Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 31

FME 522 THERMODYNAMICS V

COURSE OUTLINE
1. Thermodynamics of State
- The perfect gas equation and use of gas table at low pressures
- Other equations of state
- The principle of corresponding states and use of compressibility charts
- Algebraic Equations of state for incompressible liquids
- Differential equations of state
 Coefficients of thermal and volume expansion.
 Specific heats.
- Maxwell equations
- Joule-Thomson’s Coefficient
- Clausius-Claypeyron equation.
2. Non-reacting Mixtures and Psychometry
- Gaseous mixture and gas laws
- Perfect gas mixtures
- Ideal gas and Vapour mixtures
- Fundamental parameters of water/air mixtures
- Adiabatic saturator and psychrometric chart
- Moist air processes
- Air-Conditioning processes
- Cooling towers.
3. Fuels and Combustion
- Fundamentals of reaction mixtures and definitions
- Fuel and product analysis and chemical balance equations
- Consistent enthalpy and energy scales and use of thermochemical Tables
- Heat of reaction
- Adiabatic flame temperature
- Chemical equilibrium and dissociation
- Fuels for internal combustion engines.
4. Internal Combustion Engines
- S.I. and C.I. engine
- Performance characteristics
- Combustion in I.C.Es.
References
1. Eastop T.D. and McConkey A. (1993) Applied Thermodynamics for Engineering
Technologists, Prentice and Hall, 4th Ed.
2. Rogers G.F.C. & Mayhew Y.R. (1992) Engineering Thermodynamics, Longman
Singapore Publishers, 4th Ed.
3. Kyle, B.G. (2004) Chemical and Process Thermodynamics. Prentice Hall.
4. Perry’s Chemical Engineers Handbook, 7th Edition. McGraw Hill.

Page 1 of 31
I. THERMODYNAMICS OF STATE
Objectives:
(a) Establish how to solve thermodynamic problems for real substances given heat
capacity data and PVT data (Equation of state).
(b) How to construct thermodynamic property chart from experimental given heat
capacity data and PVT data.

1.1 Properties of real substances and mixtures – a review


Consider a thermodynamic system. The system can be defined by both intensive (mass
independent) and extensive (mass dependent) properties.

m
p, T, v
h, u
W
s
Q x

Thermodynamic system

p, v, T data is from the particulate nature of matter (fundamental physics)


u, h as properties came as a consequence of the 1st law of thermodynamics (conservation of
energy)
Q−W x c
2
i.e. =∆ E ,(∆ E=∆ u+∆ h+ ∆ +∆ gz)
ṁ 2
Entropy, s a state property came as a consequence of the 2 nd law of thermodynamics. It is a
directionality property of a process or a measure of irreversibility.
d QR
i.e. ∆ s=
T
x is a phase property and is always either a mass fraction or volume fraction of a given phase.
All the extensive properties can be reduced to intensive properties by dividing mass (i.e.
intensified, and a preceded by specific, e.g. specific volume etc.
Any thermodynamic system in which mass, heat and work flow across its boundaries, the
thermodynamic state (as described by this properties) may change owing to the flows. Hence,
we require balance equations to relate mass, heat and work flows with change of state. These
equations are generally summarized as the first and second laws of thermodynamics.
i.e.
Q−W x
=∆ E ; (1st Law) 1.1

Page 2 of 31

∮ dW ≤ 0 ; (2nd law) 1.2


single Q source

We then require definitive equations to relate the state properties,

( ∂T∂ h )
e.g. dh=c p dT ∨c p=
p

du=c dT ∨c =(
∂T )
∂u
v v ; (heat capacity coefficients) 1.3
v

pv=RT ; (equation of state) 1.4

( ∂∂ VP ) =0 ; (incompressible fluid)
T
1.5

For a single component, single phase system, the system can be described completely by
specifying the initial and final values of any two independent intensive variables (properties).
Certain intensive properties e.g. P and T are easier to measure than others. Therefore, for
most problems we seek to specify the state of the system by its temperature and pressure,
rather than its specific volume, internal energy, entropy etc. We the therefore look for
interactions between fluid properties that allow one to eliminate some thermodynamic
variables in terms of easier to measure ones.
For example, noting that h and u are functions of T only, and not p and v, the values of H and
U found in property tables are stated from a reference state for which H and U are arbitrarily
set at zero. The choice of property table or chart can be guided by properties under
consideration as follows: h-s diagram (Mollier chart) for turbines
h-p diagram for refrigeration
T-s diagram for engine cycles, etc.

For multi-component and multi-phase systems, the sixth balance equation becomes
I II
x A + x A =1
I I
x A + x B =1 1.6
Where A-B are components, and I-II are phases.
The energy equation (1.1) is valid regardless of which phase is present, hence there should be
no difficulty in its use for multi-phase or phase change systems. If θ^ is any intensive property
such as u, h etc., then the value of the property for a multi-phase system is given by
n
^ 1I θ^ I + x II2 θ^ II +…=∑ x i θ^ i
θ=x 1.7
i=1

Where xi is the mole or volume fraction of a given phase.

Page 3 of 31
Other common property expressions using θ^ and partial values, e.g. ( )
∂ θ^
∂ xi x
II
=θ^ are as given in
i

Eqn. 1.3. It should be noted that most of the derived intensive properties are actually
functions of the primary properties.
The state variables can be combined to form compound state variables as in the case of
properties that define equilibrium. These properties are:
(a) Helmholtz free energy function, A
A=U−Ts

(b) Gibbs free energy function, G


G=H −TS
Hence the complete property table is as shown in Table 1.1 below.

Table 1.1 System property table


Properties Consequence of …
T
P Particulate nature of matter
V
U Energy conservation (1st law of thermodynamics)
H
S Irreversibility of processes (2nd law of
thermodynamics
A Equilibrium functions
G

Any two of the above variables for a pure single-phase system completely defines the system.

I.2 Equations of State


(a) The PVT behaviour of fluids
Many applications of thermodynamics are concerned with the behaviour of fluids, especially
gases. To evaluate the thermodynamic property changes, the data needed is provided by use
the first and second laws of thermodynamics, which permit us to calculate heat and work
effects associated with the various processes involving fluids. Important to us the PVT data,
as well as the thermal data such as heat capacities.
A pure fluid is in an equilibrium state when two intensive properties are fixed and all
properties are uniform throughout (this also applies to non-reacting mixtures as long as the
composition is specified). To completely characterize the behaviour of a fluid, three variables
are required, i.e. a three-dimensional PVT plot. However, the PVT behaviour of a fluid is

Page 4 of 31
usually displayed on a P-V diagram on which constant temperature curves are drawn. Figure
1.1 shows a general behaviour of a pure fluid in 2-D plot.

Fig. 1.1: P-v diagram for a pure substance


At the critical point, the liquid and vapour become indistinguishable, and the length of the
horizontal segment VG-VL approach zero. Sometimes the PVT data are plotted as isochores
(lines of constant density) on a P-T diagram. Figure 1.2 shows the general behaviour of a
pure fluid on a P-T plot. Isochores are shown as dashed lines emanating from the vapour-
pressure curve, AC.

Fig. 1.2: P-T diagram for a pure substance


Heat

Let’s consider a sealed tube with a liquid-vapour mixture of a pure substance, represented by
a point A (T, P). If the temperature is raised, the point representing the state of the system
moves along the vapour-pressure line (AC) until one of the following three things happens.
(i) If the overall density is greater than that of critical point, the meniscus separating
the phases will rise in the vertically held tube until a point B, where the tube
contains all the liquid. On further heating, the system traces a path along liquid
isochore BE.
(ii) If the overall density is less than that of critical density, the meniscus falls on
heating until point D is reached, where the tube contains all vapour. Further
heating moves the system along a vapour isochore DG.

Page 5 of 31
(iii) If the tube is charged with exactly the critical density, the meniscus remains near
the centre of the tube until the critical point C is reached, where it disappears. On
further heating, the system follows a critical isochore, CF. The isochore has the
same slope as the vapour-pressure curve at the critical point.
The sealed-tube, disappearing meniscus technique has widely been used to determine the
critical temperature of substances. The fluid phase lying beyond the critical points (i.e. T>T C,
P> PC) is known as the supercritical fluid. The supercritical fluid has a density close to that of
a normal liquid but possesses a much lower viscosity and a higher diffusivity. This facilitates
mass transport, and thus supercritical fluids find application in use as extraction and leaching
solvents. Mostly, CO2 and water are used as supercritical fluids due to their availability and
low critical temperatures.
The experimental determination of PVT data is obtained by either:
(i) Measurement of pressure as a function of volume at a series of fixed temperatures,
or
(ii) Measurement of pressure as a function of temperature at a series of fixed volumes.
Though easily to conceptualise, to obtain data of sufficient accuracy for thermodynamic
calculations, a fairly complex apparatus an painstaking techniques are used (e.g. Beattie’s
PVT apparatus).

(b) Equations of State


An equation of state is an algebraic expression relating state properties. It is usually obtained
by fitting PVT data. Equations of state offer the advantages of data reduction and ease of use
in subsequent calculations. Many equations of state have been proposed, with varying
complexity and applicability. Most commonly used Equations of state include:

(i) The Ideal Gas Law


The ideal gas obeys the equation of state;
PV =nRT 1.8
Where R = 8314.3 N.m/ kmol.K. For an ideal gas, both enthalpy and internal energy depends
on temperature only. This law is limited to gases at low pressure. Note that all gases approach
the ideal gas behaviour as pressure is decreased and temperature increased.

(ii) The Van der Walls Equation


This equation of state is expressed as;

( P+
a
V )
2
( V −b )=RT 1.9

RT a
Or P= − 2 1.10
V −b V

Page 6 of 31
The constants a and b are a measure of how a fluid deviates from ideal gas assumptions (i.e.
particles are perfectly elastic and rigid, volume occupied by particles is small compared to the
total volume, attractive forces between particles are negligible). It is applicable for both
a
vapour and liquids. The term 2 accounts for attractive forces between molecules, while −b
V
is a correction for the volume occupied by the molecules. Table 1.2 below gives values for a
and b for some common gases.

Table 1.2
Gas a (m6/mol.2) b (m3/mol.) × 103
O2 0.1381 3.184
N2 0.1368 3.864
H2O 0.5542 3.051
CH4 0.2303 4.306
CO 0.1473 3.951
CO2 0.3658 4.286
NH3 0.4253 3.737
H2 0.0248 2.660
He 0.00346 2.376

(iii) Modified Van der Walls Equations


They include approximations of the van der walls equation for the fluid behaviour in two-
phase region. The mostly commonly encountered are;

(a) Redlich-Kwong Equation


RT a
P= − 1 /2
V −b T V ( V −b )
1.11
(b) Soave Equation
RT a (T )
P= − 1.12
V −b V ( V + b )

(c) Peng-Robinson Equation


RT a (T )
P= − 2
V −b V + 2Vb−b2
1.13
Equations 1.11 – 1.13 are capable of representing vapour and liquid phases reasonably well.

Page 7 of 31
(iv) The Virial Equation
The Virial equation expresses the compressibility factor Z (= PV/RT) as a power series in
density or reciprocal of volume and can be derived from statistical mechanics.
Pv B C
i.e. Z= =1+ + 2 +… 1.14
RT v v
Where B and C are called the 2 nd and 3rd virial coefficients and are functions of temperature
only.

(v) Benedict-Webb-Rubin Equation


It is a form of the Virial Equation, and as with the Van der walls equation, it has various
modifications. It is capable of closely representing both liquid and vapour phases. The
original 8-parameter equation is,

( )
3
C0 Cρ (
1+γ ρ ) e
2
2 3 6 2 −γ P
P=RTρ + β 0 RT −A 0− 2
ρ + ( bRT−a ) ρ +aα ρ + 2 1.15
T T

(c) The Compressibility Factor


The compressibility factor, Z is defined as PV/RT, and is seen as a dimensionless or reduced
quantity. Many reduced properties have been successfully correlated by the principle of
corresponding states, which states that “systems which are in the same reduced state exhibit
the same reduced properties”.
The reduced temperature, Tr and reduced pressure, Pr are defined as,
T
T r= 1.16
Tc
P
Pr = 1.17
Pc
Where Tc and Pc are critical temperature and pressure, respectively.
For a specified Tr and Pr, every gas will have the same compressibility factor, and thus a
correlation of Z in terms of T r and Pr would apply to all gases. Such correlations give very
good approximations for substances with similar chemical structure, and the situation
improves considerably when a parameter characterizing the chemical nature of substances is
incorporated into the correlation. Two characterization parameters are widely used:
(a) The eccentric factor, ω
(b) The critical compressibility factor, Zc.
Pc V c
Note that Z c = . The critical constants and characterization parameters are shown in
RT c
Appendix 1.1

Page 8 of 31
Note:
The principle of corresponding states does not specify the functionability of the relationship
between the reduced property and reduced temperature and pressure, but it simply states that
a relationship should exist. The actual relationship must be determined from experimental
observations of many compounds.
The equations PV = ZRT and PV = nRT, which represent the functionability Z = Z(T r, Pr, Zc)
constitute a generalized equation of state in which the functionability can be stated as;
f ( P , v , T ; T c , Pc , Z c )=0 1.18
Once the three characterizing parameters T c , Pc , Z c are known, the PVT data behaviour of
any substance can be estimated.
The Peng-Robinson Equation (1.13) is commonly used to fit experimental PVT data, in
which the constants are evaluated as follows.
2 2
R Tc
a ( T c )=0.45724 , and 1.19
Pc
RTc
b=0.07780 1.20
Pc
It can be observed that b is a constant, and a is a function of temperature T and eccentric
factor, ω according to the relation,
a ( T )=a ( T c ) . α ( T , ω ) 1.21

The form of the function α has been determined by fitting vapour pressure data for many non-
polar substances to be;

√ α =1+k 1−
T
[ √ ] Tc
1.22

Where k =0.37464+ 1.54226 ω−0.26992 ω2 1.23


Which yields α =1 for T =T c
The generalised Peng-Robinson equation can be stated functionally as
f ( P , v , T ; T c , Pc , ω ) =0 1.24
It is equivalent to the three-parameter, corresponding states correlation of compressibility
factor and provides estimates of comparable accuracy. The generalized equation of state has
the advantage of being in algebraic form rather than in graphical form and therefore more
useful in computerized calculations.
For computational convenience, the Peng-Robinson equation can be transformed to
Z3 + ( B−1 ) Z 2+ ( A−3 B2−2 B ) Z + ( B3 +B 2− AB )=0 1.25

Page 9 of 31
a (T ) P Pb
Where A= 2 , B=
( RT ) RT
And the parameter A and B can be generalized to
Pr
[ 1+ k (1−T )]
2
1 /2
A=0.45724 2 r
T r

Pr
B=0.07780
Tr
Note:
For cubic equations, three roots are possible. For the modified Van der walls equations, there
are three real roots to T < Tc, of which the middle one is unstable. The other two represent the
Z-values for liquid and vapour conditions.
i.e Z bp< Z dp

Example 1.1
Find the density of chlorine gas, CL2 at a pressure of 155 bar and temperature of 521 K using,
(a) Compressibility factor chart (401 kg/m3)
(b) Generalized Peng-Robinson Equation of state (394 kg/m3)

The agreement between the two methods is seen to be quite good.

I.3 Heat Capacity Data


Besides the equation of state (PVT) relationship, we also require data for c p and cv to compute
changes in thermodynamic properties.
We recall that the first law of thermodynamics,
Q−W =∆ E (realised through changes in properties)
For real fluids, c p=c p ( P , T )
They obey an equation of state (eos)
c v =c v ( P , T )

Ideal fluid, c p=f ( T )


They obey PV= RT
c v =f ( T )
Perfect fluid c p ≠ f (T )
They obey PV= RT
cv ≠ f ( T )

Two important relations for cp and cv are,

Page 10 of 31
∂ cp
| |
2
∂V
=−T 2 1.26
∂P T ∂T P

∂ cv
| |
2
∂ P
=T 2 1.27
∂V T ∂T V

Applying Eqns. 1.26 and 1.27, let’s consider the situation in which we have data for c p as
a function of temperature at pressure P 1, and we want cp as a function of temperature at a
pressure P2. From Eqn. 1.26, we have;
∂ cp
| | |
2 2
∂V ∂ V
=−T 2 or d c p=−T 2
dP
∂P T ∂T P ∂T P
Integrating between P1 and P2, at a constant temperature T, we have
P 2 ,T

∫ d c p =c p ( P 2 , T )−c p ( P1 ,T )
P1 ,T

P2 ,T

Or c p ( P2 ,T )=c p ( P1 ,T ) + ∫ d c p
P1 ,T

|
P2 ,T
∂2 V
¿ c p ( P1 , T )−T ∫ ∂T P
2
dP 1.28
P1 ,T

Similarly,

|
V 2 ,T
∂2 P
c v ( V 2 ,T )=c v ( V 1 , T )+ T ∫ ∂T V
2
dV 1.29
V 1 ,T

Note that the last term in Equations 1.28 and 1.29 comes for an EOS. In practice, heat
capacity data are tabulated for states at very low pressure or very large specific volumes
where fluids approach ideal gases.
¿ ¿
i.e. c p ( P→ 0 , T )=c p (T ), and c v ( V → ∞ , T ) =c v (T )

Thus, equations 1.28 and 1.29 become,

|
P2 ,T
∂2V
¿
c p ( P2 ,T )=c ( T )−T
p ∫ 2
∂T P
dP 1.30
P1 ,T

|
V2, T
∂2 P
¿
c v ( V 2 ,T )=c ( T ) +T
v ∫ ∂T V
2
dV 1.31
V1, T

¿ ¿
The values of c p ( T ) and c v ( T ) are available in Tables such as Perry’s. The information is
presented in the form;

Page 11 of 31
¿ 2
c p ( T ) =a0 + a1 T +a2 T + … 1.32
¿ 2
c v ( T )=b0 +b 1 T +b2 T + … 1.33

I.4 Evaluation of Thermodynamic Properties


1.4.1 Calculation of ΔH, ΔU, ΔS as functions of P, T
(a) ΔH
Our interest is to compute h ( T 1 , P1 ) to h ( T 2 , P2 ). We know that,
2

∫ dh=h ( T 2 , P2 ) −h ( T 1 , P1 )
1

Since h is a state property, it is path independent, therefore;

Constant pressure Const. temperature


h ( T 1 , P 1 ) h ( T 2 , P 1 ) h ( T 2 , P 2)
I II

T 2 , P1

Path I: ¿ ∫ c p dT
T 1 , P1

( | ) dP
T 2 , P2
∂V
Path II: ¿ ∫ V −T
∂T
T 2 , P1 P

|
P
∂2V
Also, c p=c¿p −T ∫ ∂T2 P
dP
P=0

[ | ]
T2 T2 P 2
Thus, c p dT =∫ c dT −∫
¿
p ∫ T ∂ V2 dP dT
T1 T1 P=0 ∂T P

[ | ] | )dP
T2 T2 T 2 , P2

(
P 2
∂V ∂V
∴ ∆ h=∫ c ¿p dT−∫ T ∫ 2
dP dT + ∫ V −T 1.34
T1 T1 0 ∂T P T ,P ∂T 2 1
P

From Tables From EOS


at std state

(b) ΔS
For change in entropy S (T, P), we have;

ds=
∂S
∂T P |
dT +
∂S
∂P T
dP |
¿
cp
T
dT −
∂V
∂T | dPP

Page 12 of 31
|
p 2
∂ V
Introducingc p=c ( T )−T
¿
p ∫ 2
∂T P
dP , we have;
p =0

| | dP
T 2 , P0 P0 , T 1 P2 ,T 2
c ¿p ( T ) ∂V ∂V
∆ S= ∫ T
dT − ∫
∂T
dP− ∫ ∂T
1.35
T 1 , P0 P ,T 1 1
P P0 ,T 2 P
II I III

I II III
Path: P1 ,T 1 P0 , T 1 P0 , T 2 P2 ,T 2
P0 T 2 , P0 P2 ,T 2
∂V ∂V
∫ ∂ T dP ∫ c ¿p dT ∫ ∂T
dP
P1 T 1 , P0 P 0 ,T 0

(c) ΔU
Similarly,

( | ) ( | )
T1 , V ∞ T2 T 21 ,V 2
∂P ∂P
∆U= ∫ T
∂T V
−P dV +∫ c v dT + ∫ T
¿
∂T V
−P dV 1.36
T 1 ,V 1 T T ,V 1 2 ∞

I II III
Path: V 1 ,T 1 V ∞ , T1 V ∞ , T2 V 2 ,T 2

¿ ¿
Note: we need heat capacity data c p and c v at standard ideal gas state (P = 0, v = ∞) as an
equation of state to completely compute Δh, Δu, Δs, ΔG, and ΔA.
Recall that, ∆ G=∆ h−T ∆ s
∆ A=∆ u−T ∆ s
It follows that;

( |)
T,P
∂V
h−h =
IG
∫ V −T
∂T
dP 1.37
T , P=0 P

∫ ( | )
T ,P
IG ∂V R
s−s = − dP 1.38
T , P =0 ∂T P P
Equations 1.37 and 1.38 are known as departure functions. These functions are used to
predict how a real fluid departs from an ideal fluid at any T, P.

Page 13 of 31
Thus,

Δh, Δu of Δh, Δu of an Departure from ideal gas Departure of real gas from
= + -
a real fluid ideal fluid to real gas at end state ideal gas at initial sate
state

i.e. h1 (T1, P1) h2 (T2, P2)

departure departure

h(T1, P0) h(T2, P0)

The departure functions are computed from real gas equations of state. Using the departure
functions, thermal data equations and equations of state, fluid property tables can be
constructed.

1.4.2 Property Estimation from Corresponding States


For instance, whereby, the required PVT data required for the calculation of enthalpy and
entropy changes is unavailable, or accurate values of these property changes are not needed,
approximate values can be obtained using corresponding states. We define residue property
(Δh*, Δs*) as;
¿ '
∆ h =h −h
¿ '
∆ s =s −s
Where h, s are properties at (T, P) and h' , s ' are properties at (T, P) that would be exhibited in
ideal gas. As property changes are path-independent, a similar approach as above can be
used, i.e.

I II III IV
P1 ,T 1 P1 ,T 1 P1 ,T 2 P2 ,T 2 P2 ,T 2

(real gas) (ideal gas) (ideal gas) (ideal gas) (real gas)

Process I: Real gas to ideal gas at constant P1, T1.


' ¿
∆ hI =hP 1
,T 1 −hP , T =∆ h P ,T
1 1 1 1

∆ s I =s 'P ,T −s P ,T =∆ s¿P ,T
1 1 1 1 1 1

Process II: Isobaric heating of ideal gas


T2

∆ hII =∫ c 'p dT
T1

Page 14 of 31
T2
c 'p
∆ s II =∫ dT
T1
T

Process III: Change of pressure of ideal gas at constant temperature


∆ hIII =0
P2
∆ s III =−R ln
P1

Process IV: Ideal gas to real gas at constant P2, T2


' ¿
∆ hIV =h P ,T −h P ,T =−∆ h P , T
2 2 2 2 2 2

' ¿
∆ s IV =s P 2
,T 2 −s P2 , T 2 =−∆ s P2 , T 2

Summarizing the four steps we have;


T2

∆ h=∆ h ¿
P1 , T 1 +∫ c 'p dT −∆ h ¿P ,T 2 2
1.39
T1

T2
c 'p P
∆ s=∆ s P ,T +∫
¿ ¿
dT−R ln 2 −∆ s P ,T 1.40
1 1
T1
T P1 2 2

Applying thermodynamic relations;

( ) ( ) ( )
¿ '
∂∆h ∂h ∂h
= −
∂P T ∂P T ∂P T

( ∂∂∆Ps ) =( ∂∂ sP ) −( ∂∂Ps )
¿ '

T T T

ZRT
v=
P
And using the reduced forms;
P=P c Pr dP=Pc d Pr
T =T c T r dT =T c d T r
It can be shown that;
Pr
∆ h¿
( )
∂Z d Pr
=T 2r ∫ 1.41
RTc 0
∂Tr Pr Pr

[ ( )] P
P2
∆ s¿ ∂Z d Pr
=∫ Z−1+T r 1.42
R 0
∂Tr r

Page 15 of 31
This method is not readily used due to large uncertainty in the reduced properties. The
correlations of the reduced residual properties are usually represented in charts.

1.4.2 Property Estimation via Generalized Equation of State


We consider the generalized Peng-Robinson Equation. Choosing V as the independent
variable, we have;
∆ h¿=h' −h=( u' + P v ' )−( u+ Pv )
¿ RT −Pv + ( u ' −u ) 1.43
Evaluating ( u' −u ) by means of

( ∂∂ uv ) =T ( ∂∂ PT ) −P
T V
1.44

And integrating between v and v = ∞ where ideal gas behaviour is expected, we have;

[ ( ∂∂ PT ) −P ] dV

u −u=∫ T
'

V V

Thus,

[ ( ) ]
V
∂P
∆ h =RT ( 1−Z ) +∫ P−T
¿
−P dV 1.45
∞ ∂T V
Likewise,
∆ s ¿=s' −s=( s' −s ∞ ) + ( s∞ −s )
Integrating separately, we have,
'
v ∞
∆ s =∫
¿


( ) dV +∫( ∂∂ TP ) dV
∂P
∂T V v V
1.46

Integration of the first integral over the ideal gas range where ( ∂∂ TP ) = Rv .
V

v v
R R
The problem of limits is circumvented by adding the quantity ∫ dV −∫ dV to the r.h.s.
∞ v ∞ v
of Eqn. 1.46 to obtain;

[ ( )]
v
R ∂P v
∆ s =∫
¿
− dV −R ln ' 1.47
∞ v ∂T V v

∫[ ( ) ]
v
¿ R ∂P
Or ∆s = − dV −R ln Z 1.48
∞ v ∂T V

The Peng-Robinson equation is explicit in P, and ( ∂∂ TP ) is easily obtained, resulting to;


V

Page 16 of 31
∆ h¿
RTc
=2.078 ( 1+ k ) [ 1+k ( 1−√ T r ) ] ln
[
Z + ( 1+ √ 2 ) B
Z + ( 1− √ 2 ) B ]
−T r ( Z−1 ) 1.50

∆ s¿
R
=2.078 k
[
( 1+ k )
√T r
−k ln
][
Z+ ( 1+ √ 2 ) B
Z + ( 1−√ 2 ) B ]
−ln ( Z−B ) 1.51

Where:
Z3 + ( B−1 ) Z 2+ ( A−3 B2−2 B ) Z + ( B3 +B 2− AB )=0 1.52
Pr
[ 1+ k (1−T )]
2
1 /2
A=0.45724 2 r 1.53
Tr
Pr
B=0.07780 1.54
Tr
2
k =0.37464+ 1.54226 ω−0.26992 ω 1.55
¿ ¿
Equations 1.50 to 1.55 allows for the evaluation of Z, ∆ h , and∆ s at specified sets of T and P
from a knowledge of Tc, Pc and the acentric factor, ω.

Example 1.2
Estimate ∆ h¿ and∆ s ¿for chlorine gas, Cl2 at 155 bar and 521K using,
(a) Tables (3.5 cal/gmol.K, 2.13.5 cal/gmol.K)
(b) The generalized Peng-Robinson equation of state (3.63 cal/gmol.K, 2.13
cal/gmol.K)

Homework/Project

Develop a thermodynamic property chart for oxygen over a temperature range -100°C to
150°C (at intervals of 10°C) and pressure range 1 bar to 100 bar (at 10 bar intervals).
Calculate Z, v, h and s as functions of T and P, and hence prepare a plot of P-v, p-h and T-s.
Assume oxygen obeys Peng-Robinson equation of state.

RT a(T )
i.e. P= −
v−b v ( v+ b ) + ( v−b )
2 2
R Tc
a ( T )=0.45724 α (T )
Pc

RT c
b ( T )=0.7780
Pc

[ ( √ )]
2
T
α (T )= 1+ k 1−
Tc

Page 17 of 31
2
k =0.37464+ 1.54226 ω−0.26992 ω

The departure functions for Peng-Robinson equation of state are:

[
hT , P−h T , P =R T c T r ( Z−1 )−2.078 ( 1+k ) √ ∝ ln
ideal
[ Z + ( 1+ √ 2 ) B
Z+ ( 1−√ 2 ) B ]]
ideal
[
sT ,P −hT , P =R ln ( Z−B )−2.078 k
( 1+ k
√ Tr
−k ln
)[
Z + ( 1+ √ 2 ) B
Z + ( 1− √2 ) B ]]
Pr
Where B=0.07780
Tr

¿ −2 −5 2 −9 3
c p=25.46 +1.519 ×10 T −0.715 ×10 T + 1.311×10 T

Required:
(a) Table of specific volumes
(b) Table of h
(c) Table of s
(d) Plots

I.5 Maxwell Equations/relations


The thermodynamic properties – Helmholtz free energy function and Gibb’s free energy
function defined as A=U−TS , and G=H −TS facilitate the establishment of the
thermodynamic network and treatment of equilibrium. They are also measures of the
maximum work obtainable in an isothermal process, hence they are often called work
functions.
For any process occurring in a closed system, the change in entropy
∆ S system + ∆ S surr . ≥ 0 1.56
Considering a closed system surrounded by a heat reservoir at temperature T, the entropy
change is
Q
∆ S≥ (Clausius inequality) 1.57
T
Substituting Eqn. 1.57 in the statement of the 1st law of thermodynamics, we obtain,
∆U = Q+W; Q=∆U-W
T ∆ S ≥ ∆ U −W
T ∆ S−∆U + W ≥0

Page 18 of 31
∆ U −T ∆ S−W ≤ 0

∆ U −T ∆ S−W ≤ 0 1.58
Two types of work can be identified in a system;
i) Work exchanged with the environment at a constant pressure due to changes in
the system, W.
ii) All other forms of work, w’ (shaft work, electric work etc).
Thus, we can rewrite Eqn. 1.58 as
'
∆ U −T ∆ S+ P ∆ V −W ≤ 0 1.59
Equations 1.58 and 1.59 are alternative forms of the Clausius inequality. Equation 1.58 is
restricted for constant temperature process, while eqn. 1.59 is useful when the surroundings
are at a constant T and P. Considering an infinitesimal change, we rewrite equation 1.59 as;
'
dU −TdS+ PdV −dW ≤ 0 1.60
Or dU =TdS−PdV (restricting system to PV work only) 1.61

dU =( ∂∂US ) dS+( ∂U
V ∂V )
dV
S

Equation 1.61 is a fundamental equation of thermodynamics which relates property changes.


It can be expressed in terms of H, A, and G.
e.g. H=U + PV , or dH =dU + PdV +VdP
or dH =TdS+VdP 1.62

dH =( ∂∂HS ) dS+( ∂∂ HP ) dP
P S

similarly,
A=U−TS : dA=dU−TdS−SdT
G=H −TS : dG=dH−TdS−SdT

dA=−PdV −SdT 1.63

dA= ( ∂∂ VA ) dV +( ∂∂TA ) dT
T V

dG=VdP−SdT 1.64

dG= ( ∂∂ GP ) dP+( ∂∂TG ) dT


T P

Equations 1.61 – 1.64 allows for the establishment of a network of relationships among
thermodynamic variables. They have the form of exact differentials,

Page 19 of 31
dZ= ( ∂∂ Zx ) dx +( ∂∂ Zy ) dy
y x
1.65

Where Z=Z ( x , y ) , and can be expressed as;


dZ= Mdx+ Ndy (exact differential) 1.66
A useful property of this type of equation is,

( ∂∂My ) =( ∂∂Nx )
x y
1.67

Comparing equations 1.61 – 1.64 with Eqn. 1.65, the following relations are apparent:

( ∂∂US ) =( ∂∂HS ) =T
V P
1.68

( ∂∂VU ) =( ∂∂VA ) =−P


S T
1.69

( ∂∂TG ) =( ∂∂ TA ) =−S
P V
1.70

( ∂∂ GP ) =( ∂∂ HP ) =V
T S
1.71

Applying Eqn. 1.67 to equations 1.61 – 1.64 yields what are called Maxwell relations, i.e.

( ∂∂My ) =( ∂∂Nx )
x y
1.67

( ∂∂TV ) =−( ∂∂ PS )
S V
1.72

( ∂∂ TP ) =( ∂∂VS )
S P
1.73

( ∂∂ TP ) =( ∂∂VS )
V T
1.74

( ∂∂TV ) =−( ∂∂ PS )
P T
1.75

From dH =TdS+VdP , it can be shown by imposing the conditions of constant temperature,


that

( ∂∂ HP ) =V −T ( ∂∂TV )
T P
1.75

Examples/exercises

Page 20 of 31
I.6 The Joule-Thomson Coefficient
The Joule-Thompson coefficient ( μ JT ) is defined as,

μJT = ( ∂∂ TP ) = cV ( αT −1)
H p
1.76

Where α is coefficient of thermal expansion.


The Joule-Thompson coefficient is a measure of the change in temperature which results
from a drop in pressure when a fluid flows through a constriction (e.g. orifice, valve etc)
under iso-enthalpic conditions (adiabatic conditions).
For most real gases at ambient conditions, μJT is positive (i.e. the temperature falls as the
fluid passes through the constriction), except for H2 and He at which it is negative, and hence
the temperature increases. The temperature below which a non-ideal (real) gas that is
expanding at constant enthalpy will experience a temperature decrease, and above which it
will experience a temperature increase is called the inversion temperature.
The Joule-Thompson effect finds application in:
(a) Refrigeration and air conditioning (in vapour-compression refrigeration whereby
cooling is produced in the throttling valve).
(b) The Linde technique – a standard technique in petrochemical industry used to liquefy
a gas.
(c) Other cryogenic applications. Cryogenics is the study of the production and behaviour
of materials at very low temperatures. A gas is said to be cryogenic if it can be
liquefied at or below -150°C (123K or -238°F).

Derivation of μJT
μJT =μ JT ( T , P , H ). By applying the cyclic rule in terms of these variables, a useful result is
obtained;

i.e. ( ∂∂ TP ) ( ∂∂ HT ) ( ∂∂ HP ) =−1
H P T
1.77

Each of the three partial derivatives has a meaning:

( ∂∂ TP ) =μ
1st:
H
JT

2 :(
∂T )
∂H
nd
=c p
P

3 :(
∂H)
∂P
rd
=μ , the inverse of the isothermal Joule-Thompson Coefficient. This quantity is
T
T

easier to measure than μJT .


Thus, the expression of the cyclic rule becomes:
−μT
μJT = 1.78
cp
(from which μJT is easily obtained by measuring the isothermal Joule-Thompson coefficient).

Page 21 of 31
From dH =TdS+VdP , dividing by dP holding temperature constant yields,

( ∂∂ HP ) =T ( ∂∂ PS ) +V
T T
1.79

From Maxwell relations,

( ∂∂ SP ) =−( ∂∂ VT ) =−Vα , where α is coefficient of thermal expansion.


T P

Thus, eqn. 1.79 becomes,


μT =−TVα +V 1.80
Replacing for μT in Eqn. 1.78 gives,
V
μJT = ( αT −1 ) 1.81.
cp
(note: an alternate method by prof. Paul J. Gans will provided separately)
Exercise 1
Show that for an ideal gas the Joule-Thompson coefficient is zero.
The enthalpy of an ideal gas is a function of temperature only, h=h ( T ), which requires that
the temperature remains constant when the enthalpy remains constant. Therefore, a throttling
process cannot be used to lower the temperature of an ideal gas.
i.e.
For a throttling process,

h=const .

P1 P2 P

Note:
For an iso-enthalpic (adiabatic) expansion,
For, μJT <0 , temperature increases
μJT =0, temperature remains constant
μJT >0 , temperature decreases

For a fluid undergoing a throttling process with initial states (P 1, T1), and forced to go through
a porous plug of different pore sizes, a constant enthalpy plot can be drawn in a T-P diagram
as shown in Figure 1.3.

T Constant enthalpy
line
Page 22 of 31

T1
P2 ,T 2

Figure 1.3

Plotting for several (P1, T1) states yields a plot of several constant enthalpy lines for a
substance as shown in Figure 1.4

Figure 1.4

A line joining all points of zero slope or Joule Thompson coefficient represents the inversion
curve (or line). The point at a constant enthalpy line intersects the inversion line is the
inversion temperature. With the knowledge of constant pressure specific heat, C p and PVT
behaviour of a substance, the joule-Thompson coefficient can be determined.

Exercise
1. Derive a relation for the Joule-Thompson coefficient and the inversion temperature
for a gas whose equation of state is;

( a
)
P+ 2 v=RT
v

2. Estimate the Joule-Thompson coefficient of steam at;

Page 23 of 31
(a) 3 MPa and 300°C
(b) 6 Mpa and 500°C

I.7 The Clausius-Claypeyron Equation


The Clausius-Claypeyron Equation is used to determine properties on a co-existence line.
E.g. suppose we know the melting point of a substance at atmospheric pressure, what will be
its melting point at a different pressure?
Consider Fig.1.3 below for a coexistence line for a system in equilibrium.

g1 = g2
Phase 1
b
P+dP
g1 > g2
a
P Phase 2

T T+dT

Fig. 1.3 Co-existence line


At points “a” and “b”, the two phases are in equilibrium, hence the Gibb’s free energies of
the two phases at each point are the same.
a a b b
i.e. g1 =g 2 and g1 =g 2

Thus, if dg is the difference in Gibb’s free energy between the two points, it is the same for
both phases, i.e. dg 1=dg 2.

But, dg 1=−s1 dT + v 1 dP

dg 2=−s2 dT + v 2 dP

−s1 dT + v 1 dP=−s 2 dT +v 2 dP

Thus, ( s2−s 1 ) dT =( v 2−v 1 ) dP

dP s 2−s 1
Or, =
dT v 2−v 1
1.82
Equation 1.82 is known as the Clausius – Claypeyron equation and it relates the slope along
the coexistence line with the change of entropy and volume of a substance as it crosses the
line, i.e. changes in phases.

Page 24 of 31
Since we cannot measure entropy directly, a more useful form of the equation is obtained by
Q
using ∆ S= for an isothermal process, and finding the change in entropy at a phase
T
transition from the latent heat, L.
dP L
i.e. = (for solid-liquid phase equilibrium)
dT T ∆ V
1.83
It can also be expressed as,

ln
( )
P2 −∆ H 1 1
P1
=
R

T2 T1 ( )
1.84
∆ H vap ,
Or ln P=−¿ ¿ .
+C 1.85
RT
Where ΔH can either be ΔHvap. or ΔHsub. depending on the process under discussion.
T1, T2 are in Kelvin
P1, P2 can be in any units as long as they are the same.

Note: The normal boiling point is defined as the temperature at which vapour pressure is
equal to one atmosphere.
1 atm. = 1.0133 bar;
1 bar = 105 Pa = 100 kPa.
760 Torr = 101325 Pa = 101.325 kPa.

The Clapeyron Equation


As learnt earlier, the Clapeyron equation allows for the determination of properties at a co-
existence line (i.e. during phase change). It can help us determine also the enthalpy of
vaporization (hfg) from knowledge of PVT data. Consider the Maxwell relation,

( ∂∂ TP ) =( ∂∂VS )
V T
(i)

During a phase change process, the pressure is saturation pressure, which is a function of
temperature only, and independent of the specific volume.

i.e. P=P sat .=f (T )

Page 25 of 31
Hence, the partial derivative ( ∂∂ TP )
V
can be expressed as a total derivative, ( dPdT )
sat
which is

the slope of the saturation curve on a P-T diagram. The slope ( dPdT ) is independent of volume,
and hence can be treated as a constant during the integration of eqn (i) between two saturation
states at the same temperature (e.g. isothermal saturated liquid-vapour states). For such we
can write;

s g−sf = ( dPdT ) sat


( s g−s f ) (ii)

( ) dP s fg
Or = (iii)
dT sat v fg

Also during phase change, the pressure remains constant, therefore, from;
dh=Tds+VdP And for dP=0 we have
g g

∫ dh=¿ ∫ Tds ¿
f f

Or h fg =T s fg

Thus, Equation (iii) can be rewritten as;

( dPdT )
h fg
= (iv),
sat T . v fg

Equation (iv) is another form of the Clapeyron equation.


This equation can help us determine hfg at a given temperature by just measuring the slope of
the saturation curve on a P-T diagram, and the specific volume of the saturated liquid and
saturated vapour at the given temperature.
Generally, for any phase change occurring at a constant pressure and temperature,

( dPdT )
h12
= (v)
sat T v 12

Where 1,2 denotes the two phases.


The equation be expressed in the following simplified forms.
(i) For liquid-vapour phase changes

At low pressures, v g ≫ v f , Hence v fg ≅ v g

Page 26 of 31
RT
And assuming vapour is an ideal gas, v g=
P

( dPdT )
Phfg
Thus, = 2
sat RT

( ) ( )
dP hfg dT
Or = (vi)
P sat R T2

For a small temperature difference, hfg can be treated as constant at some average value.
Integrating (vi) we have,

( ) ( )
P2 hfg 1 1
ln ≅ − (vii)
P1 sat R T1 T2 sat

Equation (vii) is the Clausius –Clapeyron Equation (Eqn. 1.84) derived earlier.

(ii) For Solid-vapour phase changes


Replacing hfg by hig in Eqn. (vii), we get

( ) ( )
P2 hig 1 1
ln ≅ − (viii)
P1 sat R T1 T2 sat

Example
Using the Clapeyron equations (iv) and (vii), estimate the value of the enthalpy of
vaporization of refrigerant 134a at 20°C, and compare with the tabulated value. Account for
the difference in the estimated values using the two equations.

Page 27 of 31
Appendix 1.1 Critical Constants for Selected Substances
Substance Molecular Critical Critical Critical Acentric
weight temperature, pressure, compressibility factor, ω
Tc(K) Pc (atm) factor, Zc
Helium He 4.00 5.3 2.26 0.300 -0.365
Neon Ne 20.18 44.5 26.9 0.307 -0.029
Argon A 39.94 151.0 48.0 0.291 0.001
Krypton Kr 83.7 209.4 54.3 0.291 0.005
Chlorine Cl2 70.91 417.0 76.1 0.276 0.090
Hydrogen H2 2.02 33.3 12.8 0.304 -0.218
Nitrogen N2 28.02 126.2 33.5 0.291 0.039
Oxygen O2 32.00 154.8 50.1 0.308 0.025
Carbon dioxide CO2 44.01 304.2 72.9 0.274 0.239
Carbon monoxide CO 28.01 133.0 34.5 0.294 0.066
Ethylene oxide C2H4O 44.05 468.0 71.0 0.255 0.202
Nitrous oxide N2O 44.02 309.7 71.7 0.272 0.165
Nitric oxide NO 30.01 180.0 64.0 0.251 0.588
Nitrogen peroxide NO2 46.01 431.0 100.0 0.232 0.834
Sulfur dioxide SO2 64.06 430.7 77.8 0.268 0.256
Sulfur trioxide SO3 80.06 491.4 83.8 0.262 0.481
Water H2O 18.02 647.4 218.3 0.230 0.344
Ammonia NH3 17.03 405.5 111.3 0.242 0.250
Hydrazine N2H4 32.05 653.0 145.0 - 0.316
Hydrogen cyanide HCN 27.03 456.7 53.2 0.197 0.388
Carbon disulfide CS2 76.13 552.0 78.0 0.293 0.109
Hydrogen sulfide H2S 34.08 373.6 88.9 0.283 0.081
Hydrogen chloride HCl 36.47 324.6 81.5 0.147 0.133
Methane CH4 16.04 191.1 45.8 0.289 0.011
Ethane C2H6 30.07 305.5 48.2 0.284 0.099
Propane C3H8 44.09 370.0 42.0 0.277 0.153
n-butane C4H10 58.12 425.2 37.5 0.274 0.199
Isobutene C4H10 58.12 408.1 36.0 0.283 0.183
n-pentane C5H12 72.15 469.8 33.3 0.269 0.251
Ethylene C2H4 28.05 282.4 50.0 0.268 0.089
Propene C3H6 42.08 365.0 45.6 0.276 0.144
Acetylene C2H2 26.04 309.0 61.6 0.274 0.190
Cyclohexane C6H12 84.16 553.0 40.0 0.272 0.212
Benzene C6H6 78.11 562.0 48.6 0.274 0.212
Toluene C7H8 92.13 594.0 41.6 0.273 0.263
Methyl alcohol CH4O 32.04 513.2 78.5 0.220 0.556
Ethyl alcohol C2H6O 46.07 516.0 63.0 0.248 0.644
n-propyl alcohol C3H8O 60.09 537.0 50.2 0.251 0.623
Isopropyl alcohol C3H8O 60.09 508.8 53.0 0.278 0.665
Dioxane C4H8O2 88.10 585.0 50.7 0.253 0.281
Acetone C3H6O 58.08 508.7 46.6 0.238 0.304
Ethyl methyl ketone C4H8O 72.10 533.0 39.5 0.262 0.320
Acetic acid C2H4O2 60.05 594.8 57.1 0.200 0.447
Ethyl acetate C4H8O2 88.10 523.3 37.8 0.252 0.362
Acetonitrile CH3CN 41.05 547.9 47.7 0.184 0.327
Methyl chloride CH3Cl 50.49 416.3 65.9 0.276 0.153
Chloroform CHCl3 119.39 536.6 54.0 0.294 0.218
Carbon tetrachloride CCl4 153.84 556.4 45.0 0.272 0.193
From K.A. Kobe and R.E. Lynn, Jr., “The Critical Properties of Elements and Compounds,” Chem. Rev., 52, 117 (1953) and
R.C. Reid, J.M. Prausnitz, and B.E. Poling, The properties of Gases and Liquids, 4th ed., McGraw-Hill, New York, 1987.

Page 28 of 31
Appendix 1.2 Generalized Equation of State Plot

Generalized compressibility factor


Zc=0.27

Page 29 of 31
Appendix 1.3 Generalized enthalpy departure from ideal gas behaviour

Page 30 of 31
Appendix 1.4 Generalized entropy departure from ideal gas behaviour

Page 31 of 31

You might also like