Вы находитесь на странице: 1из 12

See

discussions, stats, and author profiles for this publication at:


http://www.researchgate.net/publication/229265877

Multicomponent fugacity coefficients and


residual properties from pressure-explicit
equations of state
ARTICLE in CHEMICAL ENGINEERING SCIENCE JANUARY 1991
Impact Factor: 2.34 DOI: 10.1016/0009-2509(91)80161-Q

CITATION

READS

453

1 AUTHOR:
Claudio Olivera-Fuentes
Simon Bolvar University
60 PUBLICATIONS 205 CITATIONS
SEE PROFILE

Available from: Claudio Olivera-Fuentes


Retrieved on: 22 October 2015

Chemical Engineering Science,


Printed in Great Britain.

Vol. 46. No. 8, pp. 2019-2029,

199k.

CMKE-2509/91
$3.00 + 0.00
0 1991 Pergamon Press plc

MULTICOMPONENT
FUGACITY
COEFFICIENTS
AND
RESIDUAL
PROPERTIES
FROM PRESSURE-EXPLICIT
EQUATIONS
OF STATE
CLAUDIO
Departamento

de TermodinHmica

(First receiued

OLIVERA-FUENTES

y Fen6menos
de Transferencia,
Universidad
Postal 89000, Caracas 1086-A, Venezuela

Simbn

Bolivar,

19 April 1990, accepted for publication in revisedform 28 November

Apartado

1990)

Abstract-A
procedure presented by Szarawara and Gawdzik (1989) for development
of expressions for the
fugacity coefficients of components
of a fluid mixture from cubic equations of state (EOS) is extended to the
generation
of similar formulas for the residual enthalpy and entropy, and is shown to relate to a more
fundamental
Helmholtz energy approach. The method is not limited to cubic EOS, but applies in general to
any pressure-explicit
EOS with parameters
dependent
on temperature
and composition.
A general stepThe method is illustrated with three examples: the
by-step
application
sequence
is presented.
Benedict-Webb-Rubin
and Benedict-Webb-Rubin-Starling
EOS, the genera1 class of perturbed hard-core
models (with the Carnahan-Starling-van
der Waals EOS as a particular
case), and a generalized
fourparameter cubic EOS of the Schmidt-Wenzel
type. The existence of a common functional form for all the
residual properties
from each of these EOS is demonstrated.
The method is highly flexible, making it
straightforward
to insert different temperature
functions and mixing rules into a given EOS model, or to
explore new combinations
of volume functions taken from different models.

cubic EOS, mixing and combination


rules, and temperature functions that have appeared in the literature
or might be proposed in the future.
Szarawara and Gawdzik (1989) considered
a general two-parameter,
pressure-explicit
EOS having the
functional form

INTRODUCTION

Two recent papers have independently


dealt with the
generation
of thermodynamic
property
expressions
for multicomponent
systems from cubic equations of
state (EOS). Miiller et al. (1989) considered a generic
four-parameter
cubic EOS of the Schmidt and Wenzel
(1980) type

p=--

RT

u-b

v2 + ubv + wb'

p = $(T, u, 0, b)

(1)

with a, b, u and w dependent


on temperature
and
composition.
Working
in the {T, V, N,, . . , N,}
representation,
they used the thermodynamic
relation

to the general

to develop
an expression
for the total residual
Helmholtz energy A, from which the fugacity coefficients and the molar residual entropy and enthalpy of
the multicomponent
mixture were then obtained via
the thermodynamic
relations

(6)

on composition.
Confining
with a and b dependent
themselves to the computation
of fugacity coefficients,
they used a transformation
originally suggested by
Null (1970) to proceed from the standard thermodynamic formula

where .# = Aj(NRT) is a dimensionless molar residual Helmholtz


function. Extensive tables were presented by these authors for the composition
and temperature derivatives
of the EOS parameters,
a, b, u,
w that must be substituted
into these equations
in
order to make the results applicable to most of the
2019

result

In eq. (8) as written by these authors, the thermodynamic properties to be held constant in each partial
derivative are not indicated explicitly. The context of
their paper however shows clearly that

while the derivatives

of a and b with respect to xj

CLAUDIO

2020

OLIVER.&FUENTES

should be taken with temperature


and all the remaining mole fractions held constant, as will be discussed
later. It may also be pointed out that since the lower
limit for the second integral in eq. (8) is the ideal gas
volume, v* = RT/P, this integral can be rewritten as

in which form it is probably closer to usual practice.


While Szarawara and Gawdzik justified their work
mainly in terms of cubic EOS, presenting
explicit
results for five cubic EOS using van der Waals mixing
rules without binary interaction parameters, it is clear
that eq. (8) is equally applicable
to any pressureexplicit EOS, since the difference between cubic and
noncubic EOS will become important
only when the
integrals of $ and its derivatives
are evaluated. As
pointed out by these authors, extension of eq. (8) to
any number of EOS parameters should be straightforward.
It appears that the two papers referred to above,
while overlapping
to some extent, are largely complementary,
with the work of Miiller et nl. dealing
more comprehensively
with cubic EOS and covering
derived properties other than the fugacity coefficients,
and the Szarawara and Gawdzik results extending to
nuncubic EOS with an arbitrary number of adjustable
parameters.
It is the purpose of the present work to
unify both treatments,
by extending the second approach to other residual properties,
and showing its
application
to both cubic and noncubic EOS. To this
end, a general pressure-explicit
EOS is considered,
having the mathematical
form
P = $(T, 0, el,

. , em)

(9)

where e,, . . , e,,,are the EOS parameters


in general
depending on both temperature
and composition
(but
riot on pressure or density),
ek=e,(T,x,

,...,

x,)

k=l,...,

m.

(10)

The maximum number of such parameters


is five for
a cubic EOS (Abbott,
1979; Kumar and Starling,
1980) but may be much higher for noncubic EOS, e.g.
the modified virial expansions
of the BWR form.
Following Szarawara
and Gawdzik (1989), an attempt is made in eq. (9) to differentiate
between the
physical property P and its functional dependence
$.
Because of the variety of independent
variable sets
possible
in thermodynamics,
a partial derivative
which does not specify the variables held constant is
of course meaningless. Equation (9) on the other hand,
sets down a specific functional form to be taken in the
present work as synonymous
with pressure-explicit
EOS, which consists of an algebraic expression
involving temperature,
molar volume and a set of EOS
parameters.
Partial derivatives
of $ will be written
only with respect to one of these variables, constancy
of all the remaining
variables being implicitly assumed. Thus, while unannotated
thermodynamic
derivatives such as aP/& and aP/dT are ambiguous,

the derivatives

should cause no confusion. This simpler notation will


result in more compact expressions
without in any
way limiting the generality of the analysis. A similar
distinction need not be made in eq. (lo), as the ensuing simplifications
would be only minor.
RESIDUAL

PROPERTIES

AND FLJGACITY

COEFFICIENTS

The residual entropy, or entropy departure of the


multicomponent
fluid mixture from the ideal gas at
the same volume, temperature
and composition
may
be computed from the thermodynamic
relation
sr=JI[($),X-;]dL..
Equations

(9) and (10) give the partial

(11)
derivative

as

(12)
1

hence a general result follows in dimensionless

form as

The corresponding
expression
in the {T, P,
is related to the above by
x1,
,x,} representation
(Abbott and Nass, 1986)

SR
s
_=_
R

+ In Z.

(14)

Since the ideal gas enthalpy depends only on temperature, the residual enthalpy in either representation may be obtained from the thermodynamic
relation

(15)
where tht: first two terms on the right-hand side represent the residual compressibility
and the integral
corresponds
to the residual internal energy. Substitution of eqs (9) and (12) in eq. (15) gives the general
dimensionless
result

The fugacity coefficient of the multicomponent


mixture considered
as a single fluid is related to the
residual entropy and enthalpy
by (Van Ness and

fugacity coefficients and residual properties

Multicomponent
Abbott,

1982)
(17)

Substitution
the general

of eqs (13), (14) and (15) in eq. (17) yields


expression

eq. (7) may be used as a starting point for


of an expression
for the species
fugacity coefficients,
as done by Szarawara
and
Gawdzik (1989), it seems more direct to base the
derivation on an alternative thermodynamic relation
that already contains volume (rather than pressure) as
the independent
variable (Model1 and Reid, 1983),

2021

As explained by these authors, the derivatives with


respect to a mole fraction can be taken with all other
mole fractions kept constant, a mathematically
sound
and conveniently
symmetrical
operation,
even if it
corresponds
to a physically impossible process. Equations (18), (24) and (25) constitute the generalization
of
the Szarawara and Gawdzik formula, eq. (8).
Since the partial molar values of the EOS parameters must by definition satisfy the summability relation [easily proved from eq. (25)],

Although

the development

it is seen that eq. (24) for the individual fugacity coefficients complies with the basic thermodynamic
requirement (Van Ness and Abbott, 1982)
ln$=ixiln&.

(27)

With the molar


partial derivative

volume

II = V/N, eq. (9) yields the

cm
On substituting

eq. (20) in eq. (19), recognizing

0 v-dv=@v-RTa9

oD

and inserting
as

that
(21)

au

eq. (18), the fugacity

coefficient

follows

Apart from providing a check on the consistency


of
the development,
this result shows that the expressions developed for the residual entropy and enthalpy
and species fugacity coefficients, eqs (13), (16) and (24)
are not independent,
but are connected
through
eq. (17). The formulation
of separate relations
for
these properties is however a matter of practical convenience, as it is quite likely that only one of them will
be required for a given process application
(i.e., enthalpies for energy balance, entropies for second-law
analysis, fugacity coefficients
for phase equilibrium
prediction),
and it would be computationally
inefficient to have their values depend on the previous
calculation of the other properties. A possible exception is the mixture fugacity coefficient, which may be
a useful intermediate
quantity in enthalpy or fugacity
calculations.
GENERAL APPLICATION PROCEDURE

This result clearly reduces to the correct pure component formula when the EOS parameters
are independent of composition.
The composition
derivatives
of the EOS parameters appearing in eq. (22) are possibly best obtained
in terms of their partial molar values. Since the et
are intensive properties, the products Nek define the
corresponding
extensive variables with partial molar
values

The derivation
of entropy, enthalpy
expressions for a given pressure-explicit
performed in the following sequence:

1. Write the EOS in the form of eq. (9), making the


volume dependence
completely
explicit. Density-dependent mixing rules, for example, must be broken

down into the composition functions intended to apply separately in the low-density and high-density
regions,

r, N,,

T,N,,,

and fugacity
EOS may be

i.e.

(23) &(T, 0, XI,

. . . ,X,) = e;(T, XI, . . . ,&)


+ E(u) eL( T, x1, .

in terms of which eq. (22) becomes


2. Evaluate

, xn)-

the functions

j~($-~jdn=Q(T,,e,

,...,

e,)

(28)

Moreover,
the general treatment
presented by Van
Ness and Abbott (1982) for composition
derivatives
can be applied to eq. (10) to give

_,

!!k&=??

s _aell

ae,

k=l

. .

(30)

CLAUDIO OLIVERA-FUENTES

2022

which depend only on the EOS model, eq. (9), and not
on the influence of temperature
and composition
on
the EOS parameters, eq. (10). Depending on the form
and complexity of eq. (9), calculation of eqs (29) and
(30) may be easier as derivatives of 0 or as integrals of
$_ The relation of 0 and its derivatives to the fundamental Helmholtz
equation is demonstrated
in the
Appendix.
3. Obtain general expressions
for the EOS model.
Combine:

(a) eqs (13), (14), (29) and (30) for the residual
entropy, sR;
(b) eqs (9), (18) and (28) for the mixture fugacity
coefficient, In 4;
(c) eq. (17) and the results of (a) and (b) for the
residual enthalpy, hR;
(d) eqs (24), (30) and the result of (b) for the
fugacity cnefficients, In 4;.
As pointed out abov.e, these are generic results that
will be valid irrespective
of the particular
form
adopted for eq. (10).
4. For a given form of eq. (lo), compute the temperature
and composition
derivatives
of the EOS
parameters. Temperature
derivatives are usually quite
direct, and composition
derivatives may be obtained
from eqs (23) and (25). A compilation
of derivatives
for the more usual temperature
functions and mixing
rules proposed in the literature has been presented by
Miiller et al. (1989).
5. Substitute
the partial derivatives
of the EOS
parameters
as appropriate
in the generic results obtained in step 3 to generate
final thermodynamic
expressions
for the EOS. Perform algebraic reductions as necessary to simplify and compact the results.

The volume
functions
which
correspond
to
eqs (28H30) for this EOS model are collected in
Table 1. The generic property expressions
resulting
from the direct substitution
of these into eqs (13), (17),
(18) and (24) can be summarized
in the common
structure

with the general expressions for FO. . . , F5 given for


each property in Table 2.
In the original formulation
of the BWR EOS, the
five parameters had the temperature
dependence
e, =B,RT--&--$
e, = bRT-a

(33b)

e3 = aa

(33c)

-5

(334

VW

e4

es =

Example 1. The B WR and B WRS equations of state


For a straightforward
application
of the preceding
method
to a well-known
consider
the
case,
Benedict-WebbRubin
EOS (Benedict et al., 1940).
a virial-inspired
closed form that in the notation of
eq. (9) reads
P=$(T,v,e,,..

RT
el
.,e5)=-++oZ7++
V

e2

e3

T2

where &, BO, CO, a, b, c, 01,y are eight characteristic


constants for each substance, and functions of composition for mixtures according to the mixing rules
8, = i: x,&
for 0 = &,

Q=[$x~~!/]
This general procedure is next illustrated with three
examples. The first of these is intended to validate the
procedure
by rederiving
the well-established
property expressions
for the BWR and BWRS EOS. The
second example underscores
the versatility
of the
method by showing its application to the general class
of perturbed hard spheres (or repulsion-attraction)
EOS, with the Carnahan-Starling-van
der Waals
EOS as particular example. Finally, property expressions for a generalized four-parameter
cubic EOS are
obtained in the third example as direct consequences
of the preceding case.

forn=a,b,c,n.

(34a)
CO, y

(34b)

(34c)

Table 1. The generating function @(T, U,e,, . . . , e,) and its


partiat derivatives for the BWR model, eq. (31)
Expression

Function
0

Multicomponent

fugacity coefficients and residual properties

2023

Table 2. Coefficients of the generic property expression, eq. (31) for the BWR model

RTln 4

Coefficient
R In Z

FO

-RTlnZ

hR

RTln &

- RTln Z

2e,

FL

el + eli

2e, + ezi

3el - T

FL

F3

be,

6e, - T

F4

3e4

3e4 - T

2e, + &;

F5

2e5

2%--T

~5 + esi

Starling and coworkers developed a modification


of
the above parameter
expressions.
Improved temperature dependences
that incorporate
three additional
pure-component
constants
Do, E. and d were presented for e, , e2, q as (Starling, 1971)
e, =B,RT-A++$-$

e,=hRT-a--

sion for In &, which contains a few misprints in the


paper by Starling and Han (1972b) and should read
RTln&=

(B,i+Bo)RT

-RTlnZ+k

-l
-

+2iXj

Wa)

(1 -

kij)(AoiAoj)

d
T

5e3 + &
I

(1

kij)3

(1

kij)4

+22ai)1/3

(3W

tcoi:;)12
112
@oiD,y)

keeping eqs (33d, e) unchanged.


Mixing rules for application to multicomponent
systems were proposed
by Starling and Han (1972a),
R = i i xixj(l

- k,j)Y(Qfij)lz

1 i
for

fi = Ao, Co. Do, E.

(364

(where v = 1 for AD, 3 for Co, 4 for I&, and 5 for E,),
and
d=[$x,d::]l.

.td2:)

113

Wb)

maintaining
the corresponding
eqs (34a+) for the remaining parameters.
The temperature
and composition
derivatives
of
the above parameters
are shown in Tables 3 (original
BWR parameters)
and 4 (Starling modification).
Direct substitution
into eq. (32) of the coefficients defined
in Table 2 will yield the corresponding
final expressions for the thermodynamic
properties.
These are
rather lengthy formulas that need not be presented
here, as they agree of course with the equations given
by the original authors, except for the BWRS expres-

Example

2. Perturbed

hard-core

equations

of state

In the important

class of EOS models known as


augmented
or perturbed
hard-core
equations,
the
pressure is expressed as a sum of contributions
P=$.HC(T,~,e,,

. . . . e,)+A~VT,u,e,,...,e,,,)
(37)

CLAUDIO

2024
where +kHc corresponds
body fluid,
should

which

OLIVERA-FUENTES

to an EOS model for a hard-

is chosen

as reference

state

eq. (38) demands

lim

satisfy the constraint

v-m

@??cu
lim = 1
v-m RT

(38)

and A$ is a perturbation
or truncated
expansion
about that state, which contains contributions
from
attractions
and soft repulsions;
compatibility
with

Tahle 3. Temperature

and composition

original BWR EOS parameters,

derivatives of the

eqs (33) and (34)

Parameter, et

4i

hR

- L,RT-

a,

aa,+a,a-aaa
2c

--

T3

2?{
T3

.7<

Partial molar value

Mixing rule

!lfY!k= 0

FR = F:(Z)

c Xi&,

+ Fhc

+ AF

& = Boi

fii = --R + 2(Wf&)2


for Q = &, Co. y
n, = -2R + 3(RZRJ13
for Q = a, 6, c, a

Table 4. Temperature and composition


Parameter, et

derivatives of the BWRS EOS parameters, eqs (35) and (36)


%
-

&R+$_~+%$

e2

bR +;

&RT_i;_di
T

ud
e3

7.2

Mixing ruler

Partial molar value

L 1
3

.id,i3

+For e4, e5, see Table 3.


*For B,,, u, b, c, ct, y. see Table 3.

(40)

as shown in Table 5 for the residual entropy, residual


enthalpy and fugacity coefficients.
Here Fhc is the
residual property of the hard-core fluid, referred to the
ideal gas at the same volume, and AF is a perturbation. The leading term F:(Z), related to the compressibility factor, arises not from any pressure contribution, but from the change of independent
variable
from volume to pressure (Abbott and Nass, 1986). Por

(a%laT)..

el

d=

(39)

RT

In a different but related class of EOS models, the


pressure is interpreted as a sum of contributions
from
repulsive and attractive
molecular
interactions.
As
pointed
out by Abbott
(1989), these models are
formally identical to eq. (37) and no distinction
is
made between them in most practical applications.
Perturbed hard-core EOS have been an active subject of research for the last few decades, part of the
effort being addressed at finding the best combination
of reference and perturbation
functions.
Carnahan
and Starling (1972) pointed out the convenience
of
writing also the derived thermodynamic
functions as
sums of contributions,
so that the changes brought
about e.g. by combination
of different perturbation
expressions
with a given reference model could be
easily ascertained.
The present approach
lends itself
readily
to such
transformation;
inspection
of
eqs (13H15), (17), (18) and (24) shows that any residual property FR referred to the ideal gas at the same
pressure (rather than volume) will be given by the
common structure

Bo =

that

and

d, = -2d

+ 3[d2d,)3

Multicomponent
Table 5. Contributions

the unperturbed

reference
F;c

fugacity

to the general

coefficients

+ FLY.

properties

residual property for perturbed hard-body

model, clearly

= FP(Z,c)

and residual

(41)

For a particular
example,
consider
the rigidspheres van der Waals or CSVW EOS proposed by
Carnahan
and Starling (1972). It combines the analytical rigid-spheres
repulsion
model developed
by
Carnahan and Starling (1969) and the van der Waals
attraction term,

!& = Qi = 52,(zi)

for R = a, b.

(43)

With temperature
derivatives of the parameters set
to zero and partial molar values read off Table 3, and
the polynomial terms of eq. (45) reunited into a single
fraction, final expressions
For the residual properties
of the CSVW model with the rules of eqs (47), (48) can
be written as
_ ln Z + Y(3Y - 4)
(1

-=
R

(43)

The development
of general property expressions
for this model follows the procedure already outlined
and illustrated above; intermediate
steps are therefore
omitted. By interpreting
the hard-spheres
pressure
function as a polynomial
in reciprocal
powers of

eq. (40)

and took the pure-component


parameters
as constants with values determined from a classical stability
analysis of the critical point,

SR

A$@, a) = - a.
uz

2025
EOS models,

ln4=

3y3 - 9yz + 8y

-lnZ+

(1 - H3
h"

W2 - y)

- &

2a

-=3-m
RT (1-Y)

(50)

(51)

(1 - Y).
ln $, = _ ln z _ Y(~Y - 4) 1 h 2~(2 - ;)
(1
b U-Y)
the corresponding
residual
the common form
FL-c

Fo

The perturbation
mon structure,

l-y

Fl

functions

f1

contributions

F2
_
u12

can be put into

F3
+

(I--

(45)

similarly share a com-

AF = F,&.
The coefficients F,,, . . . , F4 for each property have
the expressions given in Table 6.
Carnahan
and Starling (1972) chose the linearBerthelot mixing rules

2(aia)*
-RTV.

(52)

Johnston
and Eckert (1981) also considered
the
pure-component
ai, bi to be independent of temperdependence of these was
ature (in fact, the temperature
too weak to merit analytical expression); eqs (49)-(51)
are thus also valid for their version of the CSVW
model. They used however a one fluid mixing rule
for a,
a = i i xixjaij

1 i
and an effective,
for b,

composition-independent
b z berr.

(47a)
b = i xibi
I

(53a)
value
WV

In this case, the partial molar value of a can be read


From Table 4, and that of b is easily seen from eq. (23)
to be simply b,,,. Direct substitution
into the generic

2026

CLAUDIO

Table 6. Coefficients

OLIVERA-FUENTES

of the residual property expressions for the Carnahan-Starling


Waals EOS functions, eqs (45) and (46)

Coefficient

r
(-1RT HC

-3

-1

F2

(In hLc

-3

F,

2-2;

b-3

Coefficient

Aln4

Af
!j

Ah
i RT >

AIn&,

-I_

-2

F4

formula

(In&k

J-0

and van der

and rearrangement

then gives

In tii = - In 2 + 3~~ -9~ + 8~


(1 - Y13

2 i xjaii
j
.
RTv
(54)

Additional
variations
on the CSVW model can
obviously he obtained by adopting other temperature
and composition
functions, as done e.g. by Fkrtucco et
al. (1986).

vertical asymptotes
in the P-v plane. They are more
easily obtained as integrals of $ than as derivatives of
0; the corresponding
expressions
are shown
in
Table 7.
As in the two previous examples, systematic
application of the general procedure
results in the discovery of a common structure underlying
the property relations for both contributions
to the fluid pressure. For the repulsion term this is
FL,

u-b

= F, In-

+ F,-

Example 3. The generalized SW cubic equation ofstate


While the general procedure can be applied to any
ofthe numerous cubic EOS proposed in the literature,
it is more useful and illustrative to take advantage of
their common algebraic form to develop generic formulas that can in turn be specialized
to particular
cases, as the need arises, by appropriate
specification
of the parameters
eI, . . , e,. Consider the generalized four-parameter
cubic EOS
P = i&T, u, a, b, c, d) = z

a
vz + cv + d
(55)

This model clearly belongs to the class of repulsion-attraction


and perturbed
hard-body
EOS considered in the preceding example, even if the van der
Waals repulsion term gives a poor approximation
to
the virial expansion
of the hard-spheres
fluid
(Henderson,
1979; Wong and Prausnitz,
1985). Property expressions
will therefore assume the form of eq.
(40). The volume functions, eqs (28)-(30) for the attraction term, deserve special attention, because they
depend
on the value of the discriminant

and for the attraction

(56)

which indicates whether the isotherms


of the cubic
EOS have one (6 < 0), two (6 = 0) or three (6 > 0)

(57)

term, it is

(58)
where I,, I*, I3 are the volume integrals already given
in Table 7, and the coefficients F, , . . . , Fe are defined
for each residual property in Table 8.
It is customary when presenting residual property
expressions
for cubic EOS to reintroduce
the EOS
itself in the form
b
-=Z--1+a
v-b

This is also possible in the present


nation of eqs (40) and (57)-(59)
formula

F,=F~+F,h-p

(59)

RTv2 + cv + d

v-b
V

approach; combigives the generic

+ F,(Z

- 1)

V
v2

6 = c2 - 4d

b
v-b

CD +

Multicomponent
Table

7. Volume

functions

and integrals

fugacity

coefficients

for the attraction


eq. (55)

and residual

properties

term of the generalized

2027
four-parameter

EOS,

Integral

Function
A$,-

v2 +

a
cu +

JW

da=-

13 + cv + d

afW

&do=

-al,

?!!&,=-~
I m aa

au

-=
&

(u2 +

aW
-=
ad

a
(I? + cv + d)

Cv +

8. Coefficients

0R

r
C

(In wit
-1

FZ

F4

four-

(In 9Jifc
-I

0
1

Ai
0

Coefficient

F3

6>0

of the residual property expressions


for the generalized
parameter
EOS, eqs (57) and (58)
s

Coefficient

6=0

dt0

F1

m ac

v Pf!&=aJ
s -ad

Integral

Table

Y?!?dU=al

d)2

Ahll$

-1

-1

-1
1

FS

F,

Many different temperature


and composition
characteristic
constant
dependences-including
values-have
been essayed in the literature for the
EOS parameters of eq. (56). The corresponding
residual property expressions may also look different, yet it
should be clear by now that they are but special cases
of the general functional structure described above,
obtained by appropriate substitution of the particular
temperature derivatives and partial molar values.

alnd
(->dlnT

(F, - c)

- (d; - d)

CONCLUSIONS
A general procedure has been presented for the
derivation of expressions for residual entropy, residual enthalpy, and fugacity coefficients from pressureexplicit EOS. The method is based on the separability
of the volume, temperature and composition dependences of the pressure, and can be applied in a stepwise
sequence leading first to a set of generic expressions

CLAUDIO OLIVERA-FUENTES

2028

valid for the EOS model without regard to any mixing


for the EOS pararules or temperature
functions
meters. The search for improved
temperature
and
composition
dependences
for existing EOS models is
an important
area of current research, which when
successful can greatly enhance the model performance, e.g. the well-known
case of the Soave (1972)
modification
of the Redlich-Kwong
EOS. It should
be obvious in such cases that the property expressions
need not be derived anew, but can be obtained just by
changing a set of coefficients in a more general formuta.
The present method also allows the separation
of
different contributions
to the pressure function, which
are shown to result also in separate contributions
to
the residual properties. It would be trivially easy, for
example, to combine eqs (45) and (58) from examples
2 and 3 above to derive property
expressions
for
a Carnahan-Starling-Schmidt-Wenzel
augmented
hard-spheres
model, should this be of interest. The
flexibility in EOS splicing and recombining
afforded
by the general procedure may be helpful when introducing
functional
modifications
in a previously
studied EOS model.
The derivation
of property

SW EOS parameter
molar volume
volume
SW EOS parameter
mole fraction
dimensionless
hard-core

NOTATION

a, b, c, d
A
d

molar Helmholtz function


BWR, BWRS and cubic EOS parameters
total Helmholtz energy of system
dimensionless
Helmholtz
energy,
AI(NRT)
BWR and BWRS

F1,.

9
h
11712.13
k
m
If
N
P
R
s
T

EOS parameters

generalized
pressure-explicit
EOS
parameters
general residual property
coefficients
in generalized
residual
property expression
molar Gibbs function
molar enthalpy
integrals defined in Table 7
binary interaction
parameter
number of parameters
in EOS
number of components
in system
moles
pressure
ideal gas constant
moIar entropy
temperature

PO/( RT)

Subscripts
C

HC

PI
ij
k
r
X

Superscripts
r
R

factor,

variable,

Greek symbols
BWR and BWRS EOS parameters
013Y
discriminant
of cubic EOS, cz - d
s
perturbation
operator
A
volume function in density-dependent
c
mixing rule
generalized EOS volume integral
fugacity coefficient
generalized EOS pressure function
general EOS parameter

expressions
from equations of state can involve considerable
algebraic effort,
and incorrect formulas for newly developed EOS have
been known to appear in the literature. The mathematical labor may not be reduced by adopting the
present method, but the adherence
to a systematic
procedure does reduce the possibilities
for introducing mistakes in the derivation, and is helpful in visualizing the functional
structure subjacent to an EOS
model or class of models.

b/(4@
compressibility

*
-

at the critical temperature


of hard-body
fluid
of component
i
of all components
except i
of the interaction
between components i and j
of any EOS parameter
reduced property
all mole fractions held constant

residual, referred to ideal gas at same


T, o and composition
residual, referred to ideal gas at same
T, P and composition
of ideal gas
(overbar) partial molar value

REFERENCES

Abbott, M. M., 1979, Cubic equations of state: an interpretive review, in Equations of State in Engineering and
Re~~rch (Edited bv K. C. Chao and R. L. Robinson. Jr.).
AmericanChemieai Society, Washington DC.
Abbott, M. M., 1989, Thirteen ways of looking at the van der
Waals equations of state. Chem. Engng Prog. S(2), 25-37.
Abbott, M. M. and Nass, K. K., 1986, Equations ofstate and
classical solution thermodynamics,
in Equations of State:
Theories and Applications (Edited by K. C. Chao and R. L.
Robinson, Jr.). American Chemical Society, Washington
DC.
Benedict, M., Webb, G. B. and Rubin, L. C., 1940, An
empirical equation for thermodynamic properties of light

hydrocarbons and their mixtures. 1. Methane, ethane,


propane and butane. J. Chem. Phys. 8, 334-345.
Bertucco, A., Fermeglia, M. and Kikic, I., 1986, Modified
Carnahan-Starling-van
der Waals equation
for supercritical fluid extraction.
Chem. Engng J. 32, 21-30.
Carnahan,
N. F. and Starling, K. E., 1969, Equation of state
for nonattracting rigid spheres. J. Chem. Phys. 51(2),
635-636.
Carnahan,
N. F. and Starling, K. E., 1972, Intermolecular
repulsions and the equation of state for fluids. A.I.Ch.E. J.
18(6), 1184-1189.

Multicomponent

fugacity

coefficients

Henderson,
D., 1979, Practical calculations
of the equation
of state of fluids and fluid mixtures using perturbation
theory and related theories, in Equations ofState in Engineering and Research (Edited by K. C. Chao and R. L.
Robinson, Jr.). American Chemical Society, Washington
D.C.
Johnston,
K. P. and Eckert, C. A., 1981, An analytical
Carnahan-Starling-van
der Waals model for solubility of
hydrocarbon
solids in supercritical fluids. A.1.Ch.E. .I.
27(S), 773-779.
Kumar, K. H. and Starling, K. E., 1980, Comments
on:
Cubic equations
of state: which?. Ind. Engng Chem.
Fundam. 21(3), 255-262.
Modell, M. and Reid, R. C., 1983, thermodynamics
and its
Applications,
2nd Edition.
Prentice-Hall,
Englewood
Cliffs, NJ.
Mollerup, J., 1986, Thermodynamic
properties from a cubic
equation of state. SEP 8601, lnstituttet
for Kemiteknik,
Danmarks
Tekniske HBjskole, Lyngby, Denmark.
Miiller, E. A., Olivera-Fuentes,
C. and Esttvez, L. A., 1989,
General expressions for multicomponent
fugacity coeflicients and residual properties
from cubic equations
of
state. Lat. Am. Appl. Res. 19(2), 99-109.
Null, H. R., 1970, Phase Equilibrium in Process Design.
Wiley-Interscience,
New York.
Schmidt, G. and Wenzel, H., 1980, A modified van der Waals
type equation of state. Chem Engng Sci. 35, 1503-1512.
Soave, G., 1972, Equilibrium
constants
from a modified
Redlich-Kwong
equation of state. Chem. Engng Sci. 27,
1197-1203.
Starling, K. E., 1972, Therm0 data refined for LPG. Part 1:
equation
of state and computer
prediction.
Hydroc.
Process. 50(3), 101-104.
Starling, K. E. and Han, M. S., 1972a, Therm0 data refined
for LPG. Part 14: mixtures.
Hydroc.
Process. 51(5),
129-132.
Starling, K. E. and Han, M. S., 1972b, Therm0 data refined
for LPG. Part 15: industrial applications.
Hydroc. Process.
51(6), 107-115.
Szarawara, J. and Gawdzik, A., 1989, Method of calculation
of fugacity coefficient from cubic equations of state. Chem.
Enyng Sci. 44, 1489-1494.
Van Ness, H. C. and Abbott, M. M., 1982, Classical Thermodynamics of Nonelectrolyte
Solutions. McGraw-Hill,
New

and residual

function,

properties

given in general

2029

by the equivalent

of eq. (2)

.=J;$-.=)da.
Other residual properties
derivatives of (I, e.g.

(Al)

are generated

in terms of partial

and, from eq. (3),


(A4)
For the generalized pressure-explicit
this work, eqs (28) and (Al) show that
a=

-O(T,v,e,

,...,

e,)=

- I;(

EOS considered

$ -

in

y)du. (AS)

Hence

w
1

lnr$=&l
For the species fugacity

coefficient,

-InZ-FT.
the partial

(A7)
derivatives

(AlO)
combine

to give, finally,

York.

Wong, J. 0. and Prausnitz, J. M., 1985, Comments concerning a simple equation of state of the van der Waals form.
Chem. Engng Commun. 37, 41-53.

APPENDIX:
EQUATION
The
IT,u,x,,.

THE HELMHOLTZ

FUNDAMENTAL

AND THE GENERATING

fundamental

., x.)

FUNCTION

thermodynamic
variable
in the
representation
is the molar Helmholtz

(All)
Mollerup (1986) has used this fundamental
approach
to
derive property expressions for a generalized four-parameter
cubic EOS. His results, that also include the second derivatives of the Helmholtz function, are valid however only for
positive values of the discriminant
6 of eq. (56), therefore
excluding important
cases such as the van der Waals and
Clausius families of EOS, as well as all noncubic EOS.

Вам также может понравиться