Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

3 Volumetric Properties of Pure Fluids

Property values come from experiment. For fluids, the most comprehensive procedure requires
measurements of molar volume in relation to temperature and pressure. The resulting
pressure/volume/temperature (PVT) data are then correlated by equations of state.

3.1 The phase rule


For a multiphase system at equilibrium, the number of independent variables that must
be fixed to establish its intensive state is called the number of degrees of freedom:

𝐹 = 2−𝜋+𝑁 (3. 1)

where F is the number of degrees of freedom, π is the number of phases, and N is the number
of chemical species present in the system.

A phase is a homogeneous region of matter. Various phases can coexist, but they must be in
equilibrium for the phase rule to apply. A phase can be discontinuous e.g. a gas dispersed as
bubbles in a liquid.

For a pure homogeneous fluid, π = 1, N = 1 and F = 2: the state of the fluid is fixed when two
intensive thermodynamic properties are specified. For example, the state of subcooled water is
determined by specifying its temperature and pressure.

For two phases of the same pure species in equilibrium, π = 2, N = 1 and F = 1: the state of the
system is fixed by only a single property. For example, for steam and liquid water to remain in
equilibrium, only the temperature or pressure needs to be specified.

Three phases of a substance can exist in equilibrium at its triple point, π = 3, N = 1 and F = 0.
This results in an invariant system e.g. the triple point of H2O, where liquid, vapor, and ice
exist together in equilibrium, occurs only at 0.01°C and 0.0061 bar.

Example 3.1
How many phase-rule variables must be specified to fix the thermodynamic state of each of the
following systems?
a) Liquid water in equilibrium with its vapor.
b) Liquid water in equilibrium with a mixture of water vapor and nitrogen.
c) A three-phase system of a saturated aqueous salt solution at its boiling point with excess
salt crystals present.

15
Solution 3.1
a) The system contains a single chemical species existing as two phases (one liquid and one
vapor), and F = 2 - π + N = 2 - 2 + 1 = 1
b) Two chemical species are present. There are two phases: F = 2 - π + N = 2 - 2 + 2 = 2.
c) The three phases (π = 3) are crystalline salt, the saturated aqueous solution, and vapor
generated at the boiling point. The two chemical species (N = 2) are water and salt. For this
system, F = 2 - 3 + 2 = 1

3.2 PVT behaviour of pure substances


PT diagram
Figure 3.1 displays the equilibrium conditions of P and T at which solid, liquid, and gas phases
of a pure substance exist. The three lines display conditions of P and T at which two phases
may coexist, and they divide the diagram into single-phase regions. Line 1-2, the sublimation
curve, separates the solid and gas regions; line 2-3, the fusion curve, separates the solid and
liquid regions; line 2-C, the vaporization curve, separates the liquid and gas regions. Point C
is known as the critical point; its coordinates Pc and Tc are the highest pressure and highest
temperature at which a pure chemical species is observed to exist in vapor/liquid equilibrium.

Figure 3.1: PT diagram for a pure substance.

According to the phase rule the triple point is invariant (F = 0). If the system exists along any
of the two-phase lines, it is univariant (F = 1), whereas in the single-phase regions it is
divariant (F = 2). Invariant, univariant, and divariant states appear as points, curves, and areas,
respectively, on a PT diagram.

16
Changes of state can be represented by lines on the PT diagram: a constant-T change by a
vertical line, and a constant-P change by a horizontal line. When such a line crosses a phase
boundary, an abrupt change in properties of the fluid occurs at constant T and P; for example,
vaporization for the transition from liquid to vapor. Far from point C, the properties of liquid
and vapor are very different. However, if the temperature is raised so that the equilibrium state
progresses upward along curve 2-C, the properties of the two phases become more and more
nearly alike. At point C they become identical and the meniscus disappears.

The fluid region exists at temperatures and pressures greater than Tc and Pc. At these regions,
neither condensation nor vaporization can take place. The gas region is divided into two parts
by the dotted vertical line. A gas to the left of this line, which can be condensed either by
compression or by cooling, is called a vapour. A fluid existing at a temperature greater than Tc
is said to be supercritical e.g. atmospheric air.

PV Diagram
On a PV diagram the phase boundaries in the PT diagram become regions where two phases –
solid/liquid, solid/vapor, and liquid/vapor – coexist in equilibrium. Figure 3.2(b) is an
expanded view of the liquid, liquid/vapor, and vapor regions of the PV diagram, with four
isotherms (paths of constant T). The isotherm labelled T > Tc is smooth and does not cross a
phase boundary.

Figure 3.2: PV diagrams for a pure substance. (a) Showing solid, liquid, and gas regions. (b)
Showing liquid, liquid/vapor, and vapor regions with isotherms.

17
The lines T1 and T2 are for subcritical temperatures and consist of three segments. The
horizontal segment of each isotherm represents mixtures of liquid and vapor in equilibrium,
ranging from 100% liquid at the left to 100% vapor at the right. The horizontal segments of the
isotherms in the two-phase region become progressively shorter at higher temperatures, being
ultimately reduced to a point at C. The critical isotherm Tc, exhibits a horizontal inflection at
the critical point C at the top of the dome, where the liquid and vapour phases become
indistinguishable.

The dome from B to C represents single-phase liquids at their vaporization (boiling)


temperatures while from C to D represents single-phase vapours at their condensation
temperatures. Liquids and vapours represented by BCD are said to be saturated. The
subcooled-liquid region lies to the left of the saturated-liquid curve BC, and the superheated-
vapour region lies to the right of the saturated-vapour curve CD. Subcooled liquid exists at
temperatures below, and superheated vapor, at temperatures above the boiling point for the
given pressure.

3.3 Equation of state for liquids


For a single phase, the relation 𝑓(𝑃, 𝑉, 𝑇) = 0 is known as the PVT equation of state. If V is
considered a function of T and P, then 𝑉 = 𝑉(𝑇, 𝑃), and

𝜕𝑉 𝜕𝑉
𝑑𝑉 = ( ) 𝑑𝑇 + ( ) 𝑑𝑃 (3. 2)
𝜕𝑇 𝑃 𝜕𝑃 𝑇

The partial derivatives in this equation are related to two properties, commonly tabulated for
liquids, and defined as follows:

Volume expansivity:

1 𝜕𝑉
𝛽= ( ) (3. 3)
𝑉 𝜕𝑇 𝑃

Isothermal compressibility:

1 𝜕𝑉
𝜅=− ( ) (3. 4)
𝑉 𝜕𝑃 𝑇

Combining Eqs. (3.2) through (3.4) yields:

𝑑𝑉 = 𝛽𝑉 𝑑𝑇 − 𝜅𝑉 𝑑𝑃

18
𝑑𝑉
= 𝛽 𝑑𝑇 − 𝜅 𝑑𝑃 (3. 5)
𝑉

At conditions not close to the critical point, β and κ are weak functions of temperature and
pressure. Thus, for small changes in T and P, little error is introduced if they are assumed
constant. Integration of eq. (3.5) then yields:

𝑉2 𝑇2 𝑃2
𝑑𝑉
∫ = 𝛽 ∫ 𝑑𝑇 − 𝜅 ∫ 𝑑𝑃
𝑉1 𝑉 𝑇1 𝑃1

0 𝑉2 0 𝑇2 0 𝑃2
[ln 𝑉 ] 0 = 𝛽 [𝑇] 0 − 𝜅 [𝑃] 0
0 𝑉1 0 𝑇1 0 𝑃1

ln 𝑉2 − ln 𝑉1 = 𝛽(𝑇2 − 𝑇1 ) − 𝜅(𝑃2 − 𝑃1 )

𝑉2
ln = 𝛽(𝑇2 − 𝑇1 ) − 𝜅(𝑃2 − 𝑃1 ) (3. 6)
𝑉1

Example 3.2
For liquid acetone at 20°C and 1 bar,
β = 1.487 × 10-3 °C-1 κ = 62 × 10-6 bar-1 V = 1.287 cm3⋅g-1
For acetone, find:
a) The value of (𝝏𝑷⁄𝝏𝑻)𝑽 at 20°C and 1 bar.
b) The pressure after heating at constant V from 20°C and 1 bar to 30°C.
c) The volume change when T and P go from 20°C and 1 bar to 0°C and 10 bar.
Solution 3.2
(a) The derivative (𝜕𝑃⁄𝜕𝑇 )𝑉 is evaluated when 𝑉 is constant and therefore 𝑑𝑉 = 0:
𝛽 𝑑𝑇 − 𝜅 𝑑𝑃 = 0 (const V)
𝜅 𝑑𝑃 = 𝛽 𝑑𝑇
𝑑𝑃 𝛽 1.487 × 10−3 °C−1
= = = 24 bar °C−1
𝑑𝑇 𝜅 62 × 10−6 bar −1
(b) If β and κ are assumed constant in the 10°C temperature interval, then for constant volume, the
previous equation can be rewritten as:
𝑑𝑃 ∆𝑃 𝑃2 − 𝑃1 𝛽
= = =
𝑑𝑇 ∆𝑇 𝑇2 − 𝑇1 𝜅
𝛽
𝑃2 = (𝑇 − 𝑇1 ) + 𝑃1
𝜅 2
𝑃2 = 24 bar °C−1 (30°C − 20°C) + 1 = 241 bar

19
(c) Direct substitution into eq. (3.6) gives:
𝑉2
ln = 𝛽(𝑇2 − 𝑇1 ) − 𝜅(𝑃2 − 𝑃1 )
𝑉1
𝑉2
ln = 1.487 × 10−3 °C−1 (0°C − 20°C) − 62 × 10−6 bar −1 (10 bar − 1 bar )
𝑉1
𝑉2
ln = [1.487 × 10−3 (−20)] − [62 × 10−6 (9)] = −0.0303
𝑉1
𝑉2
= 0.9702
𝑉1
𝑉2 = (0.9702)(𝑉1 ) = (0.9702)(1.287 cm3 ⋅ g −1 ) = 1.249 cm3 ⋅ g −1
∆𝑉 = 𝑉2 − 𝑉1 = 1.249 − 1.287 = −0.038 cm3 ⋅ g −1

Both (𝜕𝑉⁄𝜕𝑇)𝑃 and (𝜕𝑉⁄𝜕𝑃)𝑇 are small. Hence, both β and κ are small. This characteristic
behaviour of liquids has led to the assumption of liquids as an incompressible fluid, for which
both β and κ are zero. No real fluid is truly incompressible, but the idealization is useful,
because it provides a sufficiently realistic model of liquid behaviour for many practical
purposes e.g. fluid mechanics.

3.4 Ideal gas and ideal-gas state


The equation PV = RT is the ideal gas law. It is only valid for pressures approaching zero and
temperatures approaching infinity. At these limits, the space between gas molecules tend
towards infinity and the forces of attraction between them tend towards zero. This leads to the
concept of a hypothetical state of matter referred to as the ideal-gas state. It is the state of a gas
comprised of real molecules that have negligible molecular volume and no intermolecular
forces at all temperatures and pressures. Two equations are fundamental to this state, namely
the “ideal-gas law” and an expression showing that internal energy depends on temperature
alone:

The equation of state:

𝑃𝑉 𝑖𝑔 = 𝑅𝑇 (3. 7)

Internal energy:

𝑈 𝑖𝑔 = 𝑈(𝑇) (3. 8)

The superscript ig denotes properties for the ideal-gas state.

20
3.5 Virial equations of state
The virial equations of state provide a compact and convenient expression of the functional
relation 𝑓(𝑃, 𝑉, 𝑇) = 0 for real gases. The isotherms of vapours and gases to the right of the
saturated vapour line (Figure 3.3) can be represented as a power series of PV in P as:

𝑃𝑉 = 𝑎(1 + 𝐵′ 𝑃 + 𝐶 ′ 𝑃2 + 𝐷 ′ 𝑃3 +. . . ) (3. 9)

where B′, C′, etc., are constants which depend on temperature and substance.

At zero pressure limit, eq. (3.9) reduces to PV = a. The term a is the same function of T for all
gases; therefore, at the zero-pressure limit:

(𝑃𝑉)𝑃→0 = 𝑎 = 𝑓(𝑇) (3. 10)

Introducing the universal gas constant R as the proportionality constant:

(𝑃𝑉)𝑃→0 = 𝑎 = 𝑅𝑇 (3. 11)

A useful auxiliary thermodynamic property, compressibility factor, is defined by the equation:

𝑃𝑉 𝑉
𝑍= = 𝑖𝑔 (3. 12)
𝑅𝑇 𝑉

It is a measure of the deviation of the real-gas molar volume from its ideal-gas value. For the
ideal-gas state, Z = 1.

Eq. (3.9) can be written as:

𝑃𝑉
= 1 + 𝐵′ 𝑃 + 𝐶 ′ 𝑃2 + 𝐷 ′ 𝑃3 +. . .
𝑎

With a = RT:

𝑃𝑉
= 1 + 𝐵′ 𝑃 + 𝐶 ′ 𝑃2 + 𝐷 ′ 𝑃3 +. . .
𝑅𝑇

With Z defined by eq. (3.12):

𝑍 = 1 + 𝐵′ 𝑃 + 𝐶 ′ 𝑃2 + 𝐷 ′ 𝑃3 + . . . (3. 13)

Z can also been expressed as:

𝐵 𝐶 𝐷
𝑍 =1+ + 2 + 3 +. . . (3. 14)
𝑉 𝑉 𝑉

21
The two sets of coefficients in Eqs. (3.13) and (3.14) are related as follows:

𝐵 𝐶 − 𝐵2 𝐷 − 3𝐵𝐶 − 2𝐵3
𝐵′ = 𝐶′ = 𝐷′ = (3. 15)
𝑅𝑇 (𝑅𝑇)2 (𝑅𝑇)3

Equation (3.13) and (3.14) are known as virial expansions, and the parameters B′, C′, D′, etc.,
and B, C, D, etc., are called virial coefficients. Parameters B′ and B are second virial
coefficients; C′ and C are third virial coefficients, and so on. For a given gas the virial
coefficients are functions of temperature only. Virial coefficients have physical meanings in
statistical mechanics: the B/V term is due to interactions between pairs of molecules while the
C/V2 term accounts of three-body interactions, etc. Since two-body interactions are many times
more common than three-body interactions, and three-body interactions are many times more
numerous than four-body interactions, the contributions to Z of the successively higher-ordered
terms decrease rapidly.

Figure 3.4: Compressibility-factor graph for methane.

For gases and vapours at low to moderate pressures, the infinite series virial equations can be
truncated to two or three terms. Figure 3.3 shows a compressibility-factor graph for methane.
All isotherms are nearly straight lines at low pressures (generally below 50 bars). At these
pressures, eq. (3.13) can be truncated to two terms as:

𝑍 = 1 + 𝐵′ 𝑃

Substituting for B′ by eq. (3.15a) yields:

𝑃𝑉 𝐵𝑃
𝑍= =1+ (3. 16)
𝑅𝑇 𝑅𝑇

22
The second virial coefficient B depends on the substance and temperature. Experimental
values are available for a number of gases. Where no data are available, the second virial
coefficients can be estimated as will be discussed later.

For pressures above the range of applicability of eq. (3.16) but below the critical pressure, the
virial equation is still accurate when truncated to three terms as:

𝑃𝑉 𝐵 𝐶
𝑍= =1+ + 2 (3. 17)
𝑅𝑇 𝑉 𝑉

In eq. (3.17), the value of V is normally obtained using iteration. Values of C also depend on
the gas and on temperature. Much less data available of the third virial coefficients as
compared to the second virial coefficients. Figure 3.4 illustrates the effect of temperature on
the virial coefficients B and C for nitrogen. Coefficient B increases with T; however, at
temperatures much higher than shown, B reaches a maximum and then slowly decreases.
Coefficient C is negative at low temperatures, passes through a maximum at a temperature near
the critical temperature, and thereafter decreases slowly with increasing T.

Figure 3.5: Virial coefficients B and C for nitrogen.

Example 3.3
Reported values for the virial coefficients of isopropanol vapor at 200°C are:
B = -388 cm3·mol-1 C = -26,000 cm6·mol-2
Calculate V and Z for isopropanol vapor at 200°C and 10 bar:
(a) For the ideal-gas state; (b) By the two term virial equation; (c) By the three term virial
equation. Gas constant is R = 83.14 bar·cm3·mol−1·K−1.
Solution 3.3
B = -388 cm3·mol-1 C = -26,000 cm6·mol-2 T = 200°C = 473.15 K P = 10 bar
R = 83.14 bar·cm3·mol−1·K−1.

23
(a) For the ideal gas state, Z = 1 and
𝑅𝑇 83.14 × 473.15
𝑉 𝑖𝑔 = = = 3934 cm3 ∙ mol−1
𝑃 10
(b) Two term virial equation:
𝑃𝑉 𝐵𝑃
𝑍= =1+
𝑅𝑇 𝑅𝑇
𝑅𝑇 𝐵𝑃
𝑉= (1 + )
𝑃 𝑅𝑇
𝑅𝑇 83.14 × 473.15
𝑉= +𝐵 = − 388 = 3546 cm3 ∙ mol−1
𝑃 10
𝑉 3546
𝑍 = 𝑖𝑔 = = 0.9014
𝑉 3934
(c) Three term virial equation:
𝑃𝑉 𝐵 𝐶
𝑍= =1+ + 2
𝑅𝑇 𝑉 𝑉
𝑅𝑇 𝐵 𝐶 83.14 × 473.15 388 26,000
𝑉= (1 + + 2 ) = (1 − − )
𝑃 𝑉 𝑉 10 𝑉 𝑉2
388 26,000
𝑉 = 3934 (1 − − ) (𝐴)
𝑉 𝑉2
Solving by iteration by setting 𝑉0 = 3934 cm3 ∙ mol−1 from ideal gas law:
388 26,000
𝑉𝑖+1 = 3934 (1 − − )
𝑉𝑖 𝑉𝑖2
i Vi Vi+1
0 3934 3496
1 3496 3489
2 3489 3488
3 3488 3488

Therefore: 𝑉 = 3488 cm3 ∙ mol−1

𝑉 can also be obtained analytically by converting the virial equation (A) to a cubic equation:
1,526,392 102,284,000
𝑉 = 3934 − −
𝑉 𝑉2
𝑉 3 = 3934 𝑉 2 − 1,526,392 𝑉 − 102,284,000
𝑉 3 − 3934 𝑉 2 + 1,526,392 𝑉 + 102,284,000 = 0
𝑉1 = −58 𝑉2 = 504 V3 = 3488
Therefore: 𝑉 = 3488 cm3 ∙ mol−1

𝑉 3488
𝑍= 𝑖𝑔
= = 0.8866
𝑉 3934

24
3.6 Cubic equations of state
Polynomial equations that are cubic in molar volume are the simplest equations capable of
representing the PVT behaviour of both liquids and vapours for a wide range of temperatures,
pressures, and molar volumes.

The van der Waals Equation of State


The first practical cubic equation of state was the van der Waals equation of state:

𝑅𝑇 𝑎
𝑃= − 2 (3. 18)
𝑉−𝑏 𝑉

Here, a and b are positive constants which depend on a particular species. When they are zero,
the equation becomes the ideal-gas equation. The term a/V2 accounts for the attractive forces
between molecules which makes the pressure lower than in the ideal-gas state. Constant b
accounts for the finite size of molecules, which makes the volume larger than in the ideal-gas
state.

Given values of a and b for a particular fluid, one can calculate P as a function of V for various
values of T. Figure 3.5 is a schematic PV diagram showing three such isotherms. The equation
of state gives similar values to experimental measurements for isotherms at Tc and at T > Tc.
At T < Tc, the equation of state isotherm after crossing the saturated-liquid line, goes through
a minimum, rises to a maximum, and then decreases before crossing the saturated-vapor line.
The line follows a series of non-equilibrium or metastable states. In contrast, experimental
isotherms contain a horizontal segment within the two-phase region where saturated liquid and
saturated vapour coexist.

Figure 3.6: PV isotherms as given by a cubic equation of state

25
Cubic equations of state have three volume roots, of which two may be complex. Physically
meaningful values of V are always real, positive, and greater than constant b.

• For an isotherm at T > Tc, solution for V at any value of P yields only one root.
• For the critical isotherm (T = Tc), solution for V yields one root, except at the critical
pressure, where there are three roots, all equal to Vc.
• For isotherms at T < Tc, the equation may exhibit one or three real roots, depending on the
pressure. At the saturation pressure Psat, two of the roots represent the Vsat(liq) and
Vsat(vap), at the ends of the horizontal isotherm. The third root represents an unstable state
V between Vsat(liq) and Vsat(vap) and may not be equal to the stable experimental value. For
any pressure other than Psat, there is only a single physically meaningful root for a liquid or
a vapour molar volume.

A Generic Cubic Equation of State


The development of several cubic equations of state have produced a class of equations,
referred to as a generic cubic equation of state:

𝑅𝑇 𝑎(𝑇)
𝑃= − (3. 19)
𝑉 − 𝑏 (𝑉 + 𝜀𝑏)(𝑉 + 𝜎𝑏)

The assignment of appropriate parameters leads to the van der Waals (vdW), Redlich/Kwong
(RK), Soave/Redlich/Kwong (SRK) and the Peng/Robinson (PR) equations (Table 3.1).

For a given equation, ɛ and σ are pure numbers, the same for all substances, whereas parameters
a(T) and b are substance dependent. The temperature dependence of a(T) is specific to each
equation of state. The parameters b and a(T) are usually found from values for the critical
constants Tc and Pc and expressed as:

𝑅𝑇𝑐
𝑏=𝛺 (3. 20)
𝑃𝑐

𝛼(𝑇𝑟 , 𝜔)𝑅 2 𝑇𝑐2


𝑎(𝑇) = 𝛹 (3. 21)
𝑃𝑐

In these equations Ω and Ψ are pure numbers, independent of substance but specific to
a particular equation of state. Function α(Tr; ω) is an empirical expression, wherein by
definition Tr ≡ T/Tc, and ω is a parameter specific to a given chemical species.

26
Table 3.1: Parameter Assignments for Equations of State
Eqn. of state 𝛼(𝑇𝑟 ) 𝜎 𝜀 𝛺 𝛹
vdW (1873) 1 0 0 1/8 27/64
RK (1949) 𝑇𝑟−0.5 1 0 0.08664 0.42748
SRK (1972) 𝛼𝑆𝑅𝐾 (𝑇𝑟 , 𝜔) 1 0 0.08664 0.42748
PR (1976) 𝛼𝑃𝑅 (𝑇𝑟 , 𝜔) 1 + √2 1 − √2 0.07780 0.45724
𝛼𝑆𝑅𝐾 (𝑇𝑟 , 𝜔) = [1 + (0.480 + 1.574𝜔 − 0.176𝜔2 )(1 − 𝑇𝑟0.5 )]2
𝛼𝑃𝑅 (𝑇𝑟 , 𝜔) = [1 + (0.37464 + 1.54226𝜔 − 0.26992𝜔2 )(1 − 𝑇𝑟0.5 )]2
Example 3.4
Given that the vapor pressure of n-butane at 350 K is 9.4573 bar, find the molar volumes of (a)
saturated-vapor and (b) saturated-liquid n-butane at these conditions as given by the
Redlich/Kwong equation. Values of Tc and Pc for n-butane: 𝑻𝒄 = 𝟒𝟐𝟓. 𝟏 𝐊 𝑷𝒄 = 𝟑𝟕. 𝟗𝟔 𝐛𝐚𝐫.
3 −1 −1
Gas constant is R = 83.14 bar·cm ·mol ·K .
Solution 3.4
𝑇 = 350 K 𝑃 = 9.4573 bar 𝑇𝑐 = 425.1 K 𝑃𝑐 = 37.96 bar
R = 83.14 bar · cm3 · mol−1 · K −1
Redlich-Kwong equation:
𝑅𝑇 𝑎(𝑇) 𝑅𝑇 𝑎(𝑇)
𝑃= − = −
𝑉 − 𝑏 (𝑉 + 𝜀𝑏)(𝑉 + 𝜎𝑏) 𝑉 − 𝑏 𝑉(𝑉 + 𝑏)
𝑅𝑇𝑐 83.14 × 425.1
𝑏=𝛺 = 0.08664 = 80.67
𝑃𝑐 37.96
𝛼(𝑇𝑟 , 𝜔)𝑅 2 𝑇𝑐2 𝑇𝑟−0.5 𝑅2 𝑇𝑐2 (𝑇/𝑇𝑐 )−0.5 𝑅2 𝑇𝑐2
𝑎(𝑇) = 𝛹 =𝛹 =𝛹
𝑃𝑐 𝑃𝑐 𝑃𝑐
(350/425.1)−0.5 × 83.142 × 425.12
𝑎(𝑇) = 0.42748 = 15,502,558
37.96
𝑅𝑇 𝑎(𝑇)
𝑃= −
𝑉 − 𝑏 𝑉(𝑉 + 𝑏)
83.14 × 350 15,502,558
9.4573 = −
𝑉 − 80.67 𝑉(𝑉 + 80.67)
29,099 15,502,558
9.4573 = − 2
𝑉 − 80.67 (𝑉 + 80.67𝑉)
9.4573(𝑉 − 80.67)(𝑉 2 + 80.67𝑉) = 29,099(𝑉 2 + 80.67𝑉) − 15,502,558(𝑉 − 80.67)
29,099 2 15,502,558
(𝑉 − 80.67)(𝑉 2 + 80.67𝑉) = (𝑉 + 80.67𝑉) − (𝑉 − 80.67)
9.4573 9.4573
𝑉 3 + 80.67𝑉 2 − 80.67𝑉 2 − 6,508𝑉 = 3,076𝑉 2 + 248,212𝑉 − 1,639,216𝑉 + 132,235,560
𝑉 3 + (80.67 − 80.67 − 3,076)𝑉 2 + (−6,508 − 248,212 + 1,639,216)𝑉 − 132,235,560 = 0
𝑉 3 − 3,076𝑉 2 + 1,384,496𝑉 − 132,235,560 = 0
𝑉 = 133 cm3 ; 389 cm3 ; 2554 cm3

27
Between saturated liquid and saturated vapour, a cubic equation gives three molar volumes; the
lowest value at the saturated liquid while the highest value is at saturated vapour. Therefore,
(𝑎) 𝑉 𝑔 = 2554 cm3
(𝑏) 𝑉 𝑙 = 133 cm3
For reference, the experimental value of 𝑉 𝑔 = 2482 cm3 and 𝑉 𝑙 = 115 cm3
The values obtained from the other cubic equations of state are indicated below:
𝑉𝑔 𝑉𝑙
Exp. vdW RK SRK PR Exp. vdW RK SRK PR
2482 2667 2555 2520 2486 115.0 191.0 133.3 127.8 112.6

Acentric Factor (ω)


The acentric factor is used in the Soave/Redlich/Kwong (SRK) and the Peng/Robinson (PR)
equations. Its derivation is described below.

It has been found that the logarithm of the reduced vapour pressure of a pure fluid is
approximately linear in the reciprocal of reduced temperature.

𝑑 log 𝑃𝑟sat
=𝑆 (3. 22)
𝑑(1⁄𝑇𝑟 )

where 𝑃𝑟sat is reduced vapor pressure, 𝑇𝑟 is reduced temperature, and S is the slope of the plot.

Figure 3.7: Approximate temperature dependence of the reduced vapour pressure.

Each fluid has its own characteristic value of S. All vapor-pressure data for the simple fluids
(Ar, Kr, Xe) lie on the same line when plotted as log 𝑃𝑟sat vs. 1⁄𝑇𝑟 and the line passes through
log 𝑃𝑟sat = −1.0 at 𝑇𝑟 = 0.7 (Figure 3.6). Data for other fluids lie on other lines whose
locations can be determined from the simple fluids (SF) line by the difference:
log 𝑃𝑟sat (SF) − log 𝑃𝑟sat .

28
The acentric factor (ω) is then defined as this difference evaluated at 𝑇𝑟 = 0.7:

𝜔 = log(𝑃𝑟sat [SF]) 𝑇𝑟=0.7 − log(𝑃𝑟sat ) 𝑇𝑟=0.7

𝜔 = −1.0 − log(𝑃𝑟sat ) 𝑇𝑟 =0.7 (3. 23)

The acentric factor can be determined for any fluid from Tc, Pc, and a single vapor-pressure
measurement made at 𝑇𝑟 = 0.7. Values of ω and the critical constants Tc, Pc, and Vc are
available for a number of substances.

3.7 Generalized correlations for gases


The most popular generalized correlations are for the compressibility factor Z and for the
second virial coefficient B.

Pitzer Correlations for the Compressibility Factor


The correlation for Z is:

𝑍 = 𝑍 0 + 𝜔𝑍1 (3. 24)

where 𝑍 0 and 𝑍1 are functions of both Tr and Pr. Values of Z0 and Z1 as functions of Tr and Pr
are provided by Lee/Kesler Generalized-Correlation Tables.

• Lee/Kesler tables contain values for liquid and vapour phases. However, the boundary
between these phases may not be the same as the experimental saturation curve. As a
result, these tables should not be used for properties near the saturation curve.
• The Lee/Kesler correlation provides reliable results for gases which are nonpolar or
only slightly polar. For highly polar gases or to gases that associate, larger errors can
be expected.

Pitzer Correlations for the Second Virial Coefficient


The tabular nature of the generalized compressibility-factor correlation is a disadvantage. An
approximate analytical expression can be obtained for a limited range of pressures using the
virial equation:

𝐵𝑃 𝐵 𝑃𝑐 𝑃𝑟 𝐵𝑃𝑐 𝑃𝑟
𝑍 =1+ =1+( ∙ )= 1+( )
𝑅𝑇 𝑅 𝑇𝑐 𝑇𝑟 𝑅𝑇𝑐 𝑇𝑟

𝑃𝑟
𝑍 = 1 + 𝐵̂ (3. 25)
𝑇𝑟

29
The reduced second virial coefficient is given as:

𝐵𝑃𝑐
𝐵̂ = (3. 26)
𝑅𝑇𝑐

The Pitzer correlation for it is:

𝐵̂ = 𝐵0 + 𝜔𝐵1 (3. 27)

Combining eq. (3.25) and eq. (3.27) gives:

𝑃𝑟
𝑍 = 1 + (𝐵0 + 𝜔𝐵1 ) (3. 28)
𝑇𝑟

Comparison of this equation with eq. (3.24) provides the following identities:

𝑃𝑟 𝑃𝑟
𝑍 0 = 1 + 𝐵0 𝑍1 = 𝐵1 (3. 29)
𝑇𝑟 𝑇𝑟

Second virial coefficients are functions of temperature only, and similarly B0 and B1 are
functions of reduced temperature only. They are adequately represented by the Abbott
equations:

0.422
𝐵0 = 0.083 − (3. 30)
𝑇𝑟1.6

0.172
𝐵1 = 0.139 − (3. 31)
𝑇𝑟4.2

The simplest form of the virial equation has validity only at low to moderate pressures where
Z is linear in pressure. The generalized virial-coefficient correlation is therefore useful only
where Z0 and Z1 are approximately linear functions of reduced pressure. Figure 3.7 compares
the linear relation of Z0 by eq. (3.29) and that from the Lee/Kesler Tables. The two correlations
are very close in the region above the dashed line. For reduced temperatures greater than Tr ≈
3, all reduced pressures give accurate values. For lower values of Tr the allowable pressure
range decreases with decreasing temperature up to a Tr ≈ 0.7 at the saturation pressure.

Like the parent correlation, it is most accurate for nonpolar species and least accurate for highly
polar and associating molecules.

30
Figure 3.8: Comparison of correlations for Z0

Example 3.5
Determine the molar volume of n-butane at 510 K and 25 bar based on each of the following:
a) The ideal-gas state.
b) The generalized compressibility-factor correlation.
̂.
c) Equation (3.28), with the generalized correlation for 𝑩
Solution 3.5
𝑇 = 510 K 𝑃 = 25 bar
(a) The ideal gas equation:
𝑅𝑇 83.14 × 510
𝑉= = = 1,696 cm3 /mol
𝑃 25
(b) Pitzer Correlations for the Compressibility Factor:
𝑍 = 𝑍 0 + 𝜔𝑍1
For n-butane, 𝑇𝑐 = 425.1 K 𝑃𝑐 = 37.96 bar ω = 0.200
𝑇 510 𝑃 25
𝑇𝑟 = = = 1.2 𝑃𝑟 = = = 0.66
𝑇𝑐 425.1 𝑃𝑐 37.96
Lee/Kesler tables:
𝑃𝑟 = 0.6 𝑃𝑟 = 0.66 𝑃𝑟 = 0.8
𝑇𝑟 = 1.2 Z0 = 0.8779 ? Z0 = 0.8330
𝑇𝑟 = 1.2 Z1 = 0.0326 ? Z1 = 0.0499

31
Interpolate for Z0 at 𝑃𝑟 = 0.66:
0
𝑍𝑃𝑟=0.66 − 0.8779 0.8330 − 0.8779
= = −0.2245
0.66 − 0.6 0.8 − 0.6
0
𝑍𝑃𝑟=0.66 = 0.8779 − 0.2245(0.66 − 0.6) = 0.86443

Interpolate for Z1 at 𝑃𝑟 = 0.66:


1
𝑍𝑃𝑟=0.66 − 0.0326 0.0499 − 0.0326
= = 0.0865
0.66 − 0.6 0.8 − 0.6
1
𝑍𝑃𝑟=0.66 = 0.0326 + 0.0865(0.66 − 0.6) = 0.03779
Hence: Z0 = 0.86443 Z1 = 0.03779
𝑍 = 𝑍 0 + 𝜔𝑍1 = 0.86443 + (0.200 × 0.03779) = 0.872
𝑉𝑃
𝑍=
𝑅𝑇
𝑍𝑅𝑇 83.14 × 510
𝑉= = 0.872 × = 1479 cm3 /mol
𝑃 25
(c) Two term virial equation:
𝑃𝑟
𝑍 = 1 + (𝐵0 + 𝜔𝐵1 )
𝑇𝑟
where:
0.422 0.422
𝐵0 = 0.083 − 1.6 = 0.083 − 1.21.6 = −0.232
𝑇𝑟

0.172 0.172
𝐵1 = 0.139 − 4.2 = 0.139 − = 0.059
𝑇𝑟 1.24.2
0.66
𝑍 = 1 + (−0.232 + 0.200 × 0.059) = 0.879
1.2
𝑍𝑅𝑇 83.14 × 510
𝑉= = 0.879 × = 1491 cm3 /mol
𝑃 25
An experimental value for 𝑉 = 1480.7 cm3 /mol

3.8 Generalized correlations for liquids


Use of the cubic equations of state to calculate the molar volumes of liquids is not very accurate.
Lee/Kesler correlations can be used, but they are only suitable for nonpolar and slightly polar
fluids. For other fluids, generalized Rackett equation can be used for the estimation of molar
volumes of saturated liquids:
2
sat (1−𝑇 )7
𝑉 = 𝑉𝑐 𝑍𝑐 𝑟 (3. 32)

32
If the value of the critical volume (𝑉𝑐 ) is known, Lydersen, Greenkorn, and Hougen correlation
may be used to estimate the liquid volume:

𝜌 𝑉𝑐
𝜌𝑟 = =
𝜌𝑐 𝑉

Which can be rewritten as:

𝑉𝑐
𝑉= (3. 33)
𝜌𝑟

where 𝜌𝑟 is the reduced density and 𝜌𝑐 is the density at the critical point.

Alternatively, if the liquid volume at one state is known (𝑉1), the liquid volume at another state
(𝑉2) can be determined by the identity:

𝜌𝑟1
𝑉2 = 𝑉1 (3. 34)
𝜌𝑟2

Plots of reduced density (𝜌𝑟1 , 𝜌𝑟2 ) as a function of reduced pressure and reduced temperature
are available (Figure 3.8).

Figure 3.9: Generalized density correlation for liquids.

33
Example 3.6
For ammonia at 310 K, estimate the density of:
(a) The saturated liquid;
(b) The liquid at 100 bar.
Solution 3.6
(a) For saturated liquid, we use Rackett equation:
2
sat (1−𝑇 )7
𝑉 = 𝑉𝐶 𝑍𝐶 𝑟
For ammonia: Vc = 72.5 cm3/mol Zc = 0.242 Tc = 405.7 K
𝑇 310
𝑇𝑟 = = = 0.764
𝑇𝑐 405.7
2
(1−𝑇𝑟 )7 2/7
𝑉 sat = 𝑉𝐶 𝑍𝐶 = 72.5 × 0.242(1−0.764) = 28.34 cm3 /mol
For comparison, the experimental value is 29.14 cm3mol−1, a 2.7% difference.
(b) For subcooled liquid, we use the correlation:
𝑉𝑐
𝑉=
𝜌𝑟
For ammonia: Pc = 112.80 bar
𝑃 100
𝑃𝑟 = = = 0.887
𝑃𝑐 112.80
From Figure 3.8, 𝜌𝑟 ≈ 2.4 at 𝑇𝑟 = 0.764 and 𝑃𝑟 = 0.887.
𝑉𝑐 72.5
𝑉= = = 30.21 cm3 /mol
𝜌𝑟 2.4
In comparison with the experimental value of 28.6 cm3mol–1, this result is higher
by 6.5%.

This question can also be solved using eq. (3.34) starting with 28.34 cm3 /mol for the saturated
liquid:
𝑉1 = 28.34 cm3 /mol
For the saturated liquid ammonia, 𝑇𝑟 = 0.764, 𝜌𝑟1 ≈ 2.3
For liquid ammonia, 𝜌𝑟2 ≈ 2.4
2.3
𝑉2 = 28.34 = 27.15 cm3 /mol
2.4

3.9 Conclusion
You need to intelligently select an appropriate equation of state or generalized correlation for
application in a given situation, as indicated by the following chart:

34
(a) Ideal-gas state Do not use unless you are told

Use if B is given. Use for low pressures (below


(b) 2-term virial equation
50 bar)

(c) 3-term virial equation Use if B and C are given

Gas
Use only if you cannot use the other EOS. Do not
(d) Cubic equation of state
use for liquids unless told.

Use for non-polar molecules. Use for values far


(e) Lee/Kesler Tables
from the saturation line. Use if table is available

Gas or Liquid Use for non-polar molecules. Check applicability


(f) Abbott equations
using Fig. 3.7

(a) Incompressible liquid Do not use unless told

(b) Constant β and κ Use if β and κ are provided


Liquid

(c) Racket equation Use for saturated liquids

(d) Lydersen equation and chart Use for subcooled liquids

35

You might also like