausnitz, 1975 is used in calculating
activity coefficient for non-electrolyte systems. Further work will involve extending the
model for calculation activity coefficient for electrolyte solutions.

3
2.0 Thermodynamic Frame Work

Vapour liquid equilibria calculation requires simultaneous solution of phase and chemical
equilibria for reactive systems. Ths section shows briefly the mathematical framework for the
VLE calculations which is built around basic concepts of thermodynamic. Most of the
discussions here were taken from Elliot and Lira, 1999; Prausnitz et al., 1999; Kim, 2009;
Hartono, 2009 and Poplsteinova, 2004.

2.1 Thermodynamic Equilibrium Condition


Homogenous Closed system
A homogeneous is a system with a uniform properties ie. properties such as density is same
from point to point, in a macroscopic sense example is a phase. A closed system do not
exchange matter with the surrounding eventhough it may exchange energy (Prausnitz et al.,
1999). An equilibrium state is one with no tendency to depart spontenously (having in mind
certain permissible changes or process such as heat transfer, work of volume displacement
and for open systems mass transfer across phase boundry). It’s properties are independent of
time. A change in equilibrium state of a system is called a process and a reversible process is
one that maintains a state of virtual equilibrium throughout the process, it is often referred to
as one connecting a series of equilibrium states.

For a reversible process, the general condition for thermodynamic equilibrium is derived by
combination of first and second law of thermodynamics as:

dU  TdS  PdV 1

This condition could be expressed in terms of all extensive thermodynamic functions; internal
energy (U), enthalpy (H), Helmholtz energy (A) and Gibbs free energy (G): The condition for
equilibrium is usually maintained by keeping two of the thermodynamic variables constant.

Conditions for equilibrium is often expressed in terms of Gibbs free energy because the two
constant variables are often the temperature and pressure. Using the fundamental
thermodynamic relation for Gibbs free energy
G  U  TS  PV 2

4
Eq. 1 could be written as

dG   SdT  VdP 3

Thus at constant temperature and pressure

dG  0 4

which implies that for thermodynamic equilibrium at constant T and P, the Gibbs free energy
of a closed system reaches its minimum.

2.2 Partial Molar Gibbs Free Energy


Homogenous open System
For a closed system, U is considered a function of S and V only; that is U = U(S, V) but for
open system, there are additional indepedent variables. For these variables we use the mole
numbers of the various compnents present, thus we consider U as the function
U  U ( S , V , n1 , n2 ..., nm ) , where m is the number of components present. Total differential is
then

 U   U   U 
dU   
 S V ,ni
dS  
 V  S ,ni
dV    S 
i
dni 5
S ,V , n j

where subsript ni refers to all mole numbers and nj to all mole numbers other than ith (the
term under differenciation). Writing eq. 5 in the form

dU  TdS  PdV +   dn
i
i i 6

 
where i   U 
 ni  S ,V , n j
Eq. 6 is the fundamental equation for an open system corresponding to eq. 1 for a closed
system. The function μi is an intensive property which depends on temperature, pressure and
composition from its position in the equation as the coefficient of dni it can be refered to as
mass or chemical potential just as T (the coefficient of dS) is a thermal potential and P (the

5
coefficient of dV) is a mechanical potential. Similar expression can be derived in terms of
Gibbs free energy G (and other extentive thermodynamic properties).

dG   SdT  VdP +   dn
i
i i 7

where

 U   H   A   G 
i             8
 ni  S ,V , n j  ni  S , P , n j  ni T ,V ,n j  ni  P ,T ,n j

The quantity i is the partial molar Gibbs free energy, but not partial molar internal energy,
enthalpy, or Helmholtz energy, because the independent variables T and P chosen for defining
of partial molar quantities are also fundamental independent variables for the Gibbs free
 G 
energy G (Prausnitz et al., 1999). The quantity, i    , the partial molar Gibbs free
 ni  P ,T , n j

energy is also called the chemical potential.

2.3 Gibbs-Duhem Equation


At constant temperature and pressure eq. 7 reduces to

 
dG =   i dni  9
 i  P ,T

For equilibrium, at constant P and T, the Gibbs free energy is minimized (i.e dG = 0) (Elliot
and Lira, 1999). Also for a closed system dni = 0 thus equation 7 equal to zero at equilibrium.
From equation 9, equilibrium condition in terms of chemical potential could then be deduced

 
dG    i dni  = 0 10
 i  P ,T

This is called Gibbs-Duhem Equation. It is a thermodynamic consistency relation for a


heterogenous system that is used for experimental data evalution and theory development. It
expresses the fact that among  + 2 variables consisting of temperature, pressure and 

6
chemical potentials of each component present in the system. Only  +1 are independent
variables and the last variable is a dependent variable calculated in such a way that Gibbs-
Duhem equation is satisfied.
In terms of activity coefficient Gibbs-Duhem equation is expressed as

 x d ln 
i
i i 0 11

Equation 11 is a differential relation between the activity coeffiecients of all the components
in solution (Prausnitz et al., 1999). For a binary solution, the Gibbs-Duhem equation may be
written as:

d ln  1 d ln  2
x1  x2 12
dx1 dx2

2.4 Equilibrium in a Heterogeneous Closed System


A heterogeneous, closed system is made up of two or more phases with each phase considered
an open system within the overall closed system. If the system is in internal equilibrium with
respect to the three processes of heat transfer, boundry displacement and mass transfer,
neglecting special effects such as surface forces; semipermeable membranes; and electric,
magnetic or gravitational forces (Prausnitz et al., 1999), the general result for a closed system
consisting of  phases (where two of the phases are liquid (L) and vapour (V)) can be written
as

T v  T L  ...  T 
P v  P L  ...  P 13
 v   L  ...   

This is the basic criterion for phase equilibrium, which states that at equilibrium, the
temperature, pressure and chemical potentials of all species are uniform over the whole
system.

7
2.5 Chemical Potential
The task of phase-equilibrium thermodynamics is to describe quantitatively the distribution at
equilibrium of every component among all the phases present. Gibbs obtained the
thermodynamic solution to the phase-eqiulibrium problem by introducing the abstract concept
of chemical potential. The task is then to relate the abstract chemical potential of a substance
to physically measurable quantity such as temperature, pressure and composition.
For a pure substance i, the chemical potential is related to the temperature and pressure by the
differential equation

d i   si dT  vi dP 14

where si is the molar entropy and vi the molar volume. Integrating and solving for i at some
temperature T and pressure P, we have

T P
i (T , P)  i (T r , P r )   si dT   v dP
i 15
Tr Pr

where superscript r is an arbitrary reference state. The integrals can be solved from thermal
and volumetric data over temperature range T r to T and pressure range P r to P but the
chemical potential i (T r , P r ) is unknown. Chemical potential at T and P can thus only be

evaluated relative to an arbitrary reference states give as T r and P r these refence states are
often known as standard states.

Chemical potential does not have an immediate equivalent in the physical world and it is
desirable to express the chemical potential in terms of some auxilaiary function that might be
more easily identified with physical reality. The term fugacity (f) was introduced by G.N.
Lewis in trying to simplify the equation of chemical equilibrium by first considering the
chemical potential for a pure, ideal gas and then generalized the result to all systems
(Prausnitz et al., 1999). Auxiliary thermodynamic functions such as fugacities and activities
are often used in thermodynamic treatment of phase equilibria.

8
2.6 Fugacity and Fugacity Coefficient
The fugacity of component i in a mixture is defined as (Elliot and Lira, 1999)
RTd In fi  d i at constant T 16

where fi is the fugacity of component i in a mixture and i is the chemical potential of the
component. For a pure ideal gas, the fugacity is equal to the pressure, and for a component i in
a mixture of ideal gases, it is equal to its partial pressure yiP. The definition of fugacity is
completed by the limit:

fi
1 as P0 17
yi P

By integrating eq. 16 at constant T for any component in any system, solid, liquid or gas or
pure mixed, ideal or non-ideal. For vapour phase we have

f
   o  RT ln 18
fo

While either  o and fo is arbitrary, both may not be chosen independently; when one is
chosen, the other is fixed. Writing an analogous expression for the liquid and vapour phase
and equating the chemical potentials using equation 13 obtain:

fv
 v   L  RT ln  0 19
fL

This transformation consequently leads to additional criteria for equilibrium called


isofugacity:

f v  f L ...  f  20

This tell us that the equilibrium condition in terms of chemical potential can be replaced
without loss of generality by equation in terms of fugacity.

9
Fugacity coefficient,  is the ratio of fugacity to real gas pressure. It is a measure of non-
ideality.
fi
 i 21
yi P

It is a way of characterizing the Gibbs excess function at fixed T, P. For a mixture of ideal
gases i = 1.

2.7 Activity and Activity Coefficient


Activity concept is an alternative approach to express the chemical potential in a real solution.
The activity of component i at given temperature, pressure and composition is defined as the
ratio of the fugacity of i at these conditions to the fugacity i at standard state. Activity of a
substance gives an indication of how active a substance is relative to its standard state, it is
expressed as:

fi
ai  22
fi 0

Substituting equation 22 into 18 gives relationship between chemical potential and activity.

i  i0  RT ln ai 23

A general expression for the chemical potential in an ideal solution in terms of ideal mixing
could be written as

iid  i0  RT ln xi 24

Activity coefficient  i gives a measure of non-ideality of solution, it is the ratio of activity of


component i to its concentration, usually the mole fraction

ai
i  25
xi

10
In an ideal solution the activity is equal to the mole fraction and the activity coefficient is
equal to unity. Introducing equation (25) into (23)

i  i0  RT ln xi  RT ln  i 26

Activity coefficient relates chemical potential in an ideal solution to the chemical potential in
a real solution, thus representing a measure of non-ideality as illustrated by combining
equations (26) and (24)

i  iid  RT ln  i 27

2.7.1 Normalization of Activity Coefficient


It is convenient to define activity in such a way that for an ideal solution activity is qual to the
mole fraction or equivalently, that the activity coeffiicent is equal to unity. Because we have
two types of ideality (one leading to Raoult’s law and the other leading to Henry’s law), it
implies there will be two ways of normalizing activity coefficient.

Symmetric Convention
This convention applies when all the components both solutes and solvent at the system
temperature and pressure are liquids in their pure state (reference state). The activity
coefficient of each component i then approaches unity as its mole fraction approaches unity.
This convention leads to an ideal solution in the Raoult’s law sense. It follows that:

i  1 as xi  1 28

Unsymmetric Convention
This convention applies when pure component cannot be used as a reference state for instance
when some component are solid or gaseous at the system temperature and pressure. In this
case it is convenient to define the reference state as the infinite dilute state of the component
at system temperature and reference pressure. This convention therefore leads to an ideal
dilute solution in the sense of Henry’s law.

11
for solvent  s  1 as xs  1 29

for ionic and molecular solutes  i  1 as xi  0 30


Subscripts s and i refer to solvent and solute respectively while asterisk(*) shows that the
activity coefficient of the solute approaches unity as mole fraction approaches zero. This
convention is said to be unsymmetric because solvent and solute are not normalized in same
way.

2.8 Excess Functions


Excess functions are thermodynamic properties of a solution that are in excess of an ideal (or
ideal dilute) solution at the same conditions of temperature, pressure and composition. For an
ideal solution all excess properties are zero (Prausnitz, 1999). A general excess function is
defined as

e E  ereal  eideal 31

One particularly important excess function is the excess Gibbs energy (G E ) defined by

G E  G( actual solution at T , P and x )  G( ideal solution at the sameT , P and x ) 32

Similar definitions hold for excess volume V E , excess entropy S E , excess enthalpy H E ,
excess internal energy U E , and excess Helmholtz energy AE . Relations between these excess
functions are exactly the same as those between the total functions, for example:

G E  H E  TS E 33

Also, partial derivative of extensive excess function are analogous to those of the total
functions. For example:

 G E 
    SE 34
 T  P , x

12
Partial molar excess functions are defined in a manner similar to partial molar thermodynamic
properties. If M is an extensive thermodynamic property, then mi , the partial molar M of
component i, is defined by
 M  E  M E 
mi    similarly mi    35
 ni  P ,T ,n j  ni  P ,T , n j

From Euler’s theorem, we have that

n m n m
E
M  i i
similary ME  i i
36
i i

From excess function definition, it can be seen from eq. 27 that RT ln  i is equal to excess

chemical potential iE . Since chemical potential at constant T and P is equal to the partial
molar Gibbs free energy as shown in section 2.2, we obtain a very important relationship
between the activity coefficient and the partial molar excess Gibbs free energy:

E
g i  RT ln  i 37

Using equation (37) in (36) gives an equally important relation

g E  RT  xi ln  i 38
i

Eq. 38 forms the basis for calculation of activity coefficients from G E models as will
described in section 3.3.

3.0 Vapour Liquid Equilibrium Calculation

Vapour liquid equilibrium calculation requires simultanoeaus solution of chemical and phase
equilibria, activity and fugacity coefficients is required in chemical equilibria calculations
while fugacity cofficient is required in phase equilibria. Both are used to express chemical
potentials in liquid and vapour phase respectively

13
3.1 Chemical Equilibria
Molecular electrolytes dissociates or react in the liquid phase to produce ionic species to an
extent governed by the chemical equilibrium. Chemical reaction in the liquid phase enhances
the mass transfer rate and the solubility of CO2 and thus affects the phase equilibrium and
vice versa, the distribution of species between the two affects the chemical equilibrium in the
liquid phase. A system is in equilibrium when there is no driving force for a change of
intensive variables within the system. Chemical reaction moves towards a dynamic
equilibrium in which both reactants and products are present but have no further tendency to
undergo net change.
The generalized chemical reaction:

 1 A1   2 A2  ...   m Am   m 1 Am 1  ... 39

could be written as (Ott and Goates, 2000):

 A
i
i i 0 40

where the coefficients  i are positive for the products of the reaction and negative for the
reactants. The condition for equilibrium in a chemical reaction is given by:

 
i
i i 0 41

The Gibbs free energy change in the chemical reaction is given by:

Gr  Gro  RT ln  aivi 42


i

where Gro is the Gibbs free energy change with reactants in their standard states.

14
3.1.1 Chemical Equilibrium Constant
The chemical equilibrium is traditionally defined by a chemical equilibrium constant. At
equilibrium, Gr = 0 so that equation eq. 42 becomes, (Ott and Goates, 2000):

Gro   RT ln K 43
where K is the equilibrium constant and is given by:
K   aivi 44
i

The activities in eq. (44) are now the equilibrium activities. An equilibrium constant K
expressed in terms of activities (or fugacities) is called a thermodynamic equilibrium constant.
Equilibrium constants are required for each of the reactions occuring in solutions. They are
related to the activities of each species as:

bB  cC  dD  eE 45

aDd aEe
K  46
aBb aCc

Here the lower case letters are the stoichiometric coefficients,  i , and the capital letters are
labels for the chemical species.

3.2 Phase Equilibria


The solubility of gas in a liquid is often proportional to its partial pressure in the gas phase,
provided that the partial pressure is not large. The equation that describes this is known as
Henry’s law:

Pi  yi P  Hxi 47

where Henry’s constant (H) is the constant of proportionality for any given solute and solvent,
depending only on temperature (Prausnitz et al., 1999). At high partial pressures, Henry’s
constant must be multiplied by activity coefficient and pressure by fugacity cofficient.

15
A measure of how chemical species distributes itself between liquid and vapour phase, is the
ratio (Perry, 1997):

yi  i f i
K  48
xi i P
The condition for phase equilibrium in a closed heterogeneous system at constant temperature
and pressure is given by eq. 20.

fi v  fi L 49

Where fi v and f i L are fugacities of component i in the vapor and liquid phase respectrively.
From eq. 48 or by inserting eqs. 21, 22, and 25 into eq. 49 we obtain:

i yi P   i xi fi oL 50

where i and  i are fugacity and activity coefficient respectively. f i oL is reference state
fugacity coefficient defined either by symmertic conventions Raoult’s law or unsymmetric
convention Henry’s law.

A correction term used to relate fugacities at the different pressures is called Poynting factor,
written as and at constant temperature and pressure change from P1 to P2 can further be
shown to be:

f 2 (T , P2 )  p2 v dP 
   exp   i  51
f1 (T , P1 )  P RT 
 1 

The complete phase equilibria for vapour – liquid equilibrium of solute and solvent at systems
temperature and pressure can thus be respectively written as (Austgen 1989):

 vi  ( P  PHo2O ) 
For solute: i yi P   i xi H i exp   52
 RT 

16
 vi ( P  Pso ) 
For solvent:  s ys P   x P  exp 
o o
s s s
o
s  53
 RT 

where Hi and v i represent Henry’s law constant, partial molar volume of molecular solute i
at infinite dilution in pure water at the system temperature and at saturation water vapour
pressure, PHo2O while vs is the molar volume of pure solvent at system temperature and

saturation pressure. The exponential correction (Poyning factor) here takes into account the
fact that liquid is at a pressure P different from the saturation pressure P o . For molecular
solutes such as carbon dioxide, Henry’s constant represents reference-state fugacities.

To solve the phase equilibrium equation, we need to evaluate the fugacity and activity
coefficients. Gas phase fugacity coefficient can be calculated from equation of state such as
Soave-Redlich-Kwong equation of state. The real modeling task lies in the calculation of
activity coefficient. In this work activity coefficient calculation using, UNIQUAC is
presented.

3.3 GE Model for Activity Coefficient

Non-ideality in liquid phase is represented by activity coefficient as described in section 2.8.


They are usually obtained from excess Gibbs free energy models using eq.38. An apprioprate
excess Gibbs energy function most take into consideration the molecular interactions between
all species in the system. For electrolyte solutions, diverse species are usually present and
interactions among them must be represented. At high concentrations, interactions between
neutral molecules or between ions and neutral molecules are very short-range in character and
dominates while at low concentrations it is interactions between ions which are very strong
long-range electrostatic interactions that dominates. The usually practice is to assume that the
contributions of the various types of interactions are independent and additive (Poplsteinova,
2004). The excess Gibbs energy is then calculated as the sum of short-range and long-range
contributions:

G E  GSR
E
 GLR
E
54

17
Most modeling applications combine the Debye-Hückel electrostatic theory for the long-range
term with modifications of well known non-electrolytes models for the short-range term. In
this work the local composition model for non-electrolytes, UNIQUAC will be presented in
more detail since they were applied in this work.

3.3.1 Local Composition Models


Regular solution theory assumes that the mixture of interactions were independent of each
other such that quadratic mixing rules provide reasonable approximations. However in some
cases, the mixture interaction can be strongly coupled to mixture composition. One way of
treating this is to recognize the possibility that the local compositions in the mixture might
deviate strongly from the bulk composition (Elliot and Lira, 1999). Some of the well-known
local composition models for non-electrolytes are Wilson, NRTL, UNIFAC and UNIQUAC
for these models to be used in electrolyte solutions, several assumptions have to be made
regarding the local composition in the presence of ions.

4.0 UNIQUAC

For this work the UNIQUAC equation described by Abrams and Prausnitz, 1975 was derived,
implemented and used for calculation of activity coefficients for non-electrolyte systems.
Main advantages of UNIQUAC is that it uses only two adjustable parameters per binary to
obtain reliable estimates for both vapor-liquid and liquid-liquid equilibria for a large variety
of multicomponent systems using the same equation for the excess Gibbs energy (Abrams and
Prausnitz, 1975).

Experimental data for typical binary mixtures are usually not sufficiently plentiful or precise
to yield three meaningful binary parameters, various attempts were made (Abrams and
Prausnitz, 1975; Maurer and Prausnitz, 1978; Anderson, 1978; Kemeny and Rasmussen,
1981) to derive a two-parameter equation for g E that retains at least some advantages of
Wilson equation without restriction to completely miscible mixtures. Abrams derived an
equation that in a sense, extends the quasichemical theory of Guggenheim for
nonrandommixtures to solutions containing molecules of different size. This is called
universal quasi-chemical (UNIQUAC) theory.

18
UNIQUAC equation for g E consists of two parts combinatorial part that describe the
dominant entropic contribution, and residual part that accounts for intermolecular forces
which are responsible for the enthalpy of mixing. The combinatorial part is determined only
by composition and by the sizes and shapes of the molecules, it requires only pure-component
data. The residual part depends on intermolecular forces, thus the two adjustable binary
parameters appear only in the residual part. The UNIQUAC equation is

gE  gE   gE 
      55
RT  RT  combinatorial  RT  residual

For a binary mixture

 gE    z   
 RT   x1 ln 1  x2 ln 2   x1q1 ln 1  x2 q2 ln 2  56
  combinatorial x1 x2 2  1 2 

 gE 
    x1q1 ln(1   2 21 )  x2 q2 ln( 2  1 12 ) 57
 RT  residual

where coordination number z is set equal to 10. Segment fraction,  , and the area fractions,
 , are given by

x1r1 x2 r2
1  2  58
x1r1  x2 r2 x1r1  x2 r2

x1q1 x2 q2
1  2  59
x1q1  x2 q2 x1q1  x2 q2

The parameters r and q are the two pure-component structural parameter per component
representing volume and area respectively. They are dimensionless and are evaluated from
bond angles and bond distances. For a binary mixture, there are two adjustable parameters,
 12 and  21 . They are given in terms of charateristic energies u12  u12  u22 ; u21  u21  u11 .

 u12   u21 
 12  exp    21  exp   60
 RT   RT 

19
For many cases eq. 60 gives the primary effect of temperature on  12 and  21 . u12 and u21 are

often weakly dependent on temperature.

Writing eq. 38 in terms of mole number ni:

nT g E  RT  ni ln  i 61
i

where nT is the total number of moles. The final expression for activity coefficient is obtained

by taking the partial derivative of excess Gibbs energy g E with respect to mole number.

 n g E 
RT ln  i   T  62
 ni T , P , n j ( j ì )

For a multicomponent system, activity coefficient expression based on UNIQUAC is then


derived to be:

k z   k k
k 
ln  k  ln 1 
qk  ln   1
xk xk 2  k  k 
  63
   i ki 
 qk 1  ln    j  jk    
 j  i   j ji
 j


Detailed derivation of this UNIQUAC expression for activity coefficient is shown in


Appendix

4.1 UNIQUAC Implementation


The expression for activity coefficient was added to the thermodynamic function library
developed by Tore Haug-Warberg and Bjørn Tore Løvfall at Department of Chemical
Engineering, NTNU. The model was coded in an in-house language called RGrad supporting
automatic gradient calculations, and from there exported to C-code which is compiled into a
set of DLL’s accessible from Matlab, Octave and Ruby (Haug-Warberg, 2008).

20
It has a folder <MODEL>/src/mex that contains two makefiles which compiles and set up the
model called <model>. The model is then simply run as <model>_mexmake or
<model>_octmake depending on whether you want a matlab/MEX interface or an Octave
interface.

4.1.1 MEA-H2O System


Monoethanolamine(MEA) is an important solvent for CO2 absorption, so for this work,
UNIQUAC equation was applied in calculation of activity coefficient in binary system of
MEA and water. In the model, the parameter adjustable energy interaction parameters (uij ) of

the UNIQUAC enthalpy term are assumed to be temperature dependent and are fitted to the
following temperature function

uij  uij0  uijT (T  298.15) 64

The r and q parameters for MEA and water as well as uijo and uijT were taken from Faramarzi

et al. 2009. Other uijo parameter values were set to large values while uijT parameters were set to

zero. Result from the calculation was compared with literauture data of Tochigi, K et al.,
1999, Belabbaci, A et al. 2009 and Kim, I et al., 2008 in figure 1. Molar Excess Gibbs energy
function was also calculated and presented in figure 5.

4.1.2 Acetic Acid-H20 System


The r and q parameters as well as energy interaction parameter aij for acetic acid-water

system were taken from Prausnitz, et al. 1999. The activity coefficient plot for acetic acid-
water system is shown in figure 2. The molar excess Gibbs energy function was also
calculated and presented in figure 6.

4.1.3 Other Systems


The predicted model for formic acid – acetic acid sysems as well as acetone – chloroform
syetems are presented in figure 3 and 4 respectively. All parameters were taken from Prautnitz
et al. 1999. It has not been easy to get experimental data to use to compare the model result
from this work, however the shapes of the plots agrees very well with plots in Prausnitz, et al.
1999.

21
5.0 Results

0.9

0.8

0.7

0.6

0.5 Tochigi K. et al. 1999


Belabbaci A. et al. 2009
Kim I. et al. 2008
0.4 H20 This work
MEA This work

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


xMEA

Figure 1: Activity coefficient plot for MEA-H20 system

0.6
Ellis & Bahari 1956
Sebastiani & Lacquaniti 1967
0.5
Hansen et al. 1955
Arich & Tagliavini 1958
0.4 Marek & Standart 1954
H20 This work
ACETIC This work
0.3
log

0.2

0.1

-0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
xH20

22
Figure 2: Activity coefficient plot for Acetic Acid – H20 system.

1.35
ACETIC
FORMIC
1.3

1.25

1.2

1.15

1.1

1.05

1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
xFORMIC

Figure 3: Activity Coefficient plot for Acetic Acid – Formic Acid System

0.9
CHLOROFORM
ACETONE
0.8

0.7

0.6

0.5

0.4

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


xACETONE

Figure 4: Activity Coefficient plot for Acetone -Chloroform System

23
0

-100

-200
Excess Gibbs Energy (gE)

-300

-400

-500

-600

-700
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
xMEA

Figure 5: Excess molar Gibbs free energy plot for MEA-H20 System

250

200
Excess Gibbs Energy gE

150

100

50

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
xACETIC

Figure 6: Excess molar Gibbs free energy plot for Acetic Acid -H20 System

24
Summary

UNIQUAC activity coefficient model was derived and successfully implemented for
calculating activity coefficient for different non-electrolyte systems. Literature activity
coefficient data for MEA-H20 system appear to vary among the different authors presented,
the model appear to fit much better to the data from Tochigi K et al 1999. Activity coefficient
data for acetic acid – H20 system however is more consistent. Model prediction agrees well
with the literature values.

Future Work

Further work will involve the calculation of activity coefficient for electrolyte systems by
extending the UNIQUAC model for electrolyte system through addition of the long-range
term.

References

Abrams, Denis S.; Prausnitz, John M., 1975; Statistical thermodynamics of liquid
mixtures. New expression for the excess Gibbs energy of partly or completely miscible
systems. AIChE Journal, 21(1), 116-28.
Anderko, Andrzej; Wang, Peiming; Rafal, Marshall., 2002; Electrolyte solutions : from
thermodynamic and transport property models to the simulation of industrial processes.
Fluid Phase Equilibria, 194-197 123-142.
Anderson, T. F.; Prausnitz, J. M., 1978; Application of the UNIQUAC equation to
calculation of multicomponent phase equilibriums. 2. Liquid-liquid equilibriums.
Industrial & Engineering Chemistry Process Design and Development , 17(4), 561-7.
Arich, Guido; Tagliavini, Giuseppe., 1958; Liquid-vapor equilibrium isotherms for the
water-acetic acid system. Ricerca sci., 28 2493-500.
Belabbaci, Aouicha; Razzouk, Antonio; Mokbel, Ilham; Jose, Jacques; Negadi, Latifa., 2009;
Isothermal Vapor - Liquid Equilibria of ( Monoethanolamine + Water ) and (4-
Methylmorpholine + Water ) Binary Systems at Several Temperatures. Journal of Chemical
& Engineering Data ACS ASAP.
Elliot, J.R.; Lira, C.T., 1999; Introductory Chemical Engineering Thermodynamics; Prentice
Hall PTR: New Jersey.
Ellis, S. R. M.; Bahari, E. P., 1956 Vapor-liquid equilibrium at low concentrations; acetic
acid-water; nitric acid-water. British Chemical Engineering , 1 210-11.

25
Faramarzi, Leila; Kontogeorgis, Georgios M.; Thomsen, Kaj; Stenby, Erling H., 2009;
Extended UNIQUAC model for thermodynamic modeling of CO2 absorption in aqueous
alkanolamine solutions. Fluid Phase Equilibria (2009), 282(2), 121-132.
Hansen, Robert S.; Miller, Frederick A.; Christian, Sherril D., 1955; Activity coefficients of
components in the systems water-acetic acid, water-propionic acid, and water-butyric acid at
25 Journal of Physical Chemistry, 59 391-5.
Hartono, Ardi; 2009, Characterization of diethylenetriamine(DETA) as absorbent for carbon
dioxide. Ph.D thesis., Chemical Engineering Department, Norwegian University of Science
and Technology, Trondheim, Norway.
Kemeny, Sandor; Rasmussen, Peter., 1981; A derivation of local composition expressions
from partition functions. Fluid Phase Equilibria , 7(2), 197-203.
Kim, Inna; 2009, Heat of reaction and VLE of post combustion CO2 absorbents. Ph.D thesis.,
Chemical Engineering Department, Norwegian University of Science and Technology,
Trondheim, Norway.
Marek, Jan; Standart, George., 1954; Effect of association on liquid-vapor equilibria. I.
Equilibrium relations for systems involving an associating component. Collection of
Czechoslovak Chemical Communications, 19 1074-84.
Maurer, G., 1983; Electrolyte solutions. Fluid Phase Equilibria (1983), 13 269-96.
Maurer, G.; Prausnitz, J. M., 1978; On the derivation and extension of the UNIQUAC
equation. Fluid Phase Equilibria, 2(2), 91-9.
Ott, J. B., Boerio-Goates J., 2000: Chemical Thermodynamics: Advanced Applications.
Academic Press
Perry, R.H., Green, D.W.,1997; Perry’s Chemical Engineering Handbook (7th edition),
McGraw-Hill.
Poplsteinova, Jana; 2004, Absorption of carbon dioxide – modelling and experimental
characterization. Ph.D thesis., Chemical Engineering Department, Norwegian University of
Science and Technology, Trondheim, Norway.
Prausnitz, J. M., Lichtenthaler, R.N., and de Azevedo, E. G., 1999; Molecular
Thermodynamics of Fluid-Phase Equilibria (3rd ed.), Prentice Hall PTR; New Jersey.
Renon, Henri., 1986; Electrolyte solutions. Fluid Phase Equilibria , 30 181-95.
Sebastiani, Enzo; Lacquaniti, L. 1967; Acetic acid - water system thermodynamic
correlation of vapor-liquid equilibrium data. Chemical Engineering Science, 22(9), 1155-
62.
Thomsen, Kaj; Rasmussen, Peter., 1999; Modeling of vapor-liquid-solid equilibrium in gas-
aqueous electrolyte systems. Chemical Engineering Science , 54(12), 1787-1802.
Tochigi, Katsumi; Akimoto, Kentarou; Ochi, Kenji; Liu, Fangyhi; Kawase, Yasuhito., 1999;
Isothermal Vapor - Liquid Equilibria for Water + 2 - Aminoethanol + Dimethyl
Sulfoxide and Its Constituent Three Binary Systems. Journal of Chemical and Engineering
Data, 44(3), 588-590.
Tore Haug-Warberg, 2008; KP8108 Advanced Thermodynamics: With applications to Phase
and Reaction Equilibria, Lecture Handout, Department of Chemical Engineering, NTNU.

26
Derivation of Activity Coecient Expression
Based on UNIQUAC
Ugochukwu Aronu

July 13, 2009

UNIQUAC Expression for Excess Gibbs Energy

Combinatorial
z }| {z Residual
}| X {
E
g X i ZX i X
= xi ln + qi xi ln qi xi ln( j ji ) (1)
RT i
xi 2
i
i i j
hu i
where ji = exp ji uii
RT

q xi q ni
i = Pi = Pi (2)
j qj xj j qj nj

r xi r ni
i = Pi = Pi (3)
j rj xj j rj nj

writing eq.1 in terms of the number of moles 'n' and factoring out 'n'.
A B
z }| { z X }| { z C
}| {
ng E X i n Z i X X
= ni ln + qi ni ln qi ni ln( j ji ) (4)
RT i
ni 2
i
i i j

gE
!      
@n( RT ) @A @B @C
ln k = = + +
@nk @nk T;ni6=k @nk T;ni6=k @nk T;ni6=k
T;ni6=k
(5)

1
Simplify Eq. 4A and dierentiate with respect to mole number.
X  
i n
A= ni ln
i
ni
X
= ni (ln i n ln ni )
i
X
= ni (ln i + ln n ln ni )
i
A1
zX }| { z A}|2 {
= ni (ln i ln ni ) + n ln n
i

@A @A1 @A2
= + (6)
@nk @nk @nk

P
@A1 @[ i ni (ln i ln ni )]
=
@nk @nk

Using kronecker-delta to dierenciate


@A1 X @i 1 1

= (ln i ln ni )ik + ni ( ik )
@nk i
i @nk ni
  X ni @i
@A1 k
= ln + 1 (7)
@nk nk i
i @nk

@A2 @ (n ln n)
=
@nk @nk
@n @ ln n
= ln n +n
@nk @nk
1
= ln n + n
n
@A2
= ln n + 1 (8)
@nk
Substituting eq. 7 and eq. 8 back in eq. 6
  X ni @i
@A k
= ln + ln n +
@nk nk i
i @nk
  X ni @i
@A k
= ln + (9)
@nk xk i
i @nk

2
From 3, i = Prjirnjinj ; thus
P  P
@i j rj nj ri ik (ri ni ) j rj jk
@nk
= P 2
j rj nj

@i r ik ri ni rk
= Pi P 2 (10)
@nk j rj nj
j rj nj

Put eq. 10 into eq. 9


2 3
  X ni 6 ri ik
@A k ri ni rk 7
@nk
= ln
xk
+

4P r n P 2 5
i j j
i j
j rj nj
P P
k nr i ri ni rk
2
= ln + Pi i i P
xk i j rj nj i ( j rj nj )2
P
k n rk i ri ni rk
2
= ln + Pk P
xk k j rj nj i ( j
2 rj nj )

Factoring out k and i and simplifying


P
k nr
= ln +1 Pi i k
xk rj nj
j
k nr
= ln +1 P k
xk j rj nj

Multiplying numerator and denominator by nk and factor out k


k nrk nk
= ln +1 P :
xk j rj nj nk
 
@A k k
= ln +1 (11)
@nk T;ni6=k xk xk

3
Simplify Eq. 4B and dierentiate with respect to mole number.
zX i zX
B = qi ni ln = qi ni (ln i ln i )
2
i
i 2
i

 
@B zX @i 1 1 @i
= qi (ln i ln i )ik + ni ( )
@nk 2
i
i @nk i @nk

A B
z }| { z  }| {
@B z  z X ni @i ni @i
= qk ln k + qi (12)
@nk 2 i 2
i
i @nk i @nk

From 10 @i
@nk
= Prjirikj nj (
Prjinrijrnkj 2 ; therefore
)

ni @i ni ri ik rnr


= P Pi i k 2
i @nk i j rj nj i ( j rj nj )
nr
= ik Pi k (13)
( j rj nj )

Similarly;
P
@i ( qj nj )qi ik qi ni qk
=
j
P
@nk ( j qj nj )2
q qnq
= P i ik Pi i k 2
j qj nj ( j qj nj )
ni @i ni qi ik ni qi ni qk
= P P
i @nk i j qj nj (i j qj nj )2
ni qk
= ik P (14)
( j qj nj )

Substitute eq. 13 and eq. 14 into eq. 12B


" ! !#
zX ni qk nr
qi ik P ik Pi k
2
i j qj nj j rj nj

!
zX nr nq
= qi Pi k Pi k
2
i j rj nj j qj nj
P P !
z qnr qi ni qk
= Pi i i k Pi
2 j rj nj j qj nj

4
Multiply numerator and denominator of by nk then factor out k and i .
P P !
z qi ni rk nk qi ni qk
= Pi Pi
2 j rj nj nk j qj nj
Factor out k and i .
P !
z i qi ni k
X
= i qk
2 nk i
P !
z i qi ni k
X
= qk i
2 qk nk i
P
Factoring out k and setting i i = 1 .
 
z k
= qk 1
2 k
Therefore
@B z k z k
= qk ln + qk ( 1)
@nk 2 k 2 k
   
@B z k k
= qk ln + 1 (15)
@nk T;ni6=k 2 k k

Dierenciate Eq. 4C with respect to mole number.

X X
C = qi ni ln( j ji )
i j
2 0 1 3
@C X X @ni 1 X @j 5
= qi 4ln @ j ji A + ni P ji
@nk i j
@nk j j ji j
@nk
2 B
3
z 0
A
}| {
1 z  }| {
6 P @j 7
6 X X n j ji @nk 7
7
6q ln @   A + q i
P (16)
= 6k j jk i 7
6  j ji 7
4 j i j 5

But; j = Pqlj qnljnl . Note that subscript ' ' is used in place of subscript ' ' for
l i

clearity. P
@j ( l ql nl )qj jk
P
qj nj qk
=
@nk ( l ql nl )2

5
@j q jk q nj qk
= Pj Pj (17)
@nk l ql nl ( l ql nl )
2

Substituting eq. 17 in eq. 16B


P h q i
X j ji Pjl qjklnl (
Pqjlnqjlnqkl 2
)
qi ni P
i j j ji
h Pj qj nj qk ji i
X P
qi ni ki qk
l ql nl
qi ni
Pl qlnl 2
) P
( )

i j j ji

Factor out i then substitute and simplify further;


P
X i ki X i j ji j
) qk P qk P
j ji j ji
i j i j

Simplify further and substitute back into Eq.16


2 3
@C X X i ki X
= 4qk ln( j jk ) + qk P qk i 5
@nk j i j j ji i

P
But i i = 1
2 3
  X X i ki
@C
= qk 41 ln( j jk ) P 5 (18)
@nk T;ni6=k j i j j ji

Substituting eq. 11, eq. 15 and eq. 18 into eq. 5 and obtain the nal
expression for activity coecient based on UNIQUAC
k k
ln k = ln +1
xk xk
 
z  
+ qk ln k + k 1
2 k k
2 3
X X i ki
+ qk 41 ln( j jk ) P 5
j i j j ji

Вам также может понравиться