0% found this document useful (0 votes)
594 views65 pages

Volumetric Properties of Pure Fluids

This chapter discusses volumetric properties of pure fluids. It begins by presenting the phase rule, which relates the number of phases, chemical species, and degrees of freedom in a thermodynamic system. It then provides a qualitative description of pressure-volume-temperature behavior for pure substances using a phase diagram. Key features include vapor pressure curves separating gas, liquid, and solid phases; the critical point; and the triple point where three phases coexist. The chapter introduces the treatment of the ideal gas state and equations of state to model real fluid behavior and predict thermodynamic properties.

Uploaded by

syayaj dhini
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)
594 views65 pages

Volumetric Properties of Pure Fluids

This chapter discusses volumetric properties of pure fluids. It begins by presenting the phase rule, which relates the number of phases, chemical species, and degrees of freedom in a thermodynamic system. It then provides a qualitative description of pressure-volume-temperature behavior for pure substances using a phase diagram. Key features include vapor pressure curves separating gas, liquid, and solid phases; the critical point; and the triple point where three phases coexist. The chapter introduces the treatment of the ideal gas state and equations of state to model real fluid behavior and predict thermodynamic properties.

Uploaded by

syayaj dhini
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

Chapter 3

Volumetric Properties
of Pure Fluids

The equations of the preceding chapter provide the means for calculation of the heat and
work quantities associated with various processes, but they are useless without knowledge of
property values for internal energy or enthalpy. Such properties differ from one substance to
another, and the laws of thermodynamics themselves do not provide any description or model
of material behavior. Property values come from experiment, or from the correlated results of
experiment, or from models grounded in and validated by experiment. Because there are no
internal-energy or enthalpy meters, indirect measurement is the rule. For fluids, the most com-
prehensive procedure requires measurements of molar volume in relation to temperature and
pressure. The resulting pressure/volume/temperature (PVT ) data are most usefully correlated
by equations of state, through which molar volume (or density), temperature, and pressure are
functionally related.
In this chapter we:

∙ Present the phase rule, which relates the number of independent variables required to fix
the thermodynamic state of a system to the number of chemical species and phases present
∙ Describe qualitatively the general nature of PVT behavior of pure substances
∙ Provide a detailed treatment of the ideal-gas state
∙ Treat equations of state, which are mathematical formulations of the PVT behavior of fluids
∙ Introduce generalized correlations that allow prediction of the PVT behavior of fluids
for which experimental data are lacking

3.1  THE PHASE RULE

As indicated in Section 2.5, the state of a pure homogeneous fluid is fixed whenever two
intensive thermodynamic properties are set at specific values. In contrast, when two phases
of the same pure species are in equilibrium, the state of the system is fixed when only a sin-
gle property is specified. For example, a system of steam and liquid water in equilibrium at

68
3.1.  The Phase Rule 69

101.33 kPa can exist only at 100°C. It is impossible to change the temperature without also
changing the pressure, if equilibrium between vapor and liquid phases is to be maintained.
There is a single independent variable.
For a multiphase system at equilibrium, the number of independent variables that must
be arbitrarily fixed to establish its intensive state is called the number of degrees of freedom of
the system. This number is given by the phase rule of J. Willard Gibbs.1 It is presented here
without proof in the form applicable to nonreacting systems:2

​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.
The intensive state of a system at equilibrium is established when its temperature, pressure,
and the compositions of all phases are fixed. These are the variables of the phase rule, but they
are not all independent. The phase rule gives the number of variables from this set that must be
specified to fix all remaining intensive variables, and thus the intensive state of the system.
A phase is a homogeneous region of matter. A gas or a mixture of gases, a liquid or a
liquid solution, and a crystalline solid are examples of phases. An abrupt change in properties
always occurs at the boundary between phases. Various phases can coexist, but they must
be in equilibrium for the phase rule to apply. A phase need not be continuous; examples of
discontinuous phases are a gas dispersed as bubbles in a liquid, a liquid dispersed as droplets
in another liquid with which it is immiscible, and solid crystals dispersed in either a gas or a
liquid. In each case a dispersed phase is distributed throughout a continuous phase.
As an example, the phase rule may be applied to an aqueous solution of ethanol in equi-
librium with its vapor. Here N = 2, π = 2, and

F = 2 − π + N = 2 − 2 + 2 = 2​
This is a system in vapor/liquid equilibrium, and it has two degrees of freedom. If the system
exists at specified T and P (assuming this is possible), its liquid- and vapor-phase composi-
tions are fixed by these conditions. A more common specification is of T and the liquid-phase
composition, in which case P and the vapor-phase composition are fixed.
Intensive variables are independent of the size of the system and of the individual
phases. Thus, the phase rule gives the same information for a large system as for a small one
and for different relative amounts of the phases. Moreover, the phase rule applies only to
individual-phase compositions, and not to the overall composition of a multiphase system.
Note also that for a phase only N − 1 compositions are independent, because the mole or mass
fractions of a phase must sum to unity.
The minimum number of degrees of freedom for any system is zero. When F = 0, the
system is invariant; Eq. (3.1) becomes π = 2 + N. This value of π is the maximum number
of phases that can coexist at equilibrium for a system containing N chemical species. When
N = 1, this limit is reached for π = 3, characteristic of a triple point (Sec. 3.2). For example, the

1Josiah Willard Gibbs (1839–1903), American mathematical physicist, who deduced it in 1875. See http://en.wiki-

pedia.org/wiki/Willard_Gibbs
2The theoretical justification of the phase rule for nonreacting systems is given in Sec. 12.2, and the phase rule for

reacting systems is considered in Sec. 14.8.


70 CHAPTER 3.  Volumetric Properties of Pure Fluids

triple point of H2O, where liquid, vapor, and the common form of ice exist together in equilib-
rium, occurs at 0.01°C and 0.0061 bar. Any change from these conditions causes at least one
phase to disappear.

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.

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​
This result is in agreement with the fact that for a given pressure water has but one
boiling point. Temperature or pressure, but not both, may be specified for a system
comprised of water in equilibrium with its vapor.
(b) Two chemical species are present. Again there are two phases, and

F = 2 − π + N = 2 − 2 + 2 = 2​
The addition of an inert gas to a system of water in equilibrium with its vapor
changes the characteristics of the system. Now temperature and pressure may be
independently varied, but once they are fixed the system described can exist in
equilibrium only at a particular composition of the vapor phase. (If nitrogen is
considered negligibly soluble in water, the liquid phase is pure water.)
(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 BEHAVIOR OF PURE SUBSTANCES

Figure 3.1 displays the equilibrium conditions of P and T at which solid, liquid, and gas phases
of a pure substance exist. Lines 1-2 and 2-C represent the conditions at which solid and liq-
uid phases exist in equilibrium with a vapor phase. These vapor pressure versus temperature
3.2.  PVT Behavior of Pure Substances 71

lines describe states of solid/vapor (line 1-2) and liquid/vapor (line 2-C) equilibrium. As indicated
in Ex. 3.1(a), such systems have but a single degree of freedom. Similarly, solid/liquid
­equilibrium is represented by line 2-3. 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.
The positive slope of the fusion line (2-3) represents the behavior of the vast majority
of substances. Water, a very common substance, has some very uncommon properties, and
exhibits a fusion line with negative slope.
The three lines meet at the triple point, where the three phases coexist in equilibrium.
According to the phase rule the triple point is invariant (F = 0). If the system exists along any
of the two-phase lines of Fig. 3.1, 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.
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.

3 A Fluid region

Pc Liquid region C
Fusion curve
Pressure

Vaporization
curve
Figure 3.1: PT diagram B
for a pure substance. Solid region
Gas region

Triple Vapor
2 point region
1 Sublimation
curve

Tc
Temperature

Water in an open flask is obviously a liquid in contact with air. If the flask is sealed and
the air is pumped out, water vaporizes to replace the air, and H2O fills the flask. Though the
pressure in the flask is much reduced, everything appears unchanged. The liquid water resides
at the bottom of the flask because its density is much greater than that of water vapor (steam),
and the two phases are in equilibrium at conditions represented by a point on curve 2-C of
Fig. 3.1. Far from point C, the properties of liquid and vapor are very different. However, if
72 CHAPTER 3.  Volumetric Properties of Pure Fluids

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. One consequence is that transitions from liquid to
vapor may occur along paths that do not cross the vaporization curve 2-C, i.e., from A to B.
The transition from liquid to gas is gradual and does not include the usual vaporization step.
The region existing at temperatures and pressures greater than Tc and Pc is marked off
by dashed lines in Fig. 3.1; these do not represent phase boundaries, but rather are limits fixed
by the meanings accorded the words liquid and gas. A phase is generally considered a liquid
if vaporization results from pressure reduction at constant temperature. A phase is considered
a gas if condensation results from temperature reduction at constant pressure. Since neither
process can be initiated in the region beyond the dashed lines, it is called the fluid region.
The gas region is sometimes divided into two parts, as indicated by the dotted vertical
line of Fig. 3.1. A gas to the left of this line, which can be condensed either by compression at
constant temperature or by cooling at constant pressure, is called a vapor. A fluid existing at a
temperature greater than Tc is said to be supercritical. An example is atmospheric air.

PV Diagram
Figure 3.1 does not provide any information about volume; it merely displays the bounda-
ries between single-phase regions. On a PV diagram [Fig. 3.2(a)] these boundaries in turn
become regions where two phases—solid/liquid, solid/vapor, and liquid/vapor—coexist in
equilibrium. The curves that outline these two-phase regions represent single phases that are
in equilibrium. Their relative amounts determine the molar (or specific) volumes within the
two-phase regions. The triple point of Fig. 3.1 here becomes a triple line, where the three
phases with different values of V coexist at a single temperature and pressure.
Figure 3.2(a), like Fig. 3.1, represents the behavior of the vast majority of substances
wherein the transition from liquid to solid (freezing) is accompanied by a decrease in specific
volume (increase in density), and the solid phase sinks in the liquid. Here again water dis-
plays unusual behavior in that freezing results in an increase in specific volume (decrease in
density), and on Fig. 3.2(a) the lines labeled solid and liquid are interchanged for water. Ice
therefore floats on liquid water. Were it not so, the conditions on the earth’s surface would be
vastly different.
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) superimposed. Isotherms on Fig. 3.1
are vertical lines, and at temperatures greater than Tc do not cross a phase boundary. On
Fig. 3.2(b) the isotherm labeled T > Tc is therefore smooth.
The lines labeled T1 and T2 are for subcritical temperatures and consist of three seg-
ments. The horizontal segment of each isotherm represents all possible mixtures of liquid and
vapor in equilibrium, ranging from 100% liquid at the left end to 100% vapor at the right end.
The locus of these end points is the dome-shaped curve labeled BCD, the left half of which
(from B to C) represents single-phase liquids at their vaporization (boiling) temperatures and
the right half (from C to D) single-phase vapors at their condensation temperatures. Liquids
and vapors represented by BCD are said to be saturated, and coexisting phases are connected
by the horizontal segment of the isotherm at the saturation pressure specific to the isotherm.
Also called the vapor pressure, it is given by a point on Fig. 3.1 where an isotherm (vertical
line) crosses the vaporization curve.
3.2.  PVT Behavior of Pure Substances 73

Solid/liquid
Tc

Fluid
C C

Liquid
Pc Pc Pc
Q
N
P P

Liquid
Solid
Gas
Vap
or T Tc
Liquid/vapor
Vapor Tc Tc
J K
Liquid/vapor T1 Tc
Solid/vapor B T2 Tc
D

Vc Vc
V V

(a) (b)

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.

The two-phase liquid/vapor region lies under dome BCD; the subcooled-liquid region
lies to the left of the saturated-liquid curve BC, and the superheated-vapor region lies to the
right of the saturated-vapor curve CD. Subcooled liquid exists at temperatures below, and
superheated vapor, at temperatures above the boiling point for the given pressure. Isotherms
in the subcooled-liquid region are very steep because liquid volumes change little with large
changes in pressure.
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. Thus, the critical iso-
therm, labeled Tc, exhibits a horizontal inflection at the critical point C at the top of the dome,
where the liquid and vapor phases become indistinguishable.

Critical Behavior
Insight into the nature of the critical point is gained from a description of the changes that
occur when a pure substance is heated in a sealed upright tube of constant volume. The dotted
vertical lines of Fig. 3.2(b) indicate such processes. They may also be traced on the PT ­diagram
of Fig. 3.3, where the solid line is the vaporization curve (Fig. 3.1), and the dashed lines are
constant-volume paths in the single-phase regions. If the tube is filled with either liquid or
vapor, the heating process produces changes that lie along the dashed lines of Fig. 3.3, for
example, by the change from E to F (subcooled-liquid) and by the change from G to H (super-
heated-vapor). The corresponding vertical lines on Fig. 3.2(b) are not shown, but they lie to the
left and right of BCD respectively.
If the tube is only partially filled with liquid (the remainder being vapor in equilibrium
with the liquid), heating at first causes changes described by the vapor-pressure curve (solid
74 CHAPTER 3.  Volumetric Properties of Pure Fluids

V2l Vc

Liquid

C
V2v
V1l Q
Figure 3.3: PT diagram for a pure fluid showing
P F N the vapor-pressure curve and constant-volume
( J, K ) V1v
lines in the single-phase regions.
H
G
E

Vapor

line of Fig. 3.3). For the process indicated by line JQ on Fig. 3.2(b), the meniscus is initially
near the top of the tube (point J), and the liquid expands sufficiently upon heating to fill the
tube (point Q). On Fig. 3.3 the process traces a path from (J, K) to Q, and with further heating
departs from the vapor-pressure curve along the line of constant molar volume V​ ​​ 2l ​​.​ 
The process indicated by line KN on Fig. 3.2(b) starts with a meniscus level closer to
the bottom of the tube (point K), and heating vaporizes liquid, causing the meniscus to recede
to the bottom (point N). On Fig. 3.3 the process traces a path from (J, K) to N. With further
heating the path continues along the line of constant molar volume V​  ​​ 2v​  ​​.
For a unique filling of the tube, with a particular intermediate meniscus level, the heating
process follows a vertical line on Fig. 3.2(b) that passes through the critical point C.
Physically, heating does not produce much change in the level of the meniscus. As the ­critical
point is approached, the meniscus becomes indistinct, then hazy, and finally disappears.
On Fig. 3.3 the path first follows the vapor-pressure curve, proceeding from point (J, K) to
the critical point C, where it enters the single-phase fluid region, and follows Vc, the line of
constant molar volume equal to the critical volume of the fluid.3

PVT Surfaces
For a pure substance, existing as a single phase, the phase rule tells us that two state variables
must be specified to determine the intensive state of the substance. Any two, from among
P, V, and T, can be selected as the specified, or independent, variables, and the third can
then be regarded as a function of those two. Thus, the relationship among P, V, and T for a
pure substance can be represented as a surface in three dimensions. PT and PV diagrams like
those illustrated in Figs. 3.1, 3.2, and 3.3 represent slices or projections of the three-dimensional
PVT surface. Fig. 3.4 presents a view of the PVT surface for carbon dioxide over a region
including liquid, vapor, and supercritical fluid states. Isotherms are superimposed on this
3A video illustrating this behavior is available in the online learning center at http://highered.mheducation.com/

sites/1259696529.
3.2.  PVT Behavior of Pure Substances 75

surface. The vapor/liquid equilibrium curve is shown in white, with the vapor and liquid
­portions of it connected by the vertical segments of the isotherms. Note that for ease of
­visualization, the molar volume is given on a logarithmic scale, because the vapor volume at
low pressure is several orders of magnitude larger than the liquid volume.

5
10

4
10
V (cm3/mol)

3
10

2
10

1
10
350
0
300 20
40
250 60
T (K) 80
200 100 P (bar)

Figure 3.4: PVT surface for carbon dioxide, with isotherms shown in black and the vapor/liquid equi-
librium curve in white.

Single-Phase Regions
For the regions of the diagram where a single phase exists, there is a unique relation connect-
ing P, V, and T. Expressed analytically, as f (P, V, T) = 0, such a relation is known as a PVT
equation of state. It relates pressure, molar or specific volume, and temperature for a pure
homogeneous fluid at equilibrium. The simplest example, the equation for the ideal-gas state,
PV = RT, has approximate validity for the low-pressure gas region and is discussed in detail
in the following section.
An equation of state may be solved for any one of the three quantities P, V, or T, given
values for other two. For example, if V is considered a function of T and P, then V = V(T, P), and

( ∂ T ) ( ∂ P )
∂ V ∂ V
dV = ​​ ​​ ___ ​  ​ ​​  ​​dT + ​​ ___
​ ​​   ​ ​  ​​  ​​dP​ (3.2)
P T
76 CHAPTER 3.  Volumetric Properties of Pure Fluids

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

∙ Volume expansivity:

V ( ∂ T ) P
1 ∂ V
β ≡ ​__
​    ​  ___
​​   ​ ​  ​​  ​​​ (3.3)

∙ Isothermal compressibility:

V ( ∂ P ) T
1 ∂ V
κ ≡ − __
​ ​   ​  ___
​​   ​ ​  ​​  ​​​ (3.4)

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


dV
___

​   ​  = β dT − κ dP​ (3.5)
V
The isotherms for the liquid phase on the left side of Fig. 3.2(b) are very steep and closely
spaced. Thus both ​(​∂ V / ∂ T​)​ P​and ​(​∂ V / ∂ P​)​ T​​ are small. Hence, both β and κ are small. This char-
acteristic behavior of liquids (outside the critical region) suggests an idealization, commonly
employed in fluid mechanics and known as the incompressible fluid, for which both β and κ
are zero. No real fluid is truly incompressible, but the idealization is useful, because it pro-
vides a sufficiently realistic model of liquid behavior for many practical purposes. No equation
of state exists for an incompressible fluid, because V is independent of T and P.
For liquids, β is almost always positive (liquid water between 0°C and 4°C is an excep-
tion), and κ is necessarily positive. 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:
​V​ 2​
ln ​___
​    ​  = β​​(​T​ 2​ − ​T​ 1​)​ − κ​​(​P​ 2​ − ​P​ 1​)​​ (3.6)
​V​ 1​

This is a less restrictive approximation than the assumption of an incompressible fluid.

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 ​cm​​ 3​⋅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.
3.3.  Ideal Gas and Ideal-Gas State 77

Solution 3.2
(a) The derivative ​(​∂ P / ∂ T​)​ V​ is determined by application of Eq. (3.5) to the case
for which V is constant and dV = 0:
β dT − κ dP = 0  ​(​const V​)​​

or

( ∂ T )
∂ P β 1.487 × ​10​​ −3​
​   ​   ​ ​ ​​  = ​__
​​​ ___    ​ = ​__________
    = 24 bar ⋅ ° ​C​ −1​​
  ​  
κ 62 × ​10​​ −6​
V

(b) If β and κ are assumed constant in the 10°C temperature interval, then for con-
stant volume Eq. (3.6) can be written:
β
​P​ 2​ = ​P​ 1​ + ​__
​    ​​(​T​ 2​ − ​T​ 1​)​ = 1 bar + 24 bar ⋅ ° ​C​ −1​ × 10° C = 241 bar​
κ
(c) Direct substitution into Eq. (3.6) gives:
​V​ 2​
   ​  = ​​(​1.487 × ​10​​ −3​)​(​−20​)​ − ​​(​62 × ​10​​ −6​)​(​9​)​ = −0.0303
ln ​___
​V​ 1​

     
​ ​ ​​
​V​ 2​
___
​   ​  = 0.9702  and  ​V​ 2​ = ​​(​0.9702​)​(​1.287​)​ = 1.249​ cm​​ 3​⋅g​​ −1​
​V​ 1​
Then,
​ΔV = ​V​ 2​ − ​V​ 1​ = 1.249 − 1.287 = −0.038 ​cm​​ 3​⋅g​​ −1​​

The preceding example illustrates the fact that heating a liquid that completely fills a closed
vessel can cause a substantial rise in pressure. On the other hand, liquid volume decreases very
slowly with rising pressure. Thus, the very high pressure generated by heating a subcooled
liquid at constant volume can be relieved by a very small volume increase, or a very small leak
in the constant volume container.

3.3  IDEAL GAS AND IDEAL-GAS STATE

In the 19th century, scientists developed a rough experimental knowledge of the PVT behavior
of gases at moderate conditions of temperature and pressure, leading to the equation PV = RT,
wherein V is molar volume and R is a universal constant. This equation adequately describes
PVT behavior of gases for many practical purposes near ambient conditions of T and P. How-
ever, more precise measurements show that for pressures appreciably above, and temperatures
appreciably below, ambient conditions, deviations become pronounced. On the other hand,
deviations become ever smaller as pressure decreases and temperature increases.
The equation PV = RT is now understood to define an ideal gas and to represent a model
of behavior more or less approximating the behavior of real gases. It is called the ideal gas
law, but is in fact valid only for pressures approaching zero and temperatures approaching
78 CHAPTER 3.  Volumetric Properties of Pure Fluids

infinity. Thus, it is a law only at limiting conditions. As these limits are approached, the mol-
ecules making up a gas become more and more widely separated, and the volume of the mol-
ecules themselves becomes a smaller and smaller fraction of the total volume occupied by the
gas. Furthermore, the forces of attraction between molecules become ever smaller because
of the increasing distances between them. In the zero-pressure limit, molecules are separated
by infinite distances. Their volumes become negligible compared with the total volume of
the gas, and the intermolecular forces approach zero. The ideal gas concept extrapolates this
behavior to all conditions of temperature and pressure.
The internal energy of a real gas depends on both pressure and temperature. Pressure
dependence results from intermolecular forces. If such forces did not exist, no energy would
be required to alter intermolecular distances, and no energy would be required to bring about
pressure and volume changes in a gas at constant temperature. Thus, in the absence of inter-
molecular forces, internal energy would depend on temperature only.
These observations are the basis for the concept of a hypothetical state of matter
designated 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.
Although related to the ideal gas, it presents a different perspective. It is not the gas that is
ideal, but the state, and this has practical advantages. 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:


​P​V​ ig​ = RT​​ (3.7)
∙ Internal energy:

​U​ ig​ = U​(​T​)​​
​ (3.8)

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


The property relations for this state are very simple, and at appropriate conditions of T
and P they may serve as suitable approximations for direct application to the real-gas state.
However, they have far greater importance as part of a general three-step procedure for calcu-
lation of property changes for real gases that includes a major step in the ideal-gas state. The
three steps are as follows:

1. Evaluate property changes for the mathematical transformation of an initial real-gas


state into the ideal-gas state at the same T and P.
2. Calculate property changes in the ideal-gas state for the T and P changes of the process.
3. Evaluate property changes for the mathematical transformation of the ideal-gas state
back to the real-gas state at the final T and P.

This procedure calculates the primary property-value changes resulting from T and P
changes by simple, but exact, equations for the ideal-gas state. The property-value changes for
transitions between real and ideal-gas states are usually relatively minor corrections. These
transition calculations are treated in Chapter 6. Here, we develop property-value calculations
for the ideal-gas state.
3.3.  Ideal Gas and Ideal-Gas State 79

Property Relations for the Ideal-Gas State


The definition of heat capacity at constant volume, Eq. (2.15), leads for the ideal-gas state to
ig
the conclusion that ​C​V​ ​is a function of temperature only:

( ∂ T )
∂ ​U​ ig​ d​U​ ig(​ ​T)​ ​
  ​​  ​​  = ​______
ig ig
​C​V​ ​  ≡ ​​ ____
​ ​   ​       ​  = ​C​V​ ​ ​(​T​)​​ (3.9)
dT
V

The defining equation for enthalpy, Eq. (2.10), applied to the ideal-gas state, leads to the
­conclusion that ​H​ ig​ is also a function only of temperature:

​​H​​  ig​ ≡ ​U​ ig​ + P​V​ ig​ = ​U​ ig​ ​(​T​)​ + RT = ​H​ ig​ ​(​T​)​​ (3.10)


ig ig
The heat capacity at constant pressure ​C​P​ ​, defined by Eq. (2.19), like ​C​V​ ​, is a function of
temperature only:

( ∂ T )
∂ ​H​ ig​ d​H​ ig​ (​ ​T​)​
  ​​  ​​  = ​ ___________
ig ig
​C​P​ ​  ≡ ​​ ____
​ ​   ​     ​  = ​C​P​ ​ ​(​T​)​​ (3.11)
dT
P

ig ig
A useful relation between C​
​​ P​  ​​ and C
​​ ​V​  ​​for the ideal-gas state comes from differentiation of Eq. (3.10):

d​H​ ig​ ____ d​U​ ig​


​C​P​ ​ ≡ ​​ _____ ​
ig ig
​  
​  ​​  = ​      ​ + R = ​C​V​ ​ + R​ (3.12)
dT dT

ig ig
This equation does not mean that ​C​P​ ​ and ​C​V​ ​ are themselves constant for the ideal-gas state,
but only that they vary with temperature in such a way that their difference is equal to R. For
any change in the ideal-gas state, Eqs. (3.9) and (3.11) lead to:

 ​  C​

∫  V
ig ig
d​​ U​​  ig​ = ​CV​ ​ ​dT​  (3.13a) ​Δ​U​ ig​ = ​ ​​  ​  ​dT ​​  (3.13b)

 ​  ​ C​ ​  ​dT ​​  (3.14b)


∫  P
ig ig
​d​H​ ig​ = ​CP​ ​ ​dT​  (3.14a) ​Δ​H​ ig​ = ​

ig
Because both ​U​ ig​and ​C​V​ ​for the ideal-gas state are functions of temperature only, ​Δ​U​ ig​​
for the ideal-gas state is always given by Eq. (3.13b), regardless of the kind of process causing
the change. This is illustrated in Fig. 3.5, which shows a graph of internal energy as a function
of ​V​ ig​at two different temperatures.
The dashed line connecting points a and b represents a constant-volume process for which
the temperature increases from T1 to T2 and the internal energy changes by Δ ​ ​U​ ig​ = ​U​  2ig​  ​ − ​U​  1ig​  ​​.
ig
This change in internal energy is given by Eq. (3.13b) as Δ ​ ​U​ ig​ = ∫ ​ C​V​  ​dT​. The dashed lines
connecting points a and c and points a and d represent other processes not occurring at constant
volume but which also lead from an initial temperature T1 to a final temperature T2. The graph
shows that the change in ​U​ ig​for these processes is the same as for the constant-volume process,
ig
and it is therefore given by the same equation, namely, ​Δ​U​ ig​ = ∫ ​ C​V​ ​dT​. However, ​Δ​U​ ig​​ is not
equal to Q for these processes, because Q depends not only on T1 and T2 but also on the path of
the process. An entirely analogous discussion applies to the enthalpy ​H​ ig​in the ideal-gas state.
80 CHAPTER 3.  Volumetric Properties of Pure Fluids

b c d Figure 3.5: Internal energy changes for the


U2 T2
­ideal-gas state. Because ​U​ ig​is independent
of ​V​ ig​, the plot of ​U​ ig​vs. ​V​ ig​at constant
U temperature is a horizontal line. For different
U1 T1 temperatures, ​U​ ig​has different values, with a
a separate line for each temperature. Two such
lines are shown, one for temperature ​T​ 1​​ and
one for a higher temperature ​T​ 2​​.
V ig

Process Calculations for the Ideal-Gas State


Process calculations provide work and heat quantities. The work of a mechanically reversible
closed-system process is given by Eq. (1.3), here written:


dW = −P d​V​ ig​​ (1.3)

For the ideal-gas state in any closed-system process, the first law as given by Eq. (2.6) written
for a unit mass or a mole, may be combined with Eq. (3.13a) to give:
ig 

dQ + dW = ​C​V​  ​dT​

Substitution for dW by Eq. (1.3) and solution for dQ yields an equation valid for the ideal-gas
state in any mechanically reversible closed-system process:
ig

dQ = ​C​V​ ​dT + Pd​V​ ig​​ (3.15)

This equation contains the variables P, Vig, and T, only two of which are independent.
Working equations for dQ and dW depend on which pair of these variables is selected as
­independent; i.e., upon which variable is eliminated by Eq. (3.7). We consider two cases,
­eliminating first P, and second, Vig. With ​P = RT / ​V​ ig​, Eqs. (3.15) and (1.3) become:

d​V​ ig​ d​V ​ig​
​dQ = ​C​V​  ​dT + RT ​____ ​  (3.16) d​ W = −RT  ​____
ig
  ig ​    ig ​  ​  (3.17)
​V​  ​ ​  ​ ​
V

R dP
For ​V​ ig​ = RT / P,  d​V​ ig​ = ​ __ ​ (  dT − T​ ___ ​ )  ​. Substituting for ​d​V​ ig​and for ​C​V​ ​ = ​C​P​ ​ − R​ transforms
ig ig
P P
Eqs. (3.15) and (1.3) into:

dP dP
​dQ = ​C​P​  ​dT − RT ​___
   ​ ​  (3.18) ​dW = −RdT + RT ​___
ig
   ​ ​  (3.19)
P P

These equations apply to the ideal-gas state for various process calcula-
tions. The assumptions implicit in their derivation are that the system is
closed and the process is mechanically reversible.
3.3.  Ideal Gas and Ideal-Gas State 81

Isothermal Process
By Eqs. (3.13b) and (3.14b),
​Δ​U​ ig​ =  Δ​H​ ig​ = 0  ​(​const T​)​​
ig
​V​ ​  ​ ​P​ 1​
  2ig  ​ = RT ln ​___
By Eqs. (3.16) and (3.18), ​Q = RT ln ​___    ​ 
​V1​ ​  ​ ​P​ 2​
ig
​V​ ​  ​ ​P​ 2​
  1ig  ​ = RT ln ​___
By Eqs. (3.17) and (3.19), ​W = RT ln ​___    ​ 
​V​2​  ​ ​P​ 1​

Because Q
​  = −W​, a result that also follows from Eq. (2.3), we can write in summary:

ig
​V​ ​  ​ ​P​ 1​
​   2ig  ​ = RT ln ​_
​Q = −W = RT ln ​_    ​   ​(​const T​)​​ (3.20)
​V​1​  ​ ​P​ 2​

Isobaric Process
By Eqs. (3.13b) and (3.19) with dP = 0,

​Δ​U​ ig​  =  ​ ​    ​​C​V​  ​dT ​  and  W = −R​​(​T​ 2​ − ​T​ 1​)​​


∫ 
ig

By Eqs. (3.14b) and (3.18),

​​Q = Δ​H​  ​ = ​  ​    ​ CP​ ​  ​dT ​ 


∫ 
ig
ig ​(​const P​)​​​ (3.21)

Isochoric (Constant-V) Process


With ​d​V​ ig​ = 0, W = 0​, and by Eqs. (3.13b) and (3.16),

​​Q =  Δ​U​ ig​ = ​  ​    ​ C​V​  ​dT ​  (const ​V​ ig​)​​


∫ 
ig
(3.22)

Adiabatic Process; Constant Heat Capacities


An adiabatic process is one for which there is no heat transfer between the system and its
surroundings; i.e., dQ = 0. Each of Eqs. (3.16) and (3.18) may therefore be set equal to zero.
ig ig
Integration with ​C​V​ ​ and ​C​P​ ​ constant then yields simple relations among the variables T, P,
and ​V​  ​, valid for mechanically reversible adiabatic compression or expansion in the ideal-gas
ig

state with constant heat capacities. For example, Eq. (3.16) becomes:

dT
___ R d​V​ ig​
​   ​  = − ___
​ ​  ig  ​ ____
​  ig ​ ​
T ​CV​ ​ ​ ​V​  ​
82 CHAPTER 3.  Volumetric Properties of Pure Fluids

ig
Integration with ​C​V​ ​constant gives:
ig
ig R/​C​V​ ​

​T​ 1​ ( ​V​ig​  ​)
​ ​ 2​
T ​V​ ​  ​
___ ​​  1  ​ ​ ​​​ 
​   ​  = ​​ ___
​ ​​
2
Similarly, Eq. (3.18) leads to:
ig
R/​C​P​ ​

​T​ 1​ ( ​P​ 1​)
​T​ 2​
___ ​P​ 2​
​   ​  = ​​ ___
​ ​​    ​​ ​​​  ​​

These equations may also be expressed as:

​  ​ig)​  ​γ − 1 ​= const​  (3.23a) T


​ (​ V
T ​  ​(​ 1​  − γ​)/​γ ​= const​ 
​P (3.23b) P ​  ​ig)​  ​γ ​= const​  (3.23c)
​ (​ V

where Eq. (3.23c) results by combining Eqs. (3.23a and (3.23b) and where by definition,4

ig
​C​ ​ ​
​γ ≡ ​_
​   Pig  ​​ (3.24)
​C​V​ ​

Equations (3.23) apply for the ideal-gas state with constant heat capaci-
ties and are restricted to mechanically reversible adiabatic expansion or
compression.
The first law for an adiabatic process in a closed system combined with Eq. (3.13a)
yields:
ig

dW = dU = ​C​V​ ​ dT​
ig
For constant ​C​V​ ​​,
ig 

W = Δ​U​ ig​ = ​C​V​  ​ΔT​ (3.25)
ig
Alternative forms of Eq. (3.25) result if ​C​V​ ​is eliminated in favor of the heat-capacity ratio γ:
ig ig
​C​ ​ ​ ​C​V​ ​ + R R R
γ ≡ ​___
  Pig  ​ = ​_____ = 1 + ​___
  ig  ​   or  ​C​V​ ​ = ​___
ig
​   ig    ​        ​​  
​C​V​ ​ ​C​V​ ​ ​C​V​ ​ γ − 1

and
RΔT
W = ​C​V​ ​ΔT = ​____
ig
​     ​​  
γ − 1

ig ig
4If ​​C​ ​  ​​ and ​​C​ ​  ​​are constant, γ is necessarily constant. The assumption of constant γ is equivalent to the assumption
p v
ig ig
that the heat capacities themselves are constant. This is the only way that the ratio ​​C​p​  ​​/​​C​v​  ​​ and the difference
ig ig  ig ig
​​C​p​  ​​-​​Cv​ ​  ​​= R can both be constant. Except for the monatomic gases, both ​​Cp​ ​  ​​ and ​​Cv​ ​  ​​actually increase with temperature,
but the ratio γ is less sensitive to temperature than the heat capacities themselves.
3.3.  Ideal Gas and Ideal-Gas State 83

ig ig
Because R
​ ​T​ 1​ = ​P​ 1​V​1​  ​​and R
​ ​T​ 2​ = ​P​ 2​V​2​  ​​, this expression may be written:

ig ig
​P​ 2​V​ ​  ​ − ​P​ 1​V​1​  ​
R​T​ 2​ − R​T​ 1​ ___________
W = ​_______
​       ​     2
= ​     ​​   (3.26)
γ − 1 γ − 1

Equations (3.25) and (3.26) are general for adiabatic compression and expansion
processes in a closed system, whether reversible or not, because P, Vig, and T are state functions,
ig
independent of path. However, T2 and ​​V​2​  ​​are usually unknown. Elimination of ​​V​  2ig​  ​​ from
Eq. (3.26) by Eq. (3.23c), valid only for mechanically reversible processes, leads to the
expression:

γ − 1 [( ​P​ 1​​ ) ] γ − 1 [( ​P​ 1​​ ) ]


ig (γ − 1)/γ (γ − 1)/γ
​P​ 1​V​ ​  ​ ___ ​P​ 2​​ R​T​ 1​ ___ ​ ​ 2​​
P
​   1  ​ 
W = ​_____  ​ ​​ ​   ​  ​​​  ​ − 1 ​  = ​___
    ​ ​​ ​    ​ ​​​  ​ − 1 ​​ (3.27)

The same result is obtained when the relation between P and Vig given by Eq. (3.23c) is used
for the integration, ​W = −∫  Pd ​V​ ig​.​

Equation (3.27) is valid only for the ideal-gas state, for constant heat
capacities, and for adiabatic, mechanically reversible, closed-system
processes.
When applied to real gases, Eqs. (3.23) through (3.27) often yield satisfactory approxi-
mations, provided the deviations from ideality are relatively small. For monatomic gases, γ =
1.67; approximate values of γ are 1.4 for diatomic gases and 1.3 for simple polyatomic gases
such as CO2, SO2, NH3, and CH4.

Irreversible Processes
All equations developed in this section have been derived for mechanically reversible,
closed-system processes for the ideal-gas state. However, the equations for property changes—
dUig, dHig, ΔUig, and ΔHig—are valid for the ideal-gas state regardless of the process. They
apply equally to reversible and irreversible processes in both closed and open systems, because
changes in properties depend only on initial and final states of the system. On the other hand,
an equation for Q or W, unless it is equal to a property change, is subject to the restrictions of
its derivation.
The work of an irreversible process is usually calculated by a two-step procedure.
First, W is determined for a mechanically reversible process that accomplishes the same
change of state as the actual irreversible process. Second, this result is multiplied or divided
by an efficiency to give the actual work. If the process produces work, the absolute value
for the reversible process is larger than the value for the actual irreversible process and must
be multiplied by an efficiency. If the process requires work, the value for the reversible
process is smaller than the value for the actual irreversible process and must be divided by
an efficiency.
Applications of the concepts and equations of this section are illustrated in the examples
that follow. In particular, the work of irreversible processes is treated in Ex. 3.5.
84 CHAPTER 3.  Volumetric Properties of Pure Fluids

Example 3.3
Air is compressed from an initial state of 1 bar and 298.15 K to a final state of 3 bar
and 298.15 K by three different mechanically reversible processes in a closed system:
(a) Heating at constant volume followed by cooling at constant pressure.
(b) Isothermal compression.
(c) Adiabatic compression followed by cooling at constant volume.
These processes are shown in the figure. We assume air to be in its ideal-gas state,
ig ig
and assume constant heat capacities, ​C ​ ​P​ ​ = 29.100​  J·mol−1·K−1.
​ ​V​ ​ = 20.785​and ​C
Calculate the work required, heat transferred, and the changes in internal energy and
enthalpy of the air for each process.

c
4
P bar

2
a
2
b
1

0 5 10 15 20 25
V ig 103 m3

Solution 3.3
Choose the system as 1 mol of air. The initial and final states of the air are identi-
cal with those of Ex. 2.7. The molar volumes given there are

​V​1ig​  ​ = 0.02479​ m​​ 3​  ​  ​V​2ig​  ​ = 0.008263 ​m​ 3​​


Because T is the same at the beginning and end of the process, in all cases,

​Δ​U​ ig​ = Δ​H​ ig​ = 0​

(a) The process here is exactly that of Ex. 2.7(b), for which:

​Q = −4958 J​  and​  W = 4958 J​​
(b) Equation (3.20) for isothermal compression applies. The appropriate value of
​  = 8.314 J·​mol​​ −1​·​K​ −1​​.
R here (from Table A.2 of App. A) is R
​P​ 1​ 1
Q = −W = RT ln ​___
​    ​  = ​​(​8.314​)​(​298.15​)​ ln ​__
   ​ = −2723 J ​
​P​ 2​ 3

(c) The initial step of adiabatic compression takes the air to its final volume of
0.008263 m3. By Eq. (3.23a), the temperature at this point is:
3.3.  Ideal Gas and Ideal-Gas State 85

γ − 1

( 0.008263 )
ig 0.4

( ​V​2​  ​)
​V​ ​  ​ 0.02479
​ ​​  1ig ​ ​ ​​​ 
T′ = ​T​ 1​ ___ ​  = ​​(​298.15​)​​​​ ________
​   
 ​ ​​​​  ​  = 462.69 K​

For this step, Q = 0, and by Eq. (3.25), the work of compression is:
ig ig
W = ​C​V​ ​ ΔT = ​C​V​ ​(​T′ − ​T​ 1​)​​  = ​​(​20.785​)​ ​(​462.69 − 298.15​)​ = 3420 J​

For the constant-volume step, no work is done; the heat transfer is:
ig
Q = Δ​U​ ig​ = ​C​V​ ​(​T​ 2​ − ​T′​  ​​​​)​​  = 20.785​ ​(​298.15 − 462.69​)​ = −3420 J​

Thus for process (c),

​W = 3420 J​  and​  Q = −3420 J​​


Although the property changes ΔUig and ΔHig are zero for each process, Q and
W are path-dependent, and here Q = −W. The figure shows each process on a PVig
diagram. Because the work for each of these mechanically reversible processes is
given by ​W = −∫ Pd​ V​​ ig​, the work for each process is proportional to the total area
below the paths on the PVig diagram from 1 to 2. The relative sizes of these areas
correspond to the numerical values of W.

Example 3.4
A gas in its ideal-gas state undergoes the following sequence of mechanically
­reversible processes in a closed system:
(a) From an initial state of 70°C and 1 bar, it is compressed adiabatically to 150°C.
(b) It is then cooled from 150 to 70°C at constant pressure.
(c) Finally, it expands isothermally to its original state.
Calculate W, Q, ΔUig, and ΔHig for each of the three processes and for the entire cycle.
ig ig
Take ​C​V​ ​ = 12.471​and ​C​P​ ​ = 20.785 J·​mol​​ −1​·K​​ −1​​.

Solution 3.4
Take as a basis 1 mol of gas.
(a) For adiabatic compression, Q = 0, and
ig
Δ​U​ ig​ = W = ​C​V​ ​ ΔT = ​​(​12.471​)​(​150 − 70​)​ = 998 J

​      ig
  ​​
Δ​H​ ig​ = ​C​P​ ​ ΔT = ​​(​20.785​)​(​150 − 70​)​ = 1663 J

Pressure P2 is found from Eq. (3.23b):


γ/​​(​γ − 1​)​
150 + 273.15 2.5
( ​T​ 1​) ( 70 + 273.15 )
​T​ 2​
​P​ 2​ = ​P​ 1​ ___
​ ​​   ​ ​ ​​​  ​  = ​​(​1​)​ ___________
​​    ​​  ​​​  ​  = 1.689 bar​
86 CHAPTER 3.  Volumetric Properties of Pure Fluids

3 b 2
70°C 150 °C

P a
c

70°C
1

V ig

(b) For this constant-pressure process,


ig
Q = Δ​H​ ig​ = ​C​P​ ​ ΔT = ​​(​20.785​)​(​70 − 150​)​ = −1663 J

    
     
 ​ ΔU = ​C​ ​ ​ ΔT = ​​(​12.471​)​(​70 − 150​)​ = −998 J
ig   ​ ​​
V
  W = Δ​U​ ig​ − Q = −998 − ​​(​−1663​)​ = 665 J
(c) For this isothermal process, ΔUig and ΔHig are zero; Eq. (3.20) yields:
​P​ 3​ ​P​ 2​ 1.689
Q = −W = RT ln ​___
​    ​  = RT ln ​___
   ​  = ​​(​8.314​)​(​343.15​)​ ln ​_____
      ​  = 1495 J​
​P​ 1​ ​P​ 1​ 1
For the entire cycle,
Q = 0 − 1663 + 1495 = −168 J
W = 998 + 665 − 1495 = 168 J

   
​     
    ​  ​  ​   ​​​
Δ​U​ ig​ = 998 − 998 + 0 = 0
Δ​H​ ig​ = 1663 − 1663 + 0 = 0
The property changes ΔUig and ΔHig both are zero for the entire cycle because
the initial and final states are identical. Note also that Q = −W for the cycle. This
follows from the first law with ΔUig = 0.

Example 3.5
If the processes of Ex. 3.4 are carried out irreversibly but so as to accomplish exactly the
same changes of state—the same changes in P, T, Uig, and Hig—then different values of
Q and W result. Calculate Q and W if each step is carried out with a work efficiency of 80%.

Solution 3.5
If the same changes of state as in Ex. 3.4 are carried out by irreversible processes,
the property changes for the steps are identical with those of Ex. 3.4. However, the
values of Q and W change.
3.3.  Ideal Gas and Ideal-Gas State 87

(a) For mechanically reversible, adiabatic compression, the work is Wrev = 998 J.
If the process is 80% efficient compared with this, the actual work is larger, and
W = 998/0.80 = 1248 J. This step cannot here be adiabatic. By the first law,

Q = Δ​U​ ig​ − W = 998 − 1248 = −250 J​
(b) The work required for the mechanically reversible cooling process is 665 J. For
the irreversible process, W = 665/0.80 = 831 J. From Ex. 3.4(b), ΔUig = −998 J, and

Q = Δ​U​ ig​ − W = −998 − 831 = −1829 J​
(c) As work is done by the system in this step, the irreversible work in absolute
value is less than the reversible work of −1495 J, and the actual work done is:
  W = ​​(​0.80​)​(​−1495​)​ = −1196 J

​    ​ ​​
Q = Δ​U​ ig​ − W = 0 + 1196 = 1196 J
For the entire cycle, ΔUig and ΔHig are zero, with
Q = −250 − 1829 + 1196 = −883 J

   
​  ​​ ​
  W = 1248 + 831 − 1196 = 883 J
A summary of these results and those for Ex. 3.4 is given in the following table;
values are in joules.

Mechanically reversible, Ex. 3.4 Irreversible, Ex. 3.5

ΔUig ΔHig Q W ΔUig ΔHig Q W


(a) 998 1663 0 998 998 1663 −250 1248
(b) −998 −1663 −1663 665 −998 −1663 −1829 831
(c) 0 0 1495 −1495 0 0 1196 −1196

Cycle 0 0 −168 168 0 0 −883 883

The cycle is one which requires work and produces an equal amount of heat.
The striking feature of the comparison shown in the table is that the total work
required when the cycle consists of three irreversible steps is more than five times
the total work required when the steps are mechanically reversible, even though
each irreversible step is assumed to be 80% efficient.

Example 3.6
Air flows at a steady rate through a horizontal pipe to a partly closed valve. The pipe
leaving the valve is enough larger than the entrance pipe that the kinetic-energy
change of the air as it flows through the valve is negligible. The valve and connecting
pipes are well insulated. The conditions of the air upstream from the valve are 20°C
and 6 bar, and the downstream pressure is 3 bar. If the air is in its ideal-gas state, what
is the temperature of the air some distance downstream from the valve?
88 CHAPTER 3.  Volumetric Properties of Pure Fluids

Solution 3.6
Flow through a partly closed valve is known as a throttling process. The system
is insulated, making Q negligible; moreover, the potential-energy and kinetic-­
energy changes are negligible. No shaft work is accomplished, and Ws = 0. Hence,
Eq. (2.31) reduces to:
ig ig
​Δ​H​ ig​ = ​H2​ ​  ​ − ​H1​ ​  ​ = 0​. Because Hig is a function of temperature only, this
requires that T 2 = T 1 . The result that ΔH ig = 0 is general for a throttling
process, because the assumptions of negligible heat transfer and potential- and kinetic-­
energy changes are usually valid. For a fluid in its ideal-gas state, no ­temperature
change occurs. The throttling process is inherently irreversible, but this is immaterial to
the ­calculation because Eq. (3.14b) is valid for the ideal-gas state whatever the process.5

Example 3.7
If in Ex. 3.6 the flow rate of air is 1 mol·s–1 and if both upstream and downstream
pipes have an inner diameter of 5 cm, what is the kinetic-energy change of the air and
ig
what is its temperature change? For air, ​C​P​ ​ = 29.100 J·​mol​​ −1​ and the molar mass is
ℳ = 29 g·mol−1.

Solution 3.7
By Eq. (2.23b),
​n ​  ​n ​​V
∙ ∙
  ​ ig​
u = ​ ___  ​  = ​ ____
​  ​ 
​  
Aρ A
where

   ​D​ 2​ = ​​(_
​    ​)​(​5 × ​10​​ −2​)​ 2​ = 1.964 × ​10​​ −3​ ​m​ 2​​
π π
A = ​__

4 4
The appropriate value here of the gas constant for calculation of the upstream
molar volume is R = 83.14 × 10−6 bar·m3·mol−1·K−1. Then

(​ ​83.14 × ​10​​  ​)​(​293.15 K​)​
−6
R​T​ 1​ ___________________
​V​1​  ​ = ​___
ig
​    ​  = ​   
       ​ = 4.062 × ​10​​ −3​ ​m​ 3​·mol​​ −1​​
​P​ 1​ 6 bar

Then,
​ ​1 mol·​s​ −1)​ (​ ​4.062 × ​10​​ −3​ ​m​ 3​·mol​​ −1​)​
(
​u​ 1​ = ​___________________________
​         
    ​ = 2.069 m·​s​ −1​​
1.964 × ​10​​ −3​ m​​ 2​
If the downstream temperature is little changed from the upstream temperature,
then to a good approximation:

​V2​ig​  ​ = 2​V1​ig​  ​    and  ​  ​u​ 2​ = 2​u​ 1​ = 4.138 m·​s​ −1​​



5The throttling of real gases may result in a relatively small temperature increase or decrease, known as the Joule/

Thomson effect. A more detailed discussion is found in Chapter 7.


3.4.  Virial Equations of State 89

The rate of change in kinetic energy is therefore:

(2 ) (2 )
1 1
​m   ​Δ​ _    Δ​​ _
​   ​u​ 2​ ​= ​n ​ℳ ​   ​u​ 2​ ​
∙ ∙

​​     
​    
​  ​  ​ ​(​4.138​ ​​
​ 2​ − ​2.069​​ 2​)​m​ 2​·s​​ −2​
​ = ​(​1 × 29 × ​10​​ −3​ kg·​s​ −1​)___________________
​    
    

2
​ = 0.186 kg·​m​ 2​·s​​ −3​ = 0.186 J·​s​ −1​
In the absence of heat transfer and work, the energy balance, Eq. (2.30), becomes:

( 2 ) (2 )
1 1
 Δ​​ ​H​ ig​ + ​_
   ​u​ 2​ ​m ​ = ​m ​Δ    ​H​ ig​ + ​m   ​Δ​ _
​   ​u​ 2​ ​ = 0
∙ ∙ ∙

​​       ​  ​ ​​
(2 ) (2 )
1 1
​(​m ​  / ℳ​)​C​P​ ​ΔT + ​m ​  Δ​​ _   ​P​ ​ ΔT + ​m ​  Δ​​ _
ig ig
​   ​u​ 2​ ​= ​n ​​C ​   ​u​ 2​ ​ = 0
∙ ∙ ∙ ∙

Then

(2 )
1
​(​1​)​(​29.100​)​ΔT = −​m ​  Δ​​ _
​   ​u​ 2​ ​ = −0.186​

and
​ΔT = −0.0064 K​

Clearly, the assumption of negligible temperature change across the valve is jus-
tified. Even for an upstream pressure of 10 bar and a downstream pressure of 1 bar
and for the same flow rate, the temperature change is only −0.076 K. We conclude
that, except for very unusual conditions, ΔHig = 0 is a satisfactory energy balance.

3.4  VIRIAL EQUATIONS OF STATE

Volumetric data for fluids are useful for many purposes, from the metering of fluids to the
sizing of tanks. Data for V as a function of T and P can of course be given as tables. However,
expression of the functional relation f(P, V, T) = 0 by equations is much more compact and
convenient. The virial equations of state for gases are uniquely suited to this purpose.
Isotherms for gases and vapors, lying to the right of the saturated-vapor curve CD in
Fig. 3.2(b), are relatively simple curves for which V decreases as P increases. Here, the ­product
PV for a given T varies much more slowly than either of its members, and hence is more easily
represented analytically as a function of P. This suggests expressing PV for an isotherm by a
power series in P:
PV = a + bP + c​P​ 2​ + . . .​

If we define, b ≡ aB′ , c ≡ aC′, etc., then,

PV = a​​(​1 + B′P + C′​P​ 2​ + D′​P​ 3​ + . . .​)​​
​ (3.28)

where B′, C′, etc., are constants for a given temperature and a given substance.
90 CHAPTER 3.  Volumetric Properties of Pure Fluids

Ideal-Gas Temperature; Universal Gas Constant


Parameters B′, C′, etc., in Eq. (3.28) are species-dependent functions of temperature, but param-
eter a is found by experiment to be the same function of temperature for all chemical species.
This is shown by isothermal measurements of V as a function of P for various gases. Extrapola-
tions of the PV product to zero pressure, where Eq. (3.28) reduces to PV = a, show that a is the
same function of T for all gases. Denoting this zero-pressure limit by an asterisk provides:
​(​PV​)​ *​ = a = f ​​(​T​)​​

This property of gases serves as a basis for an absolute temperature scale. It is defined by arbi-
trary assignment of the functional relationship f (T) and the assignment of a specific value to
a single point on the scale. The simplest procedure, the one adopted internationally to define
the Kelvin scale (Sec. 1.4):

∙ Makes (PV)* directly proportional to T, with R as the proportionality constant:


​(​PV​)​ *​ = a ≡ RT​
​ (3.29)
∙ Assigns the value 273.16 K to the temperature of the triple point of water (denoted by
subscript t):
​​​(​PV​)​​  *t​  ​  = R × 273.16 K​ (3.30)
Division of Eq. (3.29) by Eq. (3.30) gives:
(​ ​PV​)​ *​
T  K = 273.16​ ______* 
​  ​​   (3.31)
​(PV)​  t​  ​
This equation provides the experimental basis for the ideal-gas temperature scale throughout
the temperature range for which values of (PV)* are experimentally accessible. The Kelvin
temperature scale is defined so as to be in as close agreement as possible with this scale.
The proportionality constant R in Eqs. (3.29) and (3.30) is the universal gas constant. Its
numerical value is found from Eq. (3.30):
(​ ​PV​)​​  *t​  ​
R = ​ ________
​   ​​  
273.16 K
The accepted experimental value of ​​​(​PV​)​​  *t​  ​​is 22,711.8 bar·cm3 · mol−1, from which6
22,711.8 bar·​cm​​ 3​·mol​​ −1​
R = ​___________________
​         ​ = 83.1446 bar·​cm​​ 3​·mol​​ −1​·K​​ −1​​

273.16 K
Through the use of conversion factors, R may be expressed in various units. Commonly used
values are given in Table A.2 of App. A.

Two Forms of the Virial Equation


A useful auxiliary thermodynamic property is defined by the equation:
PV V
​Z ≡ ​_
​     ​  = ​_
    ​​   (3.32)
RT ​V​ ig​
6See http://physics.nist.gov/constants.
3.4.  Virial Equations of State 91

This dimensionless ratio is called the compressibility factor. 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. At ­moderate
temperatures its value is usually <1, though at elevated temperatures it may be >1. Figure 3.6
shows the compressibility factor of carbon dioxide as a function of T and P. This figure presents the
same information as Fig. 3.4, except that it is plotted in terms of Z rather than V. It shows that at low
pressure, Z approaches 1, and at moderate pressures, Z decreases roughly linearly with pressure.

0.8

0.6
Z
0.4

0.2

0
350
0
300 20
40
250 60
T (K)
80 P (bar)
200 100

Figure 3.6: PZT surface for carbon dioxide, with isotherms shown in black and the vapor/liquid
­equilibrium curve in white.

With Z defined by Eq. (3.32) and with a = RT [Eq. (3.29)], Eq. (3.28) becomes:

​Z = 1 + B′P + C′​P​ 2​ + D′​P​ 3​ + . . .​​
​ (3.33)

An alternative expression for Z is also in common use:7


B C _ D
​Z = 1 + ​_
​    ​  + ​_
    ​  + ​    ​  + . . .​​ (3.34)
V ​V​ 2​ ​V​ 3​
Both of these equations 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 coeffi-
cients; C′ and C are third virial coefficients, and so on. For a given gas the virial coefficients
are functions of temperature only.
7Proposed by H. Kamerlingh Onnes, “Expression of the Equation of State of Gases and Liquids by Means of

Series,” Communications from the Physical Laboratory of the University of Leiden, no. 71, 1901.
92 CHAPTER 3.  Volumetric Properties of Pure Fluids

The two sets of coefficients in Eqs. (3.33) and (3.34) are related as follows:

B C − ​B ​2​ D − 3BC + 2​B ​3​
​ ′= ​___
B ​ ′= ​_____
    ​ ​  (3.35a) C    ​  ​ ′= ​__________
​  (3.35b) D   
    ​ ​ (3.35c)
RT (​ R 2
​ T​) ​ ​ ​ T​) ​3​
(​ R

To derive these relations, we set Z = PV/RT in Eq. (3.34) and solve for P. This allows elimina-
tion of P on the right side of Eq. (3.33). The resulting equation reduces to a power series in 1/V
which may be compared term by term with Eq. (3.34) to yield the given relations. They hold
exactly only for the two virial expansions as infinite series, but they are acceptable approxima-
tions for the truncated forms used in practice.
Many other equations of state have been proposed for gases, but the virial equations
are the only ones firmly based on statistical mechanics, which provides physical significance
to the virial coefficients. Thus, for the expansion in 1/V, the term B/V arises on account of
interactions between pairs of molecules; the C/V2 term, on account of three-body ­interactions;
etc. Because, at gas-like densities, 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.

3.5  APPLICATION OF THE VIRIAL EQUATIONS

The two forms of the virial expansion given by Eqs. (3.33) and (3.34) are infinite series. For
engineering purposes their use is practical only where convergence is very rapid, that is, where
two or three terms suffice for reasonably close approximations to the values of the series. This
is realized for gases and vapors at low to moderate pressures.
Figure 3.7 shows a compressibility-factor graph for methane. All isotherms originate at
Z = 1 and P = 0, and are nearly straight lines at low pressures. Thus the tangent to an isotherm
at P = 0 is a good approximation of the isotherm from P → 0 to some finite pressure. Differ-
entiation of Eq. (3.33) for a given temperature gives:

( ∂ P )
∂ Z
​​​ ___
​​    ​​  ​​ ​​  = B′+2C′P + 3D′​P​ 2​ + . . .​
T
from which,

( ∂ P )
∂ Z
​​​ ___
​    ​  ​​  ​​  = B′​
T; P=0

Thus the equation of the tangent line is Z = 1 + B′P, a result also given by truncating
Eq. (3.33) to two terms.
A more common form of this equation results from substitution for B′ by Eq. (3.35a):

PV BP
​Z = ​_
​     ​  = 1 + ​_
    ​​   (3.36)
RT RT
3.5.  Application of the Virial Equations 93

500 K
1.00
350 K

300 K
Figure 3.7: Compressibility-factor graph
0.75 for methane. Shown are isotherms of the
Z = PV/RT

250
K ­compressibility factor Z, as calculated
175 K from PVT data for methane by the defining
5K ­equation Z = PV/RT. They are plotted vs.
22
0.50 pressure for a number of constant temperatures,
0K and they show graphically what the virial
20
expansion in P represents analytically.

0.25
0 50 100 150 200

P (bar)

This equation expresses direct linearity between Z and P and is often applied to vapors
at subcritical temperatures up to their saturation pressures. At higher temperatures it often pro-
vides a reasonable approximation for gases up to a pressure of several bars, with the pressure
range increasing as the temperature increases.
Equation (3.34) as well may be truncated to two terms for application at low pressures:
PV B
Z = ​___
​     ​ = 1 + ​__
   ​​ (3.37)
RT V
However, Eq. (3.36) is more convenient in application and is normally at least as accurate as
Eq. (3.37). Thus when the virial equation is truncated to two terms, Eq. (3.36) is preferred.
The second virial coefficient B is substance dependent and a function of temperature.
Experimental values are available for a number of gases.8 Moreover, estimation of second
virial coefficients is possible where no data are available, as discussed in Sec. 3.7.
For pressures above the range of applicability of Eq. (3.36) but below the critical pres-
sure, the virial equation truncated to three terms often provides excellent results. In this case
Eq. (3.34), the expansion in 1/V, is far superior to Eq. (3.33). Thus when the virial equation is
truncated to three terms, the appropriate form is:
PV B C
​Z = ​_
​     ​  = 1 + ​_
   ​  + ​_
    ​​   (3.38)
RT V ​V​ 2​

This equation is explicit in pressure but is cubic in volume. Analytic solution for V is possible,
but solution by an iterative scheme, as illustrated in Ex. 3.8, is often more convenient.
Values of C, like those of B, depend on the gas and on temperature. Much less is known
about third virial coefficients than about second virial coefficients, though data for a num-
ber of gases are found in the literature. Because virial coefficients beyond the third are rarely
known and because the virial expansion with more than three terms becomes unwieldy, its use
is uncommon.
8J. H. Dymond and E. B. Smith, The Virial Coefficients of Pure Gases and Mixtures, Clarendon Press, Oxford, 1980.
94 CHAPTER 3.  Volumetric Properties of Pure Fluids

100 4000
C
2000
0 0
B

2
1

2000

C/cm 6 mol
B/cm 3 mol

100 4000 Figure 3.8: Virial


coefficients B and C
for nitrogen.
200

300
0 100 200 300 400
T/K

Figure 3.8 illustrates the effect of temperature on the virial coefficients B and C for
nitrogen; although numerical values are different for other gases, the trends are similar. The
curve of Fig. 3.8 suggests that B increases monotonically with T; however, at temperatures
much higher than shown, B reaches a maximum and then slowly decreases. The temperature
dependence of C is more difficult to establish experimentally, but its main features are clear: C
is negative at low temperatures, passes through a maximum at a temperature near the critical
temperature, and thereafter decreases slowly with increasing T.

Example 3.8
Reported values for the virial coefficients of isopropanol vapor at 200°C are:
​B = − 388 ​cm​​ 3​ · mol​​ −1​​  C = − 26,000 ​cm​​ 6​ · mol​​ −2​​

Calculate V and Z for isopropanol vapor at 200°C and 10 bar:
(a) For the ideal-gas state; (b) By Eq. (3.36); (c) By Eq. (3.38).

Solution 3.8
The absolute temperature is T = 473.15 K, and the appropriate value of the gas
constant is R = 83.14 bar·cm3·mol−1·K−1.
(a) For the ideal-gas state, Z = 1, and
RT ​(​83.14​)​(​473.15​)​
​V​ ig​ = ​___
   ​  = ​____________

​       = 3934 ​cm​​ 3​·mol​​  1​​
​  
P 10
(b) From the second equality of Eq. (3.36), we have
RT
V = ​___
​    ​  + B = 3934 − 388 = 3546 ​cm​​ 3​·mol​​ −1​​
P
3.6.  Cubic Equations of State 95

and
PV V V 3546
Z = ​___
​     ​ = ​_____ = ​___
     ​       ​  = ​_____
    
​ = 0.9014​
RT RT / P ​V​ ig​ 3934

(c) If solution by iteration is intended, Eq. (3.38) may be written:

P( ​V​ i​ ​V​i2​ ​)
RT B C
​V​ i+1​ = ​___
​    ​ ​ ​1 + ​_
    ​ + ​_
    ​  ​​

where i is the iteration number. Iteration is initiated with the ideal-gas state value
Vig. Solution yields:
V = 3488 ​cm​​ 3​·mol​​ −1​​

from which Z = 0.8866. In comparison with this result, the ideal-gas-state value is 13%
too high, and Eq. (3.36) gives a value 1.7% too high.

3.6  CUBIC EQUATIONS OF STATE

If an equation of state is to represent the PVT behavior of both liquids and vapors, it must
encompass a wide range of temperatures, pressures, and molar volumes. Yet it must not be
so complex as to present excessive numerical or analytical difficulties in application. Poly-
nomial equations that are cubic in molar volume offer a compromise between generality and
simplicity that is suitable to many purposes. Cubic equations are in fact the simplest equations
capable of representing both liquid and vapor behavior.

The van der Waals Equation of State


The first practical cubic equation of state was proposed by J. D. van der Waals9 in 1873:

RT a
​P = ​_
​ − ​_
    ​       ​​   (3.39)
V − b ​V​ 2​

Here, a and b are positive constants, specific to a particular species; when they are zero,
the equation for the ideal-gas state is recovered. The purpose of the term a/V2 is to account for
the attractive forces between molecules, which make the pressure lower than that which would
be exerted in the ideal-gas state. The purpose of constant b is to account 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 3.9 is a schematic PV diagram showing three such isotherms.

9Johannes Diderik van der Waals (1837–1923), Dutch physicist who won the 1910 Nobel Prize for physics. See: 

http://www.nobelprize.org/nobel_prizes/physics/laureates/1910/waals-bio.html.
96 CHAPTER 3.  Volumetric Properties of Pure Fluids

Superimposed is the “dome” representing states of saturated liquid and saturated vapor.10 For
the isotherm T1 > Tc, pressure decreases with increasing molar volume. The critical i­sotherm
(labeled Tc) contains the horizontal inflection at C characteristic of the critical point. For the
isotherm T2 < Tc, the pressure decreases rapidly in the subcooled-liquid region with i­ ncreasing
V; after crossing the saturated-liquid line, it goes through a minimum, rises to a maximum,
and then decreases, crossing the saturated-vapor line and continuing downward into the
­superheated-vapor region.

C
Figure 3.9: PV isotherms as
given by a cubic equation of
T1 > Tc
P state for T above, at, and below
the critical temperature. The
Tc superimposed darker curve
P sat shows the locus of saturated
vapor and liquid volumes.
T2 < Tc

V sat(liq) V sat(vap)
V

Experimental isotherms do not exhibit the smooth transition from saturated liquid to
saturated vapor characteristic of equations of state; rather, they contain a horizontal segment
within the two-phase region where saturated liquid and saturated vapor coexist in varying
proportions at the saturation (or vapor) pressure. This behavior, shown by the dashed line in
Fig. 3.9, cannot be represented by an equation of state, and we accept as inevitable the
­unrealistic behavior of equations of state in the two-phase region.
Actually, the PV behavior predicted in the two-phase region by proper cubic equations
of state is not wholly fictitious. If pressure is decreased on a saturated liquid devoid of vapor
nucleation sites in a carefully controlled experiment, vaporization does not occur, and liquid
persists alone to pressures well below its vapor pressure. Similarly, raising the pressure on
a saturated vapor in a suitable experiment does not cause condensation, and vapor persists

10Though far from obvious, the equation of state also provides the basis for calculation of the saturated liquid- and

vapor-phase volumes that determine the location of the “dome.” This is explained in Sec. 13.7.
3.6.  Cubic Equations of State 97

alone to pressures well above the vapor pressure. These nonequilibrium or metastable states
of superheated liquid and subcooled vapor are approximated by those portions of the PV
isotherm which lie in the two-phase region adjacent to the states of saturated liquid and
saturated vapor.11
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, reference to Fig. 3.9 shows that solution for V at any value of P yields
only one such root. For the critical isotherm (T = Tc), this is also true, 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. Although these roots are real
and positive, they are not physically stable states for the portion of an isotherm lying between
saturated liquid and saturated vapor (under the “dome”). Only for the saturation pressure Psat
are the roots, V sat(liq) and V sat(vap), stable states, lying at the ends of the horizontal portion of
the true isotherm. For any pressure other than P sat, there is only a single physically meaningful
root, corresponding to either a liquid or a vapor molar volume.

A Generic Cubic Equation of State


A mid-20th-century development of cubic equations of state was initiated in 1949 by publica-
tion of the Redlich/Kwong (RK) equation:12

RT a​(​ ​T)​ ​
P = ​____
​ − ​______
    ​       ​​   (3.40)
V − b V​​(​V + b​)​

Subsequent enhancements have produced an important class of equations, represented by a


generic cubic equation of state:

RT a​(​ ​T)​ ​
​P = ​_
​ − ​___________
    ​          ​​ (3.41)
V − b ​(​V + εb​)​(​V + σb​)​

The assignment of appropriate parameters leads not only to the van der Waals (vdW)
equation and the Redlich/Kwong (RK) equation, but also to the Soave/Redlich/Kwong (SRK)13
and the Peng/Robinson (PR) equations.14 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 temper-
ature dependence of a(T) is specific to each equation of state. The SRK equation is identical to
the RK equation, except for the T dependence of a(T). The PR equation takes different values
for ɛ and σ, as indicated in Table 3.1.

11The heating of liquids in a microwave oven can lead to a dangerous condition of superheated liquid, which can

“flash” explosively.
12Otto Redlich and J. N. S. Kwong, Chem. Rev., vol. 44, pp. 233–244, 1949.
13G. Soave, Chem. Eng. Sci., vol. 27, pp. 1197–1203, 1972.
14D.-Y. Peng and D. B. Robinson, Ind. Eng. Chem. Fundam., vol. 15, pp. 59–64, 1976.
98 CHAPTER 3.  Volumetric Properties of Pure Fluids

Determination of Equation-of-State Parameters


The parameters b and a(T) of Eq. (3.41) can in principle be found from PVT data, but suf-
ficient data are rarely available. They are in fact usually found from values for the critical
constants Tc and Pc. Because the critical isotherm exhibits a horizontal inflection at the critical
point, we may impose the mathematical conditions:

( ∂ V ) ( ∂ ​V​  ​)
∂ P ​∂​ 2​P
​​​ ​​ ___ ​ ​  ​​  ​​  = 0​  (3.42) ​​​ ​​ ____2  ​​  ​​  ​​  = 0​  (3.43)​​  ​​​​ 
T;cr T;cr

Subscript “cr” denotes the critical point. Differentiation of Eq. (3.41) yields
e­ xpressions for both derivatives, which may be equated to zero for P = Pc, T = Tc, and V =
Vc. The equation of state may itself be written for the critical conditions. These three equa-
tions contain five constants: Pc, Vc, Tc, a(Tc), and b. Of the several ways to treat these equa-
tions, the most suitable is elimination of Vc to yield expressions relating a(Tc) and b to Pc
and Tc. The reason is that Pc and Tc are more widely available and more accurately known
than Vc.
The algebra is intricate, but it leads eventually to the following expressions for parame-
ters b and a(Tc):

R​Tc​  ​
​b = Ω​_
​    ​ ​​ (3.44)
​Pc​  ​

and
​R​ 2​Tc​2​ ​
a​(​Tc​  ​)​ = Ψ​ _____
​    ​​  
​Pc​  ​

This result is extended to temperatures other than Tc by the introduction of a dimensionless


function α(Tr; ω) that becomes unity at the critical temperature:

α​​(​Tr​ ​; ω​)​R​ 2​T​c2​ ​
​a​(​T​)​ = Ψ​___________
​       ​​   (3.45)
​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 defini-
tion Tr ≡ T/Tc, and ω is a parameter specific to a given chemical species, defined and discussed
further below.
This analysis also shows that each equation of state implies a value of the critical com-
pressibility factor Zc that is the same for all substances. Different values are found for differ-
ent equations. Unfortunately, Zc values calculated from experimental values of Tc, Pc, and Vc
differ from one species to another, and they agree in general with none of the fixed values
predicted by common cubic equations of state. Experimental values are almost all smaller than
any of the predicted values.
3.6.  Cubic Equations of State 99

Roots of the Generic Cubic Equation of State


Equations of state are commonly transformed into expressions for the compressibility factor.
An equation for Z equivalent to Eq. (3.41) is obtained by substituting V = ZRT/P. In addition,
we define two dimensionless quantities that lead to simplification:

bP a​(​ T
​ )​ ​
​  ≡ ​___
β     ​ ​ (3.46) q​  ≡ ​____
     ​ ​ (3.47)
RT bRT

With these substitutions, Eq. (3.41) assumes the dimensionless form:


Z − β
Z = 1 + β − qβ​__________
​        ​​ (3.48)
(​ ​Z + εβ​)(​ ​Z + σβ​)​
Although the three roots of this equation may be found analytically, they are usually
calculated by iterative procedures built into mathematical software packages. Convergence
problems are most likely avoided when the equation is arranged to a form suited to the solution
for a particular root.
Equation (3.48) is particularly adapted to solving for vapor and vapor-like roots. Itera-
tive solution starts with the value Z = 1 substituted on the right side. The calculated value of
Z is returned to the right side and the process continues to convergence. The final value of Z
yields the volume root through V = ZRT/P = ZVig.
An alternative equation for Z is obtained when Eq. (3.48) is solved for the Z in the
numerator of the final fraction, yielding:

( qβ )
1 + β − Z
Z = β + ​​(​Z + εβ​)​(​Z + σβ​)​ _
​ ​      ​ ​​ (3.49)

This equation is particularly suited to solving for liquid and liquid-like roots. Iterative
solution starts with the value Z = β substituted on the right side. Once Z is known, the volume
root is again V = ZRT/P = ZVig.
Experimental compressibility-factor data show that values of Z for different fluids
exhibit similar behavior when correlated as a function of reduced temperature Tr and reduced
pressure Pr, where by definition Tr ≡ T/Tc and Pr ≡ P/Pc. It is therefore common to evalu-
ate equation-of-state parameters through these dimensionless variables. Thus, Eq. (3.46) and
Eq. (3.47) in combination with Eq. (3.44) and (3.45) yield:

​ r ​​
P Ψα​(​ T
​ r ​;​ ω​)​
​  = Ω​__
β    ​ ​  (3.50) q​  = ​_______
      ​ ​  (3.51)
​ r ​​
T Ω​Tr ​​

With parameters β and q evaluated by these equations, Z becomes a function of Tr and


Pr and the equation of state is said to be generalized because of its general applicability to all
gases and liquids. The numerical assignments for parameters ε, σ, Ω, and Ψ for the equations
of interest are summarized in Table 3.1. Expressions are also given for α(Tr; ω) for the SRK
and PR equations.
100 CHAPTER 3.  Volumetric Properties of Pure Fluids

Table 3.1: Parameter Assignments for Equations of State

Eqn. of State α(Tr) σ ε Ω Ψ Zc


vdW (1873) 1 0 0 1/8 27/64 3/8
RK (1949) ​Tr​ ​ −1/2​ 1 0 0.08664 0.42748 1/3
SRK (1972) ​ ​ SRK​(​Tr​ ​; ω​)​ †​
α 1 0 0.08664 0.42748 1/3
__ __
PR (1976) ​α​ PR(​ ​Tr​ ​; ω​)​ ‡​ 1​  + ​√ 2 ​  1​  − ​√2 
  ​  0.07780 0.45724 0.30740

​ †​α​ SRK​(​Tr​ ​; ω​)​ = ​​[​1 + (0.480 + 1.574 ω − 0.176​ ω​​  2​)(​​1 − T​ r​​​​​ 1/2​)​]​​​  2​​

​ ‡​α​ PR(​ ​Tr​ ​; ω​)​ = ​​[​1 + ​​(​0.37464 + 1.54226 ω − 0.26992​ ω​​ 2​)​(​1 − ​​Tr​ ​ 1/2​)​]​ 2​

Example 3.9
Given that the vapor pressure of n-butane at 350 K is 9.4573 bar, find the molar vol-
umes of (a) saturated-vapor and (b) saturated-liquid n-butane at these conditions as
given by the Redlich/Kwong equation.

Solution 3.9
Values of Tc and Pc for n-butane from App. B yield:
350 9.4573
​Tr​ ​ = ​_____
​   = 0.8233  and  ​Pr​ ​ = ​______
  ​      ​ 
 = 0.2491​
425.1 37.96
Parameter q is given by Eq. (3.51) with Ω, Ψ, and α(Tr) for the RK equation from
Table 3.1:
​ΨT​r−1/2
​  ​ __ Ψ 0.42748
q = ​ _______
​  ​  
  ​  ​  = ​________
= ​   ​T​r−3/2   ​(​0.8233​)​ −3/2​ = 6.6048​
  
Ω​Tr​ ​ Ω 0.08664
Parameter β is found from Eq. (3.50):
​Pr​ ​ ​(​0.08664​)​(​0.2491​)​
β = Ω​__
​    ​  = ​_______________

     ​  
= 0.026214​
​Tr​ ​ 0.8233
(a) For the saturated vapor, write the RK form of Eq. (3.48) that results upon sub-
stitution of appropriate values for ε and σ from Table 3.1:
(​ ​Z − β​)​
Z = 1 + β − qβ​______
​      ​​  
Z​(​ ​Z + β​)​
or
​(​Z − 0.026214​)​
Z = 1 + 0.026214 − ​​(​6.6048​)​(​0.026214​)_____________
​ ​   
   ​​
Z ​​(​Z + 0.026214​)​
3.6.  Cubic Equations of State 101

Solution yields Z = 0.8305, and


​(​0.8305​)​(​83.14​)​(​350​)​
ZRT _________________
​V​ v​ = ​____
​     ​  = ​  
       ​ = 2555 ​cm​​ 3​·mol​​ −1​​
P 9.4573

An experimental value is 2482 cm3·mol–1.


(b) For the saturated liquid, apply Eq. (3.49) in its RK form:

( qβ )
1 + β − Z
Z = β + Z​​(​Z + β​)​ _
​ ​      ​ ​​

or
​(​1.026214 − Z​)​
________________
Z = 0.026214 + Z​​(​Z + 0.026214​)​   
​ ​    ​​
​(​6.6048​)​(​0.026214​)​
Solution yields Z = 0.04331, and
​(​0.04331​)​(​83.14​)​(​350​)​
ZRT __________________
​V​ l​ = ​____
​     ​  = ​  
       ​ = 133.3​ cm​​ 3​·mol​​ −1​​
P 9.4573
An experimental value is 115.0 cm3·mol–1.

For comparison, values of Vv and Vl calculated for the conditions of Ex. 3.9 by all four
of the cubic equations of state considered here are summarized as follows:

Vv/cm3·mol–l Vl/cm3·mol–l

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

Corresponding States; Acentric Factor


The dimensionless thermodynamic coordinates Tr and Pr provide the basis for the simplest
corresponding-states correlations:
All fluids, when compared at the same reduced temperature and reduced
pressure, have approximately the same compressibility factor, and all
deviate from ideal-gas behavior to about the same degree.
Two-parameter corresponding-states correlations of Z require the use of only two reducing
parameters Tc and Pc. Although these correlations are very nearly exact for the simple fluids
(argon, krypton, and xenon), systematic deviations are observed for more complex fluids.
Appreciable improvement results from the introduction of a third corresponding-states param-
eter (in addition to Tc and Pc), characteristic of molecular structure. The most popular such
parameter is the acentric factor, ω, introduced by K. S. Pitzer and coworkers.15

15Fully described in K. S. Pitzer, Thermodynamics, 3rd ed., App. 3, McGraw-Hill, New York, 1995.
102 CHAPTER 3.  Volumetric Properties of Pure Fluids

The acentric factor for a pure chemical species is defined with reference to its vapor
pressure. The logarithm of the vapor pressure of a pure fluid is approximately linear in the
reciprocal of absolute temperature. This linearity can be expressed as
d log ​Pr​ ​ sat​
_________

​    ​  
= S​
d​​(​1 / ​Tr​ ​)​
where ​Pr​ ​ sat​is reduced vapor pressure, Tr is reduced temperature, and S is the slope of a plot of​
log​Pr​ ​ sat​vs. 1/Tr . Note that “log” here denotes the base 10 logarithm.
If two-parameter corresponding-states correlations were generally valid, the slope S
would be the same for all pure fluids. This is observed not to be true; within a limited range,
each fluid has its own characteristic value of S, which could in principle serve as a third
­corresponding-states parameter. However, Pitzer noted that all vapor-pressure data for the simple
fluids (Ar, Kr, Xe) lie on the same line when plotted as ​log ​Pr​ ​ sat​vs. 1/Tr and that the line passes
through l​og ​Pr​ ​ sat​ = −1.0​ at Tr = 0.7. This is illustrated in Fig. 3.10. Data for other fluids define
other lines whose locations can be fixed relative to the line for the simple fluids (SF) by the difference:

log​ P​r ​ sat​(​SF​)​ − log​ P​r ​ sat​​

The acentric factor is defined as this difference evaluated at Tr = 0.7:

​ω ≡ −1.0 − log​​(​Pr​ ​ sat​)​ ​Tr​ ​ = 0.7​​
​ (3.52)

Therefore ω 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 for a num-
ber of substances are listed in App. B.

1/ Tr
1.0 1.2 1.4 1.6 1.8 2.0
0

1 Figure 3.10: Approximate


log Prsat

Slope 2.3 temperature dependence of the


(Ar, Kr, Xe) reduced vapor pressure.

Slope 3.2
1 1
1.43 (n-Octane)
Tr 0.7

The definition of ω makes its value zero for argon, krypton, and xenon, and experimen-
tal data yield compressibility factors for all three fluids that are correlated by the same curves
when Z is plotted as a function of Tr and Pr. This is the basic premise of three-parameter
­corresponding-states correlations:
3.7.  Generalized Correlations for Gases 103

All fluids having the same value of acentric factor, when compared at the
same reduced temperature and reduced pressure, have about the same
compressibility factor, and all deviate from ideal-gas behavior to about
the same degree.
Equations of state that express Z as a function of only Tr and Pr yield ­two-parameter
corresponding-states correlations. The van der Waals and Redlich/Kwong equations are
­examples. The Soave/Redlich/Kwong (SRK) and Peng/Robinson equations, in which the acen-
tric factor enters through function α(Tr; ω) as an additional parameter, yield ­three-parameter
­corresponding-states correlations.

3.7  GENERALIZED CORRELATIONS FOR GASES

Generalized correlations find widespread use. Most popular are correlations of the kind
­developed by Pitzer and coworkers for the compressibility factor Z and for the second virial
coefficient B.16

Pitzer Correlations for the Compressibility Factor


The correlation for Z is:
Z = ​Z​ 0​ + ω​Z​ 1​​
​ (3.53)
where Z0 and Z1 are functions of both Tr and Pr . When ω = 0, as is the case for the simple fluids,
the second term disappears, and Z0 becomes identical with Z. Thus a generalized correlation
for Z as a function of Tr and Pr based only on data for argon, krypton, and xenon provides the
relationship Z 0 = F 0(Tr , Pr). By itself, this represents a two-parameter corresponding-states
correlation for Z.
Equation (3.53) is a simple linear relation between Z and ω for given values of Tr and
Pr . Experimental data for Z for nonsimple fluids plotted vs. ω at constant Tr and Pr do indeed
yield approximately straight lines, and their slopes provide values for Z1 from which the gen-
eralized function Z1 = F1(Tr , Pr) can be constructed.
Of the Pitzer-type correlations available, the one developed by Lee and Kesler17 is the
most widely used. It takes the form of tables that present values of Z 0 and Z1 as functions of
Tr and Pr . These are given in App. D as Tables D.1 through D.4. Use of these tables requires
interpolation, as demonstrated at the beginning of App. E. The nature of the correlation is
indicated by Fig. 3.11, a plot of Z 0 vs. Pr for six isotherms.
Fig. 3.12 shows Z 0 vs. Pr and Tr as a three-dimensional surface with isotherms and iso-
bars superimposed. The saturation curve, where there is a discontinuity in Z, is not precisely
defined in this plot, which is based on the data of Table D.1 and D.3. One should proceed with
caution when applying the Lee/Kesler tables near the saturation curve. Although the tables
contain values for liquid and vapor phases, the boundary between these will in general not be
the same as the saturation curve for a given real substance. The tables should not be used to
predict whether a substance is vapor or liquid at given conditions. Rather, one must know the
16See Pitzer, op. cit.
17B. I. Lee and M. G. Kesler, AIChE J., vol. 21, pp. 510–527, 1975.
104 CHAPTER 3.  Volumetric Properties of Pure Fluids

1.2

4.0

1.0

Tr = 0.7 Gases
1.5
0.8
0.9
1.2

Z 0 0.6
1.0
Two-phase
region 0.7

0.4

0.9
0.2
Compressed liquids
(Tr < 1.0)

0
0.05 0.1 0.2 0.5 1.0 0.5 5.0 10.0
Pr

Figure 3.11: The Lee/Kesler correlation for Z 0 = F 0(Tr , Pr).

phase of the substance, and then take care to interpolate or extrapolate only from points in the
table that represent the appropriate phase.
The Lee/Kesler correlation provides reliable results for gases which are nonpolar or only
slightly polar; for these, errors of no more than 2 or 3 percent are typical. When applied to
highly polar gases or to gases that associate, larger errors can be expected.
The quantum gases (e.g., hydrogen, helium, and neon) do not conform to the same
­corresponding-states behavior as do normal fluids. Their treatment by the usual correlations
is sometimes accommodated by use of temperature-dependent effective critical parameters.18
For hydrogen, the quantum gas most commonly found in chemical processing, the recom-
mended equations are:
43.6
​ ​Tc​  ​ / K = ​ __________   ​​  ​(​for H​ 2​)​​ (3.54)
21.8
1 + ​ ______   ​ 
2.016T
20.5
​Pc​  ​ / bar = ​ __________
​    ​  ​(​for H​ 2​)​​ (3.55)
44.2
______
1 + ​    ​ 
2.016T
18J. M. Prausnitz, R. N. Lichtenthaler, and E. G. de Azevedo, Molecular Thermodynamics of Fluid-Phase Equilib-

ria, 3d ed., pp. 172–173, Prentice Hall PTR, Upper Saddle River, NJ, 1999.
3.7.  Generalized Correlations for Gases 105

1.2

0.8

0
Z 0.6

0.4

0.2

0
1
10
0 4
10 3
–1 2
Pr 10
1 Tr
–2 0
10

Figure 3.12: Three-dimensional plot of the Lee/Kesler correlation for Z 0 = F 0(Tr , Pr) as given in
Table D.1 and D.3.

51.5
​Vc​  ​ / ​cm​​ 3​·mol​​ −1​ = ​ __________
​ ​(​for H​ 2​)​​
   ​​   (3.56)
9.91
______
1 − ​    ​ 
2.016T
where T is absolute temperature in kelvins. Use of these effective critical parameters for hydro-
gen requires the further specification that ω = 0.

Pitzer Correlations for the Second Virial Coefficient


The tabular nature of the generalized compressibility-factor correlation is a disadvantage, but
the complexity of the functions Z 0 and Z1 precludes their accurate representation by simple
equations. Nonetheless, we can give approximate analytical expression to these functions for
a limited range of pressures. The basis for this is Eq. (3.36), the simplest form of the virial
equation:

( R​Tc​  ​) ​Tr​ ​
BP B​Pc​  ​ ​Pr​ ​ ​Pr​ ​
Z = 1 + ​___
​     ​ = 1 + ​​ _
​    ​ __ ​    ​ = 1 + ​Bˆ​   __
​  ​​   (3.57)
RT ​Tr​ ​
106 CHAPTER 3.  Volumetric Properties of Pure Fluids

The reduced (and dimensionless) second virial coefficient and the Pitzer correlation for it are:

B​Pc​  ​ ​​ ˆ​   = ​B​ 0​ + ω​B​ 1​​  (3.59)


​​ ˆ​   ≡ ​____
B     ​  (3.58) B
R​Tc​  ​

Equations (3.57) and (3.59) together become:


​Pr​ ​ ​Pr​ ​
Z = 1 + ​B​ 0 ​ __
​ ​    ​ + ω​B​ 1 ​ __
​   ​​  
​Tr​ ​ ​Tr​ ​
Comparison of this equation with Eq. (3.53) provides the following identifications:
​Pr​ ​
​Z​ 0​ = 1 + ​B​ 0​ __
​ ​   ​​   (3.60)
​Tr​ ​
and
​Pr​ ​
​Z​ 1​ = ​B​ 1 __
​ ​    ​​
​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:19

0.422 0.172
​  ​0 ​= 0.083 − ​_____
B   1.6 ​  ​  ​1 ​= 0.139 − ​_____
​  (3.61) B   4.2 ​  ​  (3.62)
​  ​r ​ ​
T ​  ​r ​ ​
T

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 Z 0 and Z1 are at least approximately linear functions of reduced pressure. Figure 3.13
compares the linear relation of Z 0 to Pr as given by Eqs. (3.60) and (3.61) with values of
Z 0 from the Lee/Kesler compressibility-factor correlation, Tables D.1 and D.3 of App. D.
The two correlations differ by less than 2% in the region above the dashed line of the
figure. For reduced temperatures greater than Tr ≈ 3, there appears to be no limitation on
the pressure. For lower values of Tr the allowable pressure range decreases with decreasing
temperature. A point is reached, however, at Tr ≈ 0.7 where the pressure range is limited by the
saturation pressure.20 This is indicated by the left-most segment of the dashed line. The minor
contributions of Z1 to the correlations are here neglected. In view of the uncertainty associated
with any generalized correlation, deviations of no more than 2% in Z 0 are not significant.
The relative simplicity of the generalized second-virial-coefficient correlation does
much to recommend it. Moreover, temperatures and pressures of many chemical-processing

19These correlations first appeared in 1975 in the third edition of this book, attributed as a personal communication

to M. M. Abbott, who developed them.


20Although the Lee/Kesler tables of Appendix D list values for superheated vapor and subcooled liquid, they do

not provide values at saturation conditions.


3.7.  Generalized Correlations for Gases 107

4.0
1.0
2.4

1.8
0.9
Z0

Tr 0.8
0.9
1.0 1.1 1.5
0.8

1.3

0.7
0.0 0.5 1.0 1.5 2.0 2.5
Pr

Figure 3.13: Comparison of correlations for Z 0. The virial-coefficient correlation is represented


by the straight lines; the Lee/Kesler correlation, by the points. In the region above the dashed line
the two correlations differ by less than 2%.

operations lie within the region appropriate to the compressibility-factor correlation. Like the
parent correlation, it is most accurate for nonpolar species and least accurate for highly polar
and associating molecules.

Correlations for the Third Virial Coefficient


Accurate data for third virial coefficients are far less common than for second virial coef-
ficients. Nevertheless, generalized correlations for third virial coefficients do appear in the
literature.
Equation (3.38) may be written in reduced form as:
​Pr​ ​ 2
( ​Tr​ ​ Z )
​Pr​ ​
​ ​​     ​  + ​Cˆ ​​  ___
Z = 1 + ​Bˆ___ ​​    ​​  ​​​  ​​ (3.63)
​Tr​ ​ Z
where the reduced second virial coefficient ​​B ˆ​​  is defined by Eq. (3.58). The reduced (and
dimensionless) third virial coefficient and the Pitzer correlation for it are:

C​Pc ​  ​2​ ˆ 0 1
​​ ˆ  ​ ≡ ​______
C   2 2   ​ ​  (3.64) ​​C ​  = ​C​  ​ + ω​C​  ​​  (3.65)
​  ​ T
R ​ c ​  ​ ​
108 CHAPTER 3.  Volumetric Properties of Pure Fluids

An expression for C 0 as a function of reduced temperature is given by Orbey and Vera:21
0.02432 _______ 0.00313
​C ​​ 0​ = 0.01407 + ​________
​      
​  − ​  10.5    ​​   (3.66)
​Tr​ ​ ​Tr​ ​  ​
The expression for C 1 given by Orbey and Vera is replaced here by one that is algebraically
simpler, but essentially equivalent numerically:
0.05539 ________ 0.00242
​C​ 1​ = −0.02676 + ​________
​   2.7    ​  
− ​  10.5    ​​   (3.67)
​Tr​ ​  ​ ​Tr​ ​  ​
Equation (3.63) is cubic in Z, and it cannot be expressed in the form of Eq. (3.53). With
Tr and Pr specified, Z may be found by iteration. An initial value of Z = 1 on the right side of
Eq. (3.63) usually leads to rapid convergence.

The Ideal-Gas State as a Reasonable Approximation


The question often arises as to when the ideal-gas state may be a reasonable approximation to
reality. Figure 3.14 can serve as a guide.

10

Z 0 = 1.02
1

Figure 3.14: In the region lying Z 0 = 0.98


below the curves, where Z 0 Pr 0.1
lies between 0.98 and 1.02, the
ideal-gas state is a reasonable
approximation.

0.01

0.001
0 1 2 3 4
Tr

21H. Orbey and J. H. Vera, AIChE J., vol. 29, pp. 107–113, 1983.
3.7.  Generalized Correlations for Gases 109

Example 3.10
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.
​​ˆ ​​.
(c) Equation (3.57), with the generalized correlation for B 
​​ˆ ​​.
​​ˆ ​​ and C 
(d) Equation (3.63), with the generalized correlations for B 

Solution 3.10
(a) For the ideal-gas state,
RT ​(​83.14​)​(​510​)​
​V = ​___
   ​  = ​__________
      = 1696.1​ cm​​ 3​·mol​​ −1​
​  
P 25

(b) With values of Tc and Pc given in Table B.1 of App. B,

510 25
​Tr​ ​ = ​_____
  = 1.200​    ​Pr​ ​ = ​_____
  ​   = 0.659​​
     ​  
425.1 37.96

Interpolation in Tables D.1 and D.2 then provides:

​Z​ 0​ = 0.865​    ​Z​ 1​ = 0.038​

By Eq. (3.53) with ω = 0.200,

Z = ​Z​  0​ + ω​Z​ 1​ = 0.865 + ​​(​0.200​)​(​0.038​)​ = 0.873
​      ​(​0.873​)​(​83.14​)​(​510​)​
ZRT ________________ ​ ​
V = ​____
     ​  = ​      = 1480.7 ​cm​​ 3​·mol​​ −1​
​  
P 25

If Z1, the secondary term, is neglected, Z ​  = ​Z​ 0​ = 0.865.​This two-parameter


​  = 1467.1 ​cm​​ 3​·mol​​ −1​,​which is less than
­corresponding-states correlation yields V
1% lower than the value given by the three-parameter correlation.
(c) Values of B0 and B1 are given by Eqs. (3.61) and (3.62):
​B​ 0​ = −0.232​    ​B​ 1​ = 0.059​

Equations (3.59) and (3.57) then yield:

​ˆ​   = ​B​ 0​ + ω​B​ 1​ = −0.232 + ​​(​0.200​)​(​0.059​)​ = −0.220
B
​​     0.659  ​​ ​
Z = 1 + ​​(​−0.220​)_____ ​    
​ = 0.879
1.200
​  = 1489.1 ​cm​​ 3​·mol​​ −1​, a value less than 1% higher than that given by
from which V
the compressibility-factor correlation.
110 CHAPTER 3.  Volumetric Properties of Pure Fluids

(d) Values of C 0 and C1 are given by Eqs. (3.66) and (3.67):


​ ​C​ 0​ = 0.0339​    ​C​ 1​ = 0.0067​​
Equation (3.65) then yields:

​​C  ˆ ​ = ​C​ 0​ + ω​C​ 1​ = 0.0339 + ​​(​0.200​)​(​0.0067​)​ = 0.0352​
​​ˆ ​​and the value of B 
With this value of C  ​​ˆ​​ from part (c), Eq. (3.63) becomes,

( 1.200Z )
2

( 1.200Z )
0.659 0.659
Z = 1 + ​​(​−0.220​)​ _
​ ​   ​ ​ + ​​(​0.0352​)​​​ ​ ______ 
   ​ ​​​  ​​

  Solution for Z yields Z = 0.876 and V = 1485.8 cm3·mol–1. The value of


V differs from that of part (c) by about 0.2%. An experimental value for V is
1480.7 cm3·mol–1. Significantly, the results of parts (b), (c), and (d) are in excellent
agreement. Mutual agreement at these conditions is suggested by Fig. 3.13.

Example 3.11
What pressure is generated when 500 mol of methane is stored in a volume of 0.06 m3
at 50°C? Base calculations on each of the following:
(a) The ideal-gas state.
(b) The Redlich/Kwong equation.
(c) A generalized correlation.

Solution 3.11
​  = 0.06 / 500 = 0.0012 ​m​ 3​·mol​​ −1​​.
The molar volume of the methane is V
​  = 8.314 × ​10​​ −5​ bar·​m​ 3​·mol​​ −1​·K​​ −1​​:
(a) For the ideal-gas state, with R

(​ ​8.314 × ​10​​  ​)​(​323.15​)​
−5
RT __________________
P = ​___
​    ​  = ​  
       ​ = 223.9 bar​
V 0.00012
(b) The pressure as given by the Redlich/Kwong equation is:
RT a​(​ ​T)​ ​
P = ​____
​ − ​______
    ​       ​​   (3.40)
V − b V​​(​V + b​)​
Values of b and a(T) come from Eqs. (3.44) and (3.45), with Ω, Ψ, and
​α​(​Tr​ ​)​ = ​Tr​ ​ −1/2​from Table 3.1. With values of Tc and Pc from Table B.1, we have:

323.15
​Tr​ ​ = ​______
   ​ 
 = 1.695
190.6
(8.314 × ​10​​ −5​)​(​190.6​)​
_________________
​​       
    
b = 0.08664 ​  
       ​ 10​​ −5​ ​m​ 3​·mol​​ −1​
​ = 2.985 × ​ ​​
45.99
​(​1.695​)​ −0.5​(8.314 × ​10​​ −5​)​ 2​(​190.6​)​ 2​
___________________________
a = 0.42748 ​   
       ​ = 1.793 × ​10​​ −6​ bar​·m​​ 6​·mol​​ −2​
45.99
3.7.  Generalized Correlations for Gases 111

Substitution of numerical values into the Redlich/Kwong equation now yields:


(8.314 × ​10​​ −5​)​(​323.15​)​ __________________________
1.793 × ​10​​ −6​
​P = ​__________________
  
     

​ − ​

  
     ​ = 198.3 bar​
0.00012 − 2.985 × ​10​​ −5​ 0.00012(0.00012 + 2.985 × ​10​​ −5​)
(c) B
 ecause the pressure here is high, the generalized compressibility-factor
correlation is the proper choice. In the absence of a known value for Pr , an
iterative procedure is based on the following equation:
Z(8.314 × ​10​​ −5​)​(​323.15​)​
ZRT ___________________
​P = ​____
    ​  = ​  
       ​ = 223.9 Z​
V 0.00012
Because P
​  = ​Pc​  ​Pr​ ​ = 45.99 ​Pr​ ​, this equation becomes:
45.99 ​Pr​ ​ Z
​Z = ​________
     = 0.2054 ​Pr​ ​    or​    ​Pr​ ​ = ​_______
​        ​ 
223.9 0.2054
  One now assumes a starting value for Z, say Z = 1. This gives Pr = 4.68, and
allows a new value of Z to be calculated by Eq. (3.53) from values interpolated
in Tables D.3 and D.4 at the reduced temperature of Tr = 1.695. With this new
value of Z, a new value of Pr is calculated, and the procedure continues until no
significant change occurs from one step to the next. The final value of Z so found
is 0.894 at Pr = 4.35. This may be confirmed by substitution into Eq. (3.53) of
values for Z0 and Z1 from Tables D.3 and D.4 interpolated at Pr = 4.35 and Tr =
1.695. With ω = 0.012,
Z = ​Z​ 0​ + ω​Z​ 1​ = 0.891 + ​​(​0.012​)​(​0.268​)​ = 0.894

​     ​(​0.894​)​(8.314 × ​10​​ −5​)​(​323.15​)​
ZRT __________________ ​ ​​
P = ​____
     ​  = ​   
  
     ​ = 200.2​ bar​
V 0.00012
Because the acentric factor is small, the two- and three-parameter compressibility-
factor correlations are little different. The Redlich/Kwong equation and the generalized
compressibility-factor correlation give answers within 2% of the experimental value of
196.5 bar.

Example 3.12
A mass of 500 g of gaseous ammonia is contained in a vessel of 30,000 cm3 volume and
immersed in a constant-temperature bath at 65°C. Calculate the pressure of the gas by:
(a) The ideal-gas state;
(b) A generalized correlation.

SOLUTION 3.12
The molar volume of ammonia in the vessel is:

​V​ t​ ​V​ t​ 30,000


V = ​__
​    ​  = ​____
     ​  = ​_________
  = 1021.2 ​cm​​ 3​·mol​​ −1​​
  ​  
n m / ℳ 500 / 17.02
112 CHAPTER 3.  Volumetric Properties of Pure Fluids

(a) For the ideal-gas state,


RT ​(​83.14​)​(​65 + 273.15​)​
P = ​___
​    ​  = ​________________
         ​ = 27.53 bar​
V 1021.2
(b) Because the reduced pressure is low (Pr ≈ 27.53/112.8 = 0.244), the general-
ized virial-coefficient correlation should suffice. Values of B0 and B1 are given by
Eqs. (3.61) and (3.62). With Tr = 338.15/405.7 = 0.834,

​B​ 0​ = −0.482​    ​B​ 1​ = −0.232​​
Substitution into Eq. (3.59) with ω = 0.253 yields:

​ˆ​   = − 0.482 + ​​(​0.253​)​(​−0.232​)​ = −0.541
B
​​     ​Bˆ​R −​​(​0.541​)​(​83.14​)​(​405.7​)​
   ​Tc​  ​ __________________ ​ ​​
B = ​ _____  ​   = ​  
       ​ = −161.8​ cm​​ 3​·mol​​ −1​
​Pc​  ​ 112.8
By the second equality of Eq. (3.36):
RT ​(​83.14​)​(​338.15​)​
​P = ​____
  = ​_____________
  ​      
   ​ = 23.76 bar​
V − B 1021.2 + 161.8
An iterative solution is not necessary because B is independent of pressure. The c­al-
culated P corresponds to a reduced pressure of Pr = 23.76/112.8 = 0.211. Reference
to Fig. 3.13 confirms the suitability of the generalized virial-coefficient correlation.
Experimental data indicate that the pressure is 23.82 bar at the given condi-
tions. Thus the ideal-gas state yields an answer high by about 15%, whereas the
virial-coefficient correlation gives an answer in substantial agreement with exper-
iment, even though ammonia is a polar molecule.

3.8  GENERALIZED CORRELATIONS FOR LIQUIDS

Although the molar volumes of liquids can be calculated by means of generalized cubic equa-
tions of state, the results are often not of high accuracy. However, the Lee/Kesler correlation
includes data for subcooled liquids, and Fig. 3.11 illustrates curves for both liquids and gases.
Values for both phases are provided in Tables D.1 through D.4 of App. D. Recall, however,
that this correlation is most suitable for nonpolar and slightly polar fluids.
In addition, generalized equations are available for the estimation of molar volumes of
saturated liquids. The simplest equation, proposed by Rackett,22 is an example:
2/7
​V​ sat​ = ​Vc​  ​​​Zc​  ​​​​  ​(1​ −T​ r​​)​​  ​​​
​ (3.68)
An alternative form of this equation is sometimes useful:
​Pr​ ​ ​[1+​​(​1−​Tr​ ​)​​​  2/7​]​
​Z​ sat​ = ​__
​    ​Z   ​ ​  ​​ (3.69)
​Tr​ ​ c
22H. G. Rackett, J. Chem. Eng. Data, vol. 15, pp. 514–517, 1970; see also C. F. Spencer and S. B. Adler, ibid.,

vol. 23, pp. 82–89, 1978, for a review of available equations.


3.8.  Generalized Correlations for Liquids 113

The only data required are the critical constants, given in Table B.1 of App. B. Results are
usually accurate to 1 or 2%.
Lydersen, Greenkorn, and Hougen23 developed a two-parameter corresponding-states
correlation for estimation of liquid volumes. It provides a correlation of reduced density ρr as
a function of reduced temperature and pressure. By definition,
ρ ​Vc​  ​
​ρ​ r​ ≡ ​__
​     ​ = ​___
   ​ ​ (3.70)
​ρ​ c​ V
where ρc is the density at the critical point. The generalized correlation is shown by Fig. 3.15.
This figure may be used directly with Eq. (3.70) for determination of liquid volumes if the
value of the critical volume is known. A better procedure is to make use of a single known
liquid volume (state 1) by the identity,
​ρ​ ​r​  ​​
​V​ 2​ = ​V​ 1​ ___
​ ​  1  ​​ (3.71)
​ρ​ ​r​ 2​​
where
​ ​ 2​ = required volume
V
​    ​V​ 1​ =​ known volume​ ​
   ​
​ρ​ ​r​ 1​​,​ ρ​ ​r​ 2​​ = reduced densities read from Fig. 3.15

3.5
Tr = 0.3
0.4
0.5
3.0
0.6
0.7
0.8
2.5 0.9
1.0
ρr Tr = 0.95

2.0
0.97
0.99
1.5
Saturated Liquid

1.0
0 1 2 3 4 5 6 7 8 9 10
Pr

Figure 3.15: Generalized density correlation for liquids.

23A. L. Lydersen, R. A. Greenkorn, and O. A. Hougen, “Generalized Thermodynamic Properties of Pure Fluids,”

Univ. Wisconsin, Eng. Expt. Sta. Rept. 4, 1955.


114 CHAPTER 3.  Volumetric Properties of Pure Fluids

This method gives good results and requires only experimental data that are usually available.
Figure 3.15 makes clear the increasing effects of both temperature and pressure on liquid den-
sity as the critical point is approached.
Correlations for the molar densities as functions of temperature are given for many pure
liquids by Daubert and coworkers.24

Example 3.13
For ammonia at 310 K, estimate the density of:
(a) The saturated liquid;
(b) The liquid at 100 bar.

Solution 3.13
(a) Apply the Rackett equation at the reduced temperature,  ​​Tr​ ​ = 310 / 405.7 = 0.7641​.
With ​Vc​  ​ = 72.47​and ​Zc​  ​ = 0.242​(from Table B.1),
2/7 2/7
​V​ sat​ = ​Vc​  ​​​Zc​  ​​​​  ​(1−​Tr​  ​​)​​  ​​  = ​​(​72.47​)​(​0.242​)​ ​(​0.2359​)​  ​ ​= 28.33 ​cm​​ 3​·mol​​ −1​​

For comparison, the experimental value is 2​ 9.14 ​cm​​ 3​·mol​​ −1​, a 2.7% difference.
(b) The reduced conditions are:
100
​Tr​ ​ = 0.764​    ​Pr​ ​ = ​_____
  = 0.887​​
  ​  
112.8
Substituting the value, ρr = 2.38 (from Fig. 3.15), and Vc into Eq. (3.70) gives:
​Vc​  ​ 72.47
​V = ​___
   ​  = ​_____
   = 30.45 ​cm​​ 3​·mol​​ −1​
 ​ 
​ρ​ r​ 2.38
In comparison with the experimental value of 28.6 cm3 ·mol–1, this result is higher
by 6.5%.
If we start with the experimental value of 29.14 cm3·mol–1 for saturated liquid
at 310 K, Eq. (3.71) may be used. For the saturated liquid at ​Tr​ ​ = 0.764,​ ρ​ ​r​ 1​ ​ = 2.34​
(from Fig. 3.15). Substitution of known values into Eq. (3.71) gives:

( 2.38 )
​ρ​ ​r​  ​​ 2.34
​V​ 2​ = ​V​ 1___
​ ​  1  ​ = ​​(​29.14​)​ _
​  ​ ​ = 28.65 ​cm​​ 3​·mol​​ −1​​
  
​ρ​ ​r​ 2​​

This result is in essential agreement with the experimental value.


Direct application of the Lee/Kesler correlation with values of Z 0 and Z1
interpolated from Tables D.1 and D.2 leads to a value of 33.87 cm3·mol–1, which
is significantly in error, no doubt owing to the highly polar nature of ammonia.
24T. E. Daubert, R. P. Danner, H. M. Sibul, and C. C. Stebbins, Physical and Thermodynamic Properties of Pure

Chemicals: Data Compilation, Taylor & Francis, Bristol, PA, extant 1995.
3.9. Synopsis 115

3.9 SYNOPSIS

After studying this chapter, including the end-of-chapter problems, one should be able to:

∙ State and apply the phase rule for nonreacting systems


∙ Interpret PT and PV diagrams for a pure substance, identifying the solid, liquid, gas, and
fluid regions; the fusion (melting), sublimation, and vaporization curves; and the critical
and triple points
∙ Draw isotherms on a PV diagram for temperatures above and below the critical
temperature
∙ Define isothermal compressibility and volume expansivity and use them in calculations
for liquids and solids
∙ Make use of the facts that for the ideal-gas state Uig and Hig depend only on T (not on P
ig ig
and Vig), and that ​C​P​ ​ = ​C​V​ ​ + R​
∙ Compute heat and work requirements and property changes for mechanically reversible
isothermal, isobaric, isochoric, and adiabatic processes in the ideal-gas state
∙ Define and use the compressibility factor Z
∙ Intelligently select an appropriate equation of state or generalized correlation for appli-
cation in a given situation, as indicated by the following chart:

(a) Ideal-gas state

(b) 2-term virial equation


Gas
(c) Cubic equation of state

(d) Lee/Kesler tables, Appendix D


Gas or
liquid?
(e) Incompressible liquid

(f) Rackett equation, Eq. (3.68)

Liquid (g) Constant β and κ

(h) Lydersen et al. chart, Fig. 3.15

∙ Apply the two-term virial equation of state, written in terms of pressure or molar density
∙ Relate the second and third virial coefficients to the slope and curvature of a plot of the
compressibility factor versus molar density
∙ Write the van der Waals and generic cubic equations of state, and explain how the
equation-of-state parameters are related to critical properties
116 CHAPTER 3.  Volumetric Properties of Pure Fluids

∙ Define and use Tr , Pr , and ω


∙ Explain the basis for the two- and three-parameter corresponding-states correlations
∙ Compute parameters for the Redlich/Kwong, Soave/Redlich/Kwong, and Peng/Robinson
equations of state from critical properties
∙ Solve any of the cubic equations of state, where appropriate, for the vapor or vapor-like
and/or liquid or liquid-like molar volumes at given T and P
∙ Apply the Lee/Kesler correlation with data from Appendix D
∙ Determine whether the Pitzer correlation for the second virial coefficient is applicable
for given T and P, and use it if appropriate
∙ Estimate liquid-phase molar volumes by generalized correlations

3.10 PROBLEMS

How many phase rule variables must be specified to fix the thermodynamic state of
3.1.
each of the following systems?

(a) A sealed flask containing a liquid ethanol-water mixture in equilibrium with its vapor.
(b) A sealed flask containing a liquid ethanol-water mixture in equilibrium with its
vapor and nitrogen.
(c) A sealed flask containing ethanol, toluene, and water as two liquid phases plus vapor.

A renowned laboratory reports quadruple-point coordinates of 10.2 Mbar and 24.1°C


3.2.
for four-phase equilibrium of allotropic solid forms of the exotic chemical β-miasmone.
Evaluate the claim.

A closed, nonreactive system contains species 1 and 2 in vapor/liquid equilibrium.


3.3.
Species 2 is a very light gas, essentially insoluble in the liquid phase. The vapor phase
contains both species 1 and 2. Some additional moles of species 2 are added to the
system, which is then restored to its initial T and P. As a result of the process, does the
total number of moles of liquid increase, decrease, or remain unchanged?

A system comprised of chloroform, 1,4-dioxane, and ethanol exists as a two-phase


3.4.
vapor/liquid system at 50°C and 55 kPa. After the addition of some pure ethanol, the
system can be returned to two-phase equilibrium at the initial T and P. In what respect
has the system changed, and in what respect has it not changed?

For the system described in Prob. 3.4:


3.5.

(a) How many phase-rule variables in addition to T and P must be chosen so as to fix
the compositions of both phases?
(b) If the temperature and pressure are to remain the same, can the overall composi-
tion of the system be changed (by adding or removing material) without affecting
the compositions of the liquid and vapor phases?
3.10. Problems
3.10 117

Express the volume expansivity and the isothermal compressibility as functions of


3.6.
density ρ and its partial derivatives. For water at 50°C and 1 bar, κ​  = 44.18 × ​10​​ −6​  ​
bar​​ −1​. To what pressure must water be compressed at 50°C to change its density by
1%? Assume that κ is independent of P.

Generally, volume expansivity β and isothermal compressibility κ depend on T and P.


3.7.
Prove that:

( ∂ P ) ( ∂ T )
∂ β ∂ κ
​​​ ___
​    ​  ​ ​​​  =  −​​ ___
​    ​  ​​  ​​​
T P
The Tait equation for liquids is written for an isotherm as:
3.8.

( B + P )
AP
V = ​V​ 0​ ​1 − ​_
​     ​ ​​

where V is molar or specific volume, V0 is the hypothetical molar or specific volume


at zero pressure, and A and B are positive constants. Find an expression for the isother-
mal compressibility consistent with this equation.

For liquid water the isothermal compressibility is given by:


3.9.
c
κ = ​______
​      ​​ 
V​​(​P + b​)​
where c and b are functions of temperature only. If 1 kg of water is compressed iso-
thermally and reversibly from 1 to 500 bar at 60°C, how much work is required? At
60°C, b = 2700 bar and c = 0.125 cm3·g–1.

3.10. Calculate the reversible work done in compressing 1(ft)3 of mercury at a constant
temperature of 32(°F) from 1(atm) to 3000(atm). The isothermal compressibility of
mercury at 32(°F) is:

κ / ​(​atm​)​ −1​ = 3.9 × ​10​​ −6​ − 0.1 × ​10​​ −9​ P​/​(​atm​)​​

3.11. Five kilograms of liquid carbon tetrachloride undergo a mechanically reversible, iso-
baric change of state at 1 bar during which the temperature changes from 0°C to 20°C.
Determine ΔVt, W, Q, ΔHt, and ΔUt. The properties for liquid carbon tetrachloride
at 1 bar and 0°C may be assumed independent of temperature: β ​  = 1.2 × ​10​​ −3​  ​K​ −1​​,​​
−1 −1
C​P  ​ = 0.84 ​kJ·kg​​  ​·K​​  ​, and ρ −3
​  = 1590 ​kg·m​​  ​​.

3.12. Various species of hagfish, or slime eels, live on the ocean floor, where they burrow
inside other fish, eating them from the inside out and secreting copious amounts of
slime. Their skins are widely used to make eelskin wallets and accessories. Suppose
a hagfish is caught in a trap at a depth of 200 m below the ocean surface, where the
water temperature is 10°C, then brought to the surface where the temperature is 15°C.
If the isothermal compressibility and volume expansivity are assumed constant and
equal to the values for water,

​(​β = ​10​​ −4​ K​​ −1​ and κ = 4.8 × ​10​​ −5​ ​bar​​ −1​)​
118 CHAPTER 3.  Volumetric Properties of Pure Fluids

what is the fractional change in the volume of the hagfish when it is brought to the surface?

Table 3.2 provides the specific volume, isothermal compressibility, and volume expansivity
of several liquids at 20°C and 1 bar25 for use in Problems 3.13 to 3.15, where β and κ may be
assumed constant.

Table 3.2: Volumetric Properties of Liquids at 20°C

Specific Isothermal Volume


Molecular Volume Compressibility Expansivity
Formula Chemical Name V/L·kg–1 κ/10–5 bar–1 β/10–3·°C–1
C2H4O2 Acetic Acid 0.951 9.08 1.08
C6H7N Aniline 0.976 4.53 0.81
CS2 Carbon Disulfide 0.792 9.38 1.12
C6H5Cl Chlorobenzene 0.904 7.45 0.94
C6H12 Cyclohexane 1.285 11.3 1.15
C4H10O Diethyl Ether 1.401 18.65 1.65
C2H5OH Ethanol 1.265 11.19 1.40
C4H8O2 Ethyl Acetate 1.110 11.32 1.35
C8H10 m-Xylene 1.157 8.46 0.99
CH3OH Methanol 1.262 12.14 1.49
CCl4 Tetrachloromethane 0.628 10.5 1.14
C7H8 Toluene 1.154 8.96 1.05
CHCl3 Trichloromethane 0.672 9.96 1.21

3.13. For one of the substances in Table 3.2, compute the change in volume and work done
when one kilogram of the substance is heated from 15°C to 25°C at a constant pres-
sure of 1 bar.

3.14. For one of the substances in Table 3.2, compute the change in volume and work done
when one kilogram of the substance is compressed from 1 bar to 100 bar at a constant
temperature of 20°C.

3.15. For one of the substances in Table 3.2, compute the final pressure when the substance
is heated from 15°C and 1 bar to 25°C at constant volume.

3.16. A substance for which κ is a constant undergoes an isothermal, mechanically reversi-
ble process from initial state (P1, V1) to final state (P2, V2), where V is molar volume.

(a) Starting with the definition of κ, show that the path of the process is described by:
​V = A​(​ ​T)​ ​exp​(​ −κP​)​​

(b) Determine an exact expression which gives the isothermal work done on 1 mol of
this constant-κ substance.
25CRC Handbook of Chemistry and Physics, 90th Ed., pp. 6-140-6-141 and p. 15–25, CRC Press, Boca Raton,

Florida, 2010.
3.10. Problems
3.10 119

3.17. One mole of an ideal gas with CP = (7/2)R and CV = (5/2)R expands from P1 = 8 bar
and T1 = 600 K to P2 = 1 bar by each of the following paths:
(a) Constant volume;
(b) Constant temperature;
(c) Adiabatically.
Assuming mechanical reversibility, calculate W, Q, ΔU, and ΔH for each process.
Sketch each path on a single PV diagram.

3.18. One mole of an ideal gas with CP = (5/2)R and CV = (3/2)R expands from P1 = 6 bar
and T1 = 800 K to P2 = 1 bar by each of the following paths:
(a) Constant volume
(b) Constant temperature
(c) Adiabatically
Assuming mechanical reversibility, calculate W, Q, ΔU, and ΔH for each process.
Sketch each path on a single PV diagram.

3.19. An ideal gas initially at 600 K and 10 bar undergoes a four-step mechanically revers-
ible cycle in a closed system. In step 12, pressure decreases isothermally to 3 bar; in
step 23, pressure decreases at constant volume to 2 bar; in step 34, volume decreases
at constant pressure; and in step 41, the gas returns adiabatically to its initial state.
Take CP = (7/2)R and CV = (5/2)R.

(a) Sketch the cycle on a PV diagram.


(b) Determine (where unknown) both T and P for states 1, 2, 3, and 4.
(c) Calculate Q, W, ΔU, and ΔH for each step of the cycle.

3.20. An ideal gas initially at 300 K and 1 bar undergoes a three-step mechanically revers-
ible cycle in a closed system. In step 12, pressure increases isothermally to 5 bar; in
step 23, pressure increases at constant volume; and in step 31, the gas returns adiabati-
cally to its initial state. Take CP = (7/2)R and CV = (5/2)R.

(a) Sketch the cycle on a PV diagram.


(b) Determine (where unknown) V, T, and P for states 1, 2, and 3.
(c) Calculate Q, W, ΔU, and ΔH for each step of the cycle.

3.21. The state of an ideal gas with CP = (5/2)R is changed from P = 1 bar and V​  ​​ 1t ​ ​  = 12 ​m​ 3​​
​​ 2t ​ ​  = 1 ​m​ 3​by the following mechanically reversible processes:
to P2 = 12 bar and V​ 

(a) Isothermal compression.


(b) Adiabatic compression followed by cooling at constant pressure.
(c) Adiabatic compression followed by cooling at constant volume.
(d) Heating at constant volume followed by cooling at constant pressure.
(e) Cooling at constant pressure followed by heating at constant volume.

Calculate Q, W, ΔUt, and ΔHt for each of these processes, and sketch the paths of all
processes on a single PV diagram.
120 CHAPTER 3.  Volumetric Properties of Pure Fluids

3.22. The environmental lapse rate dT/dz characterizes the local variation of temperature
with elevation in the earth’s atmosphere. Atmospheric pressure varies with elevation
according to the hydrostatic formula,
dP
___
​ ​   ​  = −ℳρg​
dz
where ℳ is molar mass, ρ is molar density, and g is the local acceleration of gravity.
Asssume that the atmosphere is an ideal gas, with T related to P by the polytropic
formula:

T​P​ ​(​1−δ)​⁄ δ​​ = constant​

Develop an expression for the environmental lapse rate in relation to ℳ, g, R, and δ.

3.23. An evacuated tank is filled with gas from a constant-pressure line. Develop an expres-
sion relating the temperature of the gas in the tank to the temperature T′ of the gas in
the line. Assume the gas is ideal with constant heat capacities, and ignore heat transfer
between the gas and the tank. 

3.24. A tank of 0.1 m3 volume contains air at 25°C and 101.33 kPa. The tank is connected to
a compressed-air line which supplies air at the constant conditions of 45°C and 1500
kPa. A valve in the line is cracked so that air flows slowly into the tank until the pres-
sure equals the line pressure. If the process occurs slowly enough that the temperature
in the tank remains at 25°C, how much heat is lost from the tank? Assume air to be an
ideal gas for which CP = (7/2)R and CV = (5/2)R.

3.25. Gas at constant T and P is contained in a supply line connected through a valve to a
closed tank containing the same gas at a lower pressure. The valve is opened to allow
flow of gas into the tank, and then is shut again.

(a) Develop a general equation relating n1 and n2, the moles (or mass) of gas in the
tank at the beginning and end of the process, to the properties U1 and U2, the
internal energy of the gas in the tank at the beginning and end of the process, and
H′, the enthalpy of the gas in the supply line, and to Q, the heat transferred to the
material in the tank during the process.
(b) Reduce the general equation to its simplest form for the special case of an ideal
gas with constant heat capacities.
(c) Further reduce the equation of (b) for the case of n1 = 0.
(d) Further reduce the equation of (c) for the case in which, in addition, Q = 0.
(e) Treating nitrogen as an ideal gas for which CP = (7/2)R, apply the appropriate
equation to the case in which a steady supply of nitrogen at 25°C and 3 bar flows
into an evacuated tank of 4 m3 volume, and calculate the moles of nitrogen that
flow into the tank to equalize the pressures for two cases:

1. Assume that no heat flows from the gas to the tank or through the tank walls.
2. Assume that the tank weighs 400 kg, is perfectly insulated, has an initial tem-
perature of 25°C, has a specific heat of 0.46 kJ·kg−1·K−1, and is heated by the
gas so as always to be at the temperature of the gas in the tank.
3.10. Problems
3.10 121

3.26. Develop equations that may be solved to give the final temperature of the gas
remaining in a tank after the tank has been bled from an initial pressure P1 to a final
pressure P2. Known quantities are initial temperature, tank volume, heat capacity of
the gas, total heat capacity of the containing tank, P1, and P2. Assume the tank to be
always at the temperature of the gas remaining in the tank and the tank to be perfectly
insulated.

3.27. A rigid, nonconducting tank with a volume of 4 m3 is divided into two unequal parts
by a thin membrane. One side of the membrane, representing 1/3 of the tank, contains
nitrogen gas at 6 bar and 100°C, and the other side, representing 2/3 of the tank, is
evacuated. The membrane ruptures and the gas fills the tank.

(a) What is the final temperature of the gas? How much work is done? Is the process
reversible?
(b) Describe a reversible process by which the gas can be returned to its initial state.
How much work is done?

Assume nitrogen is an ideal gas for which CP = (7/2)R and CV = (5/2)R.

3.28. An ideal gas, initially at 30°C and 100 kPa, undergoes the following cyclic processes
in a closed system:

(a) In mechanically reversible processes, it is first compressed adiabatically to


500  kPa, then cooled at a constant pressure of 500 kPa to 30°C, and finally
expanded isothermally to its original state.
(b) The cycle traverses exactly the same changes of state, but each step is irreversible
with an efficiency of 80% compared with the corresponding mechanically revers-
ible process. Note: The initial step can no longer be adiabatic.

Calculate Q, W, ΔU, and ΔH for each step of the process and for the cycle. Take CP =
(7/2)R and CV = (5/2)R.

3.29. One cubic meter of an ideal gas at 600 K and 1000 kPa expands to five times its initial
volume as follows:

(a) By a mechanically reversible, isothermal process.


(b) By a mechanically reversible, adiabatic process.
(c) By an adiabatic, irreversible process in which expansion is against a restraining
pressure of 100 kPa.

For each case calculate the final temperature, pressure, and the work done by the gas.
Take CP = 21 J·mol−1·K−1.

3.30. One mole of air, initially at 150°C and 8 bar, undergoes the following mechanically
reversible changes. It expands isothermally to a pressure such that when it is cooled
at constant volume to 50°C its final pressure is 3 bar. Assuming air is an ideal gas for
which CP = (7/2)R and CV = (5/2)R, calculate W, Q, ΔU, and ΔH.
122 CHAPTER 3.  Volumetric Properties of Pure Fluids

3.31. An ideal gas flows through a horizontal tube at steady state. No heat is added and
no shaft work is done. The cross-sectional area of the tube changes with length, and
this causes the velocity to change. Derive an equation relating the temperature to the
velocity of the gas. If nitrogen at 150°C flows past one section of the tube at a velocity
of 2.5 m·s−1, what is its temperature at another section where its velocity is 50 m·s−1?
Assume CP = (7/2)R.

3.32. One mole of an ideal gas, initially at 30°C and 1 bar, is changed to 130°C and 10 bar
by three different mechanically reversible processes:

∙ The gas is first heated at constant volume until its temperature is 130°C; then it is
compressed isothermally until its pressure is 10 bar.
∙ The gas is first heated at constant pressure until its temperature is 130°C; then it is
compressed isothermally to 10 bar.
∙ The gas is first compressed isothermally to 10 bar; then it is heated at constant
pressure to 130°C.

Calculate Q, W, ΔU, and ΔH in each case. Take CP = (7/2)R and CV = (5/2)R. Alter-
natively, take CP = (5/2)R and CV = (3/2)R.

3.33. One mole of an ideal gas, initially at 30°C and 1 bar, undergoes the following mechan-
ically reversible changes. It is compressed isothermally to a point such that when it is
heated at constant volume to 120°C its final pressure is 12 bar. Calculate Q, W, ΔU,
and ΔH for the process.

3.34. One mole of an ideal gas in a closed system, initially at 25°C and 10 bar, is first
expanded adiabatically, then heated isochorically to reach a final state of 25°C and
1 bar. Assuming these processes are mechanically reversible, compute T and P after
the adiabatic expansion, and compute Q, W, ΔU, and ΔH for each step and for the
overall process. Take CP = (7/2)R and CV = (5/2)R.

3.35. A process consists of two steps: (1) One mole of air at T = 800 K and P = 4 bar is
cooled at constant volume to T = 350 K. (2) The air is then heated at constant pressure
until its temperature reaches 800 K. If this two-step process is replaced by a single
isothermal expansion of the air from 800 K and 4 bar to some final pressure P, what is
the value of P that makes the work of the two processes the same? Assume mechanical
reversibility and treat air as an ideal gas with CP = (7/2)R and CV = (5/2)R.

3.36. One cubic meter of argon is taken from 1 bar and 25°C to 10 bar and 300°C by each
of the following two-step paths. For each path, compute Q, W, ΔU, and ΔH for each
step and for the overall process. Assume mechanical reversibility and treat argon as an
ideal gas with CP = (5/2)R and CV = (3/2)R.

(a) Isothermal compression followed by isobaric heating.


(b) Adiabatic compression followed by isobaric heating or cooling.
(c) Adiabatic compression followed by isochoric heating or cooling.
(d) Adiabatic compression followed by isothermal compression or expansion.
3.10. Problems
3.10 123

3.37. A scheme for finding the internal volume ​V​Bt ​​ of a gas cylinder consists of the following
steps. The cylinder is filled with a gas to a low pressure P1, and connected through a
small line and valve to an evacuated reference tank of known volume ​V​At ​​.  The valve
is opened, and gas flows through the line into the reference tank. After the system
returns to its initial temperature, a sensitive pressure transducer provides a value for
the pressure change ΔP in the cylinder. Determine the cylinder volume ​V​Bt ​​  from the
following data:

​V​At ​ ​ = 256​ cm​​ 3​
∙ ​
∙ ​ΔP / ​P​ 1​ = −0.0639​

3.38. A closed, nonconducting, horizontal cylinder is fitted with a nonconducting, friction-


less, floating piston that divides the cylinder into Sections A and B. The two sections
contain equal masses of air, initially at the same conditions, T1 = 300 K and P1 =
1(atm). An electrical heating element in Section A is activated, and the air temper-
atures slowly increase: TA in Section A because of heat transfer, and TB in Section B
because of adiabatic compression by the slowly moving piston. Treat air as an ideal
gas with ​CP​  ​ = ​_ 27 ​R​, and let nA be the number of moles of air in Section A. For the pro-
cess as described, evaluate one of the following sets of quantities:

(a) TA, TB, and Q/nA, if P(final) = 1.25(atm)


(b) TB, Q/nA, and P (final), if TA = 425 K
(c) TA, Q/nA, and P (final), if TB = 325 K
(d) TA, TB, and P (final), if Q/nA = 3 kJ·mol−1

3.39. One mole of an ideal gas with constant heat capacities undergoes an arbitrary mechan-
ically reversible process. Show that:
1
​ΔU = ​___      ​Δ
 (​ ​PV​)​​
γ − 1

3.40. Derive an equation for the work of mechanically reversible, isothermal compression
of 1 mol of a gas from an initial pressure P1 to a final pressure P2 when the equation
of state is the virial expansion [Eq. (3.33)] truncated to:


Z = 1 + B′P​

How does the result compare with the corresponding equation for an ideal gas?

3.41. A certain gas is described by the equation of state:

( RT )
θ
​ PV = RT + ​​ ​b − ​_
     ​  ​P​

Here, b is a constant and θ is a function of T only. For this gas, determine expressions
for the isothermal compressibility κ and the thermal pressure coefficient ​(​​∂ P / ∂ T​)​ V​.​
These expressions should contain only T, P, θ, dθ/dT, and constants.
124 CHAPTER 3.  Volumetric Properties of Pure Fluids

3.42. For methyl chloride at 100°C the second and third virial coefficients are:


​B = −242.5 ​cm​​ 3​·mol​​ −1​  C = 25,200 ​cm​​ 6​·mol​​ −2​​

Calculate the work of mechanically reversible, isothermal compression of 1 mol of


methyl chloride from 1 bar to 55 bar at 100°C. Base calculations on the following
forms of the virial equation:
B C
(a) ​ Z = 1 + ​ __ ​   + ​ ___2  ​​ 
V ​V​  ​
(b) ​ Z = 1 + B′P + C′​P​ 2​
B C − ​B​ 2​
B′ = ​___
where ​     ​   and  C′= ​_____
   ​ 
​
RT ​(​RT​)​ 2​
Why don’t both equations give exactly the same result?

3.43. Any equation of state valid for gases in the zero-pressure limit implies a full set of
virial coefficients. Show that the second and third virial coefficients implied by the
generic cubic equation of state, Eq. (3.41), are:
a​(​ ​T)​ ​ (​ ​ε + σ​)​ba​​(​T​)​
​B = b − ​____
​    ​  ​  C = ​b​ 2​ + ​________
    ​​ 

RT RT
Specialize the result for B to the Redlich/Kwong equation of state, express it in reduced
form, and compare it numerically with the generalized correlation for B for simple flu-
ids, Eq. (3.61). Discuss what you find.

3.44. Calculate Z and V for ethylene at 25°C and 12 bar by the following equations:

(a) The truncated virial equation [Eq. (3.38)] with the following experimental values
of virial coefficients:

​B = −140 ​cm​​ 3​·mol​​ −1​  C = 7200 ​cm​​ 6​·mol​​ −2​​

(b) The truncated virial equation [Eq. (3.36)], with a value of B from the generalized
Pitzer correlation [Eqs. (3.58)–(3.62)]
(c) The Redlich/Kwong equation
(d) The Soave/Redlich/Kwong equation
(e) The Peng/Robinson equation

3.45. Calculate Z and V for ethane at 50°C and 15 bar by the following equations:

(a) The truncated virial equation [Eq. (3.38)] with the following experimental values
of virial coefficients:


​B = −156.7  ​cm​​ 3​·mol​​ −1​  C = 9650  ​cm​​ 6​·mol​​ −2​​

(b) The truncated virial equation [Eq. (3.36)], with a value of B from the generalized
Pitzer correlation [Eqs. (3.58)–(3.62)]
(c) The Redlich/Kwong equation
3.10. Problems
3.10 125

(d) The Soave/Redlich/Kwong equation


(e) The Peng/Robinson equation

3.46. Calculate Z and V for sulfur hexafluoride at 75°C and 15 bar by the following
equations:

(a) The truncated virial equation [Eq. (3.38)] with the following experimental values
of virial coefficients:


​B = −194 ​cm​​ 3​·mol​​ −1​  C = 15,300 ​cm​​ 6​·mol​​ −2​​

(b) The truncated virial equation [Eq. (3.36)], with a value of B from the generalized
Pitzer correlation [Eqs. (3.58)–(3.62)]
(c) The Redlich/Kwong equation
(d) The Soave/Redlich/Kwong equation
(e) The Peng/Robinson equation

For sulfur hexafluoride, Tc = 318.7 K, Pc = 37.6 bar, Vc = 198 cm3·mol−1, and


ω = 0.286.

3.47. Calculate Z and V for ammonia at 320 K and 15 bar by the following equations:

(a) The truncated virial equation [Eq. (3.38)] with the following values of virial
­coefficients:


​B = −208 ​cm​​ 3​·mol​​ −1​  C = 4378 ​cm​​ 6​·mol​​ −2​​

(b) The truncated virial equation [Eq. (3.36)], with a value of B from the generalized
Pitzer correlation [Eqs. (3.58)–(3.62)]
(c) The Redlich/Kwong equation
(d) The Soave/Redlich/Kwong equation
(e) The Peng/Robinson equation

3.48. Calculate Z and V for boron trichloride at 300 K and 1.5 bar by the following
equations:

(a) The truncated virial equation [Eq. (3.38)] with the following values of virial
coefficients:


​B = −724 ​cm​​ 3​·mol​​ −1​  C = −93,866 ​cm​​ 6​·mol​​ −2​​

(b) The truncated virial equation [Eq. (3.36)], with a value of B from the generalized
Pitzer correlation [Eqs. (3.58)–(3.62)]
(c) The Redlich/Kwong equation
(d) The Soave/Redlich/Kwong equation
(e) The Peng/Robinson equation

For BCl3, Tc = 452 K, Pc = 38.7 bar, and ω = 0.086.


126 CHAPTER 3.  Volumetric Properties of Pure Fluids

3.49. Calculate Z and V for nitrogen trifluoride at 300 K and 95 bar by the following
equations:

(a) The truncated virial equation [Eq. (3.38)] with the following values of virial
coefficients:


​B = −83.5 ​cm​​ 3​·mol​​ −1​  C = −5592 ​cm​​ 6​·mol​​ −2​​

(b) The truncated virial equation [Eq. (3.36)], with a value of B from the generalized
Pitzer correlation [Eqs. (3.58)–(3.62)]
(c) The Redlich/Kwong equation
(d) The Soave/Redlich/Kwong equation
(e) The Peng/Robinson equation

For NF3, Tc = 234 K, Pc = 44.6 bar, and ω = 0.126.

3.50. Determine Z and V for steam at 250°C and 1800 kPa by the following:

(a) The truncated virial equation [Eq. (3.38)] with the following experimental values
of virial coefficients:


​B = −152.5 ​cm​​ 3​·mol​​ −1​  C = −5800 ​cm​​ 6​·mol​​ −2​​

(b) The truncated virial equation [Eq. (3.36)], with a value of B from the generalized
Pitzer correlation [Eqs. (3.58)–(3.62)].
(c) The steam tables (App. E).

3.51. With respect to the virial expansions, Eqs. (3.33) and (3.34), show that:

( ∂ P ) ( ∂ ρ )
∂ Z ∂ Z
B′ = ​​ ___
​ ​    ​  ​​  ​​  and  B = ​​ ___
​    ​  ​​  ​​​
T,P=0 T,ρ=0
where ρ ≡ 1/V.

3.52. Equation (3.34) when truncated to four terms accurately represents the volumetric
data for methane gas at 0°C with:


​B = −53.4 ​cm​​ 3​·mol​​ −1​  C = 2620 ​cm​​ 6​·mol​​ −2​  D = 5000 ​cm​​ 9​·mol​​ −3​​

(a) Use these data to prepare a plot of Z vs. P for methane at 0°C from 0 to 200 bar.
(b) To what pressures do Eqs. (3.36) and (3.37) provide good approximations?

3.53. Calculate the molar volume of saturated liquid and the molar volume of saturated
vapor by the Redlich/Kwong equation for one of the following and compare results
with values found by suitable generalized correlations.

(a) Propane at 40°C where Psat = 13.71 bar


(b) Propane at 50°C where Psat = 17.16 bar
(c) Propane at 60°C where Psat = 21.22 bar
(d) Propane at 70°C where Psat = 25.94 bar
3.10. Problems
3.10 127

(e) n-Butane at 100°C where Psat = 15.41 bar


(f ) n-Butane at 110°C where Psat = 18.66 bar
(g) n-Butane at 120°C where Psat = 22.38 bar
(h) n-Butane at 130°C where Psat = 26.59 bar
(i) Isobutane at 90°C where Psat = 16.54 bar
(j) Isobutane at 100°C where Psat = 20.03 bar
(k) Isobutane at 110°C where Psat = 24.01 bar
(l) Isobutane at 120°C where Psat = 28.53 bar
(m) Chlorine at 60°C where Psat = 18.21 bar
(n) Chlorine at 70°C where Psat = 22.49 bar
(o) Chlorine at 80°C where Psat = 27.43 bar
(p) Chlorine at 90°C where Psat = 33.08 bar
(q) Sulfur dioxide at 80°C where Psat = 18.66 bar
(r) Sulfur dioxide at 90°C where Psat = 23.31 bar
(s) Sulfur dioxide at 100°C where Psat = 28.74 bar
(t) Sulfur dioxide at 110°C where Psat = 35.01 bar
(u) Boron trichloride at 400 K where Psat = 17.19 bar
For BCl3, Tc = 452 K, Pc = 38.7 bar, and ω = 0.086.
(v) Boron trichloride at 420 K where Psat = 23.97 bar
(w) Boron trichloride at 440 K where Psat = 32.64 bar
(x) Trimethylgallium at 430 K where Psat = 13.09 bar
For Ga(CH3)3, Tc = 510 K, Pc = 40.4 bar, and ω = 0.205.
(y) Trimethylgallium at 450 K where Psat = 18.27 bar
(z) Trimethylgallium at 470 K where Psat = 24.55 bar

3.54. Use the Soave/Redlich/Kwong equation to calculate the molar volumes of saturated
liquid and saturated vapor for the substance and conditions given by one of the
parts of Prob. 3.53 and compare results with values found by suitable generalized
correlations.

3.55. Use the Peng/Robinson equation to calculate the molar volumes of saturated liquid
and saturated vapor for the substance and conditions given by one of the parts of
Prob. 3.53 and compare results with values found by suitable generalized correlations.

3.56. Estimate the following:

(a) The volume occupied by 18 kg of ethylene at 55°C and 35 bar.


(b) The mass of ethylene contained in a 0.25 m3 cylinder at 50°C and 115 bar.

3.57. The vapor-phase molar volume of a particular compound is reported as 23,000


cm3·mol–1 at 300 K and 1 bar. No other data are available. Without assuming ideal-gas
behavior, determine a reasonable estimate of the molar volume of the vapor at 300 K
and 5 bar.

3.58. To a good approximation, what is the molar volume of ethanol vapor at 480°C and
6000 kPa? How does this result compare with the ideal-gas value?
128 CHAPTER 3.  Volumetric Properties of Pure Fluids

3.59. A 0.35 m3 vessel is used to store liquid propane at its vapor pressure. Safety considerations
dictate that at a temperature of 320 K the liquid must occupy no more than 80% of the
total volume of the vessel. For these conditions, determine the mass of vapor and the
mass of liquid in the vessel. At 320 K the vapor pressure of propane is 16.0 bar.

3.60. A 30 m3 tank contains 14 m3 of liquid n-butane in equilibrium with its vapor at 25°C.
Estimate the mass of n-butane vapor in the tank. The vapor pressure of n-butane at the
given temperature is 2.43 bar.

3.61. Estimate:

(a) The mass of ethane contained in a 0.15 m3 vessel at 60°C and 14,000 kPa.
(b) The temperature at which 40 kg of ethane stored in a 0.15 m3 vessel exerts a pres-
sure 20,000 kPa.

3.62. A size D compressed gas cylinder has an internal volume of 2.40 liters. Estimate the
pressure in a size D cylinder if it contains 454 g of one of the following semiconductor
process gases at 20°C:

(a) Phosphine, PH3, for which Tc = 324.8 K, Pc = 65.4 bar, and ω = 0.045
(b) Boron trifluoride, BF3, for which Tc = 260.9 K, Pc = 49.9 bar, and ω = 0.434
(c) Silane, SiH4, for which Tc = 269.7 K, Pc = 48.4 bar, and ω = 0.094
(d) Germane, GeH4, for which Tc = 312.2 K, Pc = 49.5 bar, and ω = 0.151
(e) Arsine, AsH3, for which Tc = 373 K, Pc = 65.5 bar, and ω = 0.011
(f) Nitrogen trifluoride, NF3, for which Tc = 234 K, Pc = 44.6 bar, and ω = 0.120

3.63. For one of the substances in Prob. 3.62, estimate the mass of the substance contained
in the size D cylinder at 20°C and 25 bar.

3.64. Recreational scuba diving using air is limited to depths of 40 m. Technical divers use
different gas mixes at different depths, allowing them to go much deeper. Assuming a
lung volume of 6 liters, estimate the mass of air in the lungs of:

(a) A person at atmospheric conditions.


(b) A recreational diver breathing air at a depth of 40 m below the ocean surface.
(c) A near-world-record technical diver at a depth of 300 m below the ocean surface,
breathing 10 mol% oxygen, 20 mol% nitrogen, 70 mol% helium.

3.65. To what pressure does one fill a 0.15 m3 vessel at 25°C in order to store 40 kg of
ethylene in it?

3.66. If 15 kg of H2O in a 0.4 m3 container is heated to 400°C, what pressure is developed?

3.67. A 0.35 m3 vessel holds ethane vapor at 25°C and 2200 kPa. If it is heated to 220°C,
what pressure is developed?

3.68. What is the pressure in a 0.5 m3 vessel when it is charged with 10 kg of carbon dioxide
at 30°C?
3.10. Problems
3.10 129

3.69. A rigid vessel, filled to one-half its volume with liquid nitrogen at its normal boiling
point, is allowed to warm to 25°C. What pressure is developed? The molar volume of
liquid nitrogen at its normal boiling point is 34.7 cm3·mol–1.

3.70. The specific volume of isobutane liquid at 300 K and 4 bar is 1.824 cm3·g–1. Estimate
the specific volume at 415 K and 75 bar.

3.71. The density of liquid n-pentane is 0.630 g·cm–3 at 18°C and 1 bar. Estimate its density
at 140°C and 120 bar.

3.72. Estimate the density of liquid ethanol at 180°C and 200 bar.

3.73. Estimate the volume change of vaporization for ammonia at 20°C. At this temperature
the vapor pressure of ammonia is 857 kPa.

3.74. PVT data may be taken by the following procedure: A mass m of a substance of molar
mass ℳ is introduced into a thermostated vessel of known total volume Vt. The sys-
tem is allowed to equilibrate, and the temperature T and pressure P are measured.

(a) Approximately what percentage errors are allowable in the measured variables (m,
ℳ, Vt, T, and P) if the maximum allowable error in the calculated compressibility
factor Z is ±1%?
(b) Approximately what percentage errors are allowable in the measured variables if
the maximum allowable error in calculated values of the second virial coefficient
B is ±1%? Assume that Z ≃ 0.9 and that values of B are calculated by Eq. (3.37).

3.75. For a gas described by the Redlich/Kwong equation and for a temperature greater than
Tc, develop expressions for the two limiting slopes,

​​​ lim​ ​​​(___
P→0 ∂ P ) P→∞( ∂ P )
∂ Z ∂ Z
​    ​  ​​  ​​​    ​ lim​ ​​​ ___
​    ​  ​​  ​​​​
T T
Note that in the limit as P → 0, V → ∞, and that in the limit as P → ∞, V → b.

3.76. If 140(ft)3 of methane gas at 60(°F) and 1(atm) is equivalent to 1(gal) of gasoline as
fuel for an automobile engine, what would be the volume of the tank required to hold
methane at 3000(psia) and 60(°F) in an amount equivalent to 10(gal) of gasoline?

3.77. Determine a good estimate for the compressibility factor Z of saturated hydrogen
vapor at 25 K and 3.213 bar. For comparison, an experimental value is Z = 0.7757.

3.78. The Boyle temperature is the temperature for which:

P→0( ∂ P )
∂ Z
​ lim​ ​​​ ___
​ ​    ​  ​​  ​​  = 0​
T
(a) Show that the second virial coefficient B is zero at the Boyle temperature.
(b) Use the generalized correlation for B, Eqs. (3.58)–(3.62), to estimate the reduced
Boyle temperature for simple fluids.
130 CHAPTER 3.  Volumetric Properties of Pure Fluids

3.79. Natural gas (assume pure methane) is delivered to a city via pipeline at a volumetric
rate of 150 million standard cubic feet per day. Average delivery conditions are 50(°F)
and 300(psia). Determine:

(a) The volumetric delivery rate in actual cubic feet per day.
(b) The molar delivery rate in kmol per hour.
(c) The gas velocity at delivery conditions in m·s–1.

The pipe is 24(in) schedule-40 steel with an inside diameter of 22.624(in). Standard
conditions are 60(°F) and 1(atm).

3.80. Some corresponding-states correlations use the critical compressibility factor Zc,
rather than the acentric factor ω, as a third parameter. The two types of correlation
(one based on Tc, Pc, and Zc, the other on Tc, Pc, and ω) would be equivalent were
there a one-to-one correspondence between Zc and ω. The data of App. B allow a test
of this correspondence. Prepare a plot of Zc vs. ω to see how well Zc correlates with ω.
Develop a linear correlation (Zc = a + bω) for nonpolar substances.

3.81. Figure 3.3 suggests that the isochores (paths of constant volume) are approximately
straight lines on a P-T diagram. Show that the following models imply linear isochores.

(a) Constant-β, κ equation for liquids


(b) Ideal-gas equation
(c) Van der Waals equation

3.82. An ideal gas, initially at 25°C and 1 bar, undergoes the following cyclic processes in a
closed system:

(a) In mechanically reversible processes, it is first compressed adiabatically to 5 bar,


then cooled at a constant pressure of 5 bar to 25°C, and finally expanded isother-
mally to its original pressure.
(b) The cycle is irreversible, and each step has an efficiency of 80% compared with
the corresponding mechanically reversible process. The cycle still consists of an
adiabatic compression step, an isobaric cooling step, and an isothermal expansion.

Calculate Q, W, ΔU, and ΔH for each step of the process and for the cycle. Take CP =
(7/2)R and CV = (5/2)R.

3.83. Show that the density-series second virial coefficients can be derived from isothermal
volumetric data via the expression:
lim​ ​(​Z − 1​)​ / ρ​    ρ(molar density)≡ 1 / V​​
​  ρ→0
B = ​

3.84. Use the equation of the preceding problem and data from Table E.2 of App. E to
obtain a value of B for water at one of the following temperatures:

(a) 300°C
(b) 350°C
(c) 400°C
3.10. Problems
3.10 131

3.85. Derive the values of Ω, Ψ, and Zc given in Table 3.1 for:

(a) The Redlich/Kwong equation of state.


(b) The Soave/Redlich/Kwong equation of state.
(c) The Peng/Robinson equation of state.

3.86. Suppose Z vs. Pr data are available at constant Tr . Show that the reduced density-series
second virial coefficient can be derived from such data via the expression:
ˆ​  = ​ lim​ ​(​Z − 1​)​Z​Tr​ ​ / ​Pr​ ​​
​​B 
​Pr​ ​→0
Suggestion: Base the development on the full virial expansion in density, Eq. (3.34)

3.87. Use the result of the preceding problem and data from Table D.1 of App. D to obtain a value
of ​​Bˆ​​  for simple fluids at Tr = 1. Compare the result with the value implied by Eq. (3.61).

3.88. The following conversation was overheard in the corridors of a large engineering firm.
New engineer: “Hi, boss. Why the big smile?”
Old-timer: “I finally won a wager with Harry Carey, from Research. He bet me
that I couldn’t come up with a quick but accurate estimate for the molar volume of
argon at 30°C and 300 bar. Nothing to it; I used the ideal-gas equation, and got about
83 cm3·mol−1. Harry shook his head, but paid up. What do you think about that?”
New engineer (consulting his thermo text): “I think you must be living right.”
Argon at the stated conditions is not an ideal gas. Demonstrate numerically why the
old-timer won his wager.

3.89. Five mol of calcium carbide are combined with 10 mol of water in a closed, rigid,
high-pressure vessel of 1800 cm3 internal empty volume. Acetylene gas is produced
by the reaction:

​CaC​ 2​(​s​)​ + 2​H​ 2​O​(​l​)​ → ​C​ 2​H​ 2​(​g​)​ + Ca​​(​OH​)​ 2​(​s​)​​

The vessel contains packing with a porosity of 40% to prevent explosive decomposi-
tion of the acetylene. Initial conditions are 25°C and 1 bar, and the reaction goes to
completion. The reaction is exothermic, but owing to heat transfer, the final tempera-
ture is only 125°C. Determine the final pressure in the vessel.
Note: At 125°C, the molar volume of Ca(OH)2 is 33.0 cm3·mol−1. Ignore the effects of
any gases (e.g., air) initially present in the vessel.

3.90. Storage is required for 35,000 kg of propane, received as a gas at 10°C and 1(atm).
Two proposals have been made:

(a) Store it as a gas at 10°C and 1(atm).


(b) Store it as a liquid in equilibrium with its vapor at 10°C and 6.294(atm). For this
mode of storage, 90% of the tank volume is occupied by liquid.

Compare the two proposals, discussing pros and cons of each. Be quantitative where possible.
132 CHAPTER 3.  Volumetric Properties of Pure Fluids

3.91. The definition of compressibility factor Z, Eq. (3.32), may be written in the more intu-
itive form:
V
​ Z ≡ ​__________
     ​​ 
V​​(​ideal gas​)​

where both volumes are at the same T and P. Recall that an ideal gas is a model sub-
stance comprising particles with no intermolecular forces. Use the intuitive definition
of Z to argue that:

(a) Intermolecular attractions promote values of Z < 1.


(b) Intermolecular repulsions promote values of Z > 1.
(c) A balance of attractions and repulsions implies that Z = 1. (Note that an ideal gas
is a special case for which there are no attractions or repulsions.)

3.92. Write the general form of an equation of state as:

Z = 1 + ​Zr​ ep​(​ρ​)​ − ​Za​  ttr​(​T, ρ​)​​

where Zrep(ρ) represents contributions from repulsions, and Zattr(T, ρ) represents con-
tributions from attractions. What are the repulsive and attractive contributions to the
van der Waals equation of state?

3.93. Given below are four proposed modifications of the van der Waals equation of state.
Are any of these modifications reasonable? Explain carefully; statements such as, “It
isn’t cubic in volume” do not qualify.
RT a
P = ​____
(a) ​     ​   − ​__
   ​ 
V − b V
RT a
P = ​______
(b) ​    2 ​  − ​__   ​ 
​(​V − b​)​  ​ V
RT a
P = ​______
(c) ​      ​  − ​___  2  ​ 
V​​(​V − b​)​ ​V​  ​
RT a
P = ​___
(d) ​    ​  − ​___
    ​ 
V ​V​ 2​
3.94. With reference to Prob. 2.47, assume air to be an ideal gas, and develop an expression
giving the household air temperature as a function of time.

3.95. A garden hose with the water valve shut and the nozzle closed sits in the sun, full of
liquid water. Initially, the water is at 10°C and 6 bar. After some time the temperature
of the water rises to 40°C. Owing to the increase in temperature and pressure and the
elasticity of the hose, the internal diameter of the hose increases by 0.35%. Estimate
the final pressure of the water in the hose.
Data: β(ave) = 250 × 10−6 K−1; κ(ave) = 45 × 10−6 bar−1

You might also like