0% found this document useful (0 votes)
43 views

Solt 1 Solt 1 Notes

Uploaded by

orrebhanuprasad
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
43 views

Solt 1 Solt 1 Notes

Uploaded by

orrebhanuprasad
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 38

lOMoARcPSD|35886966

SOLT 1 - SOLT 1 notes

Solution Thermodynamics (Batangas State University-Lipa Campus)

Scan to open on Studocu

Studocu is not sponsored or endorsed by any college or university


Downloaded by Orre Bhanuprasad ([email protected])
lOMoARcPSD|35886966

MODULE 1 - Framework of Solution


Thermodynamics

INTRODUCTION
Our purpose in this module is to lay the theoretical foundation for
applications of thermodynamics to gas mixtures and liquid solutions. Throughout
the chemical, energy, microelectronics, personal care, and pharmaceutical
industries, multicomponent fluid mixtures undergo composition changes brought
about by mixing and separation processes, the transfer of species from one phase
to another, and chemical reaction. Thus, measures of composition become
essential variables, along with temperature and pressure, which we already
considered in detail in Chapter 6 of Introduction to Chemical Engineering
Thermodynamics by Smith, Van ness, Abbott. This adds substantially to the
complexity of tabulating and correlating thermodynamic properties, and leads to
the introduction of a menagerie of new variables and relationships among them.
Applying these relationships to practical problems, such as phase equilibrium
calculations, requires that we first map out this “thermodynamic zoo.”

INTENDED LEARNING OUTCOMES


After completing this chapter, the students must be able to:
1. Develop a fundamental property relation that is applicable to open phases of
variable composition.
2. Define the chemical potential, a fundamental new property that facilitates
treatment of phase and chemical-reaction equilibria.
3. Introduce partial properties, a class of thermodynamic properties defined
mathematically to distribute total mixture properties among individual species as
they exist in a mixture. These are composition-dependent and distinct from the
molar properties of pure species.
4. Develop property relations for the ideal-gas-state mixture, which provide the
basis for treatment of real-gas mixtures.
5. Define yet another useful property, the fugacity. Related to the chemical
potential, it lends itself to mathematical formulation of both phase- and
chemical-reaction-equilibrium problems.
6. Introduce a useful class of solution properties, known as excess properties, in
conjunction with an idealization of solution behavior called the ideal-solution
model, which serves as a reference for real-solution behavior.

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

FUNDAMENTAL PROPERTY RELATIONS

Equation (2.6), the first law for a closed system of n moles of a substance, may be written
for the special case of a reversible process:

d(nU) = d Q rev + d W rev

Equations (1.3) and (5.1) as applied to this process are:

d W rev = − P d (nV ) d Q rev = T d(nS)

These three equations combine to give:

d (nU ) = T d (nS ) − P d (nV ) (6.1)

where U, S, and V are molar values of the internal energy, entropy, and volume. All of
the primitive thermodynamic properties—P, V, T, U, and S—are included in this equation.
It is a fundamental property relation connecting these properties for closed PVT
systems. All other equations relating properties of such systems derive from it.

Additional thermodynamic properties, beyond those appearing in Eq. (6.1), are defined
as a matter of convenience, in relation to the primary properties. Thus, the enthalpy,
defined and applied in Chap. 2, is joined here by two others. The three, all with
recognized names and useful applications, are:

Enthalpy H ≡ U + PV (6.2)

Helmholtz energy A ≡ U – TS (6.3)

Gibbs energy G ≡ U + PV – TS = H – TS (6.4)

The Helmholtz energy and Gibbs energy,1 with their origins here, find application in
phase and chemical-equilibrium calculations and in statistical thermodynamics.

Multiplication of Eq. (6.2) by n, followed by differentiation, yields the general expression

d (nH ) = d (nU ) + P d (nV ) + nV dP

Substitution for d(nU) by Eq. (6.1) reduces this result to

d (nH ) = T d (nS ) + nV dP (6.5)

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

The differentials of nA and nG are obtained similarly:

d (nA ) = − nS dT − P d (nV ) (6.6)

d (nG ) = − nS dT + nV dP (6.7)

Equations (6.1) and (6.5) through (6.7) are equivalent fundamental property relations.

Application is to any process in a closed PVT system resulting in a differential

change from one equilibrium state to another.

The system may consist of a single phase (a homogeneous system), or it may comprise
several phases (a heterogeneous system); it may be chemically inert, or it may undergo
chemical reaction. The choice of which equation to use in a particular application is
dictated by convenience.

However, the Gibbs energy G is special, because of its unique functional relation to T
and P, the variables of primary interest by virtue of the ease of their measurement and
control.

An immediate application of these equations is to one mole (or to a unit mass) of a


homogeneous fluid of constant composition. For this case, n = 1, and they simplify to:

dU = T dS − P dV (6.8)

dH = T dS + V dP (6.9)

dA = −S dT − P dV (6.10)

dG = −S dT + V dP (6.11)

Implicit in each of these equations is a functional relationship that expresses a molar (or
unit mass) property as a function of a natural or special pair of independent variables:

U = U(S, V) H = H(S, P) A = A(T, V) G = G(T, P)

These variables are said to be canonical, 2 and a thermodynamic property known as a


function of its canonical variables has a unique characteristic:

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Equations (6.8) through (6.11) lead to another set of property relations because they are
exact differential expressions. In general, if F = F(x, y), then the total differential of F is
defined as:

餸 餸

or

dF = M dx + N dy (6.12)

where

餸 餸

then

餸 餸

The order of differentiation in mixed second derivatives is immaterial, and these


equations combine to give:

Because F is a function of x and y, the right side of Eq. (6.12) is an exact differential
expression, and Eq. (6.13) correctly relates the partial derivatives. This equation serves in
fact as the criterion of exactness. For example, y dx + x dy is a simple differential
expression which is exact because the derivatives of Eq. (6.13) are equal, i.e., 1 = 1. The
function that produces this expression is clearly F(x, y) = xy. On the other hand, y dx – x
dy is an equally simple differential expression, but the derivatives of Eq. (6.13) are not
equal, i.e., 1 ≠ −1, and it is not exact. There is no function of x and y whose differential
yields the original expression.

The thermodynamic properties U, H, A, and G are known to be functions of the


canonical variables on the right sides of Eqs. (6.8) through (6.11). For each of these exact
differential expressions, we may write the relationship of Eq. (6.13), producing the
Maxwell relations.

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

ꯠ䍟

Expression of U, H, A, and G as functions of their canonical variables does not preclude


the validity of other functional relationships for application to particular systems. Indeed,
Axiom 3 of Sec. 2.5, applied to homogeneous PVT systems of constant composition, asserts their
dependence on T and P. The restrictions exclude heterogeneous and reacting systems,
except for G, for which T and P are the canonical variables. A simple example is a system
comprised of a pure liquid in equilibrium with its vapor. Its molar internal energy
depends on the relative amounts of liquid and vapor present, and this is in no way
reflected by T and P. However, the canonical variables S and V also depend on the
relative amounts of the phases, giving U = U(S, V) its greater generality. On the other
hand, T and P are the canonical variables for the Gibbs energy, and G = G(T, P) is
general. Thus G is fixed for given T and P, regardless of the relative amounts of the
phases, and provides the fundamental basis for the working equations of phase equilibria.
Equations (6.8) through (6.11) lead not only to the Maxwell relations but also to many
other equations relating thermodynamic properties. The remainder of this section
develops those most useful for evaluation of thermodynamic properties from
experimental data.

Enthalpy and Entropy as Functions of T and P

In engineering practice, enthalpy and entropy are often the thermodynamic properties of
interest, and T and P are the most common measurable properties of a substance or
system. Thus, their mathematical connections, expressing the variation of H and S with
changes in T and P, are needed. This information is contained in the derivatives
(∂H/∂T)P, (∂S/∂T)P, (∂H/∂P)T, and (∂S/∂P)T, with which we can write:

t t
t

Our goal here is to express these four partial derivatives in terms of measurable
properties. The definition of heat capacity at constant pressure is:

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Another expression for this quantity is obtained by applying Eq. (6.9) to changes with
respect to T at constant P:

Combination of this equation with Eq. (2.19) gives:

The pressure derivative of entropy results directly from Eq. (6.17):

耀䍟

The corresponding derivative for enthalpy is found by applying Eq. (6.9) to changes with
respect to P at constant T:

As a result of Eq. (6.19) this becomes:

t
彣䍟

With expressions for the four partial derivatives given by Eqs. (2.19) and (6.18) through
(6.20), we can write the required functional relations as:

t 䍟

These are general equations relating enthalpy and entropy to temperature and pressure
for homogeneous fluids of constant composition. Eqs. (6.19) and (6.20) illustrate the utility of the
Maxwell relations, particularly Eqs. (6.16) and (6.17), which relate changes in entropy that
are not experimentally accessible to PVT data that are experimentally measurable.

The Ideal-Gas State

The coefficients of dT and dP in Eqs. (6.21) and (6.22) are evaluated from heat-capacity
and PVT data. The ideal-gas state (denoted by superscript ig) provides an example of
PVT behavior:

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

t
t

Substituting these equations into Eqs. (6.21) and (6.22) reduces them to:

t t
tt t
㐲 䍟

These are restatements of equations for the ideal-gas state presented in Secs. 3.3 and 5.5.

Alternative Forms for Liquids

Alternative forms of Eqs. (6.19) and (6.20) result when (∂V/∂T)P is replaced by βV [Eq.
(3.3)]:

t

These equations incorporating β, although general, are usually applied only to liquids.
However, for liquids at conditions far from the critical point, both the volume and β are
small. Thus at most conditions pressure has little effect on the properties of liquids. The
important idealization of an incompressible fluid (Sec. 3.2) is considered in Ex. 6.2.
Replacing (∂V/∂T)P in Eqs. (6.21) and (6.22) with βV yields:

t ꯠ 䍟

Because β and V are weak functions of pressure for liquids, they are usually assumed
constant at appropriate average values for integration of the final terms.

Internal Energy as a Function of P

Internal energy is related to enthalpy by Eq. (6.2) as U = H – PV. Differentiation yields:

Then by Eq. (6.20),

An alternative form results if the derivatives on the right are replaced by βV [Eq. (3.3)]
and – [Eq. (3.4)]:

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

耀䍟

THE CHEMICAL POTENTIAL AND EQUILIBRIUM

Practical applications of the chemical potential will become clearer in later chapters that
treat chemical and phase equilibria. However, at this point one can already appreciate its
role in these analyses. For a closed, single-phase PVT system containing chemically
reactive species, Eqs. (6.7) and (10.2) must both be valid, the former simply because the
system is closed and the second because of its generality. In addition, for a closed system,
all differentials dni in Eq. (10.2) must result from chemical reaction. Comparison of these
equations shows that they can both be valid only if:

µ 彣

This equation therefore represents a general criterion for chemical-reaction equilibrium


in a single-phase closed PVT system, and is the basis for the development of working
equations for the solution of reaction-equilibrium problems.
With respect to phase equilibrium, we note that for a closed nonreacting system consisting
of two phases in equilibrium, each individual phase is open to the other, and mass transfer
between phases may occur. Equation (10.2) applies separately to each phase:

䍟 µ

䍟 µ

where superscripts α and β identify the phases. For the system to be in thermal and
mechanical equilibrium, T and P must be uniform.
The change in the total Gibbs energy of the two-phase system is the sum of the
equations for the separate phases. When each total-system property is expressed by the
equation of the form,
䍟 䍟
the sum is:

µ µ

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

because the two-phase system is closed, Eq. (6.7) is also valid. Comparison of the two
equations shows that at equilibrium:

µ µ 彣

The changes and result from mass transfer between the phases; mass
conservation therefore requires:

µ µ 彣

Quantities are independent and arbitrary, and the only way the left side of the
second equation can in general be zero is for each term in parentheses separately to be
zero. Hence,

µ µ (i = 1, 2, …, N)
where N is the number of species present in the system. Successive application of this
result to pairs of phases permits its generalization to multiple phases; for π phases:

µ µ µ (i = 1, 2, …, N) (10.6)
A similar but more comprehensive derivation shows (as we have supposed) that for
equilibrium T and P must be the same in all phases.
Thus, multiple phases at the same T and P are in equilibrium when the
chemical potential of each species is the same in all phases.
The application of Eq. (10.6) in later chapters to specific phase-equilibrium problems
requires models of solution behavior, which provide expressions for G and µi as
functions of temperature, pressure, and composition. The simplest of these, the ideal-gas
state mixture and the ideal solution, are treated in Secs. 10.4 and 10.8, respectively.

The Ideal- Gas Mixture Model

Despite its limited ability to describe actual mixture behavior, the ideal- gas mixture
model provides a conceptual basis upon which to build the structure of solution
thermodynamics. It is a useful property model because it:
 Has a molecular basis.
 Approximates reality in the well- defined limit of zero pressure.
 Is analytically simple.

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

The molar volume of an ideal gas is V= RT/P [Eq. (3.14)] regardless of the nature
of the gas. Thus all ideal gases, whether pure or mixed, have the same molar volume at
the same T and P. The partial molar volume of species i in an ideal-gas mixture is found
from Eq. (11.7) applied to the volume; superscript ig denotes the ideal-gas state:
t
t 䍟 㐲 t 䍟 㐲 㐲

where the final equality depends on the equation This result means
that for ideal gases at given T and P the partial molar volume, the pure- species molar
volume, and the mixture molar volume are identical:
t t t

彣䍟

We define the partial pressure of species i in an ideal- gas mixture as the pressure that
species i would exert if it alone occupied the molar volume of the mixture. Thus.

t 䍟

where is the mole fraction of species i. The partial pressures obviously sum to the
total pressure.
Because the ideal-gas mixture model presumes molecules of zero volume that do not
interact, the thermodynamic properties (other than molar volume) of the constituent
species are independent of one another, and each species has its own set of private
properties. This is the basis for the following statement of Gibbs’s theorem:
A partial molar property (other than volume) of a constituent species in an
ideal-gas mixture is equal to the corresponding molar property of the
species as a pure ideal gas at the mixture temperature but at a pressure
equal to its partial pressure in the mixture.
t t
This is expressed mathematically for generic partial property by the
equation:
t t

The enthalpy of an ideal gas is independent of pressure: therefore
t t t
t t t
t t
More simply, t t (11.22)
t
where t is the pure species value at the mixture T and P. An analogous equation
t
applies for and other properties that are independent pressure.
The entropy of an ideal gas does depend on pressure, and by Eq. (6.24),

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

t
㐲 ln 댳䁡 쳌 䍟
Integration from to P gives:
t t
㐲 㐲 㐲

t t
Whence, 㐲
Substituting this result into Eq. (11.21) written for entropy yields:
t

t t
or 㐲 (11.23)
t t
where is the pure- species value a t the mixture T and P.
t
For the Gibbs energy of an ideal- gas mixture, tt t
; the parallel
relation for partial properties is:
t t t
t
In combination with Eqs. (11.22) and (11.23) this becomes:
t t t
t 㐲 ln
t t t
or 㐲 ln (11.24)
Differentiation of this equation in accord with Eqs (11.18) and (11.19) confirms the
results expressed by Eqs. (11.20) and (11.23).
The summability relation, Eq. (11.11), with Eqs. (11.22), (11.23), and (11.24) yields:
t
tt t

t t

t t

t t
Equations analogous to Eq. (11.25) may be written for both and . The former
appears as Eq. (4.6), but the latter reduces to an identity because of Eq. (11.20).
When Eq. (11.25) is written,
t
tt t 彣

the difference on the left is the enthalpy change associated with a process in which
appropriate amounts of the pure species at T and P are mixed to form one mole of

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

mixture at the same T and P. For ideal gases, this enthalpy change of mixing (Sec. 12.3) is
zero.
When Eq. (11.26) is rearranged as;

t t

the left side is the entropy change of mixing for ideal gases. Because 1/ >1, this quantity is
always positive, in agreement with the second law. The mixing process is inherently
irreversible, and for ideal gases mixing at constant T and P is not accompanied by heat
transfer [Eq.(11.25)].
t t
An alternative expression for the chemical potential results when in Ew.
(11.24) is replaced by an expression giving its T and P dependence. This comes from Eq.
(6.10) written for an ideal gas:
t t 㐲
㐲 댳䁡 쳌 䍟

Integration gives:
t
Γ 㐲 (11.28)
where Γ the integration constant at constant T, is a species-dependent function of
temperature only. Equation (11.24) is now written:
t t
Γ 㐲 䍟
where the argument of the logarithm is the partial pressure. Application of the
summability relation, Eq. (11.11), produces an expression for the Gibbs energy of an
ideal- gas mixture:
t
Γ 㐲 ln 䍟

These equations, remarkable in their simplicity, provide a full description of ideal-gas


behavior. Because T, P. and are the canonical variables for the Gibbs energy, all
other equations for the ideal- gas model can be generated from them.

FUGACITY AND FUGACITY COEFFICIENT: PURE SPECIES


As evident from Eq. (11.6), the chemical potential μi provides the fundamental criterion
for phase equilibria. This is true as well for chemical-reaction equilibria. However, it
exhibits

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

A dimensional ambiguity is evident with Eq. (11.28) and with analogous equations to
follow in that P has units, whereas ln P must be dimensionless. This difficulty is more
apparent than real, because the Gibbs energy is always expressed on a relative scale,
absolute values being unknown. Thus in application only differences in Gibbs energy
appear, leading to ratios of quantities with units of pressure in the argument of the
logarithm. The only requirement is that consistency of pressure units be maintained.
Characteristics with discourage its use. The Gibbs energy, and hence μί, is defined in
relation to the internal energy and entropy. Because absolute values of internal energy are
unknown, the same is true for μί. Moreover, Eq. (11.29) shows that μίig approaches
negative infinity when either P or γί approaches zero. This is true not just for an ideal gas
but for any gas. Although these characteristics do not preclude the use of chemical
potentials, the application of equilibrium criteria is facilitated by introduction of the
fugacity,7 a property that takes the place of μί but which does not exhibit its less desirable
characteristics.
The origin of the fugacity concept resides in Eq. (11.28), valid only for pure species ί
in the ideal-gas state. For a real fluid, we write an analogous equation that defines ƒί, the
fugacity of pure species ί:
Gi≡ Гi (T) + RT ln ƒi (11.31)

This new property ƒi, with units of pressure, replaces P in Eq. ( 11.28). Clearly, if
Eq.(11.28) is a special case of Eq. (11.31), then:
ƒiig = P (11.32)
and the fugacity of pure species ί as an ideal gas is necessarily equal to its pressure.
Subtraction of Eq. (11.28) from Eq. (11.31), both written for the same T and P , gives:

Gi – Giig = RT ln ƒi
P
By the definition of Eq. (6.41), Gi – Gίig is the residual Gibbs energy, GiR ; thus,

GiR = RT ln Фί (11.33)

Where the dimensionless ratio ƒi/P has been defined as another new property, the
fugacity coefficient, given by symbol Фi :

Фi = ƒi (11.34)

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

The identification of ln Фi with GiR/ RT by Eq. (11.33) permits its evaluation by the
integral of Eq. (6.49):

(const T)

ln Фi = 0 (Zi – 1) dP
(11.35)
P

Fugacity coefficients (and therefore fugacities) for pure gases evaluated by this equation
from PVT data or from a volume- explicit equation of state.
For example, when the compressibility factor is given by Eq. (3.38).

Zi -1 = Bii P
RT

Fugacity Coefficients from the Generic Cubic Equation of State

Evaluation of fugacity coefficients through cubic equations of state (e.g., the van der
Waals, Redlich/Kwong, and Peng/Robinson equations) follows directly from
combination of Eqs. (11.33) and (6.66b):
(11.37)
ln Фi = Zi – 1 – ln (Zi – βi ) – q i ti

Where βi is given by Eq.(3.50); qi , by Eq.(3.51); and Ii , by Eq. (6.65b), all written for
pure species I [for the van der Waals equation, Ii = βi / Zi ]. Application of Eq.(11.37) at
a given T and P requires prior solution of an equation of state for Zi by Eq. (3.52) for a
vapour phase or Eq. (3.56) for a liquid phase.

Vapor/Liquid Equilibrium for Pure Species

Equation (11.31), which defines the fugacity of pure species i, may be written for species
i as a standard vapour and as a saturated liquid at the same temperature:

Giv = Гi (T ) + RT ln ƒiv Gil = Гi ( T ) + RT ln ƒil


(11.38a) (11.38b)

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Fugacity of a Pure Liquid


The fugacity of pure species is a compressed liquid may calculated from the product of
easily evaluated ratios:

ƒil (P ) = ƒiv (Pisat ) ƒil (Pisat) ƒil (P ) Pisat


Pisat ƒiv (Pisat ) ƒil(Pisat )

(A) (B) (C)

All terms are at the temperature of interest. Inspection reveals that cancelation of
numerators and
Denominators produces a mathematical identity.

Example 2.5
For H2O at a temperature of 300oC and for pressure up to 10,000 kPa (100 bar) calculate
values of ƒi and Фi from data in the steam tables and plot them vs. P.

Solution 2.5
Equation (11.31) is a written twice: first for a state at pressure P; second, for a
low-pressure reference state, denoted by *, both for temperature T:

Gi =Гi (T) + RT ln ƒi and Gi* =Гi (T) + RT ln ƒi*

Subtraction eliminated Гi (T), and yields:

ln ƒi = 1 (Gi - Gi*)
ƒ i* RT
By definition Gi=Hi – TSi and Gi* = Hi* - TSi*; substitution gives:

ln ƒi = 1 Hi - Hi* _ (Si - Si* )


ƒ i* R T

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

The lowest pressure for which data at 300oC are given in the steam tables is 1kPa. Steam
atthese conditions is for practical purposes an ideal gas, for which ƒi*= P* = 1kPa. Data
for this state provide the following reference values:

Hi* = 3, 076.8 Jg-1 Si*= 10.3450 Jg-1K-1

Equation (A) may now be applied to states of superheated steam at 300oC for various
values of P from 1 kPa to the saturation pressure of 8,592.7 kPa. For example, at P =
4,000 kPa and the 300oC:

Hi = 2, 962.0 Jg-1 Si= 6, 3642 Jg-1K-1

Values of H and S must be multiplied by the molar mass of water (18.015) to put them on
a molar basis for substitution into Eq.(A):

ln ƒi = 18.015 2.962.0 – 3,076.8 _ (6.3642 – 10.3450 ) = 8.1917


ƒ i* 8.314 573.15
Whence,
ƒi= (3,611.0) (ƒi*) = (3,611.0)(1 kPa) = 3,611.0 kPa
Thus the fugacity coefficient at 4,000 kPa is:
Фi = ƒi = 3,611.0 = 0.9028
P 4,000
Similar calculations at other pressures lead to the values plotted in Fig. 11.3 at pressures
up to the saturation pressure P isat= 8,592.7 kPa. At this pressure,

Фisat = 0.7843 and ƒisat = 6,738.9 kPa

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

According to Eqs. (11.39) and (11.41), the saturation values are unchanged by
condensation. Although the plots are therefore continuous, they do show discontinuities
in slope. Values of ƒi and Фi for liquid water at higher pressure are found by application
of Eq. (11.44), with V il equal to the molar volume of saturated liquid water at 300oC:

V il = (1.403)(18.015) = 25.28 cm3 mol-1

At 10,000kPa, for example, Eq. (11.44) becomes:

ƒi = (0.7843)(8492.7) exp (25.28)(10,000 – 8,592.7) = 6,789.8kPa


(8,314)(573.15)

The fugacity coefficient of liquid water at these conditions is:

Фi = ƒi/P = 6,789.8/10,000 = 0.6790

Such calculations allow completion of Fig. 11.3, where the solid lines show how ƒi and Фi
vary with pressure.
The curve for ƒi starts at the origin, and deviates increasing from the dashed line for
an ideal gas (ƒi= P) as the pressure rises. At Pisat there is a discontinuity in slope, and the
curve then rises very slowly with increasing pressure, indicating that the fugacity of liquid
water at 300oC is a weak function of pressure. This behaviour is characteristic of liquids at
temperatures well below the critical temperature. The fugacity coefficient Фi decreases

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

steadily from its zero-pressure value of unity as the pressure rises. Its rapid decrease in the
liquid region is a consequence of the near-constancy of the fugacity itself.

FUGACITY AND FUGACITY COEFFICIENT: SPECIES IN SOLUTION

The definition of the fugacity of a species in solution is parallel to the definition


of the pure-species fugacity. For species i in a mixture of real gases or in a solution of
liquids, the equation analogous to Eq. (10.29), the ideal-gas-state expression, is:
μi Γi(T) + RT ln i

where i is the fugacity of species i in solution, replacing the partial pressure yiP. This
definition of i does not make it a partial molar property, and it is therefore identified by
a circumflex rather than by an overbar.
A direct application of this definition indicates its potential utility. Equation
(10.6), the equality of μi in every phase, is the fundamental criterion for phase
equilibrium. Because all phases in equilibrium are at the same temperature, an alternative
and equally general criterion follows immediately from Eq. (10.46):
i = i = …= (i = 1, 2, . . . , N) (10.47)
α β π
i

Thus, multiple phases at the same T and P are in equilibrium when the fugacity
of each constituent species is the same in all phases.
This is the criterion of equilibrium most often applied in the solution of
phase-equilibrium problems. For the specific case of multicomponent vapor/liquid
equilibrium, Eq. (10.47) becomes:
l= v (i = 1, 2, . . . , N) (10.48)
i i

Equation (10.39) results as a special case when this relation is applied to the vapor/liquid
equilibrium of pure species i. The definition of a residual property is given in Sec. 6.2:
MR ≡ M − Mig
where M is the molar (or unit-mass) value of a thermodynamic property and Mig is the
value that the property would have in its ideal-gas state with the same composition at the
same T and P. The defining equation for a partial residual property i
R follows from this
equation. Multiplied by n mol of mixture, it becomes:

nMR = nM − nMig
Differentiation with respect to ni at constant T, P, and nj gives:

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

㐲䍟 䍟 t䍟

Reference to Eq. (10.7) shows that each term has the form of a partial molar property.
Thus,
i
R = I - i
ig (10.49)

Because residual properties measure departures from ideal-gas-state values, their


most logical
and common application is to gas-phase properties, but they are also valid for describing
liquid-phase properties. Written for the residual Gibbs energy, Eq. (10.49) becomes:
i
R
i i
ig (10.50)
an equation which defines the partial residual Gibbs energy.
Subtracting Eq. (10.29) from Eq. (10.46), both written for the same T and P, yields:
t
㐲 ln

This result combined with Eq. (10.50) and the identity gives:
i
R 㐲 ln (10.51)
where by definition,

(10.52)

The dimensionless is called the fugacity coefficient of species i in solution.


Although most commonly applied to gases, the fugacity coefficient may also be used for
liquids, and in this case mole fraction yi is replaced by xi, the symbol traditionally used for
mole fractions in the liquid phase. Because Eq. (10.29) for the ideal-gas state is a special
case of Eq. (10.46):

t
(10.53)

Thus, the fugacity of species i in an ideal-gas-state mixture is equal to its partial


t 㐲
pressure. Moreover, , and for the ideal-gas state, 彣.

The Fundamental Residual-Property Relation


The fundamental property relation given by Eq. (10.2) is put into an alternative form
through

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

the mathematical identity [also used to generate Eq. (6.37)]:

㐲 㐲 㐲
In this equation is eliminated by Eq. (10.2) and G is replaced by its definition,
H − TS. The result, after algebraic reduction, is:

t

(10.54)
㐲 㐲 㐲

All terms in Eq. (10.54) have the units of moles; moreover, in contrast to Eq. (10.2), the
enthalpy rather than the entropy appears on the right side. Equation (10.54) is a general
relation expressing nG/RT as a function of all of its canonical variables, T, P, and the
mole numbers. It reduces to Eq. (6.37) for the special case of 1 mol of a
constant-composition phase. Equations (6.38) and (6.39) follow from either equation,
and equations for the other thermodynamic properties then come from appropriate
defining equations. Knowledge of G/RT as a function of its canonical variables allows
evaluation of all other thermodynamic properties, and therefore implicitly contains
complete property information. Unfortunately, we do not have the experimental means
by which to exploit this characteristic. That is, we cannot directly measure G/RT as a
function of T, P, and composition. However, we can obtain complete thermodynamic
information by combining calorimetric and volumetric data. In this regard, an analogous
equation relating residual properties proves useful.
Because Eq. (10.54) is general, it may be written for the special case of the ideal-gas state:

t t t
tt
(10.54)
㐲 㐲 㐲 㐲

In view of the definitions of residual properties [Eqs. (6.41) and (10.50)], subtracting this
equation from Eq. (10.54) gives:
㐲 㐲 t㐲
ln
㐲 㐲 㐲

Equations so general as Eqs. (10.55) and (10.56) are most useful for practical application
in restricted forms. Division of Eqs. (10.55) and (10.56), first, by dP with restriction to
constant T and composition, and second, by dT and restriction to constant P and
composition

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

leads to:
㐲 㐲 t㐲


(10.57)

t㐲 㐲 t㐲


(10.58)

These equations are restatements of Eqs. (6.43) and (6.44) wherein the restriction of the
derivatives to constant composition is shown explicitly. They lead to Eqs. (6.46), (6.48),
and (6.49) for the calculation of residual properties from volumetric data. Moreover, Eq.
(10.57) is the basis for the direct derivation of Eq. (10.35), which yields fugacity
coefficients from volumetric data. It is through the residual properties that this kind of
experimental information enters into the practical application of thermodynamics.

In addition, from Eq. (10.56),


ln
㐲 t㐲
(10.59)

This equation demonstrates that the logarithm of the fugacity coefficient of


a species in solution is a partial property with respect to GR/RT.

EXAMPLE 2.6
Develop a general equation to calculate ln values from compressibility-factor
data.
Solution 2.6
For n mol of a constant-composition mixture, Eq. (6.49) becomes:

t
㐲 彣

In accord with Eq. (10.59) this equation may be differentiated with respect to ni at
constant T, P, and nj to yield:
t
ln

t
Because t and , this reduces to:

ln 彣
t (10.60)

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

where integration is at constant temperature and composition. This equation is the


partial-property analog of Eq. (10.35). It allows the calculation of values from
PVT data.

Fugacity Coefficients from the Virial Equation of State


Values of for species i in solution are readily found from equations of state.
The simplest form of the virial equation provides a useful example. Written for a
gas mixture it is exactly the same as for a pure species:

t

(3.36)
The mixture second virial coefficient B is a function of temperature and
composition. Its exact composition dependence is given by statistical mechanics,
which makes the virial equation preeminent among equations of state where it is
applicable, i.e., to gases at low to moderate pressures. The equation giving this
composition dependence is:

(10.61)

where yi and yj represent mole fractions in a gas mixture. The indices i and j identify
species, and both run over all species present in the mixture. The virial coefficient
Bij characterizes bimolecular interactions between molecules of species i and
species j, and therefore Bij = Bji. The double summation accounts for all possible
bimolecular interactions. For a binary mixture i = 1, 2 and j = 1, 2; the expansion
of Eq. (10.61) then gives:

Or
Equation (10.62) allows the derivation of expressions for ln ln for a binary
gas mixture that obeys Eq. (3.36). Written for n mol of gas mixture, it becomes:

t

Differentiation with respect to n1 gives:
t
t

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Substitution for t in Eq. (10.60) yields:

ln
㐲 彣 㐲

where the integration is elementary because B is not a function of pressure. All that
remains is evaluation of the derivative.
Equation (10.62) for the second virial coefficient may be written:

Or
with

Multiplying by n and substituting yi = ni/n gives,

By differentiation:


Therefore,

ln 䍟 (10.63a)

Similarly,

ln 䍟 (10.63b)

Equations (10.63) are readily extended for application to multicomponent gas mixtures;
the
general equation is:
(10.64)
ln 䍟

where the dummy indices i and j run over all species, and

With
彣 彣 쳌댳 쳌댳

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Example 2.7
Determine the fugacity coefficients as given by Eqs. (10.63) for nitrogen and methane in
a N2(1)/CH4(2) mixture at 200 K and 30 bar if the mixture contains 40 mol-% N2.
Experimental virial-coefficient data are as follows:
B11 = −35.2 B22 = −105.0 B12 = −59.8 cm3⋅ mol−1

Solution 2.7
By definition, . Whence,
耀 䍟 彣 彣 彣 댳 䁡
Substitution of numerical values in Eqs. (10.63) yields:

ln 彣 彣 䍟 彣彣 彣
䍟 彣彣 䍟

ln 彣 彣 彣 彣 䍟 彣
䍟 彣彣 䍟
Whence,
彣耀 彣
Note that the second virial coefficient of the mixture as given by Eq. (10.62) is
B = −72.14 cm3⋅ mol−1, and that substitution in Eq. (3.36) yields a mixture
compressibility factor,
Z = 0.870.

GENERALIZED CORRELATIONS FOR THE FUGACITY COEFFICIENT


Fugacity Coefficients for Pure Species
The generalized methods developed in Sec. 3.7 for the compressibility factor Z and in
Sec. 6.4 for the residual enthalpy and entropy of pure gases are applied here to the
fugacity coefficient. Equation (10.35) is put into generalized form by substitution of the
relations,
댳 댳

Hence,

ln 彣
t 䍟 (10.65)

Where integration is at constant . Substitution for t by Eq. (3.53) yields:

ln 彣
t彣 䍟 彣
t

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Where for simplicity subscript i is omitted. This equation may be written in alternative
form: (10.66)

ln ln ln
Where

ln 彣
t彣 䍟 and ln 彣
t

The integrals in these equations may be evaluated numerically or graphically for various
values of and from the data for t彣 and t given in Tables D.1 through D.4
(App. D). Another method, and the one adopted by Lee and Kesler to extend their
correlation to fugacity coefficients, is based on an equation of state.
Equation (10.66) may also be written,

䍟 䍟 (10.67)


and we have the option of providing correlations for and rather than for their
logarithms. This is the choice made here, and Tables D.13 through D.16 present values
for these quantities as derived from the Lee/Kesler correlation as functions of and
, thus providing a three-parameter generalized correlation for fugacity coefficients.

Tables D.13 and D.15 for can be used alone as a two-parameter correlation which
does not incorporate the refinement introduced by the acentric factor.
EXAMPLE
Estimate from Eq. (10.67) a value for the fugacity of 1-butene vapor at 200°C and 70 bar.
SOLUTION
At these conditions, with 댳 彣彣K 댳 彣 bar from Table B.1, we have:
ꯠ ꯠ 彣 耀
By interpolation in Tables D.15 and D.16 at these conditions,

彣 ꯠ and 彣耀
Equation (10.67) then gives:
彣 ꯠ䍟 彣耀 䍟彣 耀

and
彣 ꯠ彣 ꯠ bar
A useful generalized correlation for ln results when the simplest form of the virial
equation is valid. Equations (3.57) and (3.59) combine to give:


t

Substitution into Eq. (10.65) and integration yield:

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966


ln

or


exp (10.68)

This equation used in conjunction with Eqs. (3.61) and (3.62), provides reliable values of
for any nonpolar or slightly polar gas when applied at conditions where Z is
approximately linear in pressure. Figure 3.13 again serves as a guide to its applicability.
Names functions HRB (TR, PR, OMEGA) and SRB (TR, PR, OMEGA) for the
evaluation of t㐲 t㐲 댳 and 㐲
t㐲 by the generalized virial – coefficient correlation are
described in Sec. 6.4. Similarly, we introduce here a function named PHIB (TR, PR,
OMEGA) for the evaluation of :
PHIB (TR, PR, OMEGA)
It combines Eq. (10.68) with Eqs. (3.61) and (3.62) to evaluate the fugacity coefficient for
given reduced temperature, reduced pressure, and accentric factor. For example, the
value of for carbon dioxide at the conditions of Example 6.4, Step 3, is denoted as:
PHIB (0.963, 0.203, 0.224) = 0.923
EXTENSION TO MIXTURES
The generalized correlation just described is for pure gases only. The remainder of this
section shows how the virial equation may be generalized to allow the calculation of
fugacity coefficients for species in gas mixtures.
The general expression for calculation of ln from second-virial-coefficient data
is given by Eq. (10.64). Values of the pure-species virial coefficients , , etc., are
found from the generalized correlation represented by Eqs. (3.58), (3.59), (3.61), and
(3.62). The cross coefficients Bik, Bij, etc., are found from an extension of the same
correlation. For this purpose, Eq. (3.59) is rewritten in the more general form: (10.69a)

Where

㐲 (10.69b)


And and are the same functions of as given by Eqs. (3.61) and (3.62). The
combining rules proposed by Prausnitz et al. for the calculation of , 댳 and 댳

are:
(10.70) (10.71)

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

댳 댳 댳 䍟 䍟

t댳 R 댳 t댳 t댳
댳 (10.72) t댳 (10.73)

댳 댳
댳 䍟 彣ꯠ 䍟

In Eq. (10.71), is an empirical interaction parameter specific to a particular i-j


molecular pair. When i = j and for chemically similar species, = 0. Otherwise, it is a
small positive number evaluated from minimal PVT data or, in the absence of data, set
equal to zero. When i = j, all equations reduce to the appropriate values for a pure
species. When i ≠ j, these equations define a set of interaction parameters that, while they
do not have any fundamental physical significance, do provide useful estimates of
fugacity coefficients in moderately non-ideal gas mixtures. Reduced temperature is given
for each ij pair by ≡ T/ 댳 . For a mixture, values of from Eq. (10.69b)
substituted into Eq. (10.61) yield the mixture second virial coefficient B and substituted
into Eq. (10.64) [Eqs. (10.63) for a binary] they yield values of ln
The primary virtue of the generalized correlation for second virial coefficients
presented here is simplicity; more accurate, but more complex, correlations appear in the
literature.
EXAMPLE:
Estimate and by Eqs. (10.63) for an equimolar mixture of methyl ethyl ketone
(1)/toluene (2) at 彣°C and 25 kPa. Set all 彣
SOLUTION
The required data are as follows:
ij 댳 K 댳 bar 댳 댳 䁡 t댳 댳

11 535.5 41.5 267. 0.249 0.323


22 591.8 41.1 316. 0.264 0.262
12 563.0 41.3 291. 0.256 1.293

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

where values in the last row have been calculated by Eqs. (10.70) through (10.74). The

values of , together with , , and calculated for each ij pair by Eqs. (3.65),
(3.66), and (10.69), are as follows:
ij 彣
댳 䁡
11 0.603 -0.865 -1.300 -1387.
22 0.546 -1.028 -2.045 -1860.
12 0.574 -0.943 -1.632 -1611.

Calculating according to its definition gives:


ꯠ 彣 댳 䁡
Equation (10.63) then yields:

ln ꯠ 彣 䍟 彣彣

ln 彣 彣 䍟 彣彣 ꯠ

Whence,
彣耀 ꯠ and 彣耀
These results are representative of values obtained for vapor phases at typical conditions
of low-pressure vapor/liquid equilibrium.

IDEAL SOLUTION MODEL


The chemical potential as given by the ideal-gas-state mixture model,
t t t
㐲 ln (10.24)

contains a final term that gives it the simplest possible composition dependence. Indeed,
this functional form could reasonably serve as the basis for the composition dependence
of chemical potential in dense gases and liquids. In it, the composition dependence arises
only from the entropy increase due to random intermixing of molecules of different
species. This entropy increase due to mixing is the same in any random mixture, and thus
can be expected to be present in liquids and dense gases as well. However, the
t
pure-species behavior implied by the term is unrealistic except for the ideal-gas
t
state. A natural extension of Eq. (10.24) therefore replaces with , the
Gibbs energy of pure i in its real physical state of gas, liquid, or solid. Thus, we define an
ideal solution as one for which:

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

㐲 ln
(10.75)

where superscript denotes an ideal-solution property. Mole fraction is here represented


by to reflect the fact that this approach is most often applied to liquids. However, a
consequence of this definition is that an ideal-gas-state mixture is a special case, namely,
an ideal solution of gases in the ideal-gas state, for which xi is replaced by .
All other thermodynamic properties for an ideal solution follow from Eq. (10.75).
The partial volume results from differentiation with respect to pressure at constant
temperature and composition in accord with Eq. (10.18):

By Eq. (10.4), (∂ /∂ 䍟 = ; whence,

(10.76)

Similarly, as a result of Eq. (10.19),

㐲 ln

By Eq. (10.5),
㐲 ln (10.77)

Because t , substitutions by Eqs. (10.75) and (10.77) yield:


t 㐲 ln 㐲 ln
or
t t
(10.78)

The summability relation, Eq. (10.11), applied to the special case of an ideal solution,
is written:

Application to Eqs. (10.75) through (10.78) yields:


㐲 ln 㐲 ln
(10.79) (10.80)

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

(10.81) t t

If in Example 10.3 the solution formed by mixing methanol (1) and water (2) were
assumed ideal, the final volume would be given by Eq. (10.81), and the V-vs.- relation
would be a straight line connecting the pure-species volumes, at 彣 with
at For the specific calculation at 彣 , the use of and in place
of partial volumes yields:
쳌 쳌
耀 彣 ꯠ댳
Both values are about 3.4% low.

The Lewis/Randall Rule


The composition dependence of the fugacity of a species in an ideal solution is
particularly simple. Recall Eqs. (10.46) and (10.31):
Γi 㐲 ln (10.46) Γi 㐲 ln (10.31)

Subtraction yields the general equation:


Γi 㐲 ln t 䍟
For the special case of an ideal solution,
㐲 ln t )

Comparison with Eq. (10.75) gives:

(10.83)

This equation, known as the Lewis/Randall rule, applies to each species in an ideal
solution at all conditions of temperature, pressure, and composition. It shows that the
fugacity of each species in an ideal solution is proportional to its mole fraction; the
proportionality constant is the fugacity of pure species i in the same physical state as the
solution and at the same T and P.
Division of both sides of Eq. (10.83) by P and substitution of for t [Eq.
(10.52)] and of for t [Eq.(10.34)] gives an alternative form:

(10.84)

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Thus, the fugacity coefficient of species i in an ideal solution is equal to the fugacity
coefficient of pure species i in the same physical state as the solution and at the same T
and P. Phases comprised of liquids whose molecules are of similar size and similar
chemical nature approximate ideal solutions. Mixtures of isomers closely conform to
these conditions. Mixtures of adjacent members of homologous series are also examples.

EXCESS PROPERTIES

The residual Gibbs energy and the fugacity coefficient are directly related to
experimental PVT data by Eqs. (6.49), (11.35), and (11.60). Where such data can be
adequately correlated by equations of state, thermodynamic-property information is
advantageously provided by residual properties. Indeed, if convenient treatment of all
fluids by means of equation of state were possible, the thermodynamic-property relations
already presented would suffice. However, liquid solutions are often more easily dealt
with through properties that measure their departures, not from ideal-gas behavior, but
from ideal-solution behavior. Thus the mathematical formalism and of excess properties is
analogous to that of the residual properties.
If M represents the molar (or unit mass) value of any extensive thermodynamic
property (e.g. V, U, H, S, G, etc.), then an excess property ME is defined as the
difference between the actual property value of a solution and the value it would have as
an ideal solution at the same temperature, pressure, and composition. Thus,

For example,
t t t
Moreover, t (11.86)
which follows from Eq. (11.85) and Eq. (6.3), the definition of G.
The definition of ME is analogous to the definition of a residual property as a given
by Eq. (6.41). Indeed, excess properties have a simple relation to residual properties,
found by subtracting Eq. (6.41) from Eq. (11.85):
㐲 t

As already noted, an ideal-gas mixture is an ideal solution of ideal gases. Equations (11.79)
through (11.82) therefore become expressions for Mig when Mi is replaced Miig. Equation
(11.82) becomes Eq. (11.25), Eq. (11.80) becomes Eq. (11.26), and Eq. (11.79) becomes

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Eq. (11.27). The two sets of equations, for Mid and Mig, therefore provide a general
relation for the difference:
t t 㐲

wherein the summation terms with logarithms have canceled. This leads immediately to:
㐲 㐲
(11.87)
Note that excess properties have no meaning for pure species, whereas residual
properties exist for pure species as well as for mixtures.
The partial-property relation analogous to Eq. (11.49) is:
(11.88)
where is a partial excess property. The fundamental excess-property relation is
derived in exactly the same way as the fundamental residual-property relation and leads
to analogous results. Equation (11.54), written for the special case of an ideal solution, is
subtracted from Eq. (11.54) itself yielding:

t
㐲 㐲 㐲 㐲

This is the fundamental excess-property relation, analogous to Eq. (11.55), the fundamental
residual-property relation.
The exact analogy that exists between properties M, residual properties MR, and
excess properties ME is indicated by Table 11.1. All of the equations that appear are basic
property relations, although only Eqs. (11.4) and (11.5) have been shown explicitly
before.

Table 11.1: Summary of Equations for the Gibbs Energy and Related Properties
M in Relation to G MR in Relation to GR ME in Relation to GE

(11.4) 㐲


(11.5) 㐲

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

t t㐲 㐲 㐲
t

㐲 㐲
㐲 㐲 㐲
㐲 㐲

t t㐲 t

Excess-Property Relations
An alternative form of Eq. (11.89) follows by introduction of the activity coefficient
through Eq. (11.91):

t
ln
㐲 㐲 㐲
(11.93)
RT

The generality of theses equations precludes their direct practical application. Rather,
use is made of restricted forms, which is written by inspection:
㐲 t 㐲

(11.94) 㐲
(11.95)


ln
(11.96)

Equations (11.93) through (11.96) are analogs of Eqs. (11.56) through (11.59) for residual
properties. Whereas the fundamental residual-property relation derives its usefulness from
its direct relation to experimental PVT data and equations of state, the fundamental
excess-property relation is useful because ,t , and are all experimentally accessible.
Activity coefficients are found from vapor/liquid equilibrium data, and and
t values come from mixing experiments, topics treated in the following chapter.

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Equation (11.96) demonstrates that ln is a partial property with respect to

㐲 . It is the analog of Eq. (11.59), which shows the same relation of ln to



㐲 . The partial-property analogs of Eqs. (11.94)and (11.95) are:
ln ln t

(11.97) 㐲
(11.97)

These equations allow calculation of the effect of pressure and temperature on the
activity coefficient
The following forms of the summability and Gibbs/Duhem equations result from
the fact that ln is a partial property of 㐲 :

ln (11.99)

ln 彣
(11.100)

These equations find important applications in phase-equilibrium thermodynamics.


Equations (11.94) and (11.95) allow direct calculation of the effects of pressure and
temperature on the excess Gibbs energy. For example, an equimolar mixture of benzene
and cyclohexane at 25°C and 1 bar has an excess volume of about 0.65cm3 mol-1 and an
excess enthalpy of about 800J mol-1. Thus, these conditions,
㐲 彣

耀
㐲 彣彣
彣 彣
耀

The most striking observation about these results is that it takes a pressure change of
more than 40 bar to have an effect on the excess Gibbs energy equivalent to that of a
temperature change of 1 K. Similar calculations based on Eqs. (11.97) and (11.98)
produce similar results.

The Nature of Excess Properties


Peculiarities of liquid-mixture behavior are dramatically revealed in the excess properties.
Those primary interests are , t , and . The excess Gibbs energy comes from
experiment through reduction of vapor/liquid equilibrium data, and t is determined by

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

mixing experiments (Chapter 12). The excess entropy is not measured directly but is
found from Eq. (11.86), written:
t

Excess properties are often strong functions of temperature, but at normal temperatures
are not strongly influenced by pressure. Their composition dependence is illustrated in
Fig. 11.4 for six binary liquid mixtures at 50°C and approximately atmospheric pressure.
For consistency, with Eq. (11.86), the product is shown rather than itself. Although the

systems exhibit a diversity of behavior, they have common features:

1. All excess properties become zero as either species approaches purity.

2. Although vs. is approximately parabolic in shape, both t and


exhibit individualistic composition dependencies.
3. When an excess property has a single sign (as does in all six cases), the
extreme value of (maximum or minimum) often occurs near the equimolar
composition.

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

Feature 1 is a consequence of the definition of an excess property, Eq. (11.85); as any


approaches unity, both and approach , the corresponding property of pure
species . Features 2 and 3 are generalizations based on observation, and admit
exceptions (note, e.g., the behavior of t for the ethanol/water system).

END OF THE MODULE PROBLEM EXERCISES

DIRECTION: Solve the following problems. Round off your final answer to 3
decimal places.

1. What is the change in entropy when 0.6 m3 of CO2 and 0.4 m3 of N2, each at 1 bar and
25°C, blend to form a gas mixture at the same conditions? Assume ideal gases.

2. A vessel, divided into two parts by a partition, contains 5 mol of nitrogen gas at 75°C
and 30 bar on one side and 3.5 mol of argon gas at 130°C and 20 bar on the other. If
the partition is removed and the gases mix adiabatically and completely, what is the
change in entropy? Assume nitrogen to be an ideal gas with CV = (5/2)R and argon to
be an ideal gas with CV = (3/2)R.

3. A stream of nitrogen flowing at a rate of 4 kg·s−1 and a stream of hydrogen flowing


at a rate of 2 kg·s−1 mix adiabatically in a steady-flow process. If the gases are
assumed ideal, what is the rate of entropy increase as a result of the process?

4. What is the ideal work for the separation of an equimolar mixture of methane and
ethane at 155°C and 3 bar in a steady-flow process into product streams of the pure
gases at 40°C and 1 bar if Tσ = 300 K?

5. What is the work required for the separation of air (21-mol-% oxygen and 79-mol-%
nitrogen) at 25°C and 1 bar in a steady-flow process into product streams of pure
oxygen and nitrogen, also at 25°C and 1 bar, if the thermodynamic efficiency of the
process is 5% and if Tσ = 300 K?

Downloaded by Orre Bhanuprasad ([email protected])


lOMoARcPSD|35886966

6. For SO2 at 600 K and 300 bar, determine good estimates of the fugacity and of
GR/RT.

7. Estimate the fugacity of isobutylene as a gas:


(a) At 300°C and 20 bar;
(b) At 300°C and 100 bar.

8. Estimate the fugacity of one of the following:


(a) Cyclopentane at 120°C and 270 bar. At 110°C the vapor pressure of cyclopentane
is 5.267 bar.
(b) 1-Butene at 125°C and 35 bar. At 120°C the vapor pressure of 1-butene is 25.83 bar.

9. Estimate the fugacity of one of the following liquids at its normal-boiling-point


temperature and 220 bar:
(a) n-Pentane;
(b) Isobutylene;
(c) 1-Butene.

10. Two (2) kmol·hr−1 of liquid n-octane (species 1) are continuously mixed with 4
kmol·hr−1 of liquid iso-octane (species 2). The mixing process occurs at constant T and P;
mechanical power requirements are negligible.
(a) Use an energy balance to determine the rate of heat transfer.
(b) Use an entropy balance to determine the rate of entropy generation (W·K−1).
State and justify all assumptions.

REFERENCE:
Smith, J.M. et al., Introduction to Chemical Engineering Thermodynamics, 8th Edition,
McGraw-Hill Companies,Inc.,Singapore,2018.

Downloaded by Orre Bhanuprasad ([email protected])

You might also like