You are on page 1of 290

LECTURE NOTES ON

INTERMEDIATE THERMODYNAMICS
Joseph M. Powers
Department of Aerospace and Mechanical Engineering
University of Notre Dame
Notre Dame, Indiana 46556-5637
USA
updated
03 August 2012, 12:45pm

CC BY-NC-ND. 03 August 2012, J. M. Powers.

Contents
Preface

1 Review

2 Cycle analysis
2.1 Carnot . . . . . . . . . . . . . . .
2.2 Exergy . . . . . . . . . . . . . . .
2.3 Rankine . . . . . . . . . . . . . .
2.3.1 Classical . . . . . . . . . .
2.3.2 Reheat . . . . . . . . . . .
2.3.3 Regeneration . . . . . . .
2.3.4 Losses . . . . . . . . . . .
2.3.5 Cogeneration . . . . . . .
2.4 Air standard cycles . . . . . . . .
2.5 Brayton . . . . . . . . . . . . . .
2.5.1 Classical . . . . . . . . . .
2.5.2 Regeneration . . . . . . .
2.5.3 Ericsson Cycle . . . . . . .
2.5.4 Jet propulsion . . . . . . .
2.6 Reciprocating engine power cycles
2.7 Otto . . . . . . . . . . . . . . . .
2.8 Diesel . . . . . . . . . . . . . . .
2.9 Stirling . . . . . . . . . . . . . . .
2.10 Refrigeration . . . . . . . . . . .
2.10.1 Vapor-compression . . . .
2.10.2 Air standard . . . . . . . .
3 Gas
3.1
3.2
3.3

mixtures
Some general issues . . . . . . . .
Ideal and non-ideal mixtures . . .
Ideal mixtures of ideal gases . . .
3.3.1 Dalton model . . . . . . .
3.3.1.1 Binary mixtures

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

17
17
24
28
28
33
33
37
37
37
38
38
43
46
48
53
55
59
64
65
65
69

.
.
.
.
.

71
71
75
75
76
79

CONTENTS

3.4

3.3.1.2 Entropy of mixing . . . . . . . . . .


3.3.1.3 Mixtures of constant mass fraction .
3.3.2 Summary of properties for the Dalton mixture
3.3.3 Amagat model . . . . . . . . . . . . . . . . .
Gas-vapor mixtures . . . . . . . . . . . . . . . . . . .
3.4.1 First law . . . . . . . . . . . . . . . . . . . . .
3.4.2 Adiabatic saturation . . . . . . . . . . . . . .
3.4.3 Wet-bulb and dry-bulb temperatures . . . . .
3.4.4 Psychrometric chart . . . . . . . . . . . . . .

. . . .
. . . .
model
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

.
.
.
.
.
.
.
.
.

4 Mathematical foundations of thermodynamics


4.1 Exact differentials and state functions . . . . . . . . . . . . . .
4.2 Two independent variables . . . . . . . . . . . . . . . . . . . .
4.3 Legendre transformations . . . . . . . . . . . . . . . . . . . . .
4.4 Heat capacity . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.5 Van der Waals gas . . . . . . . . . . . . . . . . . . . . . . . .
4.6 Redlich-Kwong gas . . . . . . . . . . . . . . . . . . . . . . . .
4.7 Compressibility and generalized charts . . . . . . . . . . . . .
4.8 Mixtures with variable composition . . . . . . . . . . . . . . .
4.9 Partial molar properties . . . . . . . . . . . . . . . . . . . . .
4.9.1 Homogeneous functions . . . . . . . . . . . . . . . . . .
4.9.2 Gibbs free energy . . . . . . . . . . . . . . . . . . . . .
4.9.3 Other properties . . . . . . . . . . . . . . . . . . . . .
4.9.4 Relation between mixture and partial molar properties
4.10 Irreversible entropy production in a closed system . . . . . . .
4.11 Equilibrium in a two-component system . . . . . . . . . . . .
4.11.1 Phase equilibrium . . . . . . . . . . . . . . . . . . . . .
4.11.2 Chemical equilibrium: introduction . . . . . . . . . . .
4.11.2.1 Isothermal, isochoric system . . . . . . . . . .
4.11.2.2 Isothermal, isobaric system . . . . . . . . . .
4.11.3 Equilibrium condition . . . . . . . . . . . . . . . . . .
5 Thermochemistry of a single reaction
5.1 Molecular mass . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Stoichiometry . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.1 General development . . . . . . . . . . . . . . . . . . .
5.2.2 Fuel-air mixtures . . . . . . . . . . . . . . . . . . . . .
5.3 First law analysis of reacting systems . . . . . . . . . . . . . .
5.3.1 Enthalpy of formation . . . . . . . . . . . . . . . . . .
5.3.2 Enthalpy and internal energy of combustion . . . . . .
5.3.3 Adiabatic flame temperature in isochoric stoichiometric
5.3.3.1 Undiluted, cold mixture . . . . . . . . . . . .
CC BY-NC-ND. 03 August 2012, J. M. Powers.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
systems
. . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

82
85
86
92
92
95
97
99
99

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

101
101
106
109
114
118
124
125
126
129
129
129
130
132
133
136
136
139
140
145
147

.
.
.
.
.
.
.
.
.

149
149
150
150
159
161
161
163
164
164

CONTENTS

5.4
5.5

5.6

5.7

5.3.3.2 Dilute, cold mixture . . . . . . . . . . . . . .


5.3.3.3 Dilute, preheated mixture . . . . . . . . . . .
5.3.3.4 Dilute, preheated mixture with minor species
Chemical equilibrium . . . . . . . . . . . . . . . . . . . . . . .
Chemical kinetics of a single isothermal reaction . . . . . . . .
5.5.1 Isochoric systems . . . . . . . . . . . . . . . . . . . . .
5.5.2 Isobaric systems . . . . . . . . . . . . . . . . . . . . . .
Some conservation and evolution equations . . . . . . . . . . .
5.6.1 Total mass conservation: isochoric reaction . . . . . . .
5.6.2 Element mass conservation: isochoric reaction . . . . .
5.6.3 Energy conservation: adiabatic, isochoric reaction . . .
5.6.4 Energy conservation: adiabatic, isobaric reaction . . . .
5.6.5 Entropy evolution: Clausius-Duhem relation . . . . . .
Simple one-step kinetics . . . . . . . . . . . . . . . . . . . . .

6 Thermochemistry of multiple reactions


6.1 Summary of multiple reaction extensions . . . . . .
6.2 Equilibrium conditions . . . . . . . . . . . . . . . .
6.2.1 Minimization of G via Lagrange multipliers
6.2.2 Equilibration of all reactions . . . . . . . . .
6.3 Concise reaction rate law formulations . . . . . . .
6.3.1 Reaction dominant: J (N L) . . . . . .
6.3.2 Species dominant: J < (N L) . . . . . . .
6.4 Adiabatic, isochoric kinetics . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

7 Kinetics in some more detail


7.1 Isothermal, isochoric kinetics . . . . . . . . . . . . . . . . . . . . .
7.1.1 O O2 dissociation . . . . . . . . . . . . . . . . . . . . . .
7.1.1.1 Pair of irreversible reactions . . . . . . . . . . . .
7.1.1.1.1 Mathematical model . . . . . . . . . . .
7.1.1.1.2 Example calculation . . . . . . . . . . .
7.1.1.1.2.1
Species concentration versus time
7.1.1.1.2.2
Pressure versus time . . . . . . .
7.1.1.1.2.3
Dynamical system form . . . . . .
7.1.1.1.3 Effect of temperature . . . . . . . . . . .
7.1.1.2 Single reversible reaction . . . . . . . . . . . . . .
7.1.1.2.1 Mathematical model . . . . . . . . . . .
7.1.1.2.1.1
Kinetics . . . . . . . . . . . . . .
7.1.1.2.1.2
Thermodynamics . . . . . . . . .
7.1.1.2.2 Example calculation . . . . . . . . . . .
7.1.2 Zeldovich mechanism of NO production . . . . . . . . . .
7.1.2.1 Mathematical model . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

165
167
168
169
175
175
183
189
189
190
191
193
196
199

.
.
.
.
.
.
.
.

203
203
210
210
216
218
219
220
220

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

223
226
226
226
226
232
233
234
236
240
241
241
241
243
244
247
247

CC BY-NC-ND. 03 August 2012, J. M. Powers.

CONTENTS

7.2

7.1.2.1.1 Standard model form . . .


7.1.2.1.2 Reduced form . . . . . . .
7.1.2.1.3 Example calculation . . .
7.1.2.2 Stiffness, time scales, and numerics
7.1.2.2.1 Effect of temperature . . .
7.1.2.2.2 Effect of initial pressure .
7.1.2.2.3 Stiffness and numerics . .
Adiabatic, isochoric kinetics . . . . . . . . . . . . .
7.2.1 Thermal explosion theory . . . . . . . . . .
7.2.1.1 One-step reversible kinetics . . . .
7.2.1.2 First law of thermodynamics . . .
7.2.1.3 Dimensionless form . . . . . . . . .
7.2.1.4 Example calculation . . . . . . . .
7.2.1.5 High activation energy asymptotics
7.2.2 Detailed H2 O2 N2 kinetics . . . . . . .

Bibliography

CC BY-NC-ND. 03 August 2012, J. M. Powers.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

248
249
252
258
258
261
263
264
265
265
266
269
270
272
276
281

Preface
These are lecture notes for AME 50531 Intermediate Thermodynamics, the second of two
undergraduate courses in thermodynamics taught in the Department of Aerospace and Mechanical Engineering at the University of Notre Dame. Most of the students in this course
are juniors or seniors majoring in aerospace or mechanical engineering. The objective of the
course is to survey both practical and theoretical problems in classical thermodynamics.
The notes draw heavily on the text specified for the course, Borgnakke and Sonntags
(BS) Fundamentals of Thermodynamics, Seventh Edition, John Wiley, New York, 2009, especially Chapters 10-16. In general the nomenclature of BS is used, and much of the notes
follow a similar structure as that text. In addition, Abbott and van Nesss Thermodynamics,
McGraw-Hill, New York, 1972, has been used to guide some of the mathematical developments. Many example problems have been directly taken from BS and other texts; specific
citations are given where they arise. Many of the unique aspects of these notes, which focus
on thermodynamics of reactive gases with detailed finite rate kinetics, have arisen due to
the authors involvement in research supported by the National Science Foundation under
Grant No. CBET-0650843. The author is grateful for the support. Any opinions, findings,
and conclusions or recommendations expressed in this material are those of the author and
do not necessarily reflect the views of the National Science Foundation.
These notes emphasize both problem-solving as well as some rigorous undergraduatelevel development of the underlying classical theory; the student should call on textbooks
and other reference materials to fill some of the gaps. It should also be remembered that
practice is essential to the learning process; the student would do well to apply the techniques
presented here by working as many problems as possible.
The notes, along with information on the course, can be found on the world wide web at
http://www.nd.edu/powers/ame.50531. At this stage, anyone is free to make copies for
their own use. I would be happy to hear from you about suggestions for improvement.
Joseph M. Powers
powers@nd.edu
http://www.nd.edu/powers

Notre Dame, Indiana; USA


=
$
CC
BY:




03 August 2012
The content of this book is licensed under Creative Commons Attribution-Noncommercial-No Derivative
Works 3.0.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

CONTENTS

Chapter 1
Review
Review BS, Chapters 1-9.
Here is a brief review of some standard concepts from a first course in thermodynamics.
property: characterizes the thermodynamics state of the system
extensive: proportional to systems extent, upper case variables

V : total volume, SI-based units are m3 ,


U: total internal energy, SI-based units are kJ,
H: total enthalpy, SI-based units are kJ,
S: total entropy, SI-based units are kJ/K.

v:
u:
h:
s:
v:
u:
h:
s:

intensive: independent of systems extent, lower case variables (exceptions are


temperature T and pressure P , which are intensive). Intensive properties can be
on a mass basis or a molar basis:
specific volume, SI-based units are m3 /kg,
specific internal energy, SI-based units are kJ/kg,
specific enthalpy, SI-based units are kJ/kg,
specific entropy, SI-based units are kJ/kg/K.
molar specific volume, SI-based units are m3 /kmole,
molar specific internal energy, SI-based units are kJ/kmole,
molar specific enthalpy, SI-based units are kJ/kmole,
molar specific entropy, SI-based units are kJ/kmole/K.

density: = 1/v, mass per unit volume, SI-based units are kg/m3 .
equations of state: relate properties
Calorically Perfect Ideal Gas (CPIG) has P v = RT and u uo =
cv (T To ),
9

10

CHAPTER 1. REVIEW
Calorically Imperfect Ideal Gas (CIIG) has P v = RT and u uo =
RT
c (T ) dT.
To v

Non-ideal state equations have P = P (T, v), u = u(T, v).

Any intensive thermodynamic property can be expressed as a function of at most two


other intensive thermodynamic properties (for simple compressible systems).
P = RT /v: thermal equation of state for ideal gas, SI-based units are kP a.

c = kP v: sound speed for CPIG, SI-based units are m/s; k = cP /cv , the ratio
of specific heats.
energy: E = U + (1/2)m(v v) + mgz, total energy = internal+kinetic+potential.
first law: dE = Q W ; if kinetic and potential are ignored, dU = Q W ; with
Q = T dS and W = P dV , we get the Gibbs equation dU = T dS P dV .
second law: dS Q/T .
process: moving from one state to another, in general with accompanying heat transfer and work.
cycle: process which returns to initial state.
R2
specific reversible work: 1 w2 = 1 P dv; w = P dv.
R2
specific reversible heat transfer: 1 q2 = 1 T ds; q = T ds.

Figure 1.1 gives an example of an isothermal thermodynamic process going from state 1
to state 2 in various thermodynamic planes. Figure 1.2 gives a sketch of a thermodynamic
cycle.
Example 1.1
Consider an ideal gas in the T s plane. Compare the slope of an isochore to that of an isobar at
a given point.
Recall the Gibbs equation for a simple compressible substance:
T ds = du + P dv.

(1.1)

We have for the ideal gas


du = cv dT,

if ideal gas.

(1.2)

This holds for all ideal gases, be they calorically perfect or imperfect. Thus, the Gibbs equation can be
rewritten as
T ds = cv dT +P dv.
| {z }
=du

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(1.3)

11

T
v1

v2

P1
T1 = T2

P2

v
u 2- u1= q12- w12

w12= 1 P dv
P
P1

T1 = T2

v2

v1

s1
2

q12= 1 T ds

v1

P2

v2

s2

v2

P1
P2

v1

T1 = T 2

T1 = T 2

Figure 1.1: Sketch of isothermal thermodynamic process.


On an isochore, v is constant, so dv = 0. So on an isochore, we have
T ds = cv dT,

on an isochore.

(1.4)

or, using the partial derivative notation,



T
T
= .
s v
cv

(1.5)

Next recall the definition of enthalpy, h:

h = u + P v.

(1.6)

dh = du + P dv + vdP.

(1.7)

We can differentiate Eq. (1.6) to get

Substitute Eq. (1.7) into the Gibbs equation, (1.1), to eliminate du in favor of dh to get
T ds =

dh P dv vdP +P dv,
{z
}
|

(1.8)

=du

dh vdP.

(1.9)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

12

CHAPTER 1. REVIEW

v
q cycle = wcycle

Figure 1.2: Sketch of thermodynamic cycle.


For an ideal gas, dh = cP dT , where cP is at most a function of T , so
T ds = cP dT vdP.
| {z }

(1.10)

=dh

Now for the isobar, dP = 0. Thus on an isobar, we have


T ds = cP dT,

on an isobar.

(1.11)

And so the slope on an isobar in the T s plane is



T
T
.
=
s P
cP

(1.12)

Since cP > cv , (recall for an ideal gas that cP (T ) cv (T ) = R > 0), we can say that the slope of the
isochore is steeper than the isobar in the T s plane.

Example 1.2
Consider the following isobaric process for air, modeled as a CPIG, from state 1 to state 2. P1 =
100 kP a, T1 = 300 K, T2 = 400 K. Show the second law is satisfied.
Since the process is isobaric, P = 100 kP a describes a straight line in the P v and P T planes
and P2 = P1 = 100 kP a. Since we have an ideal gas, we have for the v T plane:
 
R
T,
straight lines!
(1.13)
v=
P

v1
v2

=
=

RT1
=
P1
RT2
=
P2




0.287

kJ
kg K

(300 K)

100 kP a

0.287 kgkJK (400 K)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

100 kP a

= 0.861

m3
,
kg

(1.14)

= 1.148

m3
.
kg

(1.15)

13
Since the gas is ideal:
du

= cv dT.

Since our ideal gas is also calorically perfect, cv is a constant, and we get
Z T2
Z u2
dT ,
du = cv

(1.16)

(1.17)

T1

u1

u2 u1

= cv (T2 T1 ),


kJ
=
0.7175
(400 K 300 K),
kg K
kJ
= 71.750
.
kg

(1.18)
(1.19)
(1.20)

Also we have
T ds
T ds
from ideal gas :

v
T ds
ds

s2

ds

s1

s2 s1

= du + P dv,
= cv dT + P dv,
RT
R
RT
=
:
dv = dT 2 dP,
P
P
P
RT
dP,
= cv dT + RdT
P
dP
dT
R
,
= (cv + R)
T
P
dT
dP
= (cv + cP cv )
R
,
T
P
dP
dT
R
,
= cP
T
P
Z P2
Z T2
dP
dT
R
,
= cP
T
P
P1
T1
 
 
T2
P2
= cP ln
R ln
.
T1
P1

(1.21)
(1.22)
(1.23)
(1.24)
(1.25)
(1.26)
(1.27)
(1.28)
(1.29)
(1.30)

Now since P is a constant,


s2 s1

=
=
=

Then one finds


1 w2

=
=

T2
T1

cP ln
,
 


400 K
kJ
ln
,
1.0045
kg K
300 K
kJ
0.2890
.
kg K

v2

v1

P dv = P

(1.31)
(1.32)
(1.33)

v2

dv,

(1.34)

v1

P (v2 v1 ) ,


m3
m3
,
0.861
(100 kP a) 1.148
kg
kg
kJ
28.700
.
kg

(1.35)
(1.36)
(1.37)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

14

CHAPTER 1. REVIEW
Now

= q w,
= du + w,
Z
Z 2
du +
=

du
q
2

(1.38)
(1.39)
2

w,

(1.40)

= (u2 u1 ) + 1 w2 ,
kJ
kJ
= 71.750
+ 28.700
,
kg
kg
kJ
= 100.450
.
kg

1 q2
1 q2
1 q2

(1.41)
(1.42)
(1.43)

Now in this process the gas is heated from 300 K to 400 K. One would expect at a minimum that the
surroundings were at 400 K. Check for second law satisfaction.
s2 s1

kJ
kg K

0.2890

0.2890

1 q2
?
Tsurr
100.450 kJ/kg
?
400 K

kJ
kJ
0.2511
,
kg K
kg K

CC BY-NC-ND. 03 August 2012, J. M. Powers.

yes.

(1.44)
(1.45)

(1.46)

15

T
T1 = 300 K

v1

T2 = 400 K

P1= P2 = 100 kPa

v2

T2

2
1

T1
v2

v1

v
u 2- u1= q12- w12

w12= 1 P dv
P

s1

s2
q12= 1 T ds

v1
v2
1

P1= P2 = 100 kPa

P1= P2 = 100 kPa

v2
v1

T1

T2

2
1

T1

T2

Figure 1.3: Sketch for isobaric example problem.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

16

CC BY-NC-ND. 03 August 2012, J. M. Powers.

CHAPTER 1. REVIEW

Chapter 2
Cycle analysis
Read BS, Chapters 7, 8, 10, 11, 12.

2.1

Carnot

The Carnot cycle is the most well-known thermodynamic cycle. It is a useful idealization,
but is very difficult to realize in practice. Its real value lies in serving as a standard to which
other cycles can be compared. These are usually fully covered in introductory courses. The
cycle can be considered as follows
1 2: isentropic compression,
2 3: isothermal expansion,
3 4: isentropic expansion, and
4 1: isothermal compression.
This forms a rectangle in the T s plane The P v plane is more complicated. Both
are shown in Figure 2.1. For this discussion, consider a calorically perfect ideal gas. The
isotherms are then straightforward and are hyperbolas described by
1
P = RT .
v
The slope of the isotherms in the P v plane is found by differentiation:

1
P
= RT 2 ,

v T
v
P
= .
v

(2.1)

(2.2)
(2.3)

The slope of the isentrope is found in the following way. Consider first the Gibbs equation,
Eq. (1.1):
T ds = du + P dv.
(2.4)
17

18

CHAPTER 2. CYCLE ANALYSIS

T
2

rop
isent

is e

r
nt

isentrope

erm

3
is o
the
rm

isotherm

op

3
isentrope

h
isot

isotherm

4
v

wcycle = P dv = qcycle = T ds
Figure 2.1: Sketch of P v and T s planes for a CPIG in a Carnot cycle.
Because the gas is calorically perfect, one has

du = cv dT,

(2.5)

so the Gibbs equation, Eq. (2.4), becomes

T ds = cv dT + P dv.

(2.6)

P v = RT,
P dv + vdP = RdT,
P dv + vdP
= dT.
R

(2.7)
(2.8)

Now for the ideal gas, one has

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.9)

19

2.1. CARNOT
So then
T ds = cv

T ds =


|

c


P dv + vdP
+P dv,
R
{z
}
=dT

cv
+ 1 P dv + vdP,
R

cv P
+1 P + v
,
R v

R
c
v
0 =
R

cv

+1 P
P
R
=

,
cv

v s
v
R
cv
+1 P
cP cv
,
=
cv
v
cP cv
cv + cP cv P
=
,
cv
v
cP P
,
=
cv v
P
= k .
v

(2.10)

Take ds = 0,

(2.11)
(2.12)

(2.13)
(2.14)
(2.15)
(2.16)
(2.17)

Since k > 1, the magnitude of the slope of the isentrope is greater than the magnitude of
the slope of the isotherm:



P P



(2.18)
v > v .
s
T
Example 2.1

(Adapted from BS, 7.94, p. 274) Consider an ideal gas Carnot cycle with air in a piston cylinder
with a high temperature of 1200 K and a heat rejection at 400 K. During the heat addition, the volume
triples. The gas is at 1.00 m3 /kg before the isentropic compression. Analyze.
Take state 1 to be the state before the compression. Then
T1 = 400 K,

v1 = 1.00

m3
.
kg

(2.19)

By the ideal gas law

P1 =

RT1
=
v1

kJ
kg K

0.287

1.00

(400 K)

m3
kg

= 1.148 102 kP a.

(2.20)

Now isentropically compress to state 2. By the standard relations for a CPIG, one finds
T2
=
T1

v1
v2

k1

P2
P1

 k1
k

(2.21)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

20

CHAPTER 2. CYCLE ANALYSIS


So
v2 = v1

T1
T2

1
 k1



 1
m3
m3
400 K 1.41
1.00
= 0.06415
.
kg
1200 K
kg

(2.22)

Note that v2 < v1 as is typical in a compression. The isentropic relation between pressure and volume
can be rearranged to give the standard
P2 v2k = P1 v1k .

(2.23)

Thus, one finds that


P2 = P1

v1
v2

k

m3
kg
3
0.06415 m
kg

1.00

= (1.148 102 kP a)

!1.4

= 5.36866 103 kP a.

(2.24)

Note the pressure has increased in the isentropic compression. Check to see if the ideal gas law is
satisfied at state 2:


0.287 kgkJK (1200 K)
RT2
=
P2 =
= 5.36866 103 kP a.
(2.25)
3
v2
0.06415 m
kg
This matches. Now the expansion from state 2 to 3 is isothermal. This is the heat addition step in
which the volume triples. So one gets


m3
m3
= 0.19245
.
(2.26)
v3 = 3v2 = 3 0.06415
kg
kg
T3 = T2 = 1200 K.

(2.27)

The ideal gas law then gives

P3 =

RT3
=
v3

0.287

kJ
kg K

0.19245

(1200 K)
m3
kg

= 1.78955 103 kP a.

(2.28)

Process 3 to 4 is an isentropic expansion back to 400 K. Using the isentropic relations for the CPIG,
one gets
1

 1

  k1
1200 K 1.41
m3
m3
T3
.
(2.29)
= 0.19245
= 3.000
v4 = v3
T4
kg
400 K
kg
P4 = P3
Check:

v3
v4

k

= (1.78955 10 kP a)

m3
kg
m3
kg

0.19245
3.000

!1.4

= 3.82667 101 kP a.



0.287 kgkJK (400 K)
RT4
= 3.82667 101 kP a.
=
P4 =
3
v4
3.000 m
kg

(2.30)

(2.31)

A summary of the states is given in Table 2.1.


Now calculate the work, heat transfer and efficiency. Take the adiabatic exponent for air to be
k = 1.4. Now since
cP
,
cP cv = R,
(2.32)
k=
cv
CC BY-NC-ND. 03 August 2012, J. M. Powers.

21

2.1. CARNOT
3

T (K)
P (kP a) v ( m
)
kg
2
400
1.148 10
1.000
1200 5.36886 103 0.06415
1200 1.78955 103 0.19245
400 3.82667 102
3.000

1
2
3
4

Table 2.1: State properties for Carnot cycle.

one gets

kcv

cv (k 1) =
cv

R + cv
,
cv
R + cv ,

(2.33)
(2.34)

R,

(2.35)
kJ
kg K

0.287
R
=
k1
1.4 1

= 0.7175

kJ
.
kg K

(2.36)

Recall the first law:

u2 u1 = 1 q2 1 w2 .

(2.37)

Recall also the caloric equation of state for a CPIG:

u2 u1 = cv (T2 T1 ).

(2.38)

Now process 1 2 is isentropic, so it is also adiabatic, hence 1 q2 = 0, so one has


u2 u1

1 q2 1 w2 ,
|{z}

(2.39)

=0

cv (T2 T1 ) = 1 w2 ,



kJ
0.7175
(1200 K 400 K) = 1 w2 ,
kg K
1 w2

= 5.7400 102

(2.40)
(2.41)
kJ
.
kg

(2.42)

The work is negative as work is being done on the system in the compression process.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

22

CHAPTER 2. CYCLE ANALYSIS


Process 2 3 is isothermal, so there is no internal energy change. The first law gives
u3 u2

2 w3 ,

2 q3

(2.43)

cv (T3 T2 )
|
{z
}

2 q3

2 q3

2 w3 ,
Z v3
P dv,
2 w3 =
v2
Z v3
RT
dv,
v
v2
Z v3
dv
,
RT2
v
v2
v3
RT2 ln ,
v2


0.19245
kJ
(1200 K) ln
0.287
kg K
0.06415

2 w3 ,

(2.44)

=0

=
=
=
=

2 q3

= 3.78362 102

(2.45)
(2.46)
(2.47)
(2.48)
(2.49)
m3
kg
m3
kg

kJ
.
kg

(2.50)
(2.51)

The work is positive, which is characteristic of the expansion process.


Process 3 4 is adiabatic so 3 q4 = 0. The first law analysis gives then
u4 u3

3 q4 3 w4 ,
|{z}

(2.52)

=0

cv (T4 T3 ) =



kJ
(400 K 1200 K) =
0.7175
kg K
3 w4

3 w4 ,

(2.53)

3 w4 ,

(2.54)

5.74000 102

kJ
.
kg

(2.55)

Process 4 1 is isothermal. Similar to the other isothermal process, one finds


u1 u4 =
cv (T1 T4 ) =
|
{z
}

4 q1

4 w1 ,
4 q1 4 w1 ,

(2.56)
(2.57)

4 w1 ,
Z v1
P dv,
4 w1 =
v4
Z v1
RT
dv,
v
v4
v1
RT4 ln ,
v4


1.0000
kJ
(400 K) ln
0.287
kg K
3.0000

(2.58)

=0

0 =

4 q1

=
=
=
=
=

4 q1

1.26121 102

CC BY-NC-ND. 03 August 2012, J. M. Powers.

kJ
.
kg

(2.59)
(2.60)
(2.61)
m3
kg
m3
kg

(2.62)
(2.63)

23

2.1. CARNOT
Process

kJ
kg

12
5.74000 102
23
0
3 4 5.74000 102
41
0
Total
0

kJ
kg

kJ
kg

0 5.74000 102
3.78362 102
3.78362 102
0
5.74000 102
1.26121 102 1.26121 102
2.52241 102
2.52241 102

Table 2.2: First law parameters for Carnot cycle.


Table 2.2 summarizes the first law considerations. The cycle work is found by adding the work of each
individual process:
wcycle

1 w2

+ 2 w3 + 3 w4 + 4 w1 ,

(2.64)

kJ
.
= (5.74 + 3.78362 + 5.74 1.26121) 102 = 2.52241 102
kg

(2.65)

The cycle heat transfer is


qcycle

1 q2

+ 2 q3 + 3 q4 + 4 q1 ,

(2.66)

kJ
.
= (0 + 3.78362 + 0 1.26121) 102 = 2.52241 102
kg

(2.67)

Note that
wcycle = qcycle .

(2.68)

Check now for the cycle efficiency. Recall that the thermal efficiency is
=

what you want


wcycle
.
=
what you pay for
qin

(2.69)

Here this reduces to


=

2.52241 102
wcycle
=
3.78362 102
2 q3

kJ
kg
kJ
kg

= 0.6667.

(2.70)

Recall that the efficiency for any Carnot cycle should be


=1

Tlow
.
Thigh

(2.71)

So general Carnot theory holds that the efficiency should be


=1

2
400 K
= 0.6667.
1200 K
3

(2.72)

Recall further that for a Carnot cycle, one has


Tlow
qlow
=
.
qhigh
Thigh

(2.73)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

24

CHAPTER 2. CYCLE ANALYSIS


For this problem then one has
4 q1
2 q3

Tlow
,
Thigh

(2.74)

kJ
kg
kJ
2
10 kg

400 K
,
1200 K

(2.75)

1
3

1
.
3

(2.76)

1.26121 102

3.78362

Indeed, the appropriate relation holds.

2.2

Exergy

Here we will introduce the concept of exergy. This concept is widely used in some industrial
design applications; its use in the fundamental physics literature is not extensive, likely
because it is not a thermodynamic property of a material, but also includes mechanical
properties. This system quantity is one measure of how much useful work can be extracted
from a system which is brought into equilibrium with a so-called reference rest state.
First let us imagine that the surroundings are at a reference temperature of To and a
reference pressure of Po . This yields a reference enthalpy of ho and a reference entropy of so .
We also take the surroundings to be at rest with a velocity of v = 0, and a reference height
of zo .
Consider the sketch of Fig. 2.2. Let us consider a steady flow into and out of a control
volume. The conservation of mass equation yields
dmcv
=m
in m
out .
dt
For steady flow conditions, we have d/dt = 0, so we recover
m
in = m
out = m.

The first law of thermodynamics for the control volume yields




1
dEcv
cv + m
= Q cv W
in hin + vin vin + gzin
dt
2


1
m
out hout + vout vout + gzout .
2

(2.77)

(2.78)

(2.79)

We will soon relate Q cv /m


to qH,Carnot as sketched in Fig. 2.2; this will introduce a sign
convention problem. For now, we will maintain the formal sign convention associated with
Q cv connoting heat transfer into the control volume. For steady flow conditions, we get


1

0 = Qcv Wcv + m
hin hout + (vin vin vout vout ) + g(zin zout ) .
(2.80)
2
CC BY-NC-ND. 03 August 2012, J. M. Powers.

25

2.2. EXERGY

control
volume

h + (1/2) v . v+ g z

ho + gz o

qH,Carnot

Carnot
engine

Carnot

q L,Carnot
To

Figure 2.2: Sketch of control volume balance for exergy discussion.


Now take the in state to be simply a generic state with no subscript, and the out
state to be the reference state, so


1
cv + m
0 = Q cv W
h ho + (v v) + g(z zo ) .
2

(2.81)

Let us next scale by m


so as to get
1
0 = qcv wcv + h ho + v v + g(z zo ).
2
cv /m.
Here we have taken qcv Q cv /m
and wcv = W

Now, let us insist that wcv = 0, and solve for qcv to get


1
qcv = h ho + v v + g(z zo ) .
2

(2.82)

(2.83)

Now, we imagine the working fluid to be at an elevated enthalpy, velocity, and height
relative to its rest state. Thus in the process of bringing it to its rest state, we will induce
qcv < 0. By our standard sign convention, this means that thermal energy is leaving the
system. This energy which leaves the system can be harnessed, in the best of all possible
worlds, by a Carnot engine in contact with a thermal reservoir at temperature To to generate
useful work. It is this work which represents the available energy, also known as the exergy.
For the Carnot engine, we have
wCarnot = qH,Carnot qL,Carnot .

(2.84)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

26

CHAPTER 2. CYCLE ANALYSIS

Note the standard sign convention for work and heat transfer is abandoned for the Carnot
analysis! It is wCarnot which gives the availability or exergy, which we define as :
= qH,Carnot qL,Carnot .

(2.85)

Now let us require that the heat input to the Carnot engine be the heat associated with the
heat transfer from the control volume. Because of the inconsistency in sign conventions, this
requires that
qcv = qH,Carnot .

(2.86)

Now for Carnot cycles, we know that


qH,Carnot = T (s so ),
qL,Carnot = To (s so ).

(2.87)
(2.88)

So, by substituting Eqs. (2.86) and (2.88) into Eq. (2.85), the availability is
= qcv To (s so ).
Now use Eq. (2.83) to eliminate qcv from Eq. (2.89) to get


1
=
h ho + v v + g(z zo ) To (s so ),
2


1
= h To s + v v + gz (ho To so + gzo) .
2

(2.89)

(2.90)

(2.91)

So the exergy, that is, the ability to useful work, is enhanced by


high enthalpy h, which for ideal gases implies high temperature T ,
high velocity v,
high height z, and
low entropy (or high order or structure) s.
Example 2.2
Find the exergy for a CPIG.
For a CPIG, we have
h ho

s so

cP (T To ),
 
 
P
T
R ln
.
cP ln
To
Po

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.92)
(2.93)

27

2.2. EXERGY
So we get

 
 
T
1
P
= cP (T To ) To cP ln
R ln
+ v v + g(z zo ).
To
Po
2

(2.94)

For T To P Po , we can use Taylor series to simplify this somewhat. First, we recall the general
Taylor series for a log function near unity. Consider
y(x) = ln x,

(2.95)

for x 1. For a Taylor series near x = 1, we have


y(x) y(1) +
Now for y = ln x, we have



dy
1 d2 y
(x 1)2 + . . . .
(x

1)
+
dx x=1
2 dx2 x=1
1
dy
= ,
dx
x

1
d2 y
= 2,
dx2
x

(2.96)

(2.97)

and thus
y(1) = ln(1) = 0,
So


dy
= 1,
dx x=1


d2 y
= 1
dx2 x=1

1
y(x) = ln x 0 + (x 1) (x 1)2 + . . .
2

(2.98)

(2.99)

We then expand via the following steps:

=
=
=


 
1
P
T
+ RTo ln
+ v v + g(z zo ),
(2.100)
cP (T To ) cP To ln
To
Po
2


 
 
T
k1
1
T
P
cP T o
+
+ v v + g(z zo ),
1 ln
ln
(2.101)
To
To
k
Po
2
!
!





2


1 T
T
k1 P
T
1
1
1 + ... +
1 + ...
cP T o
To
To
2 To
k
Po


1
+ v v + g(z zo ),
2
!


2
!
1 T
1
k1 P
1 + ... +
1+ ...
+ v v + g(z zo ).
cP T o
2 To
k
Po
2

(2.102)
(2.103)

Note that in the neighborhood of the ambient state, relative pressure differences are more effective than
relative temperature differences at inducing high exergy.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

28

CHAPTER 2. CYCLE ANALYSIS

2.3

Rankine

2.3.1

Classical

The Rankine cycle forms the foundation for the bulk of power generating devices which
utilize steam as a working fluid. The ideal cycle is described by
1 2: isentropic pumping process in the pump,
2 3: isobaric heat transfer in the boiler,
3 4: isentropic expansion in the turbine, and
4 1: isobaric heat transfer in the condenser.
Note that to increase cycle efficiency one can
lower the condenser pressure (increases liquid water in turbine),
superheat the steam, or
increase the pressure during heat addition.
A schematic for the Rankine cycle and the associated path in the T s plane is shown
in Figure 2.3.
T

turbine

boiler

2
1

condenser

pump
1

Figure 2.3: Schematic for Rankine cycle and the associated T s plane.

Example 2.3
(adopted from Moran and Shapiro, p. 312) Consider steam in an ideal Rankine cycle. Saturated
vapor enters the turbine at 8.0 M P a. Saturated liquid exits the condenser at P = 0.008 M P a. The
net power output of the cycle is 100 M W . Find
thermal efficiency
CC BY-NC-ND. 03 August 2012, J. M. Powers.

29

2.3. RANKINE
back work ratio

mass flow rate of steam


rate of heat transfer Q in to the fluid in the boiler
rate of heat transfer Q out in the condenser

mass flow rate of condenser cooling water if the cooling water enters at 15 C and exits at 35 C.

Use the steam tables to fix the state. At the turbine inlet, one has P3 = 8.0 M P a, and x3 = 1 (saturated
steam). This gives two properties to fix the state, so that
h3 = 2758

kJ
,
kg

s3 = 5.7432

kJ
.
kg K

(2.104)

State 4 has P4 = 0.008 M P a and s4 = s3 = 5.7432 kJ/kg/K, so the state is fixed. From the saturation
tables, it is found then that

 

kJ
kJ
5.7432

0.5926
kg K
kg K
s4 sf
=
= 0.6745.
(2.105)
x4 =
kJ
sg sf
7.6361 kg K
Note the quality is 0 x4 1, as it must be. The enthalpy is then




kJ
kJ
kJ
.
+ (0.6745) 2403.1
= 1794.8
h4 = hf + x4 hf g = 173.88
kg
kg
kg

(2.106)

State 1 is saturated liquid at 0.008 M P a, so x1 = 0, P1 = 0.008 M P a. One then gets h1 = hf =


173.88 kJ/kg, v1 = vf = 0.0010084 m3 /kg.
Now state 2 is fixed by the boiler pressure and s2 = s1 . But this requires use of the sparse
compressed liquid tables. Alternatively, the pump work is easily approximated by assuming an incompressible fluid so that
h2

h2

W
= h1 + v1 (P2 P1 ),
h1 +
m


 

kJ
m3
kJ
173.88
+ 0.0010084
(8000 kP a 8 kP a) = 181.94
.
kg
kg
kg

The net power is

cycle = W
t+W
p.
W

(2.107)
(2.108)

(2.109)

Now the first law for the turbine and pump give
t
W
= h3 h4 ,
m

The energy input that is paid for is

p
W
= h1 h2 .
m

(2.110)

Q in
= h3 h2 .
m

(2.111)

The thermal efficiency is then found by

=
=

t+W
p
(h3 h4 ) + (h1 h2 )
W
=
,

h3 h2
Q
 
 
 in
kJ
2758 kJ
kg 1794.8 kg + 173.88
 

2758 kJ
kg 181.94
0.371.

(2.112)
kJ
kg
kJ
kg

181.94


kJ
kg



(2.113)
(2.114)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

30

CHAPTER 2. CYCLE ANALYSIS


By definition the back work ratio bwr is the ratio of pump work to turbine work:

bwr



W

p
=
,
t
W


h1 h2
,

=
h3 h4
 


173.88 kJ 181.94 kJ



kg
kg
 
 ,
= 
kJ


2758 kJ
kg 1794.8 kg

(2.115)
(2.116)

(2.117)

0.00837.

(2.118)

The desired mass flow can be determined since we know the desired net power. Thus

=
=

cycle
W
,
(h3 h4 ) + (h1 h2 )

2758

kJ
kg

100 103 kW
 
1794.8 kJ
kg + 173.88


(2.119)

kJ
kg


181.94

kJ
kg

kg
,
= 104.697
s


s 
kg 
kg
3600
= 3.769 105
.
=
104.697
s
hr
hr

 ,

(2.120)
(2.121)
(2.122)

The necessary heat transfer rate in the boiler is then


Q in

= m(h
3 h2 ),
 
 


kJ
kJ
kg
2758
181.94
,
=
104.697
s
kg
kg
= 269706 kW ,

(2.123)

(2.126)

269.7 M W.

(2.124)
(2.125)

In the condenser, one finds


Q out

= m(h
1 h4 ),

 
 

kg
kJ
kJ
=
104.697
173.88
1794.8
,
s
kg
kg
= 169705 kW,
=

169.7 M W.

(2.127)
(2.128)
(2.129)
(2.130)

Note also for the cycle that one should find


cycle = Q in + Q out = (269.7 M W ) (169.7 M W ) = 100 M W.
W
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.131)

31

2.3. RANKINE
For the condenser mass flow rate now perform a mass balance:
dEcv
=
{z }
| dt
=0

0 =

m
c

=
=
=
=
=

cv +m
c (hin hout ) + m(h
4 h1 ),
Q cv W
|{z}
|{z}

(2.132)

=0

=0

m
c (hin hout ) + m(h
4 h1 ),
m(h
4 h1 )

,
hin hout

 
 
kJ
104.697 kg
1794.8
s
kg 173.88

 


kJ
62.99 kJ
kg 146.68 kg
kg
,
2027.79
s 


3600 s
kg
,
2027.79
s
hr
7.3 106

(2.133)
(2.134)

kJ
kg



(2.135)
(2.136)
(2.137)

kg
.
hr

(2.138)

The enthalpy for the cooling water was found by assuming values at the saturated state at the respective
temperatures of 15 C and 35 C.

Example 2.4
Compute the exergy at various points in the flow of a Rankine cycle as considered in the previous
example problem. For that example, we had at state 1, the pump inlet that
h1 = 173.88

kJ
,
kg

s1 = 0.5926

kJ
.
kg K

(2.139)

kJ
,
kg

s2 = 0.5926

kJ
.
kg K

(2.140)

After the pump, at state 2, we have


h2 = 181.94
After the boiler, at state 3, we have
h3 = 2758

kJ
,
kg

s3 = 5.7432

kJ
.
kg K

(2.141)

After the turbine, at state 4, we have


h4 = 1794.8

kJ
,
kg

s4 = 5.7432

kJ
.
kg K

(2.142)

Now for this example, kinetic and potential energy contributions to the exergy are negligible, so we
can say in general that

= (h To s) (ho To so ).

(2.143)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

32

CHAPTER 2. CYCLE ANALYSIS


Now for this problem, we have To = 298.15 K, ho = 104.89 kJ/kg, so = 0.3674 kJ/kg/K. We have
estimated ho and so as the enthalpy and entropy of a saturated liquid at 25 C = 298.15 K.
So the exergies are as follows. At the pump entrance, we get
1

= (h1 To s1 ) (ho To so ) ,



kJ
=
173.88
(298.15 K) 0.5926
kg



kJ

104.89
(298.15 K) 0.3674
kg
= 1.84662


kJ
kg K

kJ
,
kg K

kJ
.
kg

(2.144)

(2.145)
(2.146)

After the pump, just before the boiler, we have


2

= (h2 To s2 ) (ho To so ) ,



kJ
=
181.94
(298.15 K) 0.5926
kg



kJ
(298.15 K) 0.3674

104.89
kg
= 9.90662



kJ
kg K

kJ
,
kg K

kJ
.
kg

(2.147)

(2.148)
(2.149)

So the exergy has gone up. Note here that 2 1 = h2 h1 because the process is isentropic. Here
2 1 = h2 = h1 = 181.94 173.88 = 8.06 kJ/kg.
After the boiler, just before the turbine, we have
3

= (h3 To s3 ) (ho To so ) ,




kJ
kJ
(298.15 K) 5.7432
=
2758
kg
kg K




kJ
kJ

104.89
(298.15 K) 0.3674
,
kg
kg K
= 1050.32

kJ
.
kg

(2.150)

(2.151)
(2.152)

Relative to the pump, the boiler has added much more exergy to the fluid. After the turbine, just
before the condenser, we have
4

= (h4 To s4 ) (ho To so ) ,



kJ
=
1794.8
(298.15 K) 5.7432
kg



kJ
(298.15 K) 0.3674

104.89
kg
= 87.1152

kJ
.
kg



kJ
kg K

kJ
,
kg K

(2.153)

(2.154)
(2.155)

Note that 4 3 = h4 h3 because the process is isentropic. The actual exergy (or available work)
at the exit of the turbine is relative low, even though the enthapy state at the turbine exit remains at
an elevated value.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

33

2.3. RANKINE

2.3.2

Reheat

In a Rankine cycle with reheat, the steam is extracted from an intermediate stage of the
turbine and reheated in the boiler. It is then expanded through the turbine again to the
condenser pressure. One also avoids liquid in the turbine with this strategy. This generally
results in a gain in cycle efficiency. Geometrically, the behavior on a T s diagram looks
more like a Carnot cycle. This is often covered in an introductory thermodynamics class, so
no formal example will be given here. A schematic for the Rankine cycle with reheat and
the associated T s diagram is shown in Figure 2.4.
T
3
Turbine
Boiler

4
5
6

2
1

Condenser

Pump
1

Figure 2.4: Schematic for Rankine cycle with reheat along with the relevant T s diagram.

2.3.3

Regeneration

In a Rankine cycle with regeneration, some steam is extracted from the turbine and used
to pre-heat the liquid which is exiting the pump. This can lead to an increased thermal
efficiency, all else being equal. The analysis is complicated by the need to take care of more
complex mass and energy balances in some components. A schematic for the Rankine cycle
with regeneration and open feedwater heating is shown in Figure 2.5.
Example 2.5
(BS, Ex. 11.4, pp. 438-440) Steam leaves boiler and enters turbine at 4 M P a, 400 C. After
expansion to 400 kP a, some stream is extracted for heating feedwater in an open feedwater heater.
Pressure in feedwater heater is 400 kP a, and water leaves it at a saturated state at 400 kP a. The rest
of the steam expands through the turbine to 10 kP a. Find the cycle efficiency.
1 2: compression through Pump P 1,
2 & 6 3: mixing in open feedwater heater to saturated liquid state,
3 4: compression through Pump P 2,
4 5: heating in boiler,
CC BY-NC-ND. 03 August 2012, J. M. Powers.

34

CHAPTER 2. CYCLE ANALYSIS


5

Turbine
Boiler

7
6

4
Feedwater
Heater

Condenser

Pump
P2
3

Pump
P1

Figure 2.5: Schematic for Rankine cycle with regeneration and open feedwater heating.
5 6: partial expansion in turbine,
6 7: completion of turbine expansion, and
7 1: cooling in condenser.
From the tables, one can find
h5 = 3213.6

kJ
,
kg

h6 = 2685.6

kJ
,
kg

v1 = 0.00101

h7 = 2144.1

m3
,
kg

kJ
,
kg

h1 = 191.8

v3 = 0.001084

kJ
,
kg

h3 = 604.73

m3
.
kg

kJ
,
kg

(2.156)

First consider the low pressure pump.


h2

=
=
=

h1 + v1 (P2 P1 ),

 

kJ
m3
191.8
+ 0.00101
((400 kP a) (10 kP a)) ,
kg
kg
kJ
192.194
kg

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.157)
(2.158)
(2.159)

35

2.3. RANKINE
The pump work is
= v1 (P1 P2 ),


m3
=
0.00101
((10 kP a) (400 kP a)) ,
kg
kJ
= 0.3939
.
kg

wP 1

(2.160)
(2.161)
(2.162)

Note, a sign convention consistent with work done by the fluid is used here. At this point the text
abandons this sign convention instead.
Now consider the turbine
dmcv
= m
5m
6m
7,
(2.163)
dt }
| {z
=0

m
5

1 =
dEcv
=
dt }
| {z
=0

cv
W

m
6+m
7,
m
7
m
6
+
,
m
5
m
5

(2.164)

cv + m
Q cv W
5 h5 m
6 h6 m
7 h7 ,
|{z}

(2.166)

(2.165)

=0

m
5 h5 m
6 h6 m
7 h7 .

(2.167)

On a per mass basis, we get,


wt

cv
m
6
m
7
W
= h5
h6
h7 ,
m
5
m
5
m
5


m
6
m
6
h5
h7 ,
h6 1
m
5
m
5


m
6
m
6
h7 ,
h6 1
h5 h6 + h6
m
5
m
5

 

m
6
m
6
h5 h6 + h6 1
1
h7 ,
m
5
m
5


m
6
(h6 h7 ).
h5 h6 + 1
m
5

=
=
=
=
=

(2.168)
(2.169)
(2.170)
(2.171)
(2.172)

Now consider the feedwater heater. The first law for this device gives
dEcv
=
| dt
{z }
=0

h3

=
=


kJ
=
604.73
kg
m
6
=
m
5

cv +m
2 h2 + m
6 h6 m
3 h3 ,
Q cv W
|{z}
|{z}
=0

(2.173)

=0

m
6
m
2
h2 +
h6 ,
m
3
m
3
m
7
m
6
h2 +
h6 ,
m
5
m
5


m
6
m
6
h2 +
h6 ,
1
m
5
m
5





m
6
m
6
kJ
kJ
1
+
,
192.194
2685.6
m
5
kg
m
5
kg
0.165451.

(2.174)
(2.175)
(2.176)
(2.177)
(2.178)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

36

CHAPTER 2. CYCLE ANALYSIS


Now get the turbine work


m
6
(h6 h7 ),
= h5 h6 + 1
m
5

 

kJ
kJ
=
3213.6
2685.6
kg
kg
 


kJ
kJ
2144.1
,
+ (1 0.165451)
2685.6
kg
kg
kJ
= 979.908
.
kg

wt

(2.179)

(2.180)
(2.181)

Now get the work for the high-pressure pump


wP 2

=
=
=

v3 (P3 P4 ),


m3
0.001084
((400 kP a) (4000 kP a)) ,
kg
kJ
3.9024
.
kg

(2.182)
(2.183)
(2.184)

Now
h4

h3 + v3 (P4 P3 ),
 


m3
kJ
+ 0.001084
((4000 kP a) (400 kP a)) ,
604.73
kg
kg
kJ
608.6
.
kg

=
=
=

(2.185)
(2.186)
(2.187)

Now get the net work


net
W

wnet

=
=
=
=
=

m
5 wt + m
1 wP 1 + m
5 wP 2 ,
m
1
wP 1 + wP 2 ,
wt +
m
5
m
7
wt +
wP 1 + wP 2 ,
m
5


m
6
wP 1 + wP 2 ,
wt + 1
m
5



 

kJ
kJ
kJ
979.908
+ (1 0.165451) 0.3939
3.9024
,
kg
kg
kg
kJ
.
975.677
kg

(2.188)
(2.189)
(2.190)
(2.191)
(2.192)
(2.193)

Now for the heat transfer in the boiler, one has


qh

=
=
=

h5 h4 ,

 

kJ
kJ
3213.6
608.6
,
kg
kg
kJ
2605.0
.
kg

(2.194)
(2.195)
(2.196)

Thus, the thermal efficiency is


975.677 kJ
wnet
kg
=
= 0.375.
=
qh
2605.0 kJ
kg
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.197)

37

2.4. AIR STANDARD CYCLES

This does represent an increase of efficiency over a comparable Rankine cycle without regeneration,
which happens to be 0.369.

2.3.4

Losses

Turbine: These are typically the largest losses in the system. The turbine efficiency is
defined by
wt
h3 h4
=
=
.
(2.198)
wts
h3 h4s
Here h4s and wts are the enthalpy and work the working fluid would have achieved had
the process been isentropic. Note this is for a control volume.

Pump: Pump losses are usually much smaller in magnitude than those for turbines.
The pump efficiency for a control volume is defined by
=

h2s h1
wps
=
.
wp
h2 h1

(2.199)

Piping: pressure drops via viscous and turbulent flow effects induce entropy gains
in fluid flowing through pipes. There can also be heat transfer from pipes to the
surroundings and vice versa.
Condenser: Losses are relatively small here.
Losses will always degrade the overall thermal efficiency of the cycle.

2.3.5

Cogeneration

Often steam is extracted after the boiler for alternative uses. A good example is the Notre
Dame power plant, where steam at high pressure and temperature is siphoned from the
turbines to heat the campus in winter. The analysis for such a system is similar to that for
a system with regeneration.
A schematic for cogeneration cycle is shown in Figure 2.6.

2.4

Air standard cycles

It is useful to model several real engineering devices by what is known as an air standard
cycle. This is based on the following assumptions:
CPIG,
CC BY-NC-ND. 03 August 2012, J. M. Powers.

38

CHAPTER 2. CYCLE ANALYSIS

QH

Steam
Generator
(e.g. ND power
plant)

High pressure
Low pressure
turbine
turbine

WT

6
7
Thermal
process
steam load
(e.g. heating the
ND campus)

Liquid

Pump
P2

Condenser

1 Liquid

8
2

3
Mixer

W P2

QL
(e.g. to St. Josephs Lake)

Pump
P1

W P1

Figure 2.6: Schematic for Rankine cycle with cogeneration.


no inlet or exhaust stages,
combustion process replaced by heat transfer process,
cycle completed by heat transfer to surroundings (not exhaust), and
all process internally reversible.
In some cycles, it is common to model the working fluid as a fixed mass. In others, it is
common to model the system as a control volume.

2.5

Brayton

2.5.1

Classical

The Brayton cycle is the air standard model for gas turbine engines. It is most commonly
modeled on a control volume basis. It has the following components:
1 2: isentropic compression,
CC BY-NC-ND. 03 August 2012, J. M. Powers.

39

2.5. BRAYTON
2 3: isobaric heat transfer to combustion chamber,
3 4: isentropic expansion through turbine, and
4 1: isobaric heat exchange with surroundings.

A schematic for the Brayton cycle is shown in Figure 2.7. Diagrams for P v and T s for
Fuel

Combustion
Chamber

Compressor

Turbine

Products

Air

environmental exhaust return

Figure 2.7: Schematic for Brayton cycle.


the Brayton cycle are shown in Figure 2.8.
P

T
2

3
isentrope
isobar
2

isentrope
1

4
isobar

Figure 2.8: P v and T s diagrams for the Brayton cycle.


The efficiency of the air standard Brayton cycle is found as
=

wnet
.
qH

(2.200)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

40

CHAPTER 2. CYCLE ANALYSIS

For the cycle, the first law holds that


wnet = qnet = qH qL .

(2.201)

One notes for an isobaric control volume combustor that


dEcv
cv +mh
= Q cv W
in mh
out .
|{z}
dt
| {z }
=0

=0

So

(2.202)

Q cv =
Q cv
=
m

q =
q =

m(h
out hin ),

(2.203)

hout hin ,

(2.204)

hout hin ,
cP (Tout Tin ).

(2.205)
(2.206)

So the thermal efficiency is

T1

T4
T1

qH qL
qL
cP (T4 T1 )
wnet
.
=
=1
=1
=1 
T3
qH
qH
qH
cP (T3 T2 )
T2 T2 1

(2.207)

Now because of the definition of the process, one also has


P3
P2
=
.
P4
P1
And because 1 2 and 3 4 are isentropic, one has
P2
=
P1

T2
T1

k/(k1)

So one then has

P3
=
=
P4

(2.208)

T3
T4

k/(k1)

(2.209)

T2
T3
= .
T1
T4

(2.210)

T3
T4
= .
T2
T1

(2.211)

T4
T3
1=
1.
T2
T1

(2.212)

Cross-multiplying, one finds

Subtracting unity from both sides gives

CC BY-NC-ND. 03 August 2012, J. M. Powers.

41

2.5. BRAYTON
So the thermal efficiency takes the form
=1

1
1
T1
= 1 T2 = 1  (k1)/k .
T2
P2
T1

(2.213)

P1

Note that if a Carnot cycle were operating between the same temperature bounds that
Carnot = 1 T1 /T3 would be greater than that for the Brayton cycle.
Relative to the Rankine cycle, the Brayton cycle has a large fraction of compressor work.
So the backwork ratio bwr is larger.
One must also account for deviations from ideality. These effects are summarized in
component efficiencies. The relevant efficiencies here, assuming a control volume approach,
are
h2s h1
c =
,
(2.214)
h2 h1
h3 h4
.
(2.215)
t =
h3 h4s
For a CPIG, one gets then

T2s T1
,
T2 T1
T3 T4
.
=
T3 T4s

c =

(2.216)

(2.217)

Example 2.6
(adopted from Moran and Shapiro, Example 9.6) Air enters the compressor of an air-standard
Brayton cycle at 100 kP a, 300 K with a volumetric flow rate is 5 m3 /s. The pressure ratio in the
compressor is 10. The turbine inlet temperature is 1400 K. Find the thermal efficiency, the back work
ratio and the net power. Both the compressor and turbine have efficiencies of 0.8.
First calculate the state after the compressor if the compressor were isentropic.
k/(k1)

T2s
P2
=
= 10,
(2.218)
P1
T1
T2s

=
=

T1 (10)(k1)/k ,
(300 K)(10)(1.41)/1.4 ,

(2.219)
(2.220)

579.209 K.

(2.221)

Now
c

T2

=
=
=

T2s T1
,
T2 T1
T2s T1
T1 +
,
c
(579.209 K) (300 K)
(300 K) +
,
0.8
649.012 K.

(2.222)
(2.223)
(2.224)
(2.225)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

42

CHAPTER 2. CYCLE ANALYSIS


Now state three has T3 = 1400 K. Recall that 2 3 involves heat addition in a combustion chamber.
Now calculate T4s for the ideal turbine. Recall that the expansion is to the same pressure as the
compressor inlet so that P3 /P4 = 10. The isentropic relations give

k/(k1)
T4s
P4
,
(2.226)
=
T3
P3
 (k1)/k
P4
T4s = T3
,
(2.227)
P3
 (1.41)/1.4
1
= (1400 K)
,
(2.228)
10
= 725.126 K.
(2.229)
Now account for the actual behavior in the turbine:
t

T4

=
=
=

T3 T4
,
T3 T4s
T3 t (T3 T4s ),
(1400 K) (0.8)((1400 K) (725.126 K)),
860.101 K.

(2.230)
(2.231)
(2.232)
(2.233)

Now calculate the thermal efficiency of the actual cycle.

=
=
=

qH qL
qL
cP (T4 T1 )
T4 T1
wnet
=
=1
=1
,
=1
qH
qH
qH
cP (T3 T2 )
T3 T2
860.101 K 300 K
1
,
1400 K 649.012 K
0.254181.

(2.234)
(2.235)
(2.236)

If the cycle were ideal, one would have


ideal

wnet
qH qL
qL
cP (T4s T1 )
T4s T1
=
=1
=1
,
=1
qH
qH
qH
cP (T3 T2s )
T3 T2s
(725.126 K) (300 K)
= 1
,
(1400 K) (579.209 K)
= 0.482053.
=

(2.237)
(2.238)
(2.239)

Note also that for the ideal cycle one also has
ideal = 1

300 K
T1
=1
= 0.482053.
T2s
579.209 K

(2.240)

The back work ratio is


bwr

=
=
=
=
=

wcomp
,
wturb
cP (T2 T1 )
,
cP (T3 T4 )
T2 T1
,
T3 T4
(649.012 K) (300 K)
,
(1400 K) (860.101 K)
0.646439.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.241)
(2.242)
(2.243)
(2.244)
(2.245)

43

2.5. BRAYTON
If the process were ideal, the back work ratio would have been
bwrideal

=
=
=
=
=

wcomp
,
wturb
cP (T2s T1 )
,
cP (T3 T4s )
T2s T1
,
T3 T4s
(579.209 K) (300 K)
,
(1400 K) (725.126 K)
0.413721.

(2.246)
(2.247)
(2.248)
(2.249)
(2.250)

Now get the net work. First get the mass flow rate from the volume flow rate. This requires the specific
volume.


0.287 kgkJK (300 K)
m3
RT1
=
= 0.861
.
(2.251)
v1 =
P1
100 kP a
kg
Now
m
= 1 V =

3
5 ms
V
kg
=
.
3 = 5.8072
v1
s
0.861 m
kg

(2.252)

Recall cP = kcv = 1.4(717.5 J/kg/K) = 1004.5 J/kg/K = 1.0045 kJ/kg/K. Now


cycle
W

=
=
=

mc
P ((T3 T4 ) (T2 T1 ),
(2.253)



kJ
kg
1.0045
((1400 K) (860.101 K) (649.012 K) + (300 K)),(2.254)
5.8072
s
kg K
1113.51 kW.

(2.255)

If the cycle were ideal, one would have


ideal
W

=
=
=

mc
P ((T3 T4s ) (T2s T1 ),
(2.256)



kJ
kg
1.0045
((1400 K) (725.126 K) (579.209 K) + (300 K)),(2.257)
5.8072
s
kg K
2308.04 kW.
(2.258)

Note that the inefficiencies had a significant effect on both the back work ratio and the net work of the
engine.

2.5.2

Regeneration

One can improve cycle efficiency by regeneration. The hot gas in the turbine is used to
preheat the gas exiting the compressor before it enters the combustion chamber. In this
cycle, one takes
1 2: compression,
2 x: compression exit gas goes through regenerator (heat exchanger),
CC BY-NC-ND. 03 August 2012, J. M. Powers.

44

CHAPTER 2. CYCLE ANALYSIS


x 3: combustion chamber,
3 4: expansion in turbine, and
4 y: turbine exit gas goes through regenerator (heat exchanger).

A schematic for the Brayton cycle is shown in Figure 2.9.


Regenerator

x
2

Combustion
chamber

4
3

Compressor

Turbine

Figure 2.9: Schematic for Brayton cycle with regeneration.


For this to work, the gas at the turbine exit must have a higher temperature than the gas
at the compressor exit. If the compression ratio is high, the compressor exit temperature is
high, and there is little benefit to regeneration.
Now consider the regenerator, which is really a heat exchanger. The first law holds that
dEcv
cv +mh
= Q cv W
2 mh
x + mh
4 mh
y,
|{z}
|{z}
dt
| {z }
=0

=0

(2.259)

=0

0 = h2 hx + h4 hy ,
= cP (T2 Tx ) + cP (T4 Ty ),
= T2 Tx + T4 Ty .

(2.260)
(2.261)
(2.262)

Now consider the inlet temperatures to be known. Then the first law constrains the outlet
temperatures such that
T2 + T4 = Tx + Ty .

(2.263)

If the heat exchanger were an ideal co-flow heat exchanger, one might expect the outlet
temperatures to be the same, that is, Tx = Ty , and one would have Tx = Ty = (1/2)(T2 + T4 ).
But in fact one can do better. If a counter-flow heat exchanger is used, one could ideally
expect to find
Tx = T4 ,
CC BY-NC-ND. 03 August 2012, J. M. Powers.

Ty = T2 .

(2.264)

45

2.5. BRAYTON
The thermal efficiency is
=
=
=
=
=
=
=

wnet
,
qH
wt wc
,
qH
cP (T3 T4 ) cP (T2 T1 )
,
cP (T3 Tx )
(T3 T4 ) (T2 T1 )
,
T3 Tx
(T3 T4 ) (T2 T1 )
,
T3 T4
T2 T1
1
,
T3 T4


T2
T1 T1 1
,
1 
T3 1 TT34

 
(k1)/k
P2
1
T1
P1
1 
 (k1)/k  ,
T3 1 PP43

 
(k1)/k
P2
1
T1
P1
1 
 (k1)/k  ,
T3 1 PP12
T1
T3

T1
1
T3

= 1

 (k1)/k
(k1)/k 1 P1
P2
P2
 (k1)/k ,
P1
1 PP12

P2
P1

(k1)/k

(2.265)
(2.266)
(2.267)
(2.268)
(2.269)
(2.270)
(2.271)

(2.272)

(2.273)

(2.274)

(2.275)

Note this is the Carnot efficiency moderated by the pressure ratio. As the pressure ratio
rises, the thermal efficiency declines.
The regenerator itself will not be perfect. To achieve the performance postulated here
would require infinite time or an infinite area heat exchanger. As both increase, viscous
losses increase, and so there is a trade-off. One can summarize the behavior of an actual
regenerator by an efficiency defined as
reg =

actual
hx h2
=
.

hx h2
ideal

(2.276)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

46

CHAPTER 2. CYCLE ANALYSIS

If we have a CPIG, then


reg =

Tx T2
.
Tx T2

(2.277)

Example 2.7
(extension for Moran and Shapiro, Example 9.6) How would the addition of a regenerator affect
the thermal efficiency of the isentropic version of the previous example problem?
One may take the rash step of trusting the analysis to give a prediction from Eq. (2.275) of

=
=

 (k1)/k
T1 P2
1
,
T3 P1


300 K
(10)(1.41)/1.4 = 0.586279.
1
1400 K

(2.278)
(2.279)

Without regeneration, the thermal efficiency of the ideal Brayton cycle had a value of 0.482053. Had
the engine used a Carnot cycle between the same temperature limits, the efficiency would have been
0.785714.

2.5.3

Ericsson Cycle

If one used isothermal compression and expansion, which is slow and impractical, in place
of isentropic processes in the Brayton cycle, one would obtain the Ericsson cycle. Diagrams
for P v and T s for the Ericsson cycle are shown in Figure 2.10.
P

T
2

isotherm
isobar

isobar

isotherm
1

Figure 2.10: P v and T s diagram for the Ericsson cycle.


One can outline the Ericsson cycle as follows:
CC BY-NC-ND. 03 August 2012, J. M. Powers.

47

2.5. BRAYTON
1 2: isothermal compression,
2 3: isobaric heating in the combustion chamber,
3 4: isothermal expansion in turbine, and
4 1: isobaric heat transfer in surroundings.

One also takes the control volume approach.


First consider the work in the isothermal compressor. For the control volume, one recalls
that the work is given by the enthalpy difference, and that dh = T ds + vdP = qc wc ,
with qc = T ds and wc = vdP , so
wc =

v dP = RT1

dP
P2
= RT1 ln .
P
P1

(2.280)

Note that P2 /P1 > 1, so wc < 0; work is done on the fluid in the compressor.
For the turbine, again for the control volume, the key is the enthalpy difference and
dh = T ds + vdP = qt wt , with qt = T ds and wt = vdP . The turbine work would be
wt =

4
3

v dP = RT3

4
3

P4
dP
= RT3 ln .
P
P3

(2.281)

But because P2 = P3 and P1 = P4 , the turbine work is also


wt = RT3 ln

P1
P2
= RT3 ln .
P2
P1

(2.282)

Note that because P2 /P1 > 1, the turbine work is positive.


In the combustion chamber, one has from the first law
h3 h2 = qC
cP (T3 T2 ) = qC .

(2.283)
(2.284)

For the isothermal turbine, one has dh = cP dT = 0 = qt wt , so qt = wt . The heat


transfer necessary to keep the turbine isothermal is
qt = wt = RT3 ln

P2
.
P1

(2.285)

So the total heat that one pays for is


qH = qC + qt = cP (T3 T1 ) + RT3 ln

P2
.
P1

(2.286)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

48

CHAPTER 2. CYCLE ANALYSIS

So the thermal efficiency is


=

wt + wc
,
qH
R(T3 T1 ) ln PP12

(2.287)

,
cP (T3 T1 ) + RT3 ln PP21


T1
1 T3 ln PP12


=
,
P2
T1
k
+
ln
1

k1
T3
P1

  (k1)/k
1 TT31 ln PP12
= 

 (k1)/k ,
T1
1 T3 + ln PP21



T1

=
1

T3

1+

1 T1
3
(k1)/k
P
ln P2
1



T1
1 T3
T1

1
+ ....
1  (k1)/k
T3
ln PP12
| {z }
Carnot
{z
}
|

(2.288)

(2.289)

(2.290)

(2.291)

(2.292)

correction

This is expressed as a Carnot efficiency modified by a correction which degrades the efficiency.
Note the efficiency is less than that of a Carnot cycle. For high temperature ratios and high
pressure ratios, the efficiency approaches that of a Carnot engine.
Now if one used multiple staged intercooling on the compressors and multiple staged
expansions with reheat on the turbines, one can come closer to the isothermal limit, and
better approximate the Carnot cycle.

2.5.4

Jet propulsion

If one modifies the Brayton cycle so that the turbine work is just sufficient to drive the
compressor and the remaining enthalpy at the turbine exit is utilized in expansion in a
nozzle to generate thrust, one has the framework for jet propulsion. Here the control volume
approach is used. A schematic for a jet engine is shown in Figure 2.11. Diagrams for P v
and T s for the jet propulsion Brayton cycle are shown in Figure 2.12. An important
component of jet propulsion analysis is the kinetic energy of the flow. In the entrance region
of the engine, the flow is decelerated, inducing a ram compression effect. For high velocity
applications, this effect provides sufficient compression for the cycle and no compressor or
turbine are needed! However, such a device, known as a ramjet, is not self-starting, and so
is not practical for many applications.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

49

2.5. BRAYTON

Burner

2
1
Diffuser

a
Compressor

Turbine

4
Nozzle

environmental exhaust return

Figure 2.11: Schematic for jet propulsion.


P

T
3
2

4
isobar
isentrope
4

isentrope

2
5

a
1

5
v

isobar

Figure 2.12: P v and T s diagrams for jet propulsion Brayton cycle.


One can outline the ideal jet propulsion cycle as follows:
1 a: deceleration and ram compression in diffuser,
a 2: isentropic compression in the compressor,
2 3: isobaric heating in the combustion chamber,
3 4: isentropic expansion in turbine,
4 5: isentropic conversion of thermal energy to kinetic energy in nozzle, and
5 1: isobaric heat transfer in surroundings.
The goal of the jet propulsion cycle is to produce a thrust force which is necessary to
balance a fluid-induced aerodynamic drag force. When such a balance exists, the system is
in steady state. One can analyze such a system in the reference frame in which the engine
is stationary. In such a frame, Newtons second law gives
d
(mCV vCV ) =
}
|dt {z
=0

F
|{z}

thrust force

1 mv
5.
+ P1 A1 i P5 A5 i +mv
{z
}
|

(2.293)

small

CC BY-NC-ND. 03 August 2012, J. M. Powers.

50

CHAPTER 2. CYCLE ANALYSIS

Solving for the thrust force, neglecting the small differences in pressure, one gets,
F=m
(v5 v1 ) .

(2.294)

The work done by the thrust force is the product of this force and the air speed, v1 . This
gives
p | = |F v1 | = |m(v
|W
5 v1 ) v1 | .
(2.295)
Now the efficiency of the cycle is a bit different. The net work of the turbine and compressor
is zero. Instead, the propulsive efficiency is defined as
p|
Propulsive Power
|W
=
,
Energy Input Rate
|Q H |

p =

(2.296)

|mv
1 (v5 v1 )|
.
m|(h

3 h2 )|

(2.297)

Note the following unusual behavior for flight in an ideal inviscid atmosphere in which
the flow always remains attached. In such a flow DAlemberts paradox holds: there is
no aerodynamic drag. Consequently there is no need for thrust generation in steady state
operation. Thrust would only be needed to accelerate to a particular velocity. For such
an engine then the exit velocity would equal the entrance velocity: v5 = v1 and F = 0.
Moreover, the propulsive efficiency is p = 0.
Example 2.8

(adopted from C
engal and Boles, p. 485). A turbojet flies with a velocity of 850 fst at an altitude
where the air is at 5 psia and 40 F . The compressor has a pressure ratio of 10, and the temperature
at the turbine inlet is 2000 F . Air enters the compressor at a rate of m
= 100 lbm
s . Assuming an air
standard with CPIG air, find the

temperature and pressure at the turbine exit,


velocity of gas at nozzle exit, and
propulsive efficiency.
For the English units of this problem, one recalls that
1 Btu = 777.5 f t lbf,

gc = 32.2

lbm f t
,
lbf s2

Btu
f t2
= 25037 2 .
lbm
s

(2.298)

For air, one has


cP = 0.240

Btu
,
lbm R

R = 53.34

f t lbf
f t2
Btu
= 1717.5 2 = 0.0686
,

lbm R
s R
lbm R

k = 1.4.

(2.299)

First one must analyze the ram compression process. An energy balance gives
h1 +

va va
v1 v1
= ha +
.
2
2 }
| {z
small

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.300)

51

2.5. BRAYTON

It is usually sufficient to neglect the kinetic energy of the fluid inside the engine. While it still has
a non-zero velocity, it has been slowed enough so that enthalpy dominates kinetic energy inside the
engine. So one takes
v1 v1
,
2
v1 v1
= h1 ha +
,
2
v1 v1
,
= cP (T1 Ta ) +
2
v1 v1
= T1 +
.
2cP

ha

(2.301)

= h1 +

0
0
Ta

(2.302)
(2.303)
(2.304)

The ambient temperature is


T1 = 40 + 460 = 420 R.

(2.305)

So after ram compression, the temperature at the inlet to the compressor is

Ta = (420 R) +


850

2 0.240

ft
s

2

Btu
lbm R

Btu
lbm
2
25037 fst2

= 480 R.

(2.306)

Now get the pressure at the compressor inlet via the isentropic relations:
Pa = P1

Ta
T1

k/(k1)

= (5 psia)

480 R
420 R

1.4/(1.41)

= 8.0 psia.

(2.307)

Now consider the isentropic compressor.


P2 = 10Pa = 10(8.0 psia) = 80 psia.

(2.308)

Now get the temperature after passage through the compressor.


T2 = Ta

P2
Pa

(k1)/k

(1.41)/1.4

= (480 R) (10)

= 927 R.

(2.309)

The temperature at the entrance of the turbine is known to be 2000 F = 2460 R. Now the turbine
work is equal to the compressor work, so
wc

= wt ,

(2.310)

h2 ha = h3 h4 ,
cP (T2 Ta ) = cP (T3 T4 ),
T2 Ta

T4

(2.311)
(2.312)

= T3 T4 ,

(2.313)

= T3 T2 + Ta ,
= 2460 R 927 R + 480 R,

(2.314)
(2.315)

(2.316)

2013 R.

Use the isentropic relations to get P4 :


P4 = P3

T4
T3

k/(k1)

= (80 psia)

2013 R
2460 R

1.4/(1.41)

= 39.7 psia.

(2.317)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

52

CHAPTER 2. CYCLE ANALYSIS


Expansion through the nozzle completes the cycle. Assume an isentropic nozzle, therefore

T5

= T4
= T4

P5
P4
P1
P4

(k1)/k
(k1)/k

= = (2013 R)
= 1114 R.

(2.318)

(2.319)

5 psia
39.7 psia

(1.41)/1.4

(2.320)
(2.321)

Now consider the energy balance in the nozzle:


h5 +

v5 v5
2

h4 +

v4 v4
,
2 }
| {z

(2.322)

small

h5 h4 +

v5 v5
cP (T5 T4 ) +
,
2
p
2cP (T4 T5 ),
v
u 

u
t2 0.240 Btu
(2013 R 1114 R)
lbm R

=
v5

v52
,
2

=
=

3288

(2.323)
(2.324)
(2.325)
f t2

25037 s2
1 Btu
lbm

ft
.
s

(2.326)

(2.327)

Now find the thrust force magnitude

|F| =
=
=

m(v
5 v1 ) =
f t lbm
s2
lbm f t
lbf s2

243800
32.2




lbm
ft
ft
f t lbm
100
850
,
3288
= 243800
s
s
s
s2

(2.328)

(2.329)

7571.4 lbf.

(2.330)

The power is the product of the thrust force and the air speed, so
P| =
|W
=
=
=




Btu
ft
,
|F v1 | = (7571.4 lbf ) 850
s
777.5 f t lbf
Btu
8277
,
s
!


1.415 hp
Btu
,
8277
Btu
s
s

(2.332)

11711 hp.

(2.334)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.331)

(2.333)

2.6. RECIPROCATING ENGINE POWER CYCLES

53

Now for the heat transfer in the combustion chamber


Q H

= m(h
3 h2 ),
= mc
P (T3 T2 ),



Btu
lbm
0.240
(2460 R 927 R),
=
100
s
lbm R
Btu
= 36792
.
s

(2.335)
(2.336)
(2.337)
(2.338)

So the propulsion efficiency is


p =

P
8277 Btu
W
s
=
= 0.225.
36792 Btu
Q H
s

(2.339)

The remainder of the energy is excess kinetic energy and excess thermal energy, both of which ultimately
dissipate so as to heat the atmosphere.

Other variants of the turbojet engine include the very important turbofan engine in which
a large cowling is added to the engine and a an additional fan in front of the compressor
forces a large fraction of the air to bypass the engine. The turbofan engine is used in most
large passenger jets. Analysis reveals a significant increase in propulsive efficiency as well as
a reduction in jet noise. Other important variants include the turboprop, propfan, ramjet,
scramjet, jet with afterburners, and rocket.

2.6

Reciprocating engine power cycles

Here are some common notions for engines that depend on pistons driving in cylinders. The
piston has a bore diameter B. The piston is connected to the crankshaft, and the stroke S
of the piston is twice the radius of the crank, Rcrank :
S = 2Rcrank .

(2.340)

See Figure 2.13 for an illustration of this geometry for a piston-cylinder arrangement in two
configurations. The total displacement for all the cylinders is
Vdispl = Ncyl (Vmax Vmin ) = Ncyl Acyl S.

(2.341)

Note that

B2
.
4
The compression ratio is not the ratio of pressures; it is the ratio of volumes:
Acyl =

rv = CR =

Vmax
.
Vmin

(2.342)

(2.343)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

54

CHAPTER 2. CYCLE ANALYSIS

2Rcrank

Figure 2.13: Piston-cylinder configurations illustrating geometric definitions.


For these engines the volumes are closed in expansion and compression, so the net work is
I
wnet = P dv = Pmef f (vmax vmin ).
(2.344)
Here one has defined a mean effective pressure: Pmef f . The net work per cylinder per cycle
is
Wnet = mwnet = Pmef f m(vmax vmin ) = Pmef f (Vmax Vmin ).
(2.345)
Now consider the total power developed for all the cylinders. Assume the piston operates at
a frequency of cycles/s:
net = Ncyl Pmef f (Vmax Vmin ).
W
It is more common to express in revolutions per minute: RP M = (60
net = Pmef f Ncyl (Vmax Vmin ) RP M ,
W
s
{z
} 60 min
|

(2.346)
s
),
min

so
(2.347)

=Vdispl

= Pmef f Vdispl

RP M
s .
60 min

(2.348)

This applies for a two-stroke engine. If the engine is a four stroke engine, the net power is
reduced by a factor of 1/2.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

55

2.7. OTTO

2.7

Otto

The air-standard Otto cycle approximates the gasoline engine. It employs a fixed mass
approach.
Diagrams for P v and T s for the Otto cycle are shown in Figure 2.14.
P

T
3

isochore

isentrope
2
2

isentrope

isochore

Figure 2.14: P v and T s diagrams for the Otto cycle.


One can outline the Otto cycle as follows:
1 2: isentropic compression in the compression stroke,
2 3: isochoric heating in the combustion stroke during spark ignition,
3 4: isentropic expansion in power stroke, and
4 1: isochoric rejection of heat to the surroundings.
Note for isochoric heating, such as 2 3, in a fixed mass environment, the first law gives
u3 u2 =
u3 u2
u3 u2

2 w3 ,
Z v3
P dv,
= 2 q3
v2
Z v2
P dv ,
= 2 q3
v2
| {z }
2 q3

(2.349)
but v2 = v3 ,

(2.350)
(2.351)

=0

2 q3
2 q3

= u3 u2 ,
= cv (T3 T2 ),

if CPIG.

(2.352)
(2.353)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

56

CHAPTER 2. CYCLE ANALYSIS


The thermal efficiency is found as follows:
wnet
=
,
qH
qH qL
,
=
qH
qL
= 1
,
qH
cv (T4 T1 )
= 1
,
cv (T3 T2 )
T4 T1
,
= 1
T3 T2


T1 TT41 1
.
= 1 
T2 TT32 1

Now one also has the isentropic relations:


 k1
T2
V1
=
,
T1
V2
 k1
V4
T3
.
=
T4
V3

(2.354)
(2.355)
(2.356)
(2.357)
(2.358)
(2.359)

(2.360)
(2.361)

But V4 = V1 and V2 = V3 , so

 k1
T2
V1
T3
= .
=
T4
V2
T1
Cross multiplying the temperatures, one finds

(2.362)

T4
T3
= .
T2
T1

(2.363)

Thus the thermal efficiency reduces to


=1
In terms of the compression ratio rv =

V1
,
V2

T1
.
T2

(2.364)

one has

= 1 rv1k = 1

1
rvk1

(2.365)

Note if the compression ratio increases, the thermal efficiency increases. High compression
ratios introduce detonation in the fuel air mixture. This induces strong pressure waves in
the cylinder and subsequent engine knock. It can cause degradation of piston walls.
Some deviations of actual performance from that of the air-standard Otto cycle are as
follows:
CC BY-NC-ND. 03 August 2012, J. M. Powers.

57

2.7. OTTO
specific heats actually vary with temperature,
combustion may be incomplete (induces pollution and lowers fuel efficiency),
work of inlet and exhaust ignored, and
losses of heat transfer to engine walls ignored.

Example 2.9
(adopted from Moran and Shapiro, p. 363). The temperature at the beginning of the compression
process of an air-standard Otto cycle with a compression ratio of 8 is 540 R, the pressure is 1 atm,
and the cylinder volume is 0.02 f t3 . The maximum temperature is 3600 R. Find
temperature and pressure at each stage of the process,
thermal efficiency, and
Pmef f in atm.
For the isentropic compression,
T2

V1
V2

k1

T1

(540 R)(8)1.41 ,

(2.366)
(2.367)

1240.69 R.

(2.368)

One can use the ideal gas law to get the pressure at state 2:
P2 V2
T2

P2

=
=
=

P1 V1
,
T1
V1 T2
P1
,
V2 T1
(1 atm)(8)

(2.369)
(2.370)


1240.69 R
540 R

(2.371)

18.3792 atm.

(2.372)

Now V3 = V2 since the combustion is isochoric. And the maximum temperature is T3 = 3600 R. This
allows use of the ideal gas law to get P3 :
P3 V3
T3

P3

P2 V2
,
T2
V2 T3
P2
,
V3 T2
|{z}

(2.373)
(2.374)

=1

=
=

(18.3792 atm)(1)
53.3333 atm.

3600 R
1240.59 R

(2.375)
(2.376)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

58

CHAPTER 2. CYCLE ANALYSIS


One uses the isentropic relations for state 4:

T3
T4

T4
T3




V4
V3
V3
V4


k1
k1
V2
V1

(2.377)

(2.378)

k1

(2.379)

(2.380)

T4

T3

T4

T3

(3600 R)(8)11.4 ,

(2.381)

(2.382)

V1
V2

1k

1566.99 R.

For the pressure at state 4, use the ideal gas law:

P4 V4
T4

P4

=
=
=
=

P3 V3
,
T3
V3 T4
P3
,
V4 T3
V2 T4
P3
,
V1 T3

 
1 1566.99 R
,
(53.3333 atm)
8
3600 R
2.90184 atm.

(2.383)
(2.384)
(2.385)
(2.386)
(2.387)

The thermal efficiency is

1
,
rvk1
1
= 1 1.41 ,
8
= 0.564725.

= 1

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.388)
(2.389)
(2.390)

59

2.8. DIESEL
Now get the mean effective pressure Pmef f .
Wnet

Pmef f

=
=
=
=

Pmef f (Vmax Vmin ),


Wnet
,
Vmax Vmin
m(u3 u4 ) + m(u1 u2 )
,
V1 V2
cv T 3 T 4 + T 1 T 2
m
,
V1
1 VV21

(2.391)
(2.392)
(2.393)
(2.394)

P1 V1 cv T3 T4 + T1 T2
,
RT1 V1
1 VV2

(2.395)

P1 cv T3 T4 + T1 T2
,
T1 R
1 VV2

(2.396)

P1 1 T3 T4 + T1 T2
,
T1 k 1
1 VV2

(2.397)

2.8

1 atm
1 3600 R 1566.99 R + 540 R 1240.59 R
,

540 R 1.4 1
1 18

(2.398)

7.04981 atm.

(2.399)

Diesel

The air standard Diesel cycle approximates the behavior of a Diesel engine. It is modeled
as a fixed mass system. Here the compression is done before injection, so there is no danger
of premature ignition due to detonation. No spark plugs are used. Diagrams for P v and
T s for the Diesel cycle are shown in Figure 2.15. One can outline the Diesel cycle as
follows:
1 2: isentropic compression in the compression stroke,
2 3: isobaric heating in the combustion stroke,
3 4: isentropic expansion in power stroke, and
4 1: isochoric rejection of heat to the surroundings.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

60

CHAPTER 2. CYCLE ANALYSIS


P

3
3
isobar
isentrope

2
4

isentrope

isochore

1
v

Figure 2.15: P v and T s diagrams for the Diesel cycle.


The thermal efficiency is found as follows:
=
=
=
=
=
=

wnet
,
qH
qH qL
,
qH
qL
1
,
qH
u4 u1
1
,
h3 h2
cv (T4 T1 )
,
1
cP (T3 T2 )
1

1 T1
k T2

T4
T1
T3
T2

1
1

(2.400)
(2.401)
(2.402)
(2.403)
(2.404)
(2.405)

All else being equal, the Otto cycle will have higher efficiency than the Diesel cycle.
However, the Diesel can operate at higher compression ratios because detonation is not as
serious a problem in the compression ignition engine as it is in the spark ignition.
Example 2.10
(adopted from BS, Ex. 12.8 pp. 501-502). An air standard Diesel cycle has a compression ratio of
20, and the heat transferred to the working fluid has 1800 kJ/kg. Take air to be an ideal gas with
variable specific heat. At the beginning of the compression process, P1 = 0.1 M P a, and T1 = 15 C.
Find
Pressure and temperature at each point in the process,
Thermal efficiency,
Pmef f , the mean effective pressure.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

61

2.8. DIESEL

The enthalpies and entropies at the reference pressure for variable specific heat air are tabulated
as functions of temperature in Table A7.1. Recall that
Z T
h(T ) = href +
cP (T) dT.
(2.406)
Tref

Recall further that


T ds
ds

= dh vdP,
v
dh
dP.
=
T
T

(2.407)
(2.408)

Now for an ideal gas it can be shown that dh = cP (T )dT ; one also has v/T = R/P . Making these
substitutions into Eq. (2.408) gives
ds

cP (T )dT
dP
R
.
T
P

(2.409)

(2.410)

Integrating gives
s(T, P ) = sref +
|

P
cP (T)
,
dT R ln

P
ref
T
Tref
{z
}
T

=so
T

s(T, P ) = soT (T ) R ln

P
.
Pref

(2.411)

Here the reference pressure is Pref = 0.1 M P a. Note that this happens to be the inlet state, which
is a coincidence; so at the inlet there is no correction to the entropy for pressure deviation from its
reference value. At the inlet, one has T1 = 15 + 273.15 = 288.15 K. One then gets from interpolating
Table A7.1 that
h1 = 288.422

kJ
,
kg

soT1 = s1 = 6.82816

kJ
,
kg K

u1 = 205.756

kJ
.
kg

(2.412)

For the isentropic compression, one has


s2 = s1 = 6.82816

kJ
,
kg K

V2 =

1
V1 .
20

(2.413)

Now from the ideal gas law, one has


P2 V2
T2
P2
P1
P2
Pref

=
=
=

P1 V1
,
T1
T2 V1
,
T1 V2
T2
20
.
288 K

(2.414)
(2.415)
(2.416)

Now
s2 (T2 , P2 ) =


kJ
=
6.82816
kg K
0 =

P2
,
(2.417)
soT2 R ln
Pref

 

kJ
T2
soT2 0.287
ln 20
,
(2.418)
kg K
288 K

 
 

kJ
T2
kJ
soT2 0.287
ln 20
6.82816
F (T2 ). (2.419)
kg K
288 K
kg K
CC BY-NC-ND. 03 August 2012, J. M. Powers.

62

CHAPTER 2. CYCLE ANALYSIS


T2 (K) soT2

kJ
kg K

F (T2 )

kJ
kg K
4

8.01581
8.57198 10
8.13493
8.97387 102
7.88514 9.60091 102
7.95207 4.64783 102
8.01237 2.37449 103

900
1000
800
850
899.347

Table 2.3: Iteration for T2 .


This then becomes a trial and error search in Table A7.1 for the T2 which satisfies the previous equation. If one guesses T2 900 K, one gets soT2 = 8.01581 kJ/kg/K, and the right side evaluates to
0.000857198 kJ/kg/K. Performing a trial and error procedure, one finds the results summarized in
Table 2.3. Take T2 = 900 K as close enough!
Then
P2

T2 V1
,
T1 V2

P1

(0.1 M P a)

(2.420)

900 K
(20),
288 K
6.25 M P a.

(2.421)
(2.422)

Now at state 2, T2 = 900 K, one finds that


h2 = 933.15

kJ
,
kg

u2 = 674.82

kJ
.
kg

(2.423)

The heat is added at constant pressure, so


qH
h3

= h3 h2 ,
= h2 + qH ,

 

kJ
kJ
=
933.15
+ 1800
,
kg
kg
kJ
= 2733.15
.
kg

One can then interpolate Table A7.1 to find T3 and soT3


 


kJ

2692.31
2733.15 kJ
kg
kg
 
 ((2400 K) (2350 K)),
T3 = (2350 K) + 
kJ
kJ
2755 kg 2692.31 kg
=

soT3

2382.17 K,


kJ
=
9.16913
kg K

 

kJ

 

2733.15 kJ
kg 2692.31 kg
kJ
kJ
 

9.19586
9.16913
,
+ 
kJ
kg K
kg K
2755 kJ
kg 2692.31 kg
= 9.18633

kJ
.
kg K

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.424)
(2.425)
(2.426)
(2.427)

(2.428)
(2.429)

(2.430)
(2.431)

63

2.8. DIESEL
T4 (K)

soT4

400
800
1400
1200
1300
1250
1288

kJ
kg K

7.15926
7.88514
8.52891
8.34596
8.44046
8.39402
8.42931

F (T4 )

kJ
kg K
1

9.34561 10
4.07614 101
7.55464 102
6.31624 102
8.36535 103
2.68183 102
1.23117 104

Table 2.4: Iteration for T4 .


One also has P3 = P2 = 6.25 M P a. Now get the actual entropy at state 3:
s3 (T3 , P3 ) =
=
=

P3
soT3 R ln
,
Pref
 
 


kJ
6.25 M P a
kJ
0.287
ln
,
9.18633
kg K
kg K
0.1 M P a
kJ
7.99954
.
kg K

(2.432)
(2.433)
(2.434)

Now 3 4 is an isentropic expansion to state 4, which has the same volume as state 1; V1 = V4 . So
the ideal gas law gives
P4 V4
T4
P4
P1
P4
Pref

=
=
=

P1 V1
,
T1
T4 V1
,
T1 V4
T4 V1
.
T1 V4

(2.435)
(2.436)
(2.437)

Now consider the entropy at state 4, which must be the same as that at state 3:
s4 (T4 , P4 ) = soT4 R ln

P4
,
Pref

(2.438)

T4 V1

= soT4 R ln
T1 V4 ,
|{z}
=1


 


kJ
T4
kJ
o
= sT4 0.287
ln
,
7.99954
kg K
kg K
288 K

 
 

kJ
kJ
T4
o
0 = sT4 0.287
ln
7.99954
= F (T4 ).
kg K
288 K
kg K
s3

(2.439)

(2.440)
(2.441)

Using Table A7.1, this equation can be iterated until T4 is found. So T4 = 1288 K. At this temperature,
one has
kJ
kJ
,
u4 = 1011.98
.
(2.442)
h4 = 1381.68
kg
kg
CC BY-NC-ND. 03 August 2012, J. M. Powers.

64

CHAPTER 2. CYCLE ANALYSIS


Now one can calculate the cycle efficiency:
u4 u1
= 1
,
h3 h2
 


kJ
1011.98 kJ
kg 205.756 kg
 
,
= 1 
kJ
2733.15 kJ

933.15
kg
kg

(2.445)

 


kJ
kJ
kJ
1011.98
= 806.224
.
qL = u1 u4 = 205.756
kg
kg
kg

So,
wnet
So

 


kJ
kJ
kJ
806.224
= 993.776
.
= qH + qL = 1800
kg
kg
kg
Pmef f

=
=
=
=
=
=

2.9

(2.444)

0.552098.

Now one has

(2.443)

Wnet
,
Vmax Vmin
wnet
,
vmax vmin
wnet
,
v1 v2
wnet
,
RT2
RT1
P1 P2


0.287

(2.447)

(2.448)
(2.449)
(2.450)
(2.451)

993.776


kJ
kg K

(2.446)

kJ
kg

288 K
100 kP a

900 K
6250 kP a

1265.58 kP a.

,

(2.452)
(2.453)

Stirling

Another often-studied air-standard engine is given by the Stirling cycle. This is similar to
the Otto cycle except the adiabatic processes are replaced by isothermal ones. The efficiency
can be shown to be equal to that of a Carnot engine. Stirling engines are difficult to build.
Diagrams for P v and T s for the Stirling cycle are shown in Figure 2.16.
One can outline the Stirling cycle as follows:
1 2: isothermal compression in the compression stroke,
2 3: isochoric heat transfer in the combustion stroke,
3 4: expansion in power stroke, and
4 1: isochoric rejection of heat to the surroundings.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

65

2.10. REFRIGERATION
P

T
3
3

isotherm

isochore
isochore

isotherm

4
2

Figure 2.16: P v and T s diagrams for the Stirling cycle.

2.10

Refrigeration

A simple way to think of a refrigerator is a cyclic heat engine operating in reverse. Rather
than extracting work from heat transfer from a high temperature source and rejecting heat
to a low temperature source, the refrigerator takes a work input to move heat from a low
temperature source to a high temperature source.

2.10.1

Vapor-compression

A common refrigerator is based on a vapor-compression cycle. This is a Rankine cycle in


reverse. While one could employ a turbine to extract some work, it is often impractical.
Instead the high pressure gas is simply irreversibly throttled down to low low pressure.
One can outline the vapor-compression refrigeration cycle as follows:
1 2: isentropic compression
2 3: isobaric heat transfer to high temperature reservoir in condenser,
3 4: adiabatic expansion in throttling valve, and
4 1: isobaric (and often isothermal) heat transfer to low temperature reservoir in
evaporator.
A schematic for the vapor-compression refrigeration cycle is shown in Figure 2.17. A T s
diagram for the vapor-compression refrigeration cycle is shown in Figure 2.18.
The efficiency does not make sense for a refrigerator as 0 1. Instead a coefficient
CC BY-NC-ND. 03 August 2012, J. M. Powers.

66

CHAPTER 2. CYCLE ANALYSIS

Condenser

Compressor
Expansion
valve

Compressor

Evaporator
1

Figure 2.17: Schematic diagram for the vapor-compression refrigeration cycle.


of performance, , is defined as
what one wants
,
what one pays for
qL
=
.
wc

(2.454)
(2.455)

Note that a heat pump is effectively the same as a refrigerator, except one desires qH
rather than qL . So for a heat pump, the coefficient of performance, , is defined as
=

qH
.
wc

(2.456)

It is possible for both and to be greater than unity.


Example 2.11
(from Moran and Shapiro, p. 442) R 12 is the working fluid in an ideal vapor-compression
refrigeration cycle that communicates thermally with a cold region at 20 C and a warm region at
40 C. Saturated vapor enters the compressor at 20 C and saturated liquid leaves the condenser at
40 C. The mass flow rate of the refrigerant is 0.008 kg/s. Find
compressor power in kW ,
refrigeration capacity in ton,
coefficient of performance, and
coefficient of performance of an equivalent Carnot cycle.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

67

2.10. REFRIGERATION

T
2
3

1
s

Figure 2.18: T s diagram for the vapor-compression refrigeration cycle.


At compressor inlet for the ideal cycle, one has a saturated vapor.
T1 = 20 C,

x1 = 1.

(2.457)

The tables then give


h1 = 195.78

kJ
,
kg

s1 = 0.6884

kJ
,
kg K

P1 = 5.6729 bar.

(2.458)

Now at the end of the condenser (state 3), one has T3 = 40 C. The condenser is on an isobar, so the
pressure is the saturation pressure at the temperature, which is
P3 = P2 = 9.6065 bar.

(2.459)

So, two properties at the end of the compression are known: pressure and entropy. The allows determination of the enthalpy at the end of compression via the tables:
h2 = 205.1

kJ
.
kg

(2.460)

State 3 is at the end of the condenser, so x3 = 0, and one finds the enthalpy from the tables to be
h3 = 74.59

kJ
.
kg

(2.461)

Now the throttling device has constant enthalpy, so


h4 = h3 = 74.59

kJ
.
kg

(2.462)

However, the fluid has been throttled to a lower pressure: that of the evaporator, which is the same as
state 1, the compressor inlet, so
P4 = P1 = 5.6729 bar.
(2.463)
CC BY-NC-ND. 03 August 2012, J. M. Powers.

68

CHAPTER 2. CYCLE ANALYSIS


The compressor work is
c
W

=
=
=

m(h
2 h1 ),
 
 


kJ
kJ
kg
205.1
195.78
,
0.008
s
kg
kg
0.075 kW.

(2.464)
(2.465)
(2.466)

Now one desires the heat which leaves the cold region to be high for a good refrigerator. This is the
heat transfer at the low temperature part of the T s diagram, which here gives
Q L

=
=
=
=
=

m(h
1 h4 ),
 
 


kJ
kJ
kg
195.78
74.59
,
0.008
s
kg
kg
0.9695 kW,
!



kJ
1 ton
60 s
0.9695
,
kJ
s
min
211 min
0.276 ton.

(2.467)
(2.468)
(2.469)
(2.470)
(2.471)

The units ton is used for power and is common in the refrigeration industry. It is the power required
to freeze one short ton of water at 0 C in 24 hours. It is 12000 Btu/hr, 3.516853 kW , or 4.7162 hp.
A short ton is a U.S ton, which is 2000 lbm. A long ton is British and is 2240 lbm. A metric ton is
1000 kg and is 2204 lbm.
The coefficient of performance is

=
=
=

Q L
,
c
W
0.9695 kW
,
0.075 kW
12.93.

(2.472)
(2.473)
(2.474)

The equivalent Carnot refrigerator would have


max

=
=
=

Q L
,
c
W
Q L
,
Q H Q L
1
,
H
Q

(2.475)
(2.476)
(2.477)

=
=
=
=

TH
TL

1
,
1

TL
,
TH TL
20 + 273.15
,
40 + 273.15 (20 + 273.15)
14.6575.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(2.478)
(2.479)
(2.480)
(2.481)

69

2.10. REFRIGERATION

2.10.2

Air standard

This is effectively the inverse Brayton cycle. It is used in the liquefaction of air and other
gases. It is also used in aircraft cabin cooling. It has the following components:
1 2: isentropic compression,
2 3: isobaric heat transfer to a high temperature environment
3 4: isentropic expansion through turbine, and
4 1: isobaric heat exchange with low temperature surroundings.
A schematic for the air standard refrigeration cycle is shown in Figure 2.19. A T s diagram
QH
3
2

-w
Compressor

Expander

1
4

QL

Figure 2.19: Schematic for air standard refrigeration cycle.


for the air standard refrigeration cycle is shown in Figure 2.20.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

70

CHAPTER 2. CYCLE ANALYSIS

T
2

isobar
3
1
4

isobar
s

Figure 2.20: T s diagram for air standard refrigeration Brayton cycle.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

Chapter 3
Gas mixtures
Read BS, Chapter 13.
See for background Sandler, Chapter 7.
See for background Smith, van Ness, and Abbott, Chapter 10.
See for background Tester and Modell.
One is often faced with mixtures of simple compressible substances, and it the thermodynamics of such mixtures upon which attention is now fixed. Here a discussion of some of
the fundamentals of mixture theory will be given. In general, thermodynamics of mixtures
can be a challenging topic about which much remains to be learned. In particular, these
notes will focus on ideal mixtures of ideal gases, for which results are often consistent with
intuition. The chemical engineering literature contains a full discussion of the many nuances
associated with non-ideal mixtures of non-ideal materials.

3.1

Some general issues

Here the notation of BS will be used. There is no effective consensus on notation for mixtures.
That of BS is more unusual than most; however, the ideas are correct, which is critical.
Consider a mixture of N components, each a pure substance, so that the total mass and
total number of moles are
m = m1 + m2 + m3 + + mN =
n = n1 + n2 + n3 + + nN =

N
X

N
X

mi ,

mass, units= kg,

(3.1)

moles, units= kmole.

(3.2)

i=1

ni ,

i=1

Recall 1 mole = 6.02214179 1023 molecule. The mass fraction of component i is defined
as ci :
mi
,
ci
mass fraction, dimensionless.
(3.3)
m
71

72

CHAPTER 3. GAS MIXTURES

The mole fraction of component i is defined as yi :


yi

ni
,
n

mole fraction, dimensionless.

(3.4)

Now the molecular mass of species i is the mass of a mole of species i. It units are typically
g/mole. This is an identical unit to kg/kmole. Molecular mass is sometimes called molecular weight, but this is formally incorrect, as it is a mass measure, not a force measure.
Mathematically the definition of Mi corresponds to


mi
g
kg
Mi
.
(3.5)
,
=
ni
kmole
mole
Then one gets mass fraction in terms of mole fraction as
mi
,
m
ni Mi
=
,
m
ni Mi
,
= PN
j=1 mj
ni Mi
= PN
,
j=1 nj Mj

(3.10)

(3.11)

ci =

ni M i
,
PN n
1
n
M
j
j
j=1
n
ni M i
P N n nj M j ,
j=1 n

yi Mi
.
= PN
j=1 yj Mj

(3.6)
(3.7)
(3.8)
(3.9)

(3.12)

Similarly, one finds mole fraction in terms of mass fraction by the following:
yi =
=
=
=

CC BY-NC-ND. 03 August 2012, J. M. Powers.

ni
,
n

mi
Mi
PN mj ,
j=1 Mj
mi
Mi m
PN mj ,
j=1 Mj m
ci
Mi
P N cj .
j=1 Mj

(3.13)
(3.14)
(3.15)
(3.16)

73

3.1. SOME GENERAL ISSUES


The mixture itself has a mean molecular mass:
M
=

m
,
n
PN

(3.17)

i=1

mi

,
n
N
X
ni Mi
,
=
n
i=1
=

N
X

yi Mi .

(3.18)
(3.19)
(3.20)

i=1

Example 3.1
Air is often modeled as a mixture in the following molar ratios:
O2 + 3.76N2.

(3.21)

Find the mole fractions, the mass fractions, and the mean molecular mass of the mixture.
Take O2 to be species 1 and N2 to be species 2. Consider the number of moles of O2 to be
n1 = 1 kmole,

(3.22)

n2 = 3.76 kmole.

(3.23)

and N2 to be
The molecular mass of O2 is M1 = 32 kg/kmole. The molecular mass of N2 is M2 = 28 kg/kmole.
The total number of moles is
n = 1 kmole + 3.76 kmole = 4.76 kmole.
So the mole fractions are

1 kmole
= 0.2101,
4.76 kmole
3.76 kmole
= 0.7899.
y2 =
4.76 kmole

y1 =

Note that

N
X

yi = 1.

(3.24)

(3.25)
(3.26)

(3.27)

i=1

That is to say y1 + y2 = 0.2101 + 0.7899 = 1. Now for the masses, one has


kg
= 32 kg,
m1 = n1 M1 = (1 kmole) 32
kmole


kg
m2 = n2 M2 = (3.76 kmole) 28
= 105.28 kg.
kmole

(3.28)
(3.29)

So one has
m = m1 + m2 = 32 kg + 105.28 kg = 137.28 kg.

(3.30)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

74

CHAPTER 3. GAS MIXTURES


The mass fractions then are
c1

c2

m1
32 kg
=
= 0.2331,
m
137.28 kg
m2
105.28 kg
=
= 0.7669.
m
137.28 kg

(3.31)
(3.32)

Note that
N
X

ci = 1.

(3.33)

i=1

That is c1 + c2 = 0.2331 + 0.7669 = 1. Now for the mixture molecular mass, one has
M=

m
kg
137.28 kg
=
= 28.84
.
n
4.76 kmole
kmole

(3.34)

Check against another formula.


M=

N
X
i=1


yi Mi = y1 M1 + y2 M2 = (0.2101) 32

kg
kmole


+ (0.7899) 28

kg
kmole

= 28.84

kg
.
kmole

(3.35)

Now postulates for mixtures are not as well established as those for pure substances.
The literature has much controversial discussion of the subject. A strong advocate of the
axiomatic approach, C. A. Truesdell, proposed the following metaphysical principles for
mixtures, which are worth considering.
1. All properties of the mixture must be mathematical consequences of properties of the
constituents.
2. So as to describe the motion of a constituent, we may in imagination isolate it from the
rest of the mixture, provided we allow properly for the actions of the other constituents
upon it.
3. The motion of the mixture is governed by the same equations as is a single body.
Most important for the present discussion is the first principle. When coupled with fluid
mechanics, the second two take on additional importance. The approach of mixture theory
is to divide and conquer. One typically treats each of the constituents as a single material
and then devises appropriate average or mixture properties from those of the constituents.
The best example of this is air, which is not a single material, but is often treated as
such.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

75

3.2. IDEAL AND NON-IDEAL MIXTURES

3.2

Ideal and non-ideal mixtures

A general extensive property, such as U, for an N-species mixture will be such that
U = U(T, P, n1 , n2 , . . . , nN ).

(3.36)

A partial molar property is a generalization of an intensive property, and is defined such that
it is the partial derivative of an extensive property with respect to number of moles, with T
and P held constant. For internal energy, then the partial molar internal energy is

U
ui
(3.37)
.
ni T,P,nj ,i6=j
Pressure and temperature are held constant because those are convenient variables to control
in an experiment. One also has the partial molar volume

V
.
vi =
(3.38)
ni T,P,nj ,i6=j
It shall be soon seen that there are other natural ways to think of the volume per mole.
Now in general one would expect to find
ui = ui (T, P, n1 , n2 , . . . , nN ),
vi = vi (T, P, n1, n2 , . . . , nN ).

(3.39)
(3.40)

This is the case for what is known as a non-ideal mixture. An ideal mixture is defined as a
mixture for which the partial molar properties ui and vi are not functions of the composition,
that is
ui = ui (T, P ),
vi = vi (T, P ),

if ideal mixture,
if ideal mixture.

(3.41)
(3.42)

An ideal mixture also has the property that hi = hi (T, P ), while for a non-ideal mixture
hi = hi (T, P, n1 , . . . , nN ). Though not obvious, it will turn out that some properties of an
ideal mixture will depend on composition. For example, the entropy of a constituent of an
ideal mixture will be such that
si = si (T, P, n1, n2 , . . . , nN ).

3.3

(3.43)

Ideal mixtures of ideal gases

The most straightforward mixture to consider is an ideal mixture of ideal gases. Even here,
there are assumptions necessary that remain difficult to verify absolutely.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

76

CHAPTER 3. GAS MIXTURES

3.3.1

Dalton model

The most common model for a mixture of ideal gases is the Dalton model. Key assumptions
define this model
Each constituent shares a common temperature.
Each constituent occupies the entire volume.
Each constituent possesses a partial pressure which sums to form the total pressure of
the mixture.
The above characterize a Dalton model for any gas, ideal or non-ideal. One also takes for
convenience
Each constituent behaves as an ideal gas.
The mixture behaves as a single ideal gas.
It is more convenient to deal on a molar basis for such a theory. For the Dalton model,
additional useful quantities, the species mass concentration i , the mixture mass concentration , the species molar concentration i , and the mixture molar concentration , can be
defined. As will be seen, these definitions for concentrations are useful; however, they are not
in common usage. Following BS, the bar notation, , will be reserved for properties which
are mole-based rather than mass-based. As mentioned earlier, the notion of a partial molal
property is discussed extensively in the chemical engineering literature and has implications
beyond those considered here. For the Dalton model, in which each component occupies the
same volume, one has
Vi = V.
(3.44)
The mixture mass concentration, also called the density is simply
 
m
kg
.
= ,
V
m3

(3.45)

The mixture molar concentration is


n
= ,
V

kmole
m3

(3.46)

For species i, the equivalents are


i
i

mi
=
,
V
ni
=
,
V

CC BY-NC-ND. 03 August 2012, J. M. Powers.


kg
,
m3


kmole
.
m3

(3.47)
(3.48)

77

3.3. IDEAL MIXTURES OF IDEAL GASES

One can find a convenient relation between species molar concentration and species mole
fraction by the following operations
ni n
,
(3.49)
i =
V n
ni n
=
,
(3.50)
nV
(3.51)
= yi .
A similar relation exists between species molar concentration and species mass fraction via
ni m Mi
,
(3.52)
i =
V m Mi
=m
z }|i{
m ni Mi
,
(3.53)
=
V mMi
=ci
z}|{
mi 1
=
,
(3.54)
m Mi
ci
(3.55)
= .
Mi
The specific volumes, mass and molar, are similar. One takes
V
V
v= ,
v =
,
(3.56)
m
n
V
V
,
(3.57)
vi = .
vi =
mi
ni
Note that this definition of molar specific volume is not the partial molar volume defined in
the chemical engineering literature, which takes the form vi = V /ni |T,P,nj ,i6=j .
For the partial pressure of species i, one can say for the Dalton model
P =

N
X

Pi .

(3.58)

i=1

For species i, one has


Pi V
Pi
N
X

Pi

i=1

| {z }
=P

= ni RT,
ni RT
=
,
V
N
X
ni RT
,
=
V
i=1

N
RT X
ni .
P =
V i=1
| {z }

(3.59)
(3.60)
(3.61)

(3.62)

=n

CC BY-NC-ND. 03 August 2012, J. M. Powers.

78

CHAPTER 3. GAS MIXTURES

So, for the mixture, one has


P V = nRT.

(3.63)

One could also say


P =

n
RT = RT.
V
|{z}

(3.64)

Here n is the total number of moles in the system. Additionally R is the universal gas
constant with value

R = 8.314472

J
kJ
= 8.314472
.
kmole K
mole K

(3.65)

Sometimes this is expressed in terms of kB , the Boltzmann constant, and N , Avogadros


number:

R = kB N ,

molecule
N = 6.02214179 1023
,
mole
J
.
kB = 1.380650 1023
K molecule

(3.66)
(3.67)
(3.68)

Example 3.2
Compare the molar specific volume defined here with the partial molar volume from the chemical
engineering literature.

The partial molar volume vi , is given by



V
.
vi =
ni T,P,nj ,i6=j

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(3.69)

79

3.3. IDEAL MIXTURES OF IDEAL GASES


For the ideal gas, one has
PV
V

V
ni T,P,nj ,i6=j

= RT
=
=
=

=
vi

=
=

N
X

RT
RT

nk

P
PN

(3.70)
,

nk
k=1 ni

P
PN

(3.71)
,

(3.72)

k=1 ki
,
P

=0
=0
=1
=0
z}|{
z}|{
z}|{ z}|{
RT 1i + 2i + + ii + + N i

RT
,
P
V
PN

k=1

nk ,

k=1
PN
RT k=1

(3.73)

(3.74)
(3.75)

nk

(3.76)

V
.
n

(3.77)

Here the so-called Kronecker delta function has been employed, which is much the same as the identity
matrix:
ki

0,

ki

1,

k 6= i,

k = i.

(3.78)
(3.79)

Contrast this with the earlier adopted definition of molar specific volume
vi =

V
.
ni

(3.80)

So, why is there a difference? The molar specific volume is a simple definition. One takes the
instantaneous volume V , which is shared by all species in the Dalton model, and scales it by the
instantaneous number of moles of species i, and acquires a natural definition of molar specific volume
consistent with the notion of a mass specific volume. On the other hand, the partial molar volume
specifies how the volume changes if the number of moles of species i changes, while holding T and P
and all other species mole numbers constant. One can imagine adding a mole of species i, which would
necessitate a change in V in order to guarantee the P remain fixed.

3.3.1.1

Binary mixtures

Consider now a binary mixture of two components A and B. This is easily extended to a
general mixture of N components. First the total number of moles is the sum of the parts:
n = nA + nB .

(3.81)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

80

CHAPTER 3. GAS MIXTURES

Now, write the ideal gas law for each component:


PA VA = nA RTA ,
PB VB = nB RTB .

(3.82)
(3.83)

But by the assumptions of the Dalton model, VA = VB = V , and TA = TB = T , so


PA V
PB V

= nA RT,
= nB RT.

(3.84)
(3.85)

One also has


P V = nRT.

(3.86)

Solving for n, nA and nB , one finds


PV
,
RT
PA V
=
,
RT
PB V
=
.
RT

n =
nA
nB

(3.87)
(3.88)
(3.89)

Now n = nA + nB , so one has


PA V
PB V
PV
=
+
.
RT
RT
RT
P = PA + PB .

(3.90)
(3.91)

That is the total pressure is the sum of the partial pressures. This is a mixture rule for
pressure
One can also scale each constituent ideal gas law by the mixture ideal gas law to get
PA V
nA RT
,
=
PV
nRT
PA
nA
=
,
P
n
= yA ,
PA = yA P.

(3.94)
(3.95)

PB = yB P.

(3.96)

(3.92)
(3.93)

Likewise
Now, one also desires rational mixture rules for energy, enthalpy, and entropy. Invoke Truesdells principles on a mass basis for internal energy. Then the total internal energy U (with
CC BY-NC-ND. 03 August 2012, J. M. Powers.

81

3.3. IDEAL MIXTURES OF IDEAL GASES


units J) for the binary mixture must be
U = mu = mA uA + mB uB ,
m
mB 
A
= m
uA +
uB ,
m
m
= m (cA uA + cB uB ) ,
u = cA u A + cB u B .

(3.97)
(3.98)
(3.99)
(3.100)

For the enthalpy, one has


H = mh = mA hA + mB hB ,
m
mB 
A
hA +
hB ,
= m
m
m
= m (cA hA + cB hB ) ,
h = cA hA + cB hB .

(3.101)
(3.102)
(3.103)
(3.104)

It is easy to extend this to a mole fraction basis rather than a mass fraction basis. One can
also obtain a gas constant for the mixture on a mass basis. For the mixture, one has
PV

= nRT mRT,

PV
mR = nR,
T
= (nA + nB )R,


mB
mA
+
R,
=
MA MB


R
R
=
mA
+ mB
,
MA
MB
= (mA RA + mB RB ) ,

m
mB
A
RA +
RB ,
R =
m
m
R = (cA RA + cB RB ) .

(3.105)
(3.106)
(3.107)
(3.108)
(3.109)
(3.110)
(3.111)
(3.112)

For the entropy, one has


S = ms = = mA sA + mB sB ,
m
mB 
A
sA +
sB ,
= m
m
m
= m (cA sA + cB sB ) ,
s = cA sA + cB s B .

(3.113)
(3.114)
(3.115)
(3.116)

Note that sA is evaluated at T and PA , while sB is evaluated at T and PB . For a CPIG, one
CC BY-NC-ND. 03 August 2012, J. M. Powers.

82

CHAPTER 3. GAS MIXTURES

has
sA =

so298,A
|

+ cP A ln
{z
soT,A

so298,A

sB =

so298,B

+ cP A ln

Likewise

+ cP B ln
{z

T
To

T
To

T
To

soT,B

RA ln

PA
Po

(3.117)

RA ln

yA P
Po

(3.118)

RB ln

yB P
Po

(3.119)

Here the o denotes some reference state. As a superscript, it typically means that the
property is evaluated at a reference pressure. For example, soT,A denotes the portion of the
entropy of component A that is evaluated at the reference pressure Po and is allowed to vary
with temperature T . Note also that sA = sA (T, P, yA) and sB = sB (T, P, yB ), so the entropy
of a single constituent depends on the composition of the mixture and not just on T and P .
This contrasts with energy and enthalpy for which uA = uA (T ), uB = uB (T ), hA = hA (T ),
hB = hB (T ) if the mixture is composed of ideal gases. Occasionally, one finds hoA and hoB
used as a notation. This denotes that the enthalpy is evaluated at the reference pressure.
However, if the gas is ideal, the enthalpy is not a function of pressure and hA = hoA , hB = hoB .
If one is employing a calorically imperfect ideal gas model, then one finds for species i
that


yi P
o
si = sT,i Ri ln
,
i = A, B.
(3.120)
Po
3.3.1.2

Entropy of mixing

Example 3.3
Initially calorically perfect ideal gases A and B are segregated within the same large volume by
a thin frictionless, thermally conducting diaphragm. Thus, both are at the same initial pressure and
temperature, P1 and T1 . The total volume is thermally insulated and fixed, so there are no global heat
or work exchanges with the environment. The diaphragm is removed, and A and B are allowed to mix.
Assume A has mass mA and B has mass mB . The gases are allowed to have distinct molecular masses,
MA and MB . Find the final temperature T2 , pressure P2 , and the change in entropy.
The ideal gas law holds that at the initial state
VA1 =

mA RA T1
,
P1

VB1 =

mB RB T1
.
P1

(3.121)

At the final state one has


V2 = VA2 = VB2 = VA1 + VB1 = (mA RA + mB RB )
CC BY-NC-ND. 03 August 2012, J. M. Powers.

T1
.
P1

(3.122)

83

3.3. IDEAL MIXTURES OF IDEAL GASES


Mass conservation gives
m2 = m1 = mA + mB .

(3.123)

1W2 ,

(3.124)

One also has the first law


U2 U1

U2 U1
U2

m2 u 2
(mA + mB )u2
0
0
T2
T2

1Q2

= 0,
= U1 ,

(3.125)
(3.126)

= mA uA1 + mB uB1 ,
= mA uA1 + mB uB1 ,
= mA (uA1 u2 ) + mB (uB1 u2 ),

= mA cvA (T1 T2 ) + mB cvB (T1 T2 ),


mA cvA T1 + mB cvB T1
,
=
mA cvA + mB cvB
=

T1 .

(3.127)
(3.128)
(3.129)
(3.130)
(3.131)
(3.132)

The final pressure by Daltons law then is


P2

=
=
=
=

=
=

PA2 + PB2 ,
mB RB T2
mA RA T2
+
,
V2
V2
mA RA T1
mB RB T1
+
,
V2
V2
(mA RA + mB RB ) T1
,
V2
substitute for V2 from Eq. (3.122)
(mA RA + mB RB ) T1
,
(mA RA + mB RB ) PT11
P1 .

(3.133)
(3.134)
(3.135)
(3.136)
(3.137)
(3.138)
(3.139)

So the initial and final temperatures and pressures are identical.


Now the entropy change of gas A is




PA2
TA2
RA ln
,
sA2 sA1 = cP A ln
TA1
PA1
 


T2
yA2 P2
= cP A ln
RA ln
,
T1
yA1 P1


 
yA2 P1
T1
,
RA ln
= cP A ln
T1
yA1 P1
| {z }
=0


yA2 P1
= RA ln
,
(1)P1
= RA ln yA2 .

(3.140)
(3.141)
(3.142)

(3.143)
(3.144)

Likewise
sB2 sB1

= RB ln yB2 .

(3.145)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

84

CHAPTER 3. GAS MIXTURES


So the change in entropy of the mixture is
S

=
=
=

mA (sA2 sA1 ) + mB (sB2 sB1 ),

(3.146)

mA RA ln yA2 mB RB ln yB2 ,




R
R
(nA MA )
ln yA2 (nB MB )
ln yB2 ,
| {z } MA
| {z } MB
| {z }
| {z }
=mA
=mB
=RA

R(nA ln yA2 + nB ln yB2 ),






nA
nB

R nA ln
+nB ln
,

nA + nB
nA + nB
|
|
{z
}
{z
}

We can also scale Eq. (3.149) by Rn to get

=s

s
R

(3.149)

(3.150)

0.

1 S
n
R |{z}

(3.148)

=RB

(3.147)

nA

nB

=
n ln yA2 + n ln yB2 ,
|{z}
|{z}
=yA2

(3.151)

(3.152)

=yB2

= (yA2 ln yA2 + yB2 ln yB2 ) ,

yB2
yA2
),
+ ln yB2
= (ln yA2
yA2 yB2
= ln (yA2 yB2 ) .

(3.153)
(3.154)
(3.155)

For an N-component mixture, mixed in the same fashion such that P and T are constant,
this extends to
S = R
= R

= R

N
X
k=1
N
X

nk ln

k=1

N
X
nk

k=1
N
X

= Rn
= R

nk ln yk ,

CC BY-NC-ND. 03 August 2012, J. M. Powers.

nk

PN

i=1

{z

n ln yk ,

nk
ln yk ,
n

k=1
N
X

m
M

(3.156)

k=1

yk ln yk ,

ni

0,

(3.157)

(3.158)
(3.159)
(3.160)

85

3.3. IDEAL MIXTURES OF IDEAL GASES

= Rm

N
X

ln ykyk ,

(3.161)

k=1

yN
= Rm (ln y1y1 + ln y2y2 + + ln yN
),
y1 y2
yN
= Rm ln (y1 y2 . . . yN ) ,
!
N
Y
ykyk ,
= Rm ln

(3.162)
(3.163)
(3.164)

k=1

Dividing by m to recover an intensive property and R to recover a dimensionless property,


we get
s
s
= ln
=
R
R

N
Y

k=1

ykyk

(3.165)

Note that there is a fundamental dependency of the mixing entropy on the mole fractions.
Since 0 yk 1, the product is guaranteed to be between 0 and 1. The natural logarithm
of such a number is negative, and thus the entropy change for the mixture is guaranteed
positive semi-definite. Note also that for the entropy of mixing, Truesdells third principle
is not enforced.
Now if one mole of pure N2 is mixed with one mole of pure O2 , one certainly expects
the resulting homogeneous mixture to have a higher entropy than the two pure components.
But what if one mole of pure N2 is mixed with another mole of pure N2 . Then we would
expect no increase in entropy. However, if we had the unusual ability to distinguish N2
molecules whose origin was from each respective original chamber, then indeed there would
be an entropy of mixing. Increases in entropy thus do correspond to increases in disorder.
3.3.1.3

Mixtures of constant mass fraction

If the mass fractions, and thus the mole fractions, remain constant during a process, the
equations simplify. This is often the case for common non-reacting mixtures. Air at moderate
values of temperature and pressure behaves this way. In this case, all of Truesdells principles
can be enforced. For a CPIG, one would have
u2 u1 = cA cvA (T2 T1 ) + cB cvB (T2 T1 ),
= cv (T2 T1 ).

(3.166)
(3.167)

cv cA cvA + cB cvB .

(3.168)

h2 h1 = cA cP A (T2 T1 ) + cB cP B (T2 T1 ),
= cP (T2 T1 ).

(3.169)
(3.170)

where
Similarly for enthalpy

CC BY-NC-ND. 03 August 2012, J. M. Powers.

86

CHAPTER 3. GAS MIXTURES

where
cP cA cP A + cB cP B .

(3.171)

For the entropy


s2 s1 = cA (sA2 sA1 ) + cB (sB2 sB1 ),
(3.172)

 



 


T2
y A P2
T2
y B P2
= cA cP A ln
RA ln
+ cB cP B ln
RB ln
,
T1
y A P1
T1
y B P1
 

 
 

 
P2
T2
P2
T2
RA ln
+ cB cP B ln
RB ln
,(3.173)
= cA cP A ln
T1
P1
T1
P1
 
 
T2
P2
= cP ln
R ln
.
(3.174)
T1
P1
The mixture behaves as a pure substance when the appropriate mixture properties are defined. One can also take
cP
k= .
(3.175)
cv

3.3.2

Summary of properties for the Dalton mixture model

Listed here is a summary of mixture properties for an N-component mixture of ideal gases
on a mass basis:

M =
=

N
X

i=1
N
X

yi Mi ,

(3.176)

i ,

(3.177)

i=1

1
v = PN

1
i=1 vi

u =

h =

N
X

i=1
N
X

1
= ,

(3.178)

ci u i ,

(3.179)

ci hi ,

(3.180)

i=1

R =

R
=
M

=R

N
X

ci Ri =

i=1

N
X
i=1

z }| {
yi Mi Ri
N
X
yj Mj
j=1

CC BY-NC-ND. 03 August 2012, J. M. Powers.

| {z } =M

N
R X
yi ,
M i=1
| {z }
=1

(3.181)

87

3.3. IDEAL MIXTURES OF IDEAL GASES

cv =

N
X

ci cvi ,

(3.182)

i=1

cv = cP R,
N
X
cP =
ci cP i ,

if ideal gas,

(3.183)
(3.184)

i=1

k =

s =

cP
,
cv
N
X

(3.185)

ci si ,

(3.186)

i=1

ci =
Pi =
i =
vi =
V =
T =
hi =
hi =
hi =

yi Mi
,
M
yi P,
ci ,
1
v
= ,
ci
i
Vi ,
Ti ,
hoi ,
if ideal gas,
Pi
ui +
= ui + Pi vi = ui + Ri T ,
| {z }
i
if ideal gas
Z T
ho298,i +
cP i (T) dT,
298
|
{z
}

(3.187)
(3.188)
(3.189)
(3.190)
(3.191)
(3.192)
(3.193)
(3.194)

(3.195)

if ideal gas

si =

so298,i
|

si =

Pi

soT,i
|

 
Pi
cP i (T)
,
+
dT Ri ln

Po
T
298
{z
}
Z

=soT,i

Ri ln
{z

{z


if ideal gas

yi P
Po

if ideal gas

soT,i
|

Ri T
,
= i Ri T = Ri T ci =
vi
{z
}
|

Ri ln
{z

if ideal gas


Pi
,
Po
}

(3.196)

(3.197)

(3.198)

if ideal gas

CC BY-NC-ND. 03 August 2012, J. M. Powers.

88

CHAPTER 3. GAS MIXTURES


N
X
RT
ci
=
P = RT = RT
,
Mi
v
i=1
|
{z
}

(3.199)

if ideal gas

h =

N
X

|i=1

h = u+

ci ho298,i

cP (T ) dT ,

(3.200)

(3.201)

298

{z

if ideal gas

P
= u + P v = |u +{zRT} ,

if ideal gas

s =

N
X

ci so298,i +

|i=1

298

cP (T )
dT R ln
T
{z

P
Po

R ln

if ideal gas

N
Y
i=1

yiyi

.
}

(3.202)

These relations are not obvious. A few are derived in examples here.
Example 3.4
Derive the expression h = u + P/.
Start from the equation for the constituent hi , multiply by mass fractions, sum over all species,
and use properties of mixtures:

N
X

hi

ci h i

ci h i

ci h i

i=1

N
X
i=1

| {z }

Pi
,
i

(3.203)

Pi
,
i
N
N
X
X
Pi
ci ,
ci u i +
i
i=1
i=1
ci u i + ci

N
X

ci u i +

i=1

N
X

ci

i=1

| {z }

i Ri T
,
i

(3.204)
(3.205)

(3.206)

=u

=h

ui +

u+T

N
X

ci Ri ,

(3.207)

i=1

| {z }
=R

=
=

CC BY-NC-ND. 03 August 2012, J. M. Powers.

u + RT,

(3.208)

P
.

(3.209)

u+

89

3.3. IDEAL MIXTURES OF IDEAL GASES

Example 3.5
Find the expression for mixture entropy of the ideal gas.

s=

N
X

si

ci si

ci si

i=1

 
cP i (T)
Pi
,
dT Ri ln
Po
T
298
 
Z T
cP i (T)
Pi
o
ci s298,i + ci
,
dT ci Ri ln

P
o
T
298
 
Z T
N
N
N
X
X
cP i (T) X
Pi
o
ci
ci s298,i +
ci Ri ln
,
dT

P
o
T
298
i=1
i=1
i=1
 
Z T X
N
N
N
X
ci cP i (T) X
Pi
ci so298,i +
,
ci Ri ln
dT

P
o
T
298 i=1
i=1
i=1
 
Z T
N
cP (T) X
Pi
o
s298 +
.
ci Ri ln
dT

Po
T
298
i=1
so298,i +

(3.210)
(3.211)
(3.212)

(3.213)

(3.214)

All except the last term are natural extensions of the property for a single material. Consider now the
last term involving pressure ratios.

N
X
i=1

ci Ri ln

Pi
Po

!
P
P
ci Ri ln

+ R ln
,
R ln
Po
Po
i=1
!

N
X
Ri
Pi
P
P
R
ci
,
ln
ln
+ ln
R
Po
Po
Po
i=1
!
 
N
X
R/Mi
P
P
Pi
ci P N
R
+ ln
,
ln
ln
Po
Po
Po
j=1 cj R/Mj
i=1

!  
X

N
Pi
ci /Mi
P
P

ln
R
+ ln
,
ln
PN
P
P
Po
o
o
c
/M
i=1

j
j
j=1
{z
}
|
N
X

Pi
Po


(3.215)

(3.216)

(3.217)

(3.218)

=yi

!

P
P
Pi
+ ln
,
ln
yi ln
Po
Po
Po
i=1
!
 yi
N
X
Pi
P
P
ln
ln
,
+ ln
Po
Po
Po
i=1
!
y !
N 
Y
Pi i
Po
P
ln
+ ln
,
+ ln
Po
P
Po
i=1
!
y !
N 
P
Po Y Pi i
+ ln
,
ln
P i=1 Po
Po
!
!
N
Y
P
Po
1
yi
PN
PN
+ ln
ln
(Pi )
,
Po
P i=1 yi Po i=1 yi i=1

N
X

(3.219)

(3.220)

(3.221)

(3.222)

(3.223)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

90

CHAPTER 3. GAS MIXTURES


!

!
P
+ ln
,
= R ln
P
Po
i=1
!
y !
N 
Y
yi P i
P
= R ln
+ ln
,
P
Po
i=1
!
!
N
Y
P
yi
+ ln
yi
= R ln
.
Po
i=1
y
N 
Y
Pi i

(3.224)

(3.225)

(3.226)

So the mixture entropy becomes


s =

so298

so298
|

298

298

!
P
+ ln
,
Po
i=1
!
N
Y
P
cP (T)
yi
dT R ln
yi .
R ln
Po
T{z
i=1
} |
{z
}

N
Y
cP (T)
yiyi
dT R ln
T

classical entropy of a single body

(3.227)

(3.228)

nonTruesdellian

The extra entropy is not found in the theory for a single material, and in fact is not in the form suggested
by Truesdells postulates. While it is in fact possible to redefine the constituent entropy definition in
such a fashion that the mixture entropy in fact
 on the classical form of a single material via the
 takes
R T cP i (T)
Pi
o

definition si = s298,i + 298 T dT Ri ln Po + Ri ln yi , this has the disadvantage of predicting


no entropy change after mixing two pure substances. Such a theory would suggest that this obviously
irreversible process is in fact reversible.

On a molar basis, one has the equivalents


=

N
X

i =

i=1

1
v = PN

1
i=1 v i

u =
h =

N
X

i=1
N
X

n
=
,
V
M
=

V
1
= = vM,
n

(3.229)
(3.230)

yi ui = uM,

(3.231)

yi hi = hM,

(3.232)

i=1

cv =

N
X

yi cvi = cv M,

(3.233)

i=1

cv = cP R,
N
X
yi cP i = cP M,
cP =
i=1

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(3.234)
(3.235)

91

3.3. IDEAL MIXTURES OF IDEAL GASES


k =
s =

cP
,
cv
N
X

(3.236)
yi si = sM,

(3.237)

i=1

i
,
Mi
v
1
V
=
=
= vi Mi ,
=
ni
y
i
i
V
V
=
= v = vM ,
=

ni P,T,nj |n
{z
}

i = yi =

(3.238)

vi

(3.239)

vi

(3.240)

if ideal gas

Pi = yi P,

(3.241)

RT
,
P = RT =
{z v }
|

(3.242)

if ideal gas

RT
Pi = i RT =
,
vi
|
{z
}

(3.243)

if ideal gas

P
= u + Pv = u
| +{zRT} = hM,

h = u+

(3.244)

if ideal gas

hi =

o
hi ,

if ideal gas,

Pi
= ui + Pi v i = ui + P vi = ui + RT = hi Mi ,
| {z }
i
if ideal gas
Z T
o
= h298,i +
cP i(T ) dT = hi Mi ,
298
|
{z
}

(3.245)

hi = ui +

(3.246)

hi

(3.247)

if ideal gas

si =

so298,i
|

si



cP i (T )
yi P
,
dT R ln
+

Po
298
{z T
}
Z

=soT,i

{z


yi P
o
= si Mi ,
= sT,i R ln
Po
{z
}
|
|

if ideal gas

s =

N
X
i=1

if ideal gas

yi so298,i

298

!
 
N
Y
cP (T )
P
yi
dT R ln
R ln
= sM.
yi
Po
T
i=1
{z
}

(3.248)

(3.249)

if ideal gas

(3.250)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

92

CHAPTER 3. GAS MIXTURES

3.3.3

Amagat model

The Amagat model is an entirely different paradigm than the Dalton model. It is not used
as often. In the Amagat model,
all components share a common temperature T ,
all components share a common pressure P , and
each component has a different volume.
Consider, for example, a binary mixture of calorically perfect ideal gases, A and B. For
the mixture, one has
P V = nRT,
(3.251)
with
n = nA + nB .

(3.252)

P VA = nA RT,
P VB = nB RT.

(3.253)
(3.254)

PV
P VA P VB
=
+
.
RT
RT
RT

(3.255)

For the components one has

Then n = nA + nB reduces to

Thus
V

= VA + VB ,
VA VB
+
.
1 =
V
V

3.4

(3.256)
(3.257)

Gas-vapor mixtures

Next consider a mixture of ideal gases in which one of the components may undergo a phase
transition to its liquid state. The most important practical example is an air-water mixture.
Assume the following:
The solid or liquid contains no dissolved gases.
The gaseous phases are all well modeled as ideal gases.
When the gas mixture and the condensed phase are at a given total pressure and temperature, the equilibrium between the condensed phase and its vapor is not influenced
by the other component. So for a binary mixture of A and B where A could have both
gas and liquid components PA = Psat . That is the partial pressure of A is equal to its
saturation pressure at the appropriate temperature.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

93

3.4. GAS-VAPOR MIXTURES

Considering an air water vapor mixture, one models the water vapor as an ideal gas and
expresses the total pressure as
P = Pa + Pv .

(3.258)

Here v denotes vapor and a denotes air. A good model for the enthalpy of the water vapor
is to take
hv (T, lowP ) = hg (T ).

(3.259)

If T is given in degrees Celsius, a good model from the steam tables is




kJ
kJ
hg (T ) = 2501.3
T.
+ 1.82
kg
kg C

(3.260)

Some definitions:
absolute humidity: , the mass of water present in a unit mass of dry air, also called
humidity ratio,

=
=
=
=
=
=

mv
,
ma
Mv nv
,
Ma na
Pv V
Mv RT
,
aV
Ma PRT
Mv Pv
,
Ma Pa

kg
Pv
18.015 kmole

,
kg
28.97 kmole
Pa
Pv
0.622 ,
Pa
Pv
.
0.622
P Pv

(3.261)
(3.262)
(3.263)
(3.264)
(3.265)
(3.266)
(3.267)

dew point: temperature at which the vapor condenses when it is cooled isobarically.
saturated air: The vapor in the air-vapor mixture is at the saturation temperature and
pressure.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

94

CHAPTER 3. GAS MIXTURES


relative humidity: The ratio of the mole fraction of the vapor in the mixture to the mole
fraction of vapor in a saturated mixture at the same temperature and total pressure:

=

nv
n
ng ,
n
nv

ng

(3.268)

(3.269)

Pv V
RT
Pg V
RT

(3.270)

Pv
.
Pg

(3.271)

Note, here the subscript g denotes saturated gas values. Combining, one can relate the
relative humidity to the absolute humidity:
=

Pa
.
0.622Pg

(3.272)

Example 3.6
(from C
engal and Boles, p. 670) A 5 m 5 m 3 m room contains air at 25 C and 100 kP a at a
relative humidity of 75%. Find the
partial pressure of dry air,
absolute humidity (i.e. humidity ratio),
masses of dry air and water vapor in the room, and the
dew point.
The relation between partial and total pressure is
P = Pa + Pv .

(3.273)

Now from the definition of relative humidity,


Pv = Pg .

(3.274)

Here Pg is the saturation pressure at the same temperature, which is 25 C. At 25 C, the tables give
Pg |25

= 3.169 kP a.

(3.275)

So
Pv = 0.75(3.169 kP a) = 2.3675 kP a.

(3.276)

So from the definition of partial pressure


Pa

=
=
=

P Pv ,

100 kP a 2.3675 kP a,
97.62 kP a.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(3.277)
(3.278)
(3.279)

95

3.4. GAS-VAPOR MIXTURES


Now for the absolute humidity (or specific humidity), one has
Pv
,
Pa
2.3675 kP a
= 0.622
,
97.62 kP a
kg H2 O
.
= 0.0152
kg dry air

= 0.622

(3.280)
(3.281)
(3.282)

Now for the masses of air and water, one can use the partial pressures:
ma

=
=
=

Pa V
Ma ,
RT
Pa V
,
Ra T

(97.62 kP a) 75 m3


,
kJ
8.314 kmole
K
(298
K)
kg
28.97

(3.283)
(3.284)
(3.285)

kmole

=
mv

=
=
=

85.61 kg.

Pv V
Mv ,
RT
Pv V
,
Rv T

(2.3675 kP a) 75 m3


,
kJ
8.314 kmole
K
(298
K)
kg
18.015

(3.286)
(3.287)
(3.288)
(3.289)

kmole

1.3 kg.

(3.290)

Also one could get mv from


mv

ma ,

(3.291)

=
=

(0.0152)(85.61 kg),
1.3 kg.

(3.292)
(3.293)

Now the dew point is the saturation temperature at the partial pressure of the water vapor. With
Pv = 2.3675 kP a, the saturation tables give
Tdew

3.4.1

point

= 20.08 C.

(3.294)

First law

The first law can be applied to air-water mixtures.


CC BY-NC-ND. 03 August 2012, J. M. Powers.

96

CHAPTER 3. GAS MIXTURES

Example 3.7
(BS, Ex. 13.5, pp. 536-537.) An air-water vapor mixture enters the cooling coils of an air conditioner
unit. The inlet is at P1 = 105 kP a, T1 = 30 C, 1 = 0.80. The exit state is P2 = 100 kP a, T2 = 15 C,
2 = 0.95. Liquid water at 15 C also exits the system. Find the heat transfer per kilogram of dry air.
Mass conservation for air and water give
m
a1
m
v1

=
=

m
a2 m
a,
m
v2 + m
l2 .

(3.295)
(3.296)

Note at state 2, the mass flow of water is in both liquid and vapor form.
The first law for the control volume give
dEcv
=
| dt
{z }

cv +m
a ha1 + m
v1 hv1 m
a ha2 m
v2 hv2 m
l2 hl2 ,
Q cv W
|{z}
m
a ha2 + m
v2 hv2 + (m
v1 m
v2 )hl2 ,

(3.298)

ha2 +

(3.299)

ha2 + 2 hv2 + (1 2 )hl2 ,

(3.300)

ha2 ha1 1 hv1 + 2 hv2 + (1 2 )hl2 ,

(3.301)

cpa (T2 T1 ) 1 hv1 + 2 hv2 + (1 2 )hl2 .

(3.302)

=0

Q cv + m
a ha1 + m
v1 hv1

Qcv
m
v1
+ ha1 +
hv1
m
a
m
a
Q cv
+ ha1 + 1 hv1
m
a
Q cv
m
a

(3.297)

=0

m
v2
m
v1 m
v2
hv2 +
hl2 ,
m
a
m
a

Now at the inlet, one has from the definition of relative humidity
1

Pv1
.
Pg1

(3.303)

Here Pg is the saturated vapor pressure at the inlet temperature, T1 = 30 C. This is Pg1 = 4.246 kP a.
So one gets
Pv1

=
=

1 Pg1 ,
(0.80)(4.246 kP a),

(3.304)
(3.305)

3.397 kP a.

(3.306)

Now the absolute humidity (humidity ratio) is


1

=
=
=

Pv1
,
P1 Pv1
3.397 kP a
0.622
,
105 kP a 3.397 kP a
0.0208.

0.622

(3.307)
(3.308)
(3.309)

At the exit temperature, the saturation pressure is Pg2 = 1.705 kP a. So


Pv2

2 Pg2 ,

(3.310)

(0.95)(1.705 kP a),

(3.311)

1.620 kP a.

(3.312)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

97

3.4. GAS-VAPOR MIXTURES


Now the absolute humidity (humidity ratio) is
2

=
=
=

Pv2
,
P2 Pv2
1.620 kP a
0.622
,
100 kP a 1.620 kP a
0.0102.
0.622

Then, substituting, one gets





Q cv
kJ

(15 C 30 C) 0.0208 2556.3


=
1.004
m
a
kg K



kJ
+ (0.0208 0.0102) 62.99
+0.0102 2528.9
kg
=

3.4.2

41.77

(3.313)
(3.314)
(3.315)

kJ
kg

kJ
,
kg

kJ
.
kg dry air

(3.316)
(3.317)

Adiabatic saturation

In an adiabatic saturation process, an air-vapor mixture contacts a body of water in a well


insulated duct. If the initial humidity of the mixture is less than 100%, some water will
evaporate and join the mixture.
If the mixture leaving the duct is saturated, and the process is adiabatic, the exit temperature is the adiabatic saturation temperature. Assume the liquid water entering the system
enters at the exit temperature of the mixture.
Mass conservation for air and water and the first law for the control volume give
dmair
cv
= m
a1 m
a2 ,
dt
| {z }

air

(3.318)

=0

dmwater
cv
= m
v1 + m
l m
v2 ,
dt
| {z }

water

(3.319)

=0

dEcv
cv +m
= Q cv W
a1 ha1 + m
v1 hv1 + m
l hl m
a2 ha2 m
v2 hv2 .
|{z}
|{z}
dt
| {z }
=0

=0

(3.320)

=0

For steady state results, these reduce to

0 = m
a1 m
a2 ,
air
0 = m
v1 + m
lm
v2 ,
water
0 = m
a1 ha1 + m
v1 hv1 + m
l hl m
a2 ha2 m
v2 hv2 .

(3.321)
(3.322)
(3.323)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

98

CHAPTER 3. GAS MIXTURES

Now mass conservation for air gives


m
a1 = m
a2 m
a.

(3.324)

m
l=m
v2 m
v1 .

(3.325)

0 = m
a ha1 + m
v1 hv1 + (m
v2 m
v1 )hl m
a ha2 m
v2 hv2 ,


m
v1
m
v2 m
v1
m
v2
0 = ha1 +
hv1 +

hl ha2
hv2 ,
m
a
m
a
m
a
m
a
0 = ha1 + 1 hv1 + (2 1 ) hl ha2 2 hv2 ,
0 = ha1 ha2 + 1 (hv1 hl ) + 2 (hl hv2 ),
1 (hv1 hl ) = ha1 ha2 + 2 (hl hv2 ),
1 (hv1 hl ) = ha2 ha1 + 2 (hv2 hl ),
1 (hv1 hl ) = cpa (T2 T1 ) + 2 hf g2 .

(3.326)

Mass conservation for water gives

Then energy conservation becomes

(3.327)
(3.328)
(3.329)
(3.330)
(3.331)
(3.332)

Example 3.8
(BS, 13.7E, p. 540). The pressure of the mixture entering and leaving the adiabatic saturater is
14.7 psia, the entering temperature is 84 F , and the temperature leaving is 70 F , which is the adiabatic
saturation temperature. Calculate the humidity ratio and the relative humidity of the air-water vapor
mixture entering.
The exit state 2 is saturated, so
Pv2 = Pg2 .

(3.333)

The tables give Pv2 = Pg2 = 0.363 psia. Thus one can calculate the absolute humidity by its definition:
2

Pv2
,
P2 Pv2
0.363 psia
= 0.622
,
(14.7 psia) (0.363 psia)
lbm H2 O
.
= 0.0157485
lbm dry air

= 0.622

(3.334)
(3.335)
(3.336)

The earlier derived result from the energy balance allows calculation then of 1 :
1 (hv1 hl ) =
1

=
=
=

cpa (T2 T1 ) + 2 hf g2 ,
cpa (T2 T1 ) + 2 hf g2
,
hv1 hl
Btu
0.240 lbm
)) + 0.0157485 1054.0
F ((70 F ) (84 F


Btu
1098.1 lbm 38.1 Btu
lbm
0.012895

lbm H2 O
.
lbm dry air

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(3.337)
(3.338)

Btu
lbm

(3.339)
(3.340)

99

3.4. GAS-VAPOR MIXTURES

Here hv1 was estimated as the saturated vapor value at T = 84 F by interpolating the tables. In the
absence of more information regarding the initial vapor state, this estimate is good as any. The value
of hl is estimated as the saturated liquid value at T = 70 F . Now
1
Pv1

Pv1
,
P1 Pv1
P1 1
=
,
0.622 + 1
(14.7 psia)(0.0124895)
,
=
0.622 + 0.0124895
= 0.28936 psia.

= 0.622

(3.341)
(3.342)
(3.343)
(3.344)

For the relative humidity


1

=
=
=

3.4.3

Pv1
,
Pg1
0.28936 psia
,
0.584 psia
0.495.

(3.345)
(3.346)
(3.347)

Wet-bulb and dry-bulb temperatures

Humidity is often measured with a psychrometer, which has a wet bulb and dry bulb thermometer.
The dry bulb measures the air temperature.
The wet bulb measures the temperature of a water soaked thermometer.
If the two temperatures are equal, the air is saturated. If they are different, some of
the water on the web bulb evaporates, cooling the wet bulb thermometer.
The evaporative cooling process is commonly modeled (with some error) as an adiabatic
saturation process.
These temperatures are also influenced by non-thermodynamic issues such as heat and
mass transfer rates, which induce errors in the device.
Capacitance-based electronic devices are often used as an alternative to the traditional
psychrometer.

3.4.4

Psychrometric chart

This well-known chart summarizes much of what is important for binary mixtures of air and
water, in which the properties depend on three variables, e.g. temperature, pressure, and
composition of the mixture.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

100

CC BY-NC-ND. 03 August 2012, J. M. Powers.

CHAPTER 3. GAS MIXTURES

Chapter 4
Mathematical foundations of
thermodynamics
Read Abbott and van Ness, Chapter 3.
Read BS, 14.2-14.4, 14.9, 16.1-16.4.
See Vincenti and Kruger, Chapter 3, for more background.

4.1

Exact differentials and state functions

In thermodynamics, one is faced with many systems of the form of the well-known Gibbs
equation, Eq. (1.1):
du = T ds P dv.

(4.1)

This is known to be an exact differential with the consequence that internal energy u is a
function of the state of the system and not the details of any process which led to the state.
As a counter-example, the work,
w = P dv,

(4.2)

can be shown to be an inexact differential so that the work is indeed a function of the process
involved. Here we use the notation to emphasize that this is an inexact differential.

Example 4.1
Show the work is not a state function. If work were a state function, one might expect it to have
the form
w = w(P, v),

provisional assumption, to be tested.

(4.3)

In such a case, one would have the corresponding differential form


w =



w
w
dv
+
dP.
v P
P v

101

(4.4)

102

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


Now since w = P dv from Newtonian mechanics, one deduces that

w
= P,
v P

w
= 0.
P

(4.5)
(4.6)

Integrating Eq. (4.5), one finds

w = P v + f (P ),

(4.7)

where f (P ) is some function of P to be determined. Differentiating Eq. (4.7) with respect to P , one
gets

df (P )
w
.
(4.8)
=v+
P v
dP
Now use Eq. (4.6) to eliminate w/P |v in Eq. (4.8) so as to obtain
0

v+

df (P )
dP

v.

df (P )
,
dP

(4.9)
(4.10)

Equation (4.10) cannot be: a function of P only cannot be a function of v. So, w cannot be a state
property:
w 6= w(P, v).
(4.11)

Consider now the more general form


1 dx1 + 2 dx2 + + N dxN =

N
X

i dxi .

(4.12)

i=1

Here i and xi , i = 1, . . . , N, may be thermodynamic variables. This form is known in


mathematics as a Pfaff differential form. As formulated, one takes at this stage
xi : independent thermodynamic variables, and
i : thermodynamic variables which are functions of xi .
Now, if the differential in Eq. (4.12), when set to a differential dy, can be integrated to form
the function
y = y(x1 , x2 , . . . , xN ),
(4.13)
the differential is said to be exact. In such a case, one has
dy = 1 dx1 + 2 dx2 + + N dxN =
CC BY-NC-ND. 03 August 2012, J. M. Powers.

N
X
i=1

i dxi .

(4.14)

4.1. EXACT DIFFERENTIALS AND STATE FUNCTIONS

103

Now, if the algebraic definition of Eq. (4.13) holds, what amounts to the definition of the
partial derivative gives the parallel result that



y
y
y
dy =
dx1 +
dx2 + +
dxN .
(4.15)
x1 xj ,j6=1
x2 xj ,j6=2
xN xj ,j6=N
Now, combining Eqs. (4.14) and (4.15) to eliminate dy, one gets



y
y
y
1 dx1 + 2 dx2 + + N dxN =
dx1 +
dx2 + +
dxN .
x1 xj ,j6=1
x2 xj ,j6=2
xN xj ,j6=N

(4.16)

Rearranging, one gets


!

y
1 dx1 +
x1 xj ,j6=1

!

y
N dxN .
0=
xN xj ,j6=N
(4.17)
Since the variables xi , i = 1, . . . , N, are independent, dxi , i = 1, . . . , N, are all independent in Eq. (4.17), and in general non-zero. For equality, one must require that each of the
coefficients be zero, so

y
,
1 =
x1 xj ,j6=1

!

y
2 dx2 + +
x2 xj ,j6=2


y
2 =
,...,
x2 xj ,j6=2


y
N =
.
xN xj ,j6=N

(4.18)

So when dy is exact, one says that each of the i and xi are conjugate to each other.
From here on out, for notational ease, the j 6= 1, j 6= 2, . . . , j 6= N will be ignored in
the notation for the partial derivatives. It becomes especially confusing for higher order
derivatives, and is fairly obvious for all derivatives.
If y and all its derivatives are continuous and differentiable, then one has for all i =
1, . . . , N, and k = 1, . . . , N that
2y
2y
=
.
(4.19)
xk xi
xi xk
Now from Eq. (4.18), one has

y
,
k =
xk xj


y
l =
.
xl xj

(4.20)

Taking the partial of the first of Eq. (4.20) with respect to xl and the second with respect
to xk , one gets


2y
2y
l
k
=
=
,
.
(4.21)
xl xj
xl xk
xk xj
xk xl

CC BY-NC-ND. 03 August 2012, J. M. Powers.

104

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Since by Eq. (4.19) the order of the mixed second partials does not matter, one deduces from
Eq. (4.21) that


k
l
=
.
(4.22)
xl
xk
xj

xj

This is a necessary and sufficient condition for the exact-ness of Eq. (4.12). It is a generalization of what can be found in most introductory calculus texts for functions of two
variables.
For the Gibbs equation, (4.1), du = P dv + T ds, one has
y u,

x1 v,

x2 s,

1 P

2 T.

(4.23)

and one expects the natural, or canonical form of


u = u(v, s).

(4.24)

Here, P is conjugate to v, and T is conjugate to s. Application of the general form of


Eq. (4.22) to the Gibbs equation (4.1) gives then


T
P
=
.
v s
s v

(4.25)

Equation (4.25) is known as a Maxwell relation. Moreover, specialization of Eq. (4.20) to


the Gibbs equation (4.1) gives

u
P =
,
v s

If the general differential dy =

The path integral yB yA =

PN

i=1

(4.26)

i dxi is exact, one also can show

R B PN
A


u
T =
.
s v

i=1

i dxi is independent of the path of the integral.

The integral around a closed contour is zero:


I

dy =

I X
N

i dxi = 0.

(4.27)

i=1

The function y can only be determined to within an additive constant. That is, there
is no absolute value of y; physical significance is only ascribed to differences in y. In
fact now, other means, extraneous to this analysis, can be used to provide absolute
values of key thermodynamic variables. This will be important especially for flows
with reaction.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

105

4.1. EXACT DIFFERENTIALS AND STATE FUNCTIONS

Example 4.2
Show the heat transfer q is not a state function. Assume all processes are fully reversible. The first
law gives
du
q

= q w,
= du + w,

(4.28)
(4.29)

= du + P dv.

(4.30)

Take now the non-canonical, although acceptable, form u = u(T, v). Then one gets


u
u
dv
+
dT.
du =
v T
T v

(4.31)

So

=
=



u
u
dv
+
dT + P dv,
v T
T v




u
u
+
P
dv
+
dT.
v T
T v
| {z }
|
{z
}
M

(4.32)
(4.33)

M dv + N dT.

(4.34)

Now by Eq. (4.22), for q to be exact, one must have




N
M
=
,
T v
v T
This reduces to

This can only be true if


So dq is not exact.

P
T v

(4.35)
(4.36)


P
2u
2u
+
.
=

T v
T v
vT

= 0. But this is not the case; consider an ideal gas for which

(4.37)

P
T v

= R/v.

Example 4.3
Show conditions for ds to be exact in the Gibbs equation.
du

ds

=
=
=

T ds P dv,
du P
+ dv,
T  T


1 u
u
P
dv
+
dT
+ dv,
T v T
T v
T






P
1 u
1 u
+
dv +
dT.
T v T
T
T T v
| {z }
|
{z
}
M

(4.38)
(4.39)
(4.40)
(4.41)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

106

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


Again, invoking Eq. (4.22), one gets then



P
1 u

+
T T v T
T v


2


1 u
1 P
P
1 u
+

T T v T 2 v T
T T v T 2


1 P
P
1 u
+

2
T v T
T T v T 2


1 u
,
T T v T

1 2u
,
T vT

(4.43)

0.

(4.44)

(4.42)

This is the condition for an exact ds. Experiment can show if it is true. For example, for an ideal gas,
one finds from experiment that u = u(T ) and P v = RT , so one gets
0+

1R
1 RT
2
T v
T v
0

0,

(4.45)

0.

(4.46)

So ds is exact for an ideal gas. In fact, the relation is verified for so many gases, ideal and non-ideal,
that one simply asserts that ds is exact, rendering s to be path-independent and a state variable.

4.2

Two independent variables

Consider a general implicit function linking three variables, x, y, z:


f (x, y, z) = 0.

(4.47)

In x y z space, this will represent a surface. If the function can be inverted, it will be
possible to write the explicit forms
x = x(y, z),

y = y(x, z),

z = z(x, y).

(4.48)

Differentiating the first two of the Eqs. (4.48) gives




x
x
dx =
dy +
dz,
y z
z y


y
y
dx +
dz.
dy =
x z
z x

(4.49)
(4.50)

Now use Eq. (4.50) to eliminate dy in Eq. (4.49):






y
y
x
x
dx +
dz +
dz,
(4.51)
dx =
y z x z
z x
z y
|
{z
}
=dy




!

x
x y
x y
dx =
+
dz,
(4.52)
1
y z x z
y z z x z y



!


x y
x
x y
0dx + 0dz =
1 dx +
+
dz. (4.53)
y z x z
y z z x z y
|
{z
}
|
{z
}
=0

CC BY-NC-ND. 03 August 2012, J. M. Powers.

=0

107

4.2. TWO INDEPENDENT VARIABLES

Since x and z are independent, so are dx and dz, and the coefficients on each in Eq. (4.53)
must be zero. Therefore from the coefficient on dx in Eq. (4.53)


x y
1 = 0,
(4.54)
y z x z


x y
= 1.
(4.55)
y z x z
So


x
=
y z

1
,

y
x z

and also from the coefficient on dz in Eq. (4.53)





x
x y
+
= 0,
y z z x z y



x
x y
=
,.
z y
y z z x

(4.56)

(4.57)
(4.58)

So




x y z
= 1.
z y x z y x

If one now divides Eq. (4.49) by a fourth differential, dw, one gets


dx
x dy
x dz
=
+
.
dw
y z dw
z y dw

Demanding that z be held constant in Eq.



x
w z

x
w z

y
w z


x w
w z y z

If x = x(y, w), one then gets

(4.60) gives


x y
=
,
y z w z

x
,
=
y z

x
=
.
y z



x
x
dx =
dy +
dw.
y w
w y

(4.59)

(4.60)

(4.61)
(4.62)
(4.63)

(4.64)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

108

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Divide now by dy while holding z constant so






x
x
x w
=
+
.
y z
y w w y y z

(4.65)

These general operations can be applied to a wide variety of thermodynamic operations.


Example 4.4
Apply Eq. (4.65) to a standard P v T system and let


x
T
=
.
y z
v s

(4.66)

So T = x, v = y, and s = z. Let now u = w. So Eq. (4.65) becomes






T
T
T u
=
+
.
v s
v u
u v v s
Now by definition

cv =
so

Now by Eq. (4.26), one has

u
v s

(4.67)


u
,
T v

(4.68)


1
T
= .
u v
cv

= P , so one gets


T
P
T
=
.
v s
v u cv

For an ideal gas, u = u(T ). Inverting, one gets T = T (u), and so



P
T
= .
v s
cv

(4.69)

(4.70)

T
v u

= 0, thus
(4.71)

For an isentropic process in an ideal gas, one gets


dT
dv
dT
T

=
=
=

ln

T
To
T
To

=
=

CC BY-NC-ND. 03 August 2012, J. M. Powers.

RT
P
=
,
cv
cv v
R dv

,
cv v
dv
(k 1) ,
v
vo
(k 1) ln ,
v
 v k1
o
.
v

(4.72)
(4.73)
(4.74)
(4.75)
(4.76)

109

4.3. LEGENDRE TRANSFORMATIONS

4.3

Legendre transformations

The Gibbs equation (4.1), du = P dv + T ds, is the fundamental equation of classical


thermodynamics. It is a canonical form which suggests the most natural set of variables in
which to express internal energy u are s and v:
u = u(v, s).

(4.77)

However, v and s may not be convenient for a particular problem. There may be other
combinations of variables whose canonical form gives a more convenient set of independent
variables for a particular problem. An example is the enthalpy:
h u + P v.

(4.78)

dh = du + P dv + vdP.

(4.79)

Differentiating the enthalpy gives

Use now Eq. (4.79) to eliminate du in the Gibbs equation to give


dh
| P dv
{z vdP} = P dv + T ds.

(4.80)

=du

So

dh = T ds + vdP.

(4.81)

So the canonical variables for h are s and P . One then expects


h = h(s, P ).

(4.82)

This exercise can be systematized with the Legendre transformation,1 which defines a set
of second order polynomial combinations of variables. Consider again the exact differential
Eq. (4.14):
dy = 1 dx1 + 2 dx2 + + N dxN .

(4.83)

Two differentiable functions f and g are said to be Legendre transformations of each other if their first
derivatives are inverse functions of each other: Df = (Dg)1 . With some effort, not shown here, one can
prove that the Legendre transformations of this section satisfy this general condition.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

110

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

For N independent variables xi and N conjugate variables i , by definition there are 2N 1


Legendre transformed variables:
1
2
..
.
N
1,2
1,3
..
.

= 1 (1 , x2 , x3 , . . . , xN ) = y 1 x1 ,
= 2 (x1 , 2 , x3 , . . . , xN ) = y 2 x2 ,

(4.84)
(4.85)

= N (x1 , x2 , x3 , . . . , N ) = y N xN ,
= 1,2 (1 , 2 , x3 , . . . , xN ) = y 1 x1 2 x2 ,
= 1,3 (1 , x2 , 3 , . . . , xN ) = y 1 x1 3 x3 ,

(4.86)
(4.87)
(4.88)

1,...,N = 1,...,N (1 , 2 , 3 , . . . , N ) = y

N
X

i xi .

(4.89)

i=1

Each is a new dependent variable. Each has the property that when it is known as a
function of its N canonical variables, the remaining N variables from the original expression
(the xi and the conjugate i ) can be recovered by differentiation of . In general this is not
true for arbitrary transformations.
Example 4.5
Let y = y(x1 , x2 , x3 ). This has the associated differential form
dy = 1 dx1 + 2 dx2 + 3 dx3 .

(4.90)

Choose now a Legendre transformed variable 1 z(1 , x2 , x3 ):


z = y 1 x1 .
Then

(4.91)




z
z
z
d1 +
dx2 +
dx3 .
dz =
1 x2 ,x3
x2 1 ,x3
x3 1 ,x2

(4.92)

Now differentiating Eq. (4.91), one also gets

dz = dy 1 dx1 x1 d1 .

(4.93)

Elimination of dy in Eq. (4.93) by using Eq. (4.90) gives


dz

dx + 2 dx2 + 3 dx3 1 dx1 x1 d1 ,


|1 1
{z
}

(4.94)

=dy

x1 d1 + 2 dx2 + 3 dx3 .

Thus from Eq. (4.92), one gets



z
,
x1 =
1 x2 ,x3


z
2 =
,
x2 1 ,x3


z
3 =
.
x3 1 ,x2

(4.95)

(4.96)

So the original expression had three independent variables x1 , x2 , x3 , and three conjugate variables
1 , 2 , 3 . Definition of the Legendre function z with canonical variables 1 , x2 , and x3 allowed
determination of the remaining variables x1 , 2 , and 3 in terms of the canonical variables.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

111

4.3. LEGENDRE TRANSFORMATIONS

For the Gibbs equation, (4.1), du = P dv + T ds, one has y = u, two canonical variables,
x1 = v and x2 = s, and two conjugates, 1 = P and 2 = T . Thus N = 2, and one can
expect 22 1 = 3 Legendre transformations. They are

1,2

1 = y 1 x1 = h = h(P, s) = u + P v, enthalpy,
(4.97)
2 = y 2 x2 = a = a(v, T ) = u T s, Helmholtz free energy, (4.98)
= y 1 x1 2 x2 = g = g(P, T ) = u + P v T s, Gibbs free energy.
(4.99)

It has already been shown for the enthalpy that dh = T ds + vdP , so that the canonical
variables are s and P . One then also has


h
h
dh =
ds +
dP,
(4.100)
s P
P s

from which one deduces that


h
,
T =
s P


h
.
v=
P s

(4.101)

From Eq. (4.101), a second Maxwell relation can be deduced by differentiation of the first
with respect to P and the second with respect to s:


T
v
(4.102)
=
.
P s
s P
The relations for Helmholtz and Gibbs free energies each supply additional useful relations
including two new Maxwell relations. First consider the Helmholtz free energy
a = u T s,
da = du T ds sdT,
= (P dv + T ds) T ds sdT,
= P dv sdT.

(4.103)
(4.104)
(4.105)
(4.106)

So the canonical variables for a are v and T . The conjugate variables are P and s. Thus


a
a
da =
dv +
dT.
(4.107)
v T
T v

So one gets


a
,
P =
v T


a
.
s =
T v

(4.108)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

112

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

and the consequent Maxwell relation




P
s
=
.
T v
v T

For the Gibbs free energy

(4.109)

g = |u +{zP v} T s,

(4.110)

=h

= h T s,
dg = dh T ds sdT,
= (T ds + vdP ) T ds sdT,
{z
}
|

(4.111)
(4.112)
(4.113)

=dh

= vdP sdT.

(4.114)

So for Gibbs free energy, the canonical variables are P and T while the conjugate variables
are v and s. One then has g = g(P, T ), which gives


g
g
dg =
dP +
dT.
(4.115)
P T
T P

So one finds


g
v=
,
P T


g
s =
.
T P

The resulting Maxwell function is then

(4.116)



v
s
=
.
T P
P T

(4.117)

Example 4.6
Canonical Form
If
h(s, P ) = cP To

P
Po

R/cP

exp

s
cP

+ (ho cP To ) ,

(4.118)

and cP , To , R, Po , and ho are all constants, derive both thermal and caloric state equations P (v, T )
and u(v, T ).
Now for this material

h
s P

h
P s

= To
=

RTo
Po

CC BY-NC-ND. 03 August 2012, J. M. Powers.

P
Po


R/cP

P
Po

exp

R/cP 1

s
cP

exp

,
s
cP

(4.119)


(4.120)

113

4.3. LEGENDRE TRANSFORMATIONS


Now since

h
s P

h
P s

one has


To

RTo
Po

P
Po


T,

(4.121)

v,

(4.122)

R/cP

P
Po

exp

R/cP 1

s
cP

exp

,
s
cP

(4.123)


(4.124)

Dividing Eq. (4.123) by Eq. (4.124) gives


T
v
Pv

=
=

P
,
R
RT ,

(4.125)
(4.126)

which is the thermal equation of state. Substituting from Eq. (4.123) into the canonical equation for
h, Eq. (4.118), one also finds for the caloric equation of state
h

cP T + (ho cP To ) ,

cP (T To ) + ho ,

which is useful in itself. Substituting in for T and To ,




P v Po vo
+ ho .

h = cP
R
R
Using, Eq. (4.78), h u + P v, we get


Pv
Po vo
u + P v = cP
+ uo + Po vo .

R
R

(4.127)
(4.128)

(4.129)

(4.130)

So
u =
u =
u =
u =
u =
u =

c


c

P
1 Pv
1 Po vo + uo ,
R
 cR

P
1 (P v Po vo ) + uo ,
 cR

P
1 (RT RTo ) + uo ,
R
(cP R) (T To ) + uo ,
P

(cP (cP cv )) (T To ) + uo ,
cv (T To ) + uo .

(4.131)
(4.132)
(4.133)
(4.134)
(4.135)
(4.136)

So one canonical equation gives us all the information one needs. Often, it is difficult to do a single
experiment to get the canonical form.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

114

4.4

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Heat capacity

Recall that
cv =
cP =


u
,
T v

h
.
T

(4.137)
(4.138)

Then perform operations on the Gibbs equation

Likewise,

du = T ds P dv,


u
s
= T
,
T v
T v

s
cv = T
.
T v
dh = T ds + vdP,


s
h
= T
,
T P
T P

s
cP = T
.
T P

(4.139)
(4.140)
(4.141)

(4.142)
(4.143)
(4.144)

One finds further useful relations by operating on the Gibbs equation:

So one can then say

For an ideal gas, one has

du = T ds P dv,


u
s
= T
P,
v T
v T

P
P.
= T
T v
u = u(T, v),


u
u
dT +
dv,
du =
T v
v T



P
= cv dT + T
P dv.
T v



 
P
RT
R
u
= T
P =T

,


v T
T v
v
v
= 0.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(4.145)
(4.146)
(4.147)

(4.148)
(4.149)
(4.150)

(4.151)
(4.152)

115

4.4. HEAT CAPACITY

Consequently, u is not a function of v for an ideal gas, so u = u(T ) alone. Since Eq. (4.78),
h = u + P v, for an ideal gas reduces to h = u + RT
h = u(T ) + RT = h(T ).

(4.153)

Now return to general equations of state. With s = s(T, v) or s = s(T, P ), one gets


s
s
ds =
dT +
dv,
(4.154)
T v
v T


s
s
ds =
dT +
dP.
(4.155)
T P
P T
Now using Eqs. (4.102, 4.117, 4.141, 4.144) one gets


P
cv
dv,
dT +
ds =
T
T v

cP
v
ds =
dT
dP.
T
T P

(4.156)
(4.157)

Subtracting Eq. (4.157) from Eq. (4.156), one finds



P
cv cP
v
dT +
0 =
dv +
dP,
T
T v
T P


v
P
dv + T
dP.
(cP cv )dT = T
T v
T P

(4.158)
(4.159)

Now divide both sides by dT and hold either P or v constant. In either case, one gets


P v
cP cv = T
.
(4.160)
T v T P
Also, since P/T |v = (P/v|T )(v/T |P ), Eq. (4.160) can be rewritten as
cP cv = T

2

v
P
.
T P
v T

(4.161)

Now since T > 0, (v/T |P )2 > 0, and for all known materials P/v|T < 0, we must have
c P > cv .

(4.162)

Example 4.7
For an ideal gas find cP cv .
CC BY-NC-ND. 03 August 2012, J. M. Powers.

116

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


For the ideal gas, P v = RT , one has

P
R
= ,

T v
v


v
R
= .

T P
P

So, from Eq. (4.160), we have

cP cv

=
=
=

RR
,
vP
R2
,
T
RT
R.

(4.163)

(4.164)
(4.165)
(4.166)

This holds even if the ideal gas is calorically imperfect. That is


cP (T ) cv (T ) = R.

(4.167)

For the ratio of specific heats for a general material, one can use Eqs. (4.141) and (4.144)
to get

s
T T
cP
P , then apply Eq. (4.56) to get
k=
(4.168)
=
s
cv
T T v


s T
, then apply Eq. (4.58) to get
(4.169)
=
T P s v





s P
T v
=

,
(4.170)
P T T s
v s s T




v s
P T
=
.
(4.171)
s T P T
T s v s
So for general materials



v P
.
k=
P T v s

(4.172)

The first term can be obtained from P v T data. The second term is related to the
isentropic sound speed of the material, which is also a measurable quantity.
Example 4.8
For a calorically perfect ideal gas with gas constant R and specific heat at constant volume cv ,
find expressions for the thermodynamic variable s and thermodynamic potentials u, h, a, and g, as
functions of T and P .
CC BY-NC-ND. 03 August 2012, J. M. Powers.

117

4.4. HEAT CAPACITY


First get the entropy:
du =
T ds =

T ds P dv,
du + P dv,

(4.173)
(4.174)

T ds =

cv dT + P dv,
dT
P
cv
+
dv,
T
T
dv
dT
+R ,
cv
T
vZ
Z
dT
dv
cv
+ R ,
T
v
v
T
+ R ln ,
cv ln
T0
v0
T
R
RT /P
ln
+
ln
,
T0
cv RT0 /P0
R/cv

 
T P0
T
,
+ ln
ln
T0
T0 P
 1+R/cv
 R/cv
T
P0
ln
,
+ ln
T0
P
 1+(cp cv )/cv
 (cp cv )/cv
T
P0
ln
,
+ ln
T0
P
 k1
 k
P0
T
.
+ ln
ln
T0
P

(4.175)

ds

=
=

ds

s s0

s s0
cv

=
=
=
=
=

(4.176)
(4.177)
(4.178)
(4.179)
(4.180)
(4.181)
(4.182)
(4.183)
(4.184)

So
s = s0 + cv ln

T
T0

k

+ cv ln

P0
P

k1

(4.185)

Now, for the calorically perfect ideal gas, one has


u = uo + cv (T To ).

(4.186)

For the enthalpy, one gets


h

= u + P v,

(4.187)

= u + RT,
= uo + cv (T To ) + RT,

(4.188)
(4.189)

= uo + cv (T To ) + RT + RTo RTo ,
= uo + RTo +cv (T To ) + R(T To ),
| {z }

(4.190)
(4.191)

=ho

= ho + (cv + R)(T To ).
| {z }

(4.192)

h = ho + cp (T To ).

(4.193)

=cp

So

CC BY-NC-ND. 03 August 2012, J. M. Powers.

118

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


For the Helmholtz free energy, one gets
a

u T s.

(4.194)

Thus,
a = uo + cv (T To ) T

s0 + cv ln

T
T0

k

+ cv ln

P0
P

k1 !

(4.195)

For the Gibbs free energy, one gets


g

h T s.

(4.196)

Thus
g = ho + cp (T To ) T

4.5

s0 + cv ln

T
T0

k

+ cv ln

P0
P

k1 !

(4.197)

Van der Waals gas

A van der Waals gas is a common model for a non-ideal gas. It can capture some of the
behavior of a gas as it approaches the vapor dome. Its form is
P (T, v) =

a
RT
2,
vb v

(4.198)

where b accounts for the finite volume of the molecules, and a accounts for intermolecular
forces.
Example 4.9
Find a general expression for u(T, v) if
P (T, v) =

RT
a
.
v b v2

(4.199)

Proceed as before: First we have


du =
recalling that

u
= cv ,
T v



u
u
dT
+
dv,
T v
v T

CC BY-NC-ND. 03 August 2012, J. M. Powers.



u
P
=
T
P.
v T
T v

(4.200)

(4.201)

119

4.5. VAN DER WAALS GAS


Now for the van der Waals gas, we have

P
T v

R
,
vb
RT
P,
vb


a
a
RT
RT

2 = 2.
vb
vb v
v


P
T
P
T v

=
=

(4.202)
(4.203)
(4.204)

So we have

u
v T

a
,
v2
a
+ f (T ).
v

u(T, v) =

(4.205)
(4.206)

Here f (T ) is some as-of-yet arbitrary function of T . To evaluate f (T ), take the derivative with respect
to T holding v constant:

u
df
=
= cv .
T v
dT

(4.207)

Since f is a function of T at most, here cv can be a function of T at most, so we allow cv = cv (T ).


Integrating, we find f (T ) as
f (T ) = C +

cv (T)dT,

(4.208)

To

where C is an integration constant. Thus u is


u(T, v) = C +

To

a
cv (T)dT .
v

(4.209)

Taking C = uo + a/vo , we get

u(T, v) = uo +

To

cv (T)dT + a

1
1

vo
v

(4.210)

We also find

1
1
+ P v,

h = u + P v = uo +
vo
v
To


Z T
RT v
1
a
1
+

.
cv (T)dT + a
h(T, v) = uo +
vo
v
vb v
To
Z

cv (T)dT + a

(4.211)
(4.212)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

120

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Example 4.10
A van der Waals gas with
R =
a

b =
cv

J
,
kg K
P a m6
,
150
kg 2
m3
0.001
,
kg

 

J
J
350
+ 0.2
(T (300 K)),
kg K
kg K 2
200

(4.213)
(4.214)
(4.215)
(4.216)

begins at T1 = 300 K, P1 = 1 105 P a. It is isothermally compressed to state 2 where P2 = 1 106 P a.


It is then isochorically heated to state 3 where T3 = 1000 K. Find 1 w3 , 1 q3 , and s3 s1 . Assume the
surroundings are at 1000 K.
Recall
P =
so at state 1


10 P a =
Expanding, one gets


200

a
RT
,
v b v2

J
kg K

(4.217)

(300 K) 150 P a m
kg2


.
2
3
v1
v1 0.001 m
kg







kg 2
kg 3
kg
2
0.15 + 150 3 v1 60100 6 v1 + 100000 9 v13 = 0.
m
m
m

(4.218)

(4.219)

This is a cubic equation which has three solutions:


v1

v1

v1

0.598

m3
,
kg

physical,

m3
kg
m3
0.00125 + 0.0097i
kg
0.00125 0.0097i

(4.220)
not physical,

(4.221)

not physical.

(4.222)

Now at state 2, P2 and T2 are known, so v2 can be determined:




6
200 kgJ K (300 K) 150 P a m

kg2


.
106 P a =
3
v22
v2 0.001 m
kg

(4.223)

The physical solution is v2 = 0.0585 m3 /kg. Now at state 3 it is known that v3 = v2 and T3 = 1000 K.
Determine P3 :


6
200 kgJ K (1000 K)
150 P akgm
2
 

P3 = 
(4.224)
2 ,
3
3
3
0.0585 m
0.001 m
0.0585 m
kg
kg
kg

(3478261 P a) (43831 P a),


3434430 P a.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(4.225)
(4.226)

121

4.5. VAN DER WAALS GAS

R2
R3
R2
Now 1 w3 = 1 w2 + 2 w3 = 1 P dv + 2 P dv = 1 P dv since 2 3 is at constant volume. So

Z v2 
RT
a
=
dv,
(4.227)

1 w3
v b v2
v1
Z v2
Z v2
dv
dv
= RT1
,
(4.228)
a
2
v1 v
v1 v b




1
1
v2 b
+a
,
(4.229)

= RT1 ln
v1 b
v2
v1
 


3
3


0.0585 m
0.001 m
kg
kg
J
 

=
200
(300 K) ln 
m3
m3
kg K
0.598 kg 0.001 kg
!


1
1
P a m6
,
(4.230)
+ 150
3
3
kg 2
0.0585 m
0.598 m
kg
kg

 

J
J
=
140408
+ 2313
,
(4.231)
kg
kg
J
= 138095
,
(4.232)
kg
=

138.095

kJ
.
kg

(4.233)

The gas is compressed, so the work is negative. Since u is a state property:




Z T3
1
1
.

cv (T )dT + a
u3 u1 =
v1
v3
T1
Now
cv

=
=

so
u3 u1

=
=
=

=
=
=


 

J
J
350
+ 0.2
(T (300 K)),
kg K
kg K 2

 

J
J
290
+ 0.2
T.
kg K
kg K 2

 
 


J
1
1
J
+ 0.2
T
dT
+
a
,

kg K
kg K 2
v1
v3
T1







J
J
1
1
2
2
290
(T3 T1 ) + 0.1
,
T3 T1 + a

kg K
kg K 2
v1
v3





J
J
((1000 K) (300 K)) + 0.1
(1000 K)2 (300 K)2
290
kg K
kg K 2
!


P a m6
1
1
+ 150
,
3
3
kg 2
0.598 m
0.0585 m
kg
kg

 
 

J
J
J
203000
+ 91000
2313
,
kg
kg
kg
J
,
291687
kg
kJ
292
.
kg
Z

T3


290

(4.234)

(4.235)
(4.236)

(4.237)
(4.238)

(4.239)
(4.240)
(4.241)
(4.242)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

122

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


Now from the first law
u3 u1
1 q3
1 q3

1 q3

= 1 q3 1 w3 ,
= u3 u1 + 1 w3 ,
 


kJ
kJ
138
,
=
292
kg
kg
=

154

kJ
.
kg

(4.243)
(4.244)
(4.245)
(4.246)

The heat transfer is positive as heat was added to the system.


Now find the entropy change. Manipulate the Gibbs equation:
T ds =
ds

ds

ds

ds

s3 s1

=
=

=
=

du + P dv,
1
P
du + dv,
T
T
a  P
1
cv (T )dT + 2 dv + dv,
T
v
T

1
a  1
RT
a
cv (T )dT + 2 dv +
dv,

T
v
T v b v2
cv (T )
R
dT +
dv,
T
vb
Z T3
cv (T )
v3 b
dT + R ln
,
T
v
1b
T1





Z 1000
290 kgJ K
J
dT + R ln v3 b ,

+ 0.2
2
T
kg
K
v1 b
300

 
 

J
J
1000 K
290
ln
+ 0.2
((1000 K) (300 K))
kg K
300 K
kg K 2




m3
m3


0.0585

0.001
kg
kg
J
 
,
ln 
+ 200
m3
m3
kg K
0.598 kg 0.001 kg
 
 


J
J
J
+ 140
468
,
349
kg K
kg K
kg K
J
21
,
kg K

0.021

kJ
.
kg K

(4.247)
(4.248)
(4.249)
(4.250)
(4.251)
(4.252)

(4.253)

(4.254)

(4.255)
(4.256)
(4.257)

Is the second law satisfied for each portion of the process?


First look at 1 2
u2 u1
1 q2

=
=

1 q2

1 w2 ,
u2 u1 + 1 w2 ,



! 


Z T2
1
1
1
1
v2 b
cv (T )dT + a
+a
+ RT1 ln
.

v1
v2
v1 b
v2
v1
T1

1 q2

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(4.258)
(4.259)
(4.260)

123

4.5. VAN DER WAALS GAS


Recalling that T1 = T2 and canceling the terms in a, one gets
1 q2


v2 b
,
RT1 ln
v1 b
 


m3
m3


0.0585

0.001
kg
kg
J
 
 ,
(300 K) ln 
200
m3
m3
kg K
0.598
0.001


kg

(4.261)

(4.262)

kg

J
.
140408
kg

(4.263)

Since the process is isothermal,





v2 b
v1 b

3

s2 s1 = R ln

 

m3


0.0585
0.001 m
kg
kg
J

 
 ,
= 200
ln 
m3
m3
kg K
0.598 kg 0.001 kg
= 468.0

(4.264)

(4.265)

J
.
kg K

(4.266)

Entropy drops because heat was transferred out of the system.

Check the second law. Note that in this portion of the process in which the heat is transferred out
of the system, that the surroundings must have Tsurr 300 K. For this portion of the process let us
take Tsurr = 300 K.
s2 s1
J
kg K
J
468.0
kg K

468.0

1 q2

?
T
140408

300 K

468.0

(4.267)
J
kg

(4.268)

J
kg K

ok.

(4.269)

Next look at 2 3
2 q3
2 q3

since isochoric

2 q3

= u3 u2 + 2 w3 ,
! Z v3


Z T3
1
1
+

P dv ,
=
cv (T )dT + a
v2
v3
v2
T2
Z T3
cv (T )dT,
=
=

T2
Z 1000 K
300 K

= 294000


290

J
.
kg

J
kg K


+ 0.2

J
kg K 2

 
T dT,

(4.270)
(4.271)
(4.272)
(4.273)
(4.274)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

124

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


Now look at the entropy change for the isochoric process:
s3 s2

T3

cv (T )
dT ,
T
T2





Z T3
290 kgJ K
J
dT ,

+ 0.2
=
T
kg K 2
T2

 
 

J
J
1000 K
=
290
ln
+ 0.2
((1000 K) (300 K)),
kg K
300 K
kg K 2
J
= 489
.
kg K
=

(4.275)

(4.276)
(4.277)
(4.278)

Entropy rises because heat transferred into system.


In order to transfer heat into the system we must have a different thermal reservoir. This one must
have Tsurr 1000 K. Assume here that the heat transfer was from a reservoir held at 1000 K to assess
the influence of the second law.
s3 s2

J
kg K
J
489
kg K

489

4.6

2 q3

?
T
294000

(4.279)
J
kg

,
1000 K
J
294
,
kg K

(4.280)
ok.

(4.281)

Redlich-Kwong gas

The Redlich-Kwong equation of state is


P =

RT
a

.
v b v(v + b)T 1/2

(4.282)

It is modestly more accurate than the van der Waals equation in predicting material behavior.
Example 4.11
For the case in which b = 0, find an expression for u(T, v) consistent with the Redlich-Kwong state
equation.
Here the equation of state is now
P =

a
RT
2 1/2 .
v
v T

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(4.283)

125

4.7. COMPRESSIBILITY AND GENERALIZED CHARTS


Proceeding as before, we have


u
P
= T
P,
v T
T v
 


a
a
RT
R
+ 2 3/2
2 1/2 ,
= T
v
v
2v T
v T
3a
=
.
2v 2 T 1/2

(4.284)
(4.285)
(4.286)

Integrating, we find
u(T, v) =

3a
+ f (T ).
2vT 1/2

(4.287)

Here f (T ) is a yet-to-be-specified function of temperature only. Now the specific heat is found by the
temperature derivative of u:

3a
df
u
=
.
(4.288)
+
cv (T, v) =

3/2
T v
dT
4vT

Obviously, for this material, cv is a function of both T and v.


Let us define cvo (T ) via
df
cvo (T ).
dT

(4.289)

(4.290)

Integrating, then one gets


f (T ) = C +

cvo (T) dT.

To

1/2

Let us take C = uo +3a/2/vo/To . Thus we arrive at the following expressions for cv (T, v) and u(T, v):
cv (T, v) =

u(T, v) = uo +

To

4.7

cvo (T ) +

3a
cvo (T) dT +
2

3a
,
4vT 3/2

1
1/2

vo To

vT 1/2

(4.291)
!

(4.292)

Compressibility and generalized charts

A simple way to quantify the deviation from ideal gas behavior is to determine the so-called
compressibility Z, where
Pv
.
(4.293)
Z
RT
CC BY-NC-ND. 03 August 2012, J. M. Powers.

126

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

For an ideal gas, Z = 1. For substances with a simple molecular structure, Z can be
tabulated as functions of the so-called reduced pressure Pr and reduced temperature Tr . Tr
and Pr are dimensionless variables found by scaling their dimensional counterparts by the
specific materials temperature and pressure at the critical point, Tc and Pc :
Tr

T
,
Tc

Pr

P
.
Pc

(4.294)

Often charts are available which give predictions of all reduced thermodynamic properties.
These are most useful to capture the non-ideal gas behavior of materials for which tables are
not available.

4.8

Mixtures with variable composition

Consider now mixtures of N species. The focus here will be on extensive properties and
molar properties.
Assume that each species has ni moles, and the the total number of moles
PN
is n = i=1 ni . Now one might expect the extensive energy to be a function of the entropy,
the volume and the number of moles of each species:
U = U(V, S, ni ).

(4.295)

The extensive version of the Gibbs law in which all of the ni are held constant is
dU = P dV + T dS.
Thus


U
= P,
V S,ni

(4.296)


U
= T.
S V,ni

(4.297)

In general, since U = U(S, V, ni ), one should expect, for systems in which the ni are allowed
to change that



N
X
U
U
U
dU =
dV +
dS +
dni .
(4.298)
V S,ni
S V,ni
ni S,V,nj
i=1
Defining the new thermodynamics property, the chemical potential i , as

U
,
i
ni S,V,nj

(4.299)

one has the important Gibbs equation for multicomponent systems:


dU = P dV + T dS +
CC BY-NC-ND. 03 August 2012, J. M. Powers.

N
X
i=1

i dni .

(4.300)

127

4.8. MIXTURES WITH VARIABLE COMPOSITION

Obviously, by its definition, i is on a per mole basis, so it is given the appropriate overline
notation. In Eq. (4.300), the independent variables and their conjugates are
x1
x2
x3
x4
xN +2

=
=
=
=

V,
S,
n1 ,
n2 ,
..
.
= nN ,

1 = P,
2 = T,
3 = 1 ,
4 = 2 ,
..
.
N +2 = N .

(4.301)
(4.302)
(4.303)
(4.304)
(4.305)

Equation (4.300) has 2N +1 1 Legendre functions. Three are in wide usage: the extensive
analog to those earlier found. They are
H = U + P V,
A = U T S,
G = U + P V T S.

(4.306)
(4.307)
(4.308)

A set of non-traditional, but perfectly acceptable additional Legendre functions would be


formed from U 1 n1 . Another set would be formed from U + P V 2 n2 . There are
many more, but one in particular is sometimes noted in the literature: the so-called grand
potential, . The grand potential is defined as
U TS

N
X

i ni .

(4.309)

i=1

Differentiating each defined Legendre function, Eqs. (4.306-4.309), and combining with
Eq. (4.300), one finds
dH = T dS + V dP +

N
X

i dni ,

i=1
N
X

dA = SdT P dV +
dG = SdT + V dP +

d = P dV SdT

i=1
N
X

i=1
N
X

(4.310)

i dni ,

(4.311)

i dni ,

(4.312)

ni di .

(4.313)

i=1

Thus, canonical variables for H are H = H(S, P, ni). One finds a similar set of relations as
CC BY-NC-ND. 03 August 2012, J. M. Powers.

128

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

before from each of the differential forms:




U
H
T =
=
,
S V,ni
S P,ni



A

U
=
=
,
P =
V S,ni
V T,ni
V T,i


G
H
=
,
V =
P S,ni
P T,ni



G

A
=
=
S =
,
T V,ni
T P,ni
T V,i


.
ni =
i V,T,j





H
A
G
U
=
=
=
i =
ni S,V,nj
ni S,P,nj
ni T,V,nj
ni T,P,nj

(4.314)
(4.315)
(4.316)
(4.317)
(4.318)

(4.319)

Each of these induces a corresponding Maxwell relation, obtained by cross differentiation.


These are


T
P
=
,
(4.320)
V S,ni
S V,ni


V
T
=
,
(4.321)
P S,ni
S P,ni


S
P
=
,
(4.322)
T V,ni
V T,ni


V
S
=
,
(4.323)
T P,ni
P T,ni


S
i
=
,
(4.324)
T P,nj
ni V,nj


i
V
=
,
(4.325)
P T,nj
ni V,nj


l
k
=
,
(4.326)
nk T,P,nj
nl T,P,nj


P
S
=
,
(4.327)
V T,i
T V,i


ni
nk
=
(4.328)
k V,T,j ,j6=k
i V,T,j ,j6=i
CC BY-NC-ND. 03 August 2012, J. M. Powers.

129

4.9. PARTIAL MOLAR PROPERTIES

4.9

Partial molar properties

4.9.1

Homogeneous functions

In mathematics, a homogeneous function f (x1 , . . . , xN ) of order m is one such that


f (x1 , . . . , xN ) = m f (x1 , . . . , xN ).

(4.329)

f (x1 , . . . , xN ) = f (x1 , . . . , xN ).

(4.330)

If m = 1, one has
Thermodynamic variables are examples of homogeneous functions.

4.9.2

Gibbs free energy

Consider an extensive property, such as the Gibbs free energy G. One has the canonical
form
G = G(T, P, n1 , n2 , . . . , nN ).
(4.331)
One would like to show that if each of the mole numbers ni is increased by a common factor,
say , with T and P constant, that G increases by the same factor :
G(T, P, n1 , n2 , . . . , nN ) = G(T, P, n1 , n2 , . . . , nN ).

(4.332)

Differentiate both sides of Eq. (4.332) with respect to , while holding P , T , and nj constant,
to get
G(T, P, n1 , n2 , . . . , nN ) =



G
G
G
d(n1 )
d(n2 )
d(nN )
+
++
,



(n1 ) nj ,P,T d
(n2 ) nj ,P,T d
(nN ) nj ,P,T d



G
G
G
=
n1 +
n2 + +
nN ,
(n1 )
(n2 )
(nN )
nj ,P,T

nj ,P,T

(4.333)
(4.334)

nj ,P,T

This must hold for all , including = 1, so one requires





G
G
G
n1 +
n2 + +
nN ,
G(T, P, n1, n2 , . . . , nN ) =
n1 nj ,P,T
n2 nj ,P,T
nN nj ,P,T

N
X
G
ni .
=

n
i
n
,P,T
j
i=1

(4.335)

(4.336)

Recall now the definition partial molar property, the derivative of an extensive variable with
respect to species ni holding nj , i 6= j, T , and P constant. Because the result has units per
CC BY-NC-ND. 03 August 2012, J. M. Powers.

130

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

mole, an overline superscript is utilized. The partial molar Gibbs free energy of species i, g i
is then

G
,
(4.337)
gi
ni nj ,P,T
so that

G=

N
X

g i ni .

(4.338)

i=1

Using the definition of chemical potential, Eq. (4.319), one also notes then that
G(T, P, n1 , n2 , . . . , nN ) =

N
X

i ni .

(4.339)

i=1

The temperature and pressure dependence of G must lie entirely within i (T, P, nj ), which
one notes is also allowed to be a function of nj as well. Consequently, one also sees that the
Gibbs free energy per unit mole of species i is the chemical potential of that species:
g i = i .

(4.340)

Using Eq. (4.338) to eliminate G in Eq. (4.308), one recovers an equation for the energy:
U = P V + T S +

4.9.3

N
X

i ni .

(4.341)

i=1

Other properties

A similar result also holds for any other extensive property such as V , U, H, A, or S. One
can also show that

N
X
V
,
(4.342)
ni
V =

n
i
n
,A,T
j
i=1

N
X U

U =
ni
(4.343)

n
i
n
,V,S
j
i=1

N
X H

,
(4.344)
H =
ni

n
i
nj ,P,S
i=1

N
X
A
,
(4.345)
A =
ni
ni nj ,T,V
i=1

N
X
S
ni
S =
.
(4.346)
ni nj ,U,T
i=1
CC BY-NC-ND. 03 August 2012, J. M. Powers.

131

4.9. PARTIAL MOLAR PROPERTIES

Note that these expressions do not formally involve partial molar properties since P and T
are not constant.
Take now the appropriate partial molar derivatives of G for an ideal mixture of ideal
gases to get some useful relations:
G = H T S,



G
H
S
=
T
.
ni T,P,nj
ni T,P,nj
ni T,P,nj

(4.347)
(4.348)

Now from the definition of an ideal mixture hi = hi (T, P ), so one has

H =

N
X

nk hk (T, P ),

k=1

H
=

ni T,P,nj
ni
=

N
X

N
X

(4.349)
!

nk hk (T, P ) ,

k=1

nk
hk (T, P ),
ni
k=1 |{z}

N
X

(4.350)
(4.351)

=ik

ik hk (T, P ),

(4.352)

k=1

= hi (T, P ).

(4.353)

Here, the Kronecker delta function ki has been again used. Now for an ideal gas one further
has hi = hi (T ). The analysis is more complicated for the entropy, in which

S =
=
=
=

 

Pk
o
nk sT,k R ln
,
Po
k=1



N
X
yk P
o
,
nk sT,k R ln
P
o
k=1
!!
 
N
X
P
nk
o
nk sT,k R ln
,
R ln PN
Po
q=1 nq
k=1
!

 
N
N
X
X
P
n
k
nk soT,k R ln
R
nk ln PN
,
Po
q=1 nq
k=1
k=1

N
X

(4.354)
(4.355)
(4.356)
(4.357)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

132

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS




 
N
X
P
S
o
nk sT,k R ln
=

ni T,P,nj
ni k=1
Po
!!

N
X
nk

R
,
nk ln PN
ni T,P,nj
q=1 nq
k=1

 
N
X
nk
P
o
=
sT,k R ln
ni
Po
k=1 |{z}

(4.358)

=ik

!!

N
X
nk

R
,
nk ln PN
ni T,P,nj
n
q
q=1
k=1

 

N
X

P
o
sT,i R ln
=
nk ln
R
Po
ni T,P,nj k=1

(4.359)
nk

PN

q=1

nq

!!

.(4.360)

Evaluation of the final term on the right side requires closer examination, and in fact, after
tedious but straightforward analysis, yields a simple result which can easily be verified by
direct calculation:
!!
!

N
X
nk
ni

nk ln PN
= ln PN
.
(4.361)
ni T,P,nj k=1
q=1 nq
q=1 nq
So the partial molar entropy is in fact
!

 
S
P
n
i
= soT,i R ln
R ln PN
,
ni T,P,nj
Po
q=1 nq
 
P
o
R ln yi ,
= sT,i R ln
Po
 
Pi
o
= sT,i R ln
,
Po
= si

(4.362)
(4.363)
(4.364)
(4.365)

Thus, one can in fact claim for the ideal mixture of ideal gases that
g i = hi T si .

4.9.4

(4.366)

Relation between mixture and partial molar properties

A simple analysis shows how the partial molar property for an individual species is related
to the partial molar property for the mixture. Consider, for example, the Gibbs free energy.
The mixture-averaged Gibbs free energy per unit mole is
g=
CC BY-NC-ND. 03 August 2012, J. M. Powers.

G
.
n

(4.367)

4.10. IRREVERSIBLE ENTROPY PRODUCTION IN A CLOSED SYSTEM


Now take a partial molar derivative and analyze


g
1 G
=

ni T,P,nj
n ni T,P,nj

1 G

=
n ni T,P,nj

1 G
=

n ni T,P,nj

1 G

=
n ni T,P,nj

1 G

=
n ni T,P,nj
1
g
gi .
=
n
n

133

to get

G n
,
n2 ni T,P,nj
!

N
X
G
nk ,
n2 ni T,P,nj k=1

N
G X nk
,
n2
ni T,P,nj
G
n2

k=1
N
X

ik ,

(4.368)
(4.369)
(4.370)
(4.371)

k=1

G
,
n2

(4.372)
(4.373)

Multiplying by n and rearranging, one gets



g
.
gi = g + n
ni T,P,nj

(4.374)

A similar result holds for other properties.

4.10

Irreversible entropy production in a closed system

Consider a multicomponent thermodynamic system closed to mass exchanges with its surroundings coming into equilibrium. Allow the system to be exchanging work and heat with
its surroundings. Assume the temperature difference between the system and its surroundings is so small that both can be considered to be at temperature T . If Q is introduced
into the system, then the surroundings suffer a loss of entropy:
dSsurr =

Q
.
T

(4.375)

The systems entropy S can change via this heat transfer, as well as via other internal
irreversible processes, such as internal chemical reaction. The second law of thermodynamics
requires that the entropy change of the universe be positive semi-definite:
dS + dSsurr 0.

(4.376)

Eliminating dSsurr , one requires for the system that


dS

Q
.
T

(4.377)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

134

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Consider temporarily the assumption that the work and heat transfer are both reversible.
Thus, the irreversible entropy production must be associated with internal chemical reaction.
Now the first law for the entire system gives
dU = Q W,
= Q P dV,
Q = dU + P dV.

(4.378)
(4.379)
(4.380)

Note because the system is closed, there can be no species entering or exiting, and so there
is no change dU attributable to dni . While within the system the dni may not be 0, the net
contribution to the change in total internal energy is zero. A non-zero dni within the system
simply re-partitions a fixed amount of total energy from one species to another. Substituting
Eq. (4.380) into Eq. (4.377) to eliminate Q, one gets
dS

1
(dU + P dV ),
{z
}
T|

(4.381)

=Q

T dS dU P dV
dU T dS + P dV

0,
0.

(4.382)
(4.383)

Eq. (4.383) involves properties only and need not require assumptions of reversibility for
processes in its derivation. In special cases, it reduces to simpler forms.
For processes which are isentropic and isochoric, the second law expression, Eq. (4.383),
reduces to
dU|S,V 0.
(4.384)
For processes which are isoenergetic and isochoric, the second law expression, Eq. (4.383),
reduces to
dS|U,V 0.
(4.385)
Now using Eq. (4.300) to eliminate dS in Eq. (4.385), one can express the second law as


!
N

1
P
1X

0,
(4.386)
dU + dV
i dni

T
T
T i=1
{z
}
|
=dS

1
T

U,V

N
X
i=1

{z

i dni

0.

(4.387)

irreversible entropy production

The irreversible entropy production associated with the internal chemical reaction must be
the left side of Eq. (4.387). Often the irreversible entropy production is defined as , with
CC BY-NC-ND. 03 August 2012, J. M. Powers.

4.10. IRREVERSIBLE ENTROPY PRODUCTION IN A CLOSED SYSTEM

135

the second law requiring d 0. Equation (4.387) in terms of d is


N
1X
dni 0.
d =
T i=1 i

(4.388)

Now, while most standard texts focusing on equilibrium thermodynamics go to great lengths
to avoid the introduction of time, it really belongs in a discussion describing the approach
to equilibrium. One can divide Eq. (4.387) by a positive time increment dt to get

N
1 X dni
0.

T i=1 i dt

(4.389)

Since T 0, one can multiply Eq. (4.389) by T to get


N
X

i=1

dni
0.
dt

(4.390)

i
This will hold if a model for dn
is employed which guarantees that the left side of Eq. (4.390)
dt
is negative semi-definite. One will expect then for dni /dt to be related to the chemical
potentials i .
Elimination of dU in Eq. (4.383) in favor of dH from dH = dU + P dV + V dP gives

|dH P dV
{z V dP} T dS + P dV

0,

(4.391)

dH V dP T dS 0.

(4.392)

=dU

Thus, one finds for isobaric, isentropic equilibration that


dH|P,S 0.

(4.393)

For the Helmholtz and Gibbs free energies, one analogously finds
dA|T,V 0,
dG|T,P 0.

(4.394)
(4.395)

The expression of the second law in terms of dG is especially useful as it may be easy in an
experiment to control so that P and T are constant. This is especially true in an isobaric
phase change, in which the temperature is guaranteed to be constant as well.
Now one has
G =
=

N
X

i=1
N
X

ni g i ,

(4.396)

ni i .

(4.397)

i=1

CC BY-NC-ND. 03 August 2012, J. M. Powers.

136

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

P
One also has from Eq. (4.312): dG = SdT + V dP + N
i=1 i dni , holding T and P constant
that
N
X
dG|T,P =
i dni .
(4.398)
i=1

Here the dni are associated entirely with internal chemical reactions. Substituting Eq. (4.398)
into Eq. (4.395), one gets the important version of the second law which holds that
dG|T,P =

N
X
i=1

i dni 0.

(4.399)

In terms of time rates of change, one can divide Eq. (4.399) by a positive time increment
dt > 0 to get

N
X
dni
G
i
(4.400)
=
0.

t T,P
dt
i=1

4.11

Equilibrium in a two-component system

A major task of non-equilibrium thermodynamics is to find a functional form for dni /dt
which guarantees satisfaction of the second law, Eq. (4.400) and gives predictions which
agree with experiment. This will be discussed in more detail in the following chapter on
thermochemistry. At this point, some simple examples will be given in which a nave but
useful functional form for dni /dt is posed which leads at least to predictions of the correct
equilibrium values. A much better model which gives the correct dynamics in the time
domain of the system as it approaches equilibrium will be presented in the chapter on
thermochemistry.

4.11.1

Phase equilibrium

Here, consider two examples describing systems in phase equilibrium.


Example 4.12
Consider an equilibrium two-phase mixture of liquid and vapor H2 O at T = 100 C, x = 0.5. Use
the steam tables to check if equilibrium properties are satisfied.
In a two-phase gas liquid mixture one can expect the following reaction:
H2 O(l) H2 O(g) .

(4.401)

That is one mole of liquid, in the forward phase change, evaporates to form one mole of gas. In the
reverse phase change, one mole of gas condenses to form one mole of liquid.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

137

4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM

Because T is fixed at 100 C and the material is a two phase mixture, the pressure is also fixed at a
constant. Here there are two phases at saturation; g for gas and l for liquid. Equation (4.399) reduces
to
l dnl + g dng 0.
(4.402)

Now for the pure H2 O if a loss of moles from one phase must be compensated by the addition to
another. So one must have
dnl + dng = 0.
(4.403)
Hence
So Eq. (4.402), using Eq. (4.404) becomes

dng = dnl .

l dnl g dnl

0,

dnl (l g ) 0.

(4.404)

(4.405)
(4.406)

At this stage of the analysis, most texts, grounded in equilibrium thermodynamics, assert that l = g ,
ignoring the fact that they could be different but dnl could be zero. That approach will not be taken
here. Instead divide Eq. (4.406) by a positive time increment, dt 0 to write the second law as
dnl
( g )
dt l

0.

(4.407)

One convenient, albeit nave, way to guarantee second law satisfaction is to let
dnl
= (l g ),
dt

0,

convenient, but nave model

(4.408)

Here is some positive semi-definite scalar rate constant which dictates the time scale of approach to
equilibrium. Note that Eq. (4.408) is just a hypothesized model. It has no experimental verification; in
fact, other more complex models exist which both agree with experiment and satisfy the second law.
For the purposes of the present argument, however, Eq. (4.408) will suffice. With this assumption, the
second law reduces to
(l g )2 0,
0,
(4.409)
which is always true.
Eq. (4.408) has two important consequences:

Differences in chemical potential drive changes in the number of moles.

The number of moles of liquid, nl , increases when the chemical potential of the liquid is less than that
of the gas, l < g . That is to say, when the liquid has a lower chemical potential than the gas, the
gas is driven towards the phase with the lower potential. Because such a phase change is isobaric and
isothermal, the Gibbs free energy is the appropriate variable to consider, and one takes = g. When
this is so, the Gibbs free energy of the mixture, G = nl l + ng g is being driven to a lower value. So
when dG = 0, the system has a minimum G.

The system is in equilibrium when the chemical potentials of liquid and gas are equal: l = g .
The chemical potentials, and hence the molar specific Gibbs free energies must be the same for
each constituent of the binary mixture at the phase equilibrium. That is
gl = gg .

(4.410)

Now since both the liquid and gas have the same molecular mass, one also has the mass specific Gibbs
free energies equal at phase equilibrium:
gl = gg .

(4.411)
CC BY-NC-ND. 03 August 2012, J. M. Powers.

138

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


This can be verified from the steam tables, using the definition g = h T s. From the tables


kJ
kJ
kJ
((100 + 273.15) K) 1.3068
,
= 68.6
gl = hl T sl = 419.02
kg
kg K
kg


kJ
kJ
kJ
gg = hg T sg = 2676.05
= 68.4
((100 + 273.15) K) 7.3548
.
kg
kg K
kg

(4.412)
(4.413)

The two values are essentially the same; the difference is likely due to table inaccuracies.

Example 4.13
This example is from BS, 16.20, p. 700. A container has liquid water at 20 C, 100 kP a, in
equilibrium with a mixture of water vapor and dry air, also at 20 C, 100 kP a. Find the water vapor
pressure and the saturated water vapor pressure.
Now at this temperature, the tables easily show that the pressure of a saturated vapor is Psat =
2.339 kP a . From the previous example, it is known that for the water liquid and vapor in equilibrium,
one has
gliq = gvap .
(4.414)
Now if both the liquid and the vapor were at the saturated state, they would be in phase equilibrium
and that would be the end of the problem. But they have slight deviations from the saturated state.
One can estimate these deviations with the standard formula
dg = sdT + vdP.

(4.415)

The tables will be used to get values at 20 C, for which one can take dT = 0. This allows the
approximation of dg v dP . So for the liquid,
Z
gliq gf =
v dP vf (P Psat ),
(4.416)
gliq

= gf + vf (P Psat ).

For the vapor, approximated here as an ideal gas, one has


Z
gvap gg =
v dP,
Z
dP
,
= RT
P
Pvap
= RT ln
,
Psat
Pvap
gvap = gg + RT ln
.
Psat

(4.417)

(4.418)
(4.419)
(4.420)
(4.421)

Here, once again, one allows for deviations of the pressure of the vapor from the saturation pressure.
Now at equilibrium, one enforces gliq = gvap , so one has
Pvap
.
gf + vf (P Psat ) = gg + RT ln
Psat
{z
}
|
{z
}
|
=gliq
=gvap

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(4.422)

139

4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM

Pvap(Pa)

10

10

10

10

10

10

10

10

P (Pa)

Figure 4.1: Pressure of water vapor as a function of total pressure for example problem.
Now gf = gg , so one gets
vf (P Psat ) = RT ln

Pvap
.
Psat

(4.423)

Solving for Pvap , one gets


Pvap

=
=


vf (P Psat )
,
Psat exp
RT



3
0.001002 m
kg (100 kP a 2.339 kP a)
,


(2.339 kP a) exp
0.4615 kgkJK (293.15 K)


2.3407 kP a.

(4.424)

(4.425)
(4.426)

The pressure is very near the saturation pressure. This justifies assumptions that for such mixtures, one
can take the pressure of the water vapor to be that at saturation if the mixture is in equilibrium. If the
pressure is higher, the pressure of the vapor becomes higher as well. Figure 4.1 shows how the pressure
of the equilibrium vapor pressure varies with total pressure. Clearly, a very high total pressure, on
the order of 1 GP a is needed to induce the vapor pressure to deviate significantly from the saturation
value.

4.11.2

Chemical equilibrium: introduction

Here consider two examples which identify the equilibrium state of a chemically reactive
system.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

140

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

4.11.2.1

Isothermal, isochoric system

The simplest system to consider is isothermal and isochoric. The isochoric assumption
implies there is no work in coming to equilibrium.
Example 4.14
At high temperatures, collisions between diatomic nitrogen molecules induce the production of
monatomic nitrogen molecules. The chemical reaction can be described by the model
N2 + N2 2N + N2 .

(4.427)

Here one of the N2 molecules behaves as an inert third body. An N2 molecule has to collide with
something, to induce the reaction. Some authors leave out the third body and write instead N2 2N ,
but this does not reflect the true physics as well. The inert third body is especially important when
the time scales of reaction are considered. It plays no role in equilibrium chemistry.
Consider 1 kmole of N2 and 0 kmole of N at a pressure of 100 kP a and a temperature of 6000 K.
Using the ideal gas tables, find the equilibrium concentrations of N and N2 if the equilibration process
is isothermal and isochoric.
The ideal gas law can give the volume.
P1
V

nN2 RT
,
V
nN2 RT
,
=
P1
kJ
(1 kmole) 8.314 kmole
=
100 kP a
= 498.84 m3 .

(4.428)

(4.429)
K

(6000 K)

(4.430)
(4.431)

Initially, the mixture is all N2 , so its partial pressure is the total pressure, and the initial partial pressure
of N is 0.
Now every time an N2 molecule reacts and thus undergo a negative change, 2 N molecules are
created and thus undergo a positive change, so
dnN2 =

1
dnN .
2

(4.432)

This can be parameterized by a reaction progress variable , also called the degree of reaction, defined
such that
= dnN2 ,
1
=
dnN .
2

d
d

(4.433)
(4.434)

As an aside, one can integrate this, taking = 0 at the initial state to get

nN2 |t=0 nN2 ,


1
nN .
2

(4.435)

nN2 |t=0 ,

(4.437)

(4.436)

Thus
nN 2

nN

CC BY-NC-ND. 03 August 2012, J. M. Powers.

2.

(4.438)

4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM

141

One can also eliminate to get nN in terms of nN2 :


nN = 2 ( nN2 |t=0 nN2 ) .

(4.439)

Now for the reaction, one must have, for second law satisfaction, that
N2 dnN2 + N dnN
N2 (d) + N (2d)

N2 + 2N d
 d
N2 + 2N
dt

0,

(4.440)

0,
0

(4.441)
(4.442)

0.

(4.443)

In order to satisfy the second law, one can usefully, but navely, hypothesize that the non-equilibrium
reaction kinetics are given by
d
= k(N2 + 2N ),
dt

k 0,

convenient, but nave model

(4.444)

Note there are other ways to guarantee second law satisfaction. In fact, a more complicated model is
well known to fit data well, and will be studied later. For the present purposes, this nave model will
suffice. With this assumption, the second law reduces to
k(N2 + 2N )2 0,

k 0,

(4.445)

which is always true. Obviously, the reaction ceases when d/dt = 0, which holds only when
2N = N2 .

(4.446)

Away from equilibrium, for the reaction to go forward, one must expect d/dt > 0, and then one
must have
N2 + 2N 0,
2N N2 .

(4.447)
(4.448)

The chemical potentials are the molar specific Gibbs free energies; thus, for the reaction to go forward,
one must have
2gN g N2 .
(4.449)
Substituting using the definitions of Gibbs free

2 h N T sN




yN P
2 hN T soT,N R ln
Po


2 hN T soT,N hN2 T soT,N2

energy, one gets




2 hN T soT,N + hN2 T soT,N2




2 hN T soT,N + hN2 T soT,N2




2 hN T soT,N + hN2 T soT,N2

h N 2 T sN 2 ,



yN2 P
hN2 T soT,N2 R ln
,
Po




yN P
yN2 P
2RT ln
+ RT ln
,
Po
Po




yN P
yN2 P
2RT ln
RT ln
,
Po
Po

 2 2
yN P Po
,
RT ln
Po2 P yN2

 2
yN P
.
RT ln
yN2 Po

(4.450)
(4.451)
(4.452)
(4.453)
(4.454)
(4.455)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

142

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


At the initial state, one has yN = 0, so the right hand side approaches , and the inequality holds.
At equilibrium, one has equality.


2 hN T soT,N + hN2 T soT,N2
=

RT ln

2
P
yN
yN2 Po

(4.456)

Taking numerical values from Table A.9:





kJ
kJ
5
+
2 5.9727 10
(6000 K) 216.926
kmole
kg K



kJ
kJ
2.05848 105
(6000 K) 292.984
kmole
kg K
=
2.87635
| {z } =
ln KP

0.0563399
| {z } =
KP


 2


kJ
yN P
(6000 K) ln
8.314
,
kmole K
yN2 Po
 2

yN P
ln
,
(4.457)
yN2 Po
2
yN
P
,
yN2 Po

nN
nN +nN2

(4.458)
2

nN2
nN +nN2

 (nN + nN2 )

RT
,
Po V

(4.459)

n2N RT
,
nN2 Po V

(2 ( nN2 |t=0 nN2 )) RT


,
nN 2
Po V

(2 (1 kmole nN2 ))2 (8.314)(6000)


. (4.462)
nN 2
(100)(498.84)

(4.460)
2

(4.461)

This is a quadratic equation for nN2 . It has two roots


nN2 = 0.888154 kmole

physical;

nN2 = 1.12593 kmole,

non-physical

(4.463)

The second root generates more N2 than at the start, and also yields non-physically negative nN =
0.25186 kmole. So at equilibrium, the physical root is
nN = 2(1 nN2 ) = 2(1 0.888154) = 0.223693 kmole.

(4.464)

The diatomic species is preferred.


Note in the preceding analysis, the term KP was introduced. This is the so-called equilibrium
constant which is really a function of temperature. It will be described in more detail later, but one
notes that it is commonly tabulated for some reactions. Its tabular value can be derived from the more
fundamental quantities shown in this example. BSs Table A.11 gives for this reaction at 6000 K the
value of ln KP = 2.876. Note that KP is fundamentally defined in terms of thermodynamic properties
for a system which may or may not be at chemical equilibrium. Only at chemical equilibrium, can it
can further be related to mole fraction and pressure ratios.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

143

4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM


The pressure at equilibrium is
P2

=
=
=

(nN2 + nN )RT
,
V
(0.888154 kmole + 0.223693 kmole) 8.314
498.84
111.185 kP a.

(4.465)
kJ
kmole K

(6000 K)

(4.466)
(4.467)

The pressure has increased because there are more molecules with the volume and temperature being
equal.
The molar concentrations i at equilibrium, are
N

kmole
0.223693 kmole
mole
= 4.484 104
= 4.484 107
,
3
3
498.84 m
m
cm3

(4.468)

N 2

kmole
0.888154 kmole
mole
= 1.78044 103
= 1.78044 106
.
3
3
498.84 m
m
cm3

(4.469)

Now consider the heat transfer. One knows for the isochoric process that 1Q2 = U2 U1 . The
initial energy is given by
U1

nN2 uN2 ,

(4.470)

nN2 (hN2 RT ),

(1 kmole) 2.05848 105

(4.471)

=
=

1.555964 105 kJ.




kJ
kJ
(6000 K) ,
8.314
kmole
kmole K

(4.472)
(4.473)

The energy at the final state is


U2

=
=
=

nN 2 u N 2 + nN u N ,
nN2 (hN2 RT ) + nN (hN RT ),




kJ
kJ
(6000 K)
8.314
(0.888154 kmole) 2.05848 105
kmole
kmole K




kJ
kJ
+(0.223693 kmole) 5.9727 105
(6000 K) ,
8.314
kmole
kmole K
2.60966 105 kJ.

(4.474)
(4.475)
(4.476)
(4.477)
(4.478)

So
1Q2

=
=
=

U2 U1 ,
2.60966 105 kJ 1.555964 105 kJ,
1.05002 105 kJ.

(4.479)
(4.480)
(4.481)

Heat needed to be added to keep the system at the constant temperature. This is because the nitrogen
dissociation process is endothermic.
One can check for second law satisfaction in two ways. Fundamentally, one can demand that
Eq. (4.377), dS Q/T , be satisfied for the process, giving
S2 S1

Q
.
T

(4.482)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

144

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS


For this isothermal process, this reduces to
S2 S1

1Q2

(4.483)

(4.484)

(4.485)

(4.486)

(4.487)

(nN2 sN2 + nN sN )|2


(nN2 sN2 + nN sN )|1








yN2 P
yN P
o
o

+ nN sT,N R ln
nN2 sT,N2 R ln

Po
Po






 2

yN2 P
yN P

nN2 soT,N2 R ln
+ nN soT,N R ln

Po
Po






 1

PN2
PN

nN2 soT,N2 R ln
+ nN soT,N R ln

Po
Po


 2





PN2
PN

+ nN soT,N R ln
nN2 soT,N2 R ln

Po
Po
1








nN2 RT
nN RT
o
o

nN2 sT,N2 R ln
+ nN sT,N R ln

Po V
Po V
2








n
n
RT
RT
N
N
2
o
o

+ nN sT,N R ln
nN2 sT,N2 R ln

|{z}
Po V
Po V

=0

1Q2

1Q2

1Q2

1Q2

Now at the initial state nN = 0 kmole, and RT /Po /V has a constant value of

kJ
8.314 kmole
RT
1
K (6000 K)
=
=1
,
Po V
(100 kP a)(498.84 m3 )
kmole
so Eq. (4.487) reduces to



nN2 soT,N2 R ln

 n


nN2 

N
+ nN soT,N R ln

1 kmole
1 kmole
2



 n

N
2
nN2 soT,N2 R ln

1 kmole
1

(4.488)

19.4181

kJ
kJ
17.5004
.
K
K

(4.489)

((0.888143) (292.984 8.314 ln (0.88143)) + (0.223714) (216.926 8.314 ln (0.223714)))|2

((1) (292.984 8.314 ln (1)))|1

1Q2

105002
,
6000
(4.490)

Indeed, the second law is satisfied. Moreover the irreversible entropy production of the chemical reaction
is 19.4181 17.5004 = +1.91772 kJ/K.
For the isochoric, isothermal process, it is also appropriate to use Eq. (4.394), dA|T,V 0, to check
for second law satisfaction. This turns out to give an identical result. Since by Eq. (4.307), A = U T S,
A2 A1 = (U2 T2 S2 ) (U1 T1 S1 ). Since the process is isothermal, A2 A1 = U2 U1 T (S2 S1 ).
For A2 A1 0, one must demand U2 U1 T (S2 S1 ) 0, or U2 U1 T (S2 S1 ), or
S2 S1 (U2 U1 )/T . Since 1Q2 = U2 U1 for this isochoric process, one then recovers S2 S1 1Q2 /T.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

145

4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM


4.11.2.2

Isothermal, isobaric system

Allowing for isobaric rather than isochoric equilibration introduces small variation in the
analysis.

Example 4.15
Consider the same reaction
N2 + N2 2N + N2 .

(4.491)

for an isobaric and isothermal process. That is, consider 1 kmole of N2 and 0 kmole of N at a pressure
of 100 kP a and a temperature of 6000 K. Using the ideal gas tables, find the equilibrium concentrations
of N and N2 if the equilibration process is isothermal and isobaric.
The initial volume is the same as from the previous example:
V1

= 498.84 m3 .

(4.492)

Note the volume will change in this isobaric process. Initially, the mixture is all N2 , so its partial
pressure is the total pressure, and the initial partial pressure of N is 0.
A few other key results are identical to the previous example:
nN = 2 ( nN2 |t=0 nN2 ) ,

(4.493)

2gN g N2 .

(4.494)

and

Substituting using the definitions of Gibbs free energy, one gets



2 h N T sN





y
P
N

2 hN T soT,N R ln
Po


2 hN T soT,N hN2 T soT,N2




2 hN T soT,N + hN2 T soT,N2




2 hN T soT,N + hN2 T soT,N2

h N 2 T sN 2 ,



yN2 P
o
hN2 T sT,N2 R ln
,
Po




yN P
yN2 P
+ RT ln
,
2RT ln
Po
Po




yN P
yN2 P
2RT ln
RT ln
,
Po
Po
 2 2

yN P Po
RT ln
.
Po2 P yN2

(4.495)
(4.496)
(4.497)
(4.498)
(4.499)

In this case Po = P , so one gets




2 hN T soT,N + hN2 T soT,N2

RT ln

2
yN
yN2

(4.500)

At the initial state, one has yN = 0, so the right hand side approaches , and the inequality holds.
At equilibrium, one has equality.


2 hN T soT,N + hN2 T soT,N2

= RT ln

2
yN
yN2

(4.501)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

146

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

Taking numerical values from Table A.9:





kJ
kJ
2 5.9727 105
+
(6000 K) 216.926
kmole
kg K



kJ
kJ
(6000 K) 292.984
2.05848 105
kmole
kg K



 2 
kJ
yN
,
8.314
(6000 K) ln
kmole K
yN2
 2 
yN
2.87635
(4.502)
| {z } = ln yN2 ,
=

ln KP

0.0563399
| {z } =
KP

2
yN
,
yN2

nN
nN +nN2

(4.503)

2

nN2
nN +nN2

,

(4.504)

n2N
,
nN2 (nN + nN2 )

(2 ( nN2 |t=0 nN2 ))2


, (4.506)
nN2 (2 ( nN2 |t=0 nN2 ) + nN2 )

(4.505)

(2 (1 kmole nN2 ))
. (4.507)
nN2 (2 (1 kmole nN2 ) + nN2 )

This is a quadratic equation for nN2 . It has two roots


nN2 = 0.882147 kmole

physical;

nN2 = 1.11785 kmole,

non-physical

(4.508)

The second root generates more N2 than at the start, and also yields non-physically negative nN =
0.235706 kmole. So at equilibrium, the physical root is
nN = 2(1 nN2 ) = 2(1 0.882147) = 0.235706 kmole.

(4.509)

Again, the diatomic species is preferred. As the temperature is raised, one could show that the
monatomic species would come to dominate.
The volume at equilibrium is
(nN2 + nN )RT
,
P
(0.882147 kmole + 0.235706 kmole) 8.314
=
100 kP a
3
= 557.630 m .

V2

(4.510)
kJ
kmole K

(6000 K)

(4.511)
(4.512)

The volume has increased because there are more molecules with the pressure and temperature being
equal.
The molar concentrations i at equilibrium, are
N

0.235706 kmole
mole
kmole
= 4.227 104
= 4.227 107
,
557.636 m3
m3
cm3

(4.513)

N 2

kmole
0.882147 kmole
mole
= 1.58196 103
= 1.58196 106
.
3
3
557.636 m
m
cm3

(4.514)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

4.11. EQUILIBRIUM IN A TWO-COMPONENT SYSTEM

147

The molar concentrations are a little smaller than for the isochoric case, mainly because the volume is
larger at equilibrium.
Now consider the heat transfer. One knows for the isobaric process that Q = H2 H1 . The initial
enthalpy is given by


kJ
= 2.05848 105 kJ.
(4.515)
H1 = nN2 hN2 = (1 kmole) 2.05848 105
kmole
The enthalpy at the final state is
H2

=
=
=

nN2 hN2 + nN hN ,


+ (0.235706 kmole) 5.9727 105

(4.516)

kJ
(4.517)
,
kmole

= H2 H1 = 3.22389 105 kJ 2.05848 105 kJ = 1.16520 105 kJ.

(4.519)


(0.882147 kmole) 2.05848 105

kJ
kmole

3.22368 105 kJ.

(4.518)

So
1Q2

Heat needed to be added to keep the system at the constant temperature. This is because the nitrogen
dissociation process is endothermic. Relative to the isochoric process, more heat had to be added to
maintain the temperature. This to counter the cooling effect of the expansion.
Lastly, it is a straightforward exercise to show that the second law is satisfied for this process.

4.11.3

Equilibrium condition

The results of both of the previous examples, in which a functional form of a progress
variables time variation, d/dt, was postulated in order to satisfy the second law gave a
condition for equilibrium. This can be generalized so as to require at equilibrium that
N
X

i i = 0.

(4.520)

i=1

| {z }

Here i represents the net number of moles of species i generated in the forward reaction.
This negation of the term on the left side of Eq (4.520) is sometimes defined as the chemical
affinity, :
N
X

i i .
(4.521)
i=1

So in the phase equilibrium example, Eq. (4.520) becomes


l (1) + g (1) = 0.

(4.522)

In the nitrogen chemistry example, Eq. (4.520) becomes


N2 (1) + N (2) = 0.

(4.523)

This will be discussed in detail in the following chapter.


CC BY-NC-ND. 03 August 2012, J. M. Powers.

148

CHAPTER 4. MATHEMATICAL FOUNDATIONS OF THERMODYNAMICS

CC BY-NC-ND. 03 August 2012, J. M. Powers.

Chapter 5
Thermochemistry of a single reaction
Read BS, Chapters 15, 16
See Abbott and Van Ness, Chapter 7
See Kondepudi and Prigogine, Chapters 4, 5, 7, 9
See Turns, Chapter 2
See Kuo, Chapters 1, 2
This chapter will further develop notions associated with the thermodynamics of chemical
reactions. The focus will be on single chemical reactions.

5.1

Molecular mass

The molecular mass of a molecule is a straightforward notion from chemistry. One simply
sums the product of the number of atoms and each atoms atomic mass to form the molecular
mass. If one defines L as the number of elements present in species i, li as the number of
moles of atomic element l in species i, and Ml as the atomic mass of element l, the molecular
mass Mi of species i
L
X
Mi =
Ml li ,
i = 1, . . . , N.
(5.1)
l=1

In vector form, one would say


MT = MT ,

or

M = T M.

(5.2)

Here M is the vector of length N containing the molecular masses, M is the vector of length
L containing the elemental masses, and is the matrix of dimension L N containing the
number of moles of each element in each species. Generally, is full rank. If N > L, has
rank L. If N < L, has rank N. In any problem involving an actual chemical reaction, one
will find N L, and in most cases N > L. In isolated problems not involving a reaction,
one may have N < L. In any case, M lies in the column space of T , which is the row space
of .
149

150

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

Example 5.1
Find the molecular mass of H2 O. Here, one has two elements H and O, so L = 2, and one species,
so N = 1; thus, in this isolated problem, N < L. Take i = 1 for species H2 O. Take l = 1 for
element H. Take l = 2 for element O. From the periodic table, one gets M1 = 1 kg/kmole for H,
M2 = 16 kg/kmole for O. For element 1, there are 2 atoms, so 11 = 2. For element 2, there is 1 atom
so 21 = 1. So the molecular mass of species 1, H2 O is
 
 11
,
(5.3)
M 1 = M1 M2
21
=

=
=

M1 11 + M2 21 ,




kg
kg
(2) + 16
(1),
1
kmole
kmole
18

kg
.
kmole

(5.4)

(5.5)
(5.6)

Example 5.2
Find the molecular masses of the two species C8 H18 and CO2 . Here, for practice, the vector matrix
notation is exercised. In a certain sense this is overkill, but it is useful to be able to understand a
general notation.
One has N = 2 species, and takes i = 1 for C8 H18 and i = 2 for CO2 . One also has L = 3 elements
and takes l = 1 for C, l = 2 for H, and l = 3 for O. Now for each element, one has M1 = 12 kg/kmole,
M2 = 1 kg/kmole, M3 = 16 kg/kmole. The molecular masses are then given by

11 12


M1 M2
(5.7)
= M1 M2 M3 21 22 ,
31 32

8 1

= 12 1 16 18 0 ,
(5.8)
0 2

114 44 .
=
(5.9)

That is for C8 H18 , one has molecular mass M1 = 114 kg/kmole. For CO2 , one has molecular mass
M2 = 44 kg/kmole.

5.2
5.2.1

Stoichiometry
General development

Stoichiometry represents a mass balance on each element in a chemical reaction. For example,
in the simple global reaction
2H2 + O2 2H2 O,
(5.10)
CC BY-NC-ND. 03 August 2012, J. M. Powers.

151

5.2. STOICHIOMETRY

one has 4 H atoms in both the reactant and product sides and 2 O atoms in both the reactant
and product sides. In this section stoichiometry will be systematized.
Consider now a general reaction with N species. This reaction can be represented by
N
X

i i

i=1

N
X

i i .

(5.11)

i=1

Here i is the ith chemical species, i is the forward stoichiometric coefficient of the ith
reaction, and i is the reverse stoichiometric coefficient of the ith reaction. Both i and i
are to be interpreted as pure dimensionless numbers.
In Equation (5.10), one has N = 3 species. One might take 1 = H2 , 2 = O2 , and
3 = H2 O. The reaction is written in more general form as
(2)1 + (1)2 + (0)3 (0)1 + (0)2 + (2)3 ,
(2)H2 + (1)O2 + (0)H2 O (0)H2 + (0)O2 + (2)H2 O.

(5.12)
(5.13)

Here, one has


1 = 2,
2 = 1,
3 = 0,

1 = 0,
2 = 0,
3 = 2.

(5.14)
(5.15)
(5.16)

It is common and useful to define another pure dimensionless number, the net stoichiometric coefficients for species i, i . Here i represents the net production of number if the
reaction goes forward. It is given by
i = i i .

(5.17)

For the reaction 2H2 + O2 2H2 O, one has


1 = 1 1 = 0 2 = 2,
2 = 2 2 = 0 1 = 1,
3 = 3 3 = 2 0 = 2.

(5.18)
(5.19)
(5.20)

With these definitions, it is possible to summarize a chemical reaction as


N
X

i i = 0.

(5.21)

i=1

In vector notation, one would say


T = 0.

(5.22)

2H2 O2 + 2H2 O = 0.

(5.23)

For the reaction of this section, one might write the non-traditional form

CC BY-NC-ND. 03 August 2012, J. M. Powers.

152

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

It remains to enforce a stoichiometric balance. This is achieved if, for each element, l =
1, . . . , L, one has the following equality:
N
X

li i = 0,

l = 1, . . . , L.

(5.24)

i=1

That is to say, for each element, the sum of the product of the net species production and
the number of elements in the species must be zero. In vector notation, this becomes
= 0.

(5.25)

One may recall from linear algebra that this demands that lie in the right null space of .
Example 5.3
Show stoichiometric balance is achieved for 2H2 O2 + 2H2 O = 0.
Here again, the number of elements L = 2, and one can take l = 1 for H and l = 2 for O. Also
the number of species N = 3, and one takes i = 1 for H2 , i = 2 for O2 , and i = 3 for H2 O. Then for
element 1, H, in species 1, H2 , one has
11 = 2, H in H2 .

(5.26)

12
13

= 0,
= 2,

H in O2 ,
H in H2 O,

(5.27)
(5.28)

21
22

= 0,
= 2,

O in H2 ,
O in O2 ,

(5.29)
(5.30)

23

= 1,

O in H2 O.

(5.31)

Similarly, one gets

In matrix form then,

PN

i=1

li i = 0 gives

2
0



 

0 2 1
0
2 =
.
2 1
0
3

(5.32)

This is two equations in three unknowns. Thus, it is formally under-constrained. Certainly the trivial
solution 1 = 2 = 3 = 0 will satisfy, but one seeks non-trivial solutions. Assume 3 has a known value
3 = . Then, the system reduces to


  
2
2 0
1
=
.
(5.33)

2
0 2
The inversion here is easy, and one finds 1 = , 2 = 21 . Or in vector form,

1
2 = 1 ,
2

3

1
R1 .
= 21 ,
1
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.34)

(5.35)

153

5.2. STOICHIOMETRY

Again, this amounts to saying the solution vector (1 , 2 , 3 )T lies in the right null space of the coefficient
matrix li .
Here is any real scalar. If one takes = 2, one gets

1
2
2 = 1 ,
3
2

(5.36)

2H2 O2 + 2H2 O = 0.

(5.37)

This simply corresponds to

If one takes = 4, one still achieves stoichiometric balance, with



1
4
2 = 2 ,
3
4

(5.38)

4H2 + 2O2 4H2 O = 0.

(5.39)

which corresponds to the equally valid

In summary, the stoichiometric coefficients are non-unique but partially constrained by mass conservation. Which set is chosen is to some extent arbitrary, and often based on traditional conventions from
chemistry. But others are equally valid.

There is a small issue with units here, which will be seen to be difficult to reconcile.
In practice, it will have little to no importance. In the previous example, one might be
tempted to ascribe units of kmoles to i . Later, it will be seen that in classical reaction
kinetics, i is best interpreted as a pure dimensionless number, consistent with the definition
of this section. So in the context of the previous example, one would then take to be
dimensionless as well, which is perfectly acceptable for the example. In later problems, it
will be more useful to give the units of kmoles. Note that multiplication of by any scalar,
e.g. kmole/(6.02 1026 ), still yields an acceptable result.
Example 5.4
Balance an equation for hypothesized ethane combustion
1 C2 H6 + 2 O2 3 CO2 + 4 H2 O.

(5.40)

One could also say in terms of the net stoichiometric coefficients


1 C2 H6 + 2 O2 + 3 CO2 + 4 H2 O = 0.

(5.41)

Here one takes 1 = C2 H6 , 2 = O2 , 3 CO2 , and 4 = H2 O. So there are N = 4 species. There are
CC BY-NC-ND. 03 August 2012, J. M. Powers.

154

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


also L = 3 elements: l = 1 : C, l = 2 : H, l = 3 : O. One then has
11
12

=
=

2,
0,

C in C2 H6 ,
C in O2 ,

(5.42)
(5.43)

13
14

=
=

1,
0,

C in CO2 ,
C in H2 O,

(5.44)
(5.45)

21

6,

H in C2 H6 ,

(5.46)

22
23

=
=

0,
0,

H in O2 ,
H in CO2 ,

(5.47)
(5.48)

24
31

=
=

2,
0,

H in H2 O,
O in C2 H6 ,

(5.49)
(5.50)

32
33

=
=

2,
2,

O in O2 ,
O in CO2 ,

(5.51)
(5.52)

34 = 1,
O in H2 O.
PN
So the stoichiometry equation, i=1 li i = 0, is given by

1

2 0 1 0
0

6 0 0 2 2 = 0 .
3
0 2 2 1
0
4

(5.53)

(5.54)

Here, there are three equations in four unknowns, so the system is under-constrained. There are many
ways to address this problem. Here, choose the robust way of casting the system into row echelon form.
This is easily achieved by Gaussian elimination. Row echelon form seeks to have lots of zeros in the
lower left part of the matrix. The lower left corner has a zero already, so that is useful. Now, multiply
the top equation by 3 and subtract the result from the second to get

1

2 0 1 0
0

0 0 3 2 2 = 0 .
(5.55)
3
0 2 2 1
0
4

Next switch the last two equations to get

2
0
0

0 1 0 1
0

2
= 0 .
2 2 1
3
0 3 2
0
4

Now divide the first by 2, the second by 2 and the third by 3 to get unity in the diagonal:


0 1
0
1 0 12
1 2
0 .
0 1 1
=
2
3
0
0 0 1 32
4

(5.56)

(5.57)

So-called bound variables have non-zero coefficients on the diagonal, so one can take the bound variables
to be 1 , 2 , and 3 . The remaining variables are free variables. Here one takes the free variable to be
4 . So, set 4 = , and rewrite the system as

0
1
1 0 12
0 1 1 2 = 1 .
(5.58)
2
2

0 0 1
3
3

CC BY-NC-ND. 03 August 2012, J. M. Powers.

155

5.2. STOICHIOMETRY
Solving, one finds
1
1
3
1
3
2 7
7
= 26 = 26 ,
3

3
3
4
1

R1 .

(5.59)

Again, one finds a non-unique solution in the right null space of . If one chooses = 6, then one gets

1
2
2 7
= ,
3 4
4
6

(5.60)

which corresponds to the stoichiometrically balanced reaction


2C2 H6 + 7O2 4CO2 + 6H2 O.

(5.61)

In this example, is dimensionless.

Example 5.5
Consider stoichiometric balance for a propane oxidation reaction which may produce carbon monoxide and hydroxyl in addition to carbon dioxide and water.

The hypothesized reaction takes the form


1 C3 H8 + 2 O2 3 CO2 + 4 CO + 5 H2 O + 6 OH.

(5.62)

In terms of net stoichiometric coefficients, this becomes


1 C3 H8 + 2 O2 + 3 CO2 + 4 CO + 5 H2 O + 6 OH = 0.

(5.63)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

156

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


There are N = 6 species and L = 3 elements. One then has
11

3,

C in C3 H8 ,

(5.64)

12
13

=
=

0,
1,

C in O2 ,
C in CO2 ,

(5.65)
(5.66)

14
15

=
=

1,
0,

C in CO,
C in H2 O,

(5.67)
(5.68)

16
21

=
=

0,
8,

C in OH,
H in C3 H8 ,

(5.69)
(5.70)

22
23

=
=

0,
0,

H in O2 ,
H in CO2 ,

(5.71)
(5.72)

24
25

=
=

0,
2,

H in CO,
H in H2 O,

(5.73)
(5.74)

26

1,

H in OH,

(5.75)

31
32

=
=

0,
2,

O in C3 H8 ,
O in O2 ,

(5.76)
(5.77)

33
34

=
=

2,
1,

O in CO2 ,
O in CO,

(5.78)
(5.79)

35
36

=
=

1,
1,

O in H2 O,
O in OH.

(5.80)
(5.81)

The equation = 0, then becomes

3 0
8 0
0 2

1 1
0 0
2 1


1
2

0
0
3
= 0 .
1
4

1
0
5
6

0
2
1

(5.82)

Multiplying the first equation by 8/3 and adding it to the second gives

3 0
0 0
0 2

1
38
2

1
83
1

0
2
1


1
2

0
0
3

1
4 = 0 .

1
0
5
6

(5.83)

0
1
2


1
2

0
0
3
= 0 .
1
4

1
0
5
6

(5.84)

Trading the second and third rows gives

3 0
0 2
0 0

1
2
38

1
1
83

CC BY-NC-ND. 03 August 2012, J. M. Powers.

157

5.2. STOICHIOMETRY
Dividing the first row by 3, the second by 2 and the third by 8/3 gives

1

2
1
1

1 0 3 3
0
0
0

1
1 3
0 1 1 1
0 .

=
2
2
2
4

0 0 1 1 43 38
0
5
6

(5.85)

Take bound variables to be 1 , 2 , and 3 and free variables to be 4 , 5 , and 6 . So set 4 = 1 ,


5 = 2 , and 6 = 3 , and get

1
31
1 0 31
0 1 1 2 = 1 2 3 .
(5.86)
2
2
2
0 0 1
3
1 + 34 2 + 83 3
Solving, one finds


1
2
3

1
8 (22 3 )
1 (41 102 73 ) .
8
1
8 (81 + 62 + 33 )

For all the coefficients, one then has


1
0
2
1
1
8 (22 3 )
4
10
7
2
1 (41 102 73 )


18

= 8 (81 + 62 + 33 ) = 1 8 + 2 6 + 3 3 .

8
0

8
8
8
1

0
8
0

2
0
0
8
6
3

(5.87)

(5.88)

Here, one finds three independent vectors in the right null space. To simplify the notation, take
1 = 1 /8, 2 = 2 /8, and 3 = 3 /8. Then,

1
0
2
1
2
4
10
7

= 1 8 + 2 6 + 3 3 .
(5.89)
4
8
0
0

5
0
8
0
6
0
0
8

The most general reaction that can achieve a stoichiometric balance is given by

(22 3 )C3 H8 + (41 102 73 )O2 + (81 + 62 + 33 )CO2 + 81 CO + 82 H2 O + 83 OH = 0. (5.90)


Rearranging, one gets
(22 + 3 )C3 H8 + (41 + 102 + 73 )O2 (81 + 62 + 33 )CO2 + 81 CO + 82 H2 O + 83 OH. (5.91)
This will be balanced for all 1 , 2 , and 3 . The values that are actually achieved in practice depend
on the thermodynamics of the problem. Stoichiometry only provides some limitations.
A slightly more familiar form is found by taking 2 = 1/2 and rearranging, giving
(1 + 3 ) C3 H8 + (5 41 + 73 ) O2 (3 81 + 33 ) CO2 + 4 H2 O + 81 CO + 83 OH.

(5.92)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

158

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


One notes that often the production of CO and OH will be small. If there is no production of CO or
OH, 1 = 3 = 0 and one recovers the familiar balance of
C3 H8 + 5 O2 3 CO2 + 4 H2 O.

(5.93)

One also notes that stoichiometry alone admits unusual solutions. For instance, if 1 = 100, 2 = 1/2,
and 3 = 1, one has
2 C3 H8 + 794 CO2 388 O2 + 4 H2 O + 800 CO + 8 OH.

(5.94)

This reaction is certainly admitted by stoichiometry but is not observed in nature. To determine
precisely which of the P
infinitely many possible final states are realized requires a consideration of the
N
equilibrium condition i=1 i i = 0.
Looked at in another way, we can think of three independent classes of reactions admitted by the
stoichiometry, one for each of the linearly independent null space vectors. Taking first 1 = 1/4, 2 = 0,
3 = 0, one gets, after rearrangement
2CO + O2 2CO2 ,

(5.95)

as one class of reaction admitted by stoichiometry. Taking next 1 = 0, 2 = 1/2, 3 = 0, one gets
C3 H8 + 5O2 3CO2 + 4H2 O,

(5.96)

as the second class admitted by stoichiometry. The third class is given by taking 1 = 0, 2 = 0, 3 = 1,
and is
C3 H8 + 7O2 3CO2 + 8OH.
(5.97)
In this example, both and are dimensionless.

In general, one can expect to find the stoichiometric coefficients for N species composed
of L elements to be of the following form:
i =

N
L
X
k=1

Dik k ,

i = 1, . . . , N.

(5.98)

Here Dik is a dimensionless component of a full rank matrix of dimension N (N L)


and rank N L, and k is a dimensionless component of a vector of parameters of length
N L. The matrix whose components are Dik are constructed by populating its columns
with vectors which lie in the right null space of li . Note that multiplication of k by any
constant gives another set of i , and mass conservation for each element is still satisfied.
For later use, we associate i with the number of moles ni , and consider a differential
change in the number of moles dni from Eq. (5.98) and arrive at
dni =

N
L
X
k=1

Dik dk ,

CC BY-NC-ND. 03 August 2012, J. M. Powers.

i = 1, . . . , N.

(5.99)

159

5.2. STOICHIOMETRY

5.2.2

Fuel-air mixtures

In combustion with air, one often models air as a simple mixture of diatomic oxygen and
inert diatomic nitrogen:
(O2 + 3.76N2 ).
(5.100)
The air-fuel ratio, A and its reciprocal, the fuel-air ratio, F , can be defined on a mass
and mole basis.
mair
nair
Amass =
(5.101)
,
Amole =
.
mf uel
nf uel
Via the molecular masses, one has
Amass =

mair
nair Mair
Mair
=
= Amole
.
mf uel
nf uel Mf uel
Mf uel

(5.102)

If there is not enough air to burn all the fuel, the mixture is said to be rich. If there is
excess air, the mixture is said to be lean. If there is just enough, the mixture is said to be
stoichiometric. The equivalence ratio, , is defined as the actual fuel-air ratio scaled by the
stoichiometric fuel-air ratio:

Factual

Fstoichiometric

Astoichiometric
.
Aactual

(5.103)

The ratio is the same whether F s are taken on a mass or mole basis, because the ratio of
molecular masses cancel.
Example 5.6
Calculate the stoichiometry of the combustion of methane with air with an equivalence ratio of
= 0.5. If the pressure is 0.1 M P a, find the dew point of the products.
First calculate the coefficients for stoichiometric combustion:
1 CH4 + 2 (O2 + 3.76N2 ) 3 CO2 + 4 H2 O + 5 N2 ,

(5.104)

1 CH4 + 2 O2 + 3 CO2 + 4 H2 O + (5 + 3.762 )N2 = 0.

(5.105)

or
Here one has N = 5 species and L = 4 elements. Adopting a slightly more intuitive procedure for
variety, one writes a conservation equation for each element to get

In matrix form this becomes

1
4

0
0

1 + 3

= 0,

C,

(5.106)

41 + 24
22 + 23 + 4

= 0,
= 0,

H,
O,

(5.107)
(5.108)

3.762 + 5

= 0,

N.

(5.109)

0
0
2
3.76

1
0
2
0

0
2
1
0


1
0
0
2
0
0
3 = .
0
0
4
1
0
5

(5.110)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

160

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


Now, one might expect to have one free variable, since one has five unknowns in four equations.
While casting the equation in row echelon form is guaranteed to yield a proper solution, one can often
use intuition to get a solution more rapidly. One certainly expects that CH4 will need to be present
for the reaction to take place. One might also expect to find an answer if there is one mole of CH4 . So
take 1 = 1. Realize that one could also get a physically valid answer by assuming 1 to be equal to
any scalar. With 1 = 1, one gets

2
1
0
1 0 0
3 4
0
0
2
0
= .

(5.111)
2
2 1 0 4 0
5
0
3.76 0 0 1
One easily finds the unique inverse does exist, and that the solution is


2
2
3 1

=
4 2 .
5
7.52

(5.112)

If there had been more than one free variable, the four by four matrix would have been singular, and
no unique inverse would have existed.
In any case, the reaction under stoichiometric conditions is
CH4 2O2 + CO2 + 2H2 O + (7.52 + (3.76)(2))N2

0,

CH4 + 2(O2 + 3.76N2) CO2 + 2H2 O + 7.52N2.

(5.113)
(5.114)

For the stoichiometric reaction, the fuel-air ratio on a mole basis is


Fstoichiometric =

1
= 0.1050.
2 + 2(3.76)

(5.115)

Now = 0.5, so
Factual

= Fstoichiometric ,

(5.116)

= (0.5)(0.1050),

(5.117)

= 0.0525.

(5.118)

By inspection, one can write the actual reaction equation as


CH4 + 4(O2 + 3.76N2 ) CO2 + 2H2 O + 2O2 + 15.04N2.

(5.119)

Check:
Factual =

1
= 0.0525.
4 + 4(3.76)

(5.120)

For the dew point of the products, one needs the partial pressure of H2 O. The mole fraction of
H2 O is
2
= 0.0499
(5.121)
yH2 O =
1 + 2 + 2 + 15.04
So the partial pressure of H2 O is
PH2 O = yH2 O P = 0.0499(100 kP a) = 4.99 kP a.

(5.122)

From the steam tables, the saturation temperature at this pressure is Tsat = Tdew point = 32.88 C . If
the products cool to this temperature in an exhaust device, the water could condense in the apparatus.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS

5.3

161

First law analysis of reacting systems

One can easily use the first law to learn much about chemically reacting systems.

5.3.1

Enthalpy of formation

The enthalpy of formation is the enthalpy that is required to form a molecule from combining
its constituents at P = 0.1 MP a and T = 298 K. Consider the reaction (taken here to be
irreversible)
C + O2 CO2 .
(5.123)
In order to maintain the process at constant temperature, it is found that heat transfer to
the volume is necessary. For the steady constant pressure process, one has
U2 U1 =

1W2 ,
Z 2
= 1Q2
P dV,
1Q2

(5.124)
(5.125)

1Q2

1Q2

= 1Q2 P (V2 V1 ),
= U2 U1 + P (V2 V1 ),
= H2 H1 ,
= Hproducts Hreactants .

(5.126)
(5.127)
(5.128)
(5.129)

So
1Q2 =

products

ni hi

ni hi .

(5.130)

reactants

In this reaction, one measures that 1Q2 = 393522 kJ for the reaction of 1 kmole of C
and O2 . That is the reaction liberates such energy to the environment. So measuring the
heat transfer can give a measure of the enthalpy difference between reactants and products.
Assign a value of enthalpy zero to elements in their standard state at the reference state.
Thus, C and O2 both have enthalpies of 0 kJ/kmole at T = 298 K, P = 0.1 MP a. This
enthalpy is designated, for species i,
o

hf,i = h298,i ,

(5.131)

and is called the enthalpy of formation. So the energy balance for the products and reactants,
here both at the standard state, becomes
o

(5.132)
= nCO2 hf,CO2 nC hf,C nO2 hf,O2 ,




kJ
kJ
o
(1 kmole) 0
.
393522 kJ = (1 kmole)hf,CO2 (1 kmole) 0
kmole
kmole
(5.133)
1Q2

CC BY-NC-ND. 03 August 2012, J. M. Powers.

162

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


o

Thus, the enthalpy of formation of CO2 is hf,CO2 = 393522 kJ/kmole, since the reaction
involved creation of 1 kmole of CO2 .
Often values of enthalpy are tabulated in the forms of enthalpy differences hi . These
are defined such that
o

hi = hf,i + (hi hf,i ),


| {z }

(5.134)

=hi

o
hf,i

+ hi .

(5.135)

Lastly, one notes for an ideal gas that the enthalpy is a function of temperature only,
and so does not depend on the reference pressure; hence
o

hi = hi ,

hi = hi ,

if ideal gas.

(5.136)

Example 5.7
(BS, Ex. 15.6 pp. 629-630). Consider the heat of reaction of the following irreversible reaction in a
steady state, steady flow process confined to the standard state of P = 0.1 M P a, T = 298 K:
CH4 + 2O2 CO2 + 2H2 O().

(5.137)

The first law holds that


Qcv

products

ni hi

ni hi .

(5.138)

reactants

All components are at their reference states. Table A.10 gives properties, and one finds
Qcv

=
=

nCO2 hCO2 + nH2 O hH2 O nCH4 hCH4 nO2 hO2 ,






kJ
kJ
(1 kmole) 393522
+ (2 kmole) 285830
kmole
kmole




kJ
kJ
(1 kmole) 74873
(2 kmole) 0
,
kmole
kmole
890309 kJ.

(5.139)

(5.140)
(5.141)

A more detailed analysis is required in the likely case in which the system is not at the
reference state.
Example 5.8
(adopted from Moran and Shapiro, p. 619) A mixture of 1 kmole of gaseous methane and 2 kmole
of oxygen initially at 298 K and 101.325 kP a burns completely in a closed, rigid, container. Heat
transfer occurs until the final temperature is 900 K. Find the heat transfer and the final pressure.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

163

5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS

The combustion is stoichiometric. Assume that no small concentration species are generated. The
global reaction is given by
CH4 + 2O2 CO2 + 2H2 O.
(5.142)
The first law analysis for the closed system is slightly different:
U2 U1 = 1Q2 1W2 .

(5.143)

Since the process is isochoric, 1W2 = 0. So


1Q2

=
=
=
=
=
=
=

U2 U1 ,
nCO2 uCO2 + nH2 O uH2 O nCH4 uCH4 nO2 uO2 ,

(5.144)
(5.145)

nCO2 (hCO2 RT2 ) + nH2 O (hH2 O RT2 ) nCH4 (hCH4 RT1 ) nO2 (hO2 RT1 ), (5.146)
hCO2 + 2hH2 O hCH4 2hO2 3R(T2 T1 ),
(5.147)
o

(hCO2 ,f + hCO2 ) + 2(hH2 O,f + hH2 O ) (hCH4 ,f + hCH4 ) 2(hO2 ,f + hO2 )

3R(T2 T1 ),

(393522 + 28030) + 2(241826 + 21937) (74873 + 0) 2(0 + 0)


3(8.314)(900 298),
745412 kJ.

(5.148)
(5.149)
(5.150)

For the pressures, one has


P1 V1

(nCH4 + nO2 )RT1 ,

V1

(nCH4 + nO2 )RT1


,
P1

=
=

(5.151)
(5.152)


(1 kmole + 2 kmole) 8.314

kJ
kg K

101.325 kP a

(298 K)
,

73.36 m3 .

(5.153)
(5.154)

Now V2 = V1 , so
P2

=
=
=

(nCO2 + nH2 O )RT2


,
V2

(5.155)


(1 kmole + 2 kmole) 8.314
73.36 m3

kJ
kg K

(900 K)
,

306.0 kP a.

(5.156)
(5.157)

The pressure increased in the reaction. This is entirely attributable to the temperature rise, as the
number of moles remained constant here.

5.3.2

Enthalpy and internal energy of combustion

The enthalpy of combustion is the difference between the enthalpy of products and reactants
when complete combustion occurs at a given pressure and temperature. It is also known
CC BY-NC-ND. 03 August 2012, J. M. Powers.

164

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

as the heating value or the heat of reaction. The internal energy of combustion is related
and is the difference between the internal energy of products and reactants when complete
combustion occurs at a given volume and temperature.
The term higher heating value refers to the energy of combustion when liquid water is in
the products. Lower heating value refers to the energy of combustion when water vapor is
in the product.

5.3.3

Adiabatic flame temperature in isochoric stoichiometric systems

The adiabatic flame temperature refers to the temperature which is achieved when a fuel and
oxidizer are combined with no loss of work or heat energy. Thus, it must occur in a closed,
insulated, fixed volume. It is generally the highest temperature that one can expect to
achieve in a combustion process. It generally requires an iterative solution. Of all mixtures,
stoichiometric mixtures will yield the highest adiabatic flame temperatures because there is
no need to heat the excess fuel or oxidizer.
Here four examples will be presented to illustrate the following points.
The adiabatic flame temperature can be well over 5000 K for seemingly ordinary
mixtures.
Dilution of the mixture with an inert diluent lowers the adiabatic flame temperature.
The same effect would happen in rich and lean mixtures.
Preheating the mixture, such as one might find in the compression stroke of an engine,
increases the adiabatic flame temperature.
Consideration of the presence of minor species lowers the adiabatic flame temperature.
5.3.3.1

Undiluted, cold mixture

Example 5.9
A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2 and 1 kmole
of O2 at 100 kP a and 298 K. Find the adiabatic flame temperature and final pressure assuming the
irreversible reaction
2H2 + O2 2H2 O.
(5.158)
The volume is given by
V

=
=
=

(nH2 + nO2 )RT1


,
P1
(2 kmole + 1 kmole) 8.314
100 kP a
74.33 m3 .

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.159)
kJ
kmole K

(298 K)

(5.160)
(5.161)

165

5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS


The first law gives
U2 U1

nH2 O uH2 O nH2 uH2

U2 U1
nO2 uO2

nH2 O (hH2 O RT2 ) nH2 (hH2 RT1 ) nO2 (hO2 RT1 )


2hH2 O 2 hH2 hO2 +R(2T2 + 3T1 )
|{z} |{z}

1Q2

= 0,
= 0,

1W2 ,

(5.162)
(5.163)
(5.164)

= 0,
= 0,

(5.165)
(5.166)

2hH2 O + (8.314) ((2) T2 + (3) (298)) = 0,


hH2 O 8.314T2 + 3716.4 = 0,

(5.167)
(5.168)

241826 + hH2 O 8.314T2 + 3716.4 = 0,

(5.170)

=0

=0

hf,H2 O + hH2 O 8.314T2 + 3716.4 = 0,

(5.169)

238110 + hH2 O 8.314T2

(5.171)

= 0.

At this point, one begins an iteration process, guessing a value of T2 and an associated hH2 O . When
T2 is guessed at 5600 K, the left side becomes 6507.04. When T2 is guessed at 6000 K, the left side
becomes 14301.4. Interpolate then to arrive at
T2 = 5725 K.

(5.172)

This is an extremely high temperature. At such temperatures, in fact, one can expect other species to
co-exist in the equilibrium state in large quantities. These may include H, OH, O, HO2 , and H2 O2 ,
among others.
The final pressure is given by
P2

=
=
=

nH2 O RT2
,
V
kJ
(2 kmole) 8.314 kmole
74.33 m3
1280.71 kP a.

(5.173)
K

(5725 K)

(5.174)
(5.175)

The final concentration of H2 O is


H2 O =

5.3.3.2

kmole
2 kmole
= 2.69 102
.
74.33 m3
m3

(5.176)

Dilute, cold mixture

Example 5.10
Consider a variant on the previous example in which the mixture is diluted with an inert, taken
here to be N2 . A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2 ,
1 kmole of O2 , and 8 kmole of N2 at 100 kP a and 298 K. Find the adiabatic flame temperature and
the final pressure, assuming the irreversible reaction
2H2 + O2 + 8N2 2H2 O + 8N2 .

(5.177)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

166

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


The volume is given by
V

=
=
=

(nH2 + nO2 + nN2 )RT1


,
P1
(2 kmole + 1 kmole + 8 kmole) 8.314
100 kP a
272.533 m3 .

(5.178)
kJ
kmole K

(298 K)

(5.179)
(5.180)

The first law gives


U2 U1
U2 U1

=
=

nH2 O uH2 O nH2 uH2 nO2 uO2 + nN2 (uN2 2 uN2 1 ) =


RT1 ) nO2 (hO2 RT1 ) + nN2 ((hN2 2 RT2 ) (hN2 1 RT1 )) =

nH2 O (hH2 O RT2 ) nH2 (hH2

2hH2 O 2 hH2 hO2 +R(10T2 + 11T1 ) + 8( hN2 2 hN2 1 ) =


|{z} |{z}
| {z } | {z }
=0

=0

=hN2

1Q2

0,
0,
0,
0,

=0

2hH2 O + (8.314) (10T2 + (11)(298)) + 8hN2 2

0,

2hH2 O 83.14T2 + 27253.3 + 8hN2 2

0,

0,

2(241826) + 2hH2 O 83.14T2 + 27253.3 + 8hN2 2


456399 + 2hH2 O 83.14T2 + 8hN2 2

=
=

0,
0.

o
2hf,H2 O

1W2 ,

+ 2hH2 O 83.14T2 + 27253.3 + 8hN2 2

At this point, one begins an iteration process, guessing a value of T2 and an associated hH2 O . When
T2 is guessed at 2000 K, the left side becomes 28006.7. When T2 is guessed at 2200 K, the left side
becomes 33895.3. Interpolate then to arrive at
T2 = 2090.5 K.

(5.181)

The inert diluent significantly lowers the adiabatic flame temperature. This is because the N2 serves as
a heat sink for the energy of reaction. If the mixture were at non-stoichiometric conditions, the excess
species would also serve as a heat sink, and the adiabatic flame temperature would be lower than that
of the stoichiometric mixture.
The final pressure is given by
P2

=
=
=

(nH2 O + nN2 )RT2


,
V
kJ
(2 kmole + 8 kmole) 8.314 kmole
272.533 m3
637.74 kP a.

(5.182)
K

(2090.5 K)

(5.183)
(5.184)

The final concentrations of H2 O and N2 are


H2 O

N2

2 kmole
kmole
= 7.34 103
,
272.533 m3
m3
kmole
8 kmole
= 2.94 102
.
3
272.533 m
m3

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.185)
(5.186)

167

5.3. FIRST LAW ANALYSIS OF REACTING SYSTEMS


5.3.3.3

Dilute, preheated mixture

Example 5.11
Consider a variant on the previous example in which the diluted mixture is preheated to 1000 K.
One can achieve this via an isentropic compression of the cold mixture, such as might occur in an engine.
To simplify the analysis here, the temperature of the mixture will be increased, while the pressure will
be maintained. A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2 ,
1 kmole of O2 , and 8 kmole of N2 at 100 kP a and 1000 K. Find the adiabatic flame temperature and
the final pressure, assuming the irreversible reaction
2H2 + O2 + 8N2 2H2 O + 8N2 .

(5.187)

The volume is given by


V

=
=
=

(nH2 + nO2 + nN2 )RT1


,
P1
(2 kmole + 1 kmole + 8 kmole) 8.314
100 kP a
914.54 m3 .

(5.188)
kJ
kmole K

(1000 K)

(5.189)
(5.190)

The first law gives


U2 U1

1Q2

U2 U1 =
+ nN2 (uN2 2 uN2 1 ) =

0,
0,

2(241826 + hH2 O ) 2(20663) 22703 + (8.314) (10T2 + (11)(1000)) + 8hN2 2 8(21463) =


2hH2 O 83.14T2 627931 + 8hN2 2 =

0,
0.

nH2 O uH2 O nH2 uH2 nO2 uO2

nH2 O (hH2 O RT2 ) nH2 (hH2 RT1 ) nO2 (hO2 RT1 ) + nN2 ((hN2 2 RT2 ) (hN2 1 RT1 )) =
2hH2 O 2hH2 hO2 + R(10T2 + 11T1 ) + 8(hN2 2 hN2 1 ) =

1W2 ,

0,
0,

At this point, one begins an iteration process, guessing a value of T2 and an associated hH2 O . When
T2 is guessed at 2600 K, the left side becomes 11351. When T2 is guessed at 2800 K, the left side
becomes 52787. Interpolate then to arrive at
T2 = 2635.4 K.

(5.191)

The preheating raised the adiabatic flame temperature. Note that the preheating was by (1000 K)
(298 K) = 702 K. The new adiabatic flame temperature is only (2635.4 K) (2090.5 K) = 544.9 K
greater.
The final pressure is given by
P2

=
=
=

(nH2 O + nN2 )RT2


,
V
kJ
(2 kmole + 8 kmole) 8.314 kmole
914.54 m3
239.58 kP a.

(5.192)
K

(2635.4 K)

(5.193)
(5.194)

The final concentrations of H2 O and N2 are


H2 O

N 2

kmole
2 kmole
= 2.19 103
,
3
914.54 m
m3
8 kmole
kmole
= 8.75 103
.
3
914.54 m
m3

(5.195)
(5.196)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

168

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

5.3.3.4

Dilute, preheated mixture with minor species

Example 5.12
Consider a variant on the previous example. Here allow for minor species to be present at equilibrium. A closed, fixed, adiabatic volume contains a stoichiometric mixture of 2 kmole of H2 , 1 kmole
of O2 , and 8 kmole of N2 at 100 kP a and 1000 K. Find the adiabatic flame temperature and the final
pressure, assuming reversible reactions. Here, the details of the analysis are postponed, but the result
is given which is the consequence of a calculation involving detailed reactions rates. One can also solve
an optimization problem to minimize the Gibbs free energy of a wide variety of products to get the
same answer. In this case, the equilibrium temperature and pressure are found to be

T = 2484.8 K,

P = 227.89 kP a.

(5.197)

Equilibrium species concentrations are found to be

H2

1.3 104

minor product

1.9 105

minor product

5.7 106

O2

3.6 105

OH

5.9 105

major product

H2 O

2.0 103

trace product

HO2

1.1 108

H2 O2

1.2 109

minor product

minor product
minor product

trace product

CC BY-NC-ND. 03 August 2012, J. M. Powers.

kmole
,
m3
kmole
,
m3
kmole
,
m3
kmole
,
m3
kmole
,
m3
kmole
,
m3
kmole
,
m3
kmole
,
m3

(5.198)
(5.199)
(5.200)
(5.201)
(5.202)
(5.203)
(5.204)
(5.205)

169

5.4. CHEMICAL EQUILIBRIUM


kmole
,
m3
kmole
,
3.7 1010
m3
kmole
,
1.5 1010
m3
kmole
3.1 1010
,
m3
kmole
1.0 1010
,
m3
kmole
,
3.1 106
m3
kmole
,
5.3 109
m3
kmole
2.6 109
,
m3
kmole
1.7 109
,
m3
kmole
8.7 103
.
m3

= 1.7 109

(5.206)

N H

(5.207)

trace product

N H2

trace product

N H3

N N H

minor product

N O

trace product

N O2

trace product

N 2 O

HN O

N 2

trace product
trace product

trace product

trace product
major product

(5.208)
(5.209)
(5.210)
(5.211)
(5.212)
(5.213)
(5.214)
(5.215)

Note that the concentrations of the major products went down when the minor species were considered.
The adiabatic flame temperature also went down by a significant amount: 2635 2484.8 = 150.2 K.
Some thermal energy was necessary to break the bonds which induce the presence of minor species.

5.4

Chemical equilibrium

Often reactions are not simply unidirectional, as alluded to in the previous example. The
reverse reaction, especially at high temperature, can be important.
Consider the four species reaction
1 1 + 2 2 3 3 + 4 4 .

(5.216)

In terms of the net stoichiometric coefficients, this becomes


1 1 + 2 2 + 3 3 + 4 4 = 0.

(5.217)

One can define a variable , the reaction progress. Take the dimension of to be kmoles.
When t = 0, one takes = 0. Now as the reaction goes forward, one takes d > 0. And a
forward reaction will decrease the number of moles of 1 and 2 while increasing the number
of moles of 3 and 4 . This will occur in ratios dictated by the stoichiometric coefficients of
the problem:
CC BY-NC-ND. 03 August 2012, J. M. Powers.

170

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

dn1
dn2
dn3
dn4

=
=
=
=

1 d,
2 d,
+3 d,
+4 d.

(5.218)
(5.219)
(5.220)
(5.221)

Note that if ni is taken to have units of kmoles, i , and i are taken as dimensionless, then
must have units of kmoles. In terms of the net stoichiometric coefficients, one has
dn1
dn2
dn3
dn4

=
=
=
=

1 d,
2 d,
3 d,
4 d.

(5.222)
(5.223)
(5.224)
(5.225)

Again, for arguments sake, assume that at t = 0, one has


n1 |t=0
n2 |t=0
n3 |t=0
n4 |t=0

=
=
=
=

n1o ,
n2o ,
n3o ,
n4o .

(5.226)
(5.227)
(5.228)
(5.229)

Then after integrating, one finds


n1
n2
n3
n4

=
=
=
=

1 + n1o ,
2 + n2o ,
3 + n3o ,
4 + n4o .

(5.230)
(5.231)
(5.232)
(5.233)

One can also eliminate the parameter in a variety of fashions and parameterize the
reaction one of the species mole numbers. Choosing, for example, n1 as a parameter, one
gets
n1 n1o
=
.
(5.234)
1
Eliminating then one finds all other mole numbers in terms of n1 :
n1 n1o
+ n2o ,
1
n1 n1o
+ n3o ,
= 3
1
n1 n1o
= 4
+ n4o .
1

n2 = 2

(5.235)

n3

(5.236)

n4

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.237)

171

5.4. CHEMICAL EQUILIBRIUM


Written another way, one has
n1 n1o
n2 n2o
n3 n3o
n4 n4o
=
=
=
= .
1
2
3
4
P
For an N-species reaction, N
i=1 i i = 0, one can generalize to say
dni = = i d,
ni = i + nio ,
ni nio
= .
i

(5.238)

(5.239)
(5.240)
(5.241)

Note that
dni
= i .
d

(5.242)

Now, from the previous chapter, one manifestation of the second law is Eq. (4.399):
dG|T,P =

N
X
i=1

i dni 0.

(5.243)

Now, one can eliminate dni in Eq. (5.243) by use of Eq. (5.239) to get
dG|T,P =

G
=
T,P

N
X

i=1
N
X
i=1

i i d 0,

(5.244)

i i 0,

(5.245)

= 0.

(5.246)

Then for the reaction to go forward, one must require that the affinity be positive:
0.

(5.247)

One also knows from the previous chapter that the irreversible entropy production takes the
form of Eq. (4.387):
N
1X

dni 0,
T i=1 i

(5.248)

1 X
i i 0,
d
T
i=1

N
1 d X
i 0.

T dt i=1 i

(5.249)
(5.250)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

172

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

In terms of the chemical affinity, =

PN

i=1

i i , Eq. (5.250) can be written as

1 d
0.
(5.251)
T dt
Now one straightforward, albeit nave, way to guarantee positive semi-definiteness of the
irreversible entropy production and thus satisfaction of the second law is to construct the
chemical kinetic rate equation so that
N

X
d
= k
i i = k,
dt
i=1

k 0,

provisional, nave assumption

(5.252)

This provisional assumption of convenience will be supplanted later by a form which agrees
well with experiment. Here k is a positive semi-definite scalar. In general, it is a function of
temperature, k = k(T ), so that reactions proceed rapidly at high temperature and slowly at
low temperature. Then certainly the reaction progress variable will cease to change when
the equilibrium condition
N
X
i i = 0,
(5.253)
i=1

is met. This is equivalent to requiring

= 0.

(5.254)

Now, while Eq. (5.253) is the most compact form of the equilibrium condition, it is not
the most commonly used form. One can perform the following analysis to obtain the form
in most common usage. Start by equating the chemical potential with the Gibbs free energy
per unit mole for each species i: i = g i . Then employ the definition of Gibbs free energy
for an ideal gas, and carry out a set of operations:
N
X

g i i = 0,

at equilibrium,

(5.255)

(hi T si )i = 0,

at equilibrium.

(5.256)

i=1

N
X
i=1

For the ideal gas, one can substitute for hi (T ) and si (T, P ) and write the equilibrium condition as



Z
Z
N

T
T
X o

o
c P i (T )
yi P

dT R ln
cP i (T ) dT T s298,i +
i = 0.
h298,i +

Po
T
298
|
i=1
}
{z
}
| 298 {z

o
o
=s

=hT,i
T,i

{z
}
|
|
{z
}
o

=hT,i =hT,i

=sT,i

(5.257)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

173

5.4. CHEMICAL EQUILIBRIUM

Now writing the equilibrium condition in terms of the enthalpies and entropies referred to
the standard pressure, one gets



N 
X
yi P
o
o
i = 0,
hT,i T sT,i R ln
Po
i=1

(5.258)



N 
N

X
X
yi P
o
o
,
hT,i T sT,i i =
RT i ln
Po
{z
}
i=1 |
i=1

(5.259)

=g oT,i =oT,i

N
X

g oT,i i

|i=1 {z

Go

= RT

N
X
i=1

ln

yi P
Po

i



N
X
Go
yi P i
,

=
ln
P
RT
o
i=1
 !
N 
Y
yi P i
= ln
,
P
o
i=1



N
Y y i P  i
Go
=
,
exp
Po
RT
i=1
|
{z
}
KP

KP =


N 
Y
yi P i
i=1

KP =

P
Po

,
Po
PNi=1 i Y
n

yii .

(5.260)

(5.261)
(5.262)
(5.263)

(5.264)
(5.265)

i=1

So
KP =


N 
Y
Pi i
i=1

Po

, at equilibrium.

(5.266)

Here KP is what is known as the pressure-based equilibrium constant. It is dimensionless.


Despite its name, it is not a constant. It is defined in terms of thermodynamic properties,
and for the ideal gas is a function of T only:


Go
KP exp
,
RT

generally valid.

(5.267)

Only at equilibrium does the property KP also equal the product of the partial pressures
as in Eq. (5.266). The subscript P for pressure comes about because it is also related to
CC BY-NC-ND. 03 August 2012, J. M. Powers.

174

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

the product of the ratio of the partial pressure to the reference pressure raised to the net
stoichiometric coefficients. Also, the net change in Gibbs free energy of the reaction at the
reference pressure, Go , which is a function of T only, has been defined as
o

N
X

g oT,i i .

(5.268)

i=1

The term Go has units of kJ/kmole; it traditionally does not get an overbar. If Go > 0,
one has 0 < KP < 1, and reactants are favored over products. If Go < 0, one gets KP > 1,
and products are favored over reactants. One can also deduce that higher pressures P push
the equilibrium in such a fashion that fewer moles are present, all else being equal. One can
also define Go in terms of the chemical affinity, referred to the reference pressure, as
Go = o .

(5.269)

One can also define another convenient thermodynamic property, which for an ideal gas
is a function of T alone, the equilibrium constant Kc :
Kc

Po
RT

PNi=1 i



Go
,
exp
RT

generally valid.

(5.270)

This property is dimensional, and the


units depend on the stoichiometry of the reaction.
PN
3
i
i=1
The units of Kc will be (kmole/m )
.
The equilibrium condition, Eq. (5.266), is often written in terms of molar concentrations
and Kc . This can be achieved by the operations, valid only at an equilibrium state:
i
N 
Y
i RT
KP =
,
(5.271)
Po
i=1
PNi=1 i Y



N
RT
Go
i i ,
(5.272)
=
exp
Po
RT
i=1

PNi=1 i


N
o
Y
Po
G
i i .
(5.273)
=
exp
RT
RT
i=1
|
{z
}
Kc

So

Kc =

N
Y

i i ,

at equilibrium.

(5.274)

i=1

One must be careful to distinguish between the general definition of Kc as given in Eq. (5.270),
and the fact that at equilibrium it is driven to also have the value of product of molar species
concentrations, raised to the appropriate stoichiometric power, as given in Eq. (5.274).
CC BY-NC-ND. 03 August 2012, J. M. Powers.

5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION

5.5

175

Chemical kinetics of a single isothermal reaction

In the same fashion in ordinary mechanics that an understanding of statics enables an understanding of dynamics, an understanding of chemical equilibrium is necessary to understand
to more challenging topic of chemical kinetics. Chemical kinetics describes the time-evolution
of systems which may have an initial state far from equilibrium; it typically describes the
path of such systems to an equilibrium state. Here gas phase kinetics of ideal gas mixtures
that obey Daltons law will be studied. Important topics such as catalysis and solid or liquid
reactions will not be considered.
Further, this section will be restricted to strictly isothermal systems. This simplifies the
analysis greatly. It is straightforward to extend the analysis of this system to non-isothermal
systems. One must then make further appeal to the energy equation to get an equation for
temperature evolution.
The general form for evolution of species is taken to be
d
dt

 
i
i
= .

(5.275)

Multiplying both sides of Eq. (5.275) by molecular mass Mi and using the definition of mass
fraction ci then gives the alternate form
dci
i Mi
=
.
dt

5.5.1

(5.276)

Isochoric systems

Consider the evolution of species concentration in a system which is isothermal, isochoric


and spatially homogeneous. The system is undergoing a single chemical reaction involving
N species of the familiar form
N
X

i i = 0.

(5.277)

i=1

Because the density is constant for the isochoric system, Eq. (5.275) reduces to
di
= i .
dt

(5.278)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

176

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

Then, experiment, as well as a more fundamental molecular collision theory, shows that the
evolution of species concentration i is given by
z

}|

E
di
1 Y k

k
k ,
= i aT exp
1
dt
Kc

RT
|
{z
} | k=1{z }
| k=1
{z }
k(T )
reverse reaction
forward reaction
|
{z
}
N
Y

isochoric system. (5.279)

This relation actually holds for isochoric, non-isothermal systems as well, which will not be
considered in any detail here. Here some new variables are defined as follows:
a: a kinetic rate constant called the collision frequency factor. Its units will depend
on the actual reaction and could involve various combinations of length, time, and
temperature. It is constructed so
that di /dt has units of kmole/m3 /s; this requires it
PN

to have units of (kmole/m3 )(1 k=1 k ) /s/K .


: a dimensionless parameter whose value is set by experiments, sometimes combined
with guiding theory, to account for weak temperature dependency of reaction rates.
E: the activation energy. It has units of kJ/kmole, though others are often used, and is
fit by both experiment and fundamental theory to account for the strong temperature
dependency of reaction.
Note that in Eq. (5.279) that molar concentrations are raised to the k and k powers. As
it does not make sense to raise a physical quantity to a power with units, one traditionally
interprets the values of k , k , as well as k to be dimensionless pure numbers. They are
also interpreted in a standard fashion: the smallest integer values that actually correspond
to the underlying molecular collision which has been modeled. While stoichiometric balance
can be achieved by a variety of k values, the kinetic rates are linked to one particular set
which is defined by the community.
Equation (5.279) is written in such a way that the species concentration production rate
increases when
The net number of moles generated in the reaction, measured by i increases,
The temperature increases; here, the sensitivity may be very high, as one observes in
nature,
The species concentrations of species involved in the forward reaction increase; this
embodies the principle that the collision-based reaction rates are enhanced when there
are more molecules to collide,
CC BY-NC-ND. 03 August 2012, J. M. Powers.

5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION

177

The species concentrations of species involved in the reverse reaction decrease.


Here, three intermediate variables which are in common usage have been defined. First one
takes the reaction rate to be

!
 Y

N
N

1 Y k
E
k

r aT exp
k
k ,
(5.280)
1
Kc k=1

RT
k=1
|
{z
} | {z }
| {z }
k(T )

reverse reaction

forward reaction

or

N
1 Y k

r = aT exp

k .
Kc

k=1
|
{z
{z }
| {z }
| k=1
k(T ), Arrhenius rate
| forward reaction{z reverse reaction }

E
RT

N
Y

k k

(5.281)

law of mass action

The reaction rate r has units of kmole/m3 /s.


The temperature-dependency of the reaction rate is embodied in k(T ) is defined by what
is known as an Arrhenius rate law:


E

.
k(T ) aT exp
(5.282)
RT
This equation was advocated by vant Hoff in 1884; in 1889 Arrhenius gave a physical justification. The units of k(T ) actually depend on the reaction. This is a weakness of the
theory,
P
3 (1 N
k )
k=1
/s.
and precludes a clean non-dimensionalization. The units must be (kmole/m )
In terms of reaction progress, one can also take
r=

1 d
.
V dt

(5.283)

The factor of 1/V is necessary because r has units of molar concentration per time and
has units of kmoles. The over-riding importance of the temperature sensitivity is illustrated
as part of the next example. The remainder of the expression involving the products of the
species concentrations is the defining characteristic of systems which obey the law of mass
action. Though the history is complex, most attribute the law of mass action to Waage and
Guldberg in 1864.1
Last, the overall molar production rate of species i, is often written as i , defined as
i i r.

(5.284)

P. Waage and C. M. Guldberg, 1864, Studies Concerning Affinity, Forhandlinger: Videnskabs-Selskabet


i Christiania, 35. English translation: Journal of Chemical Education, 63(12): 1044-1047.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

178

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

As i is considered to be dimensionless, the units of i must be kmole/m3 /s.


Example 5.13
Study the nitrogen dissociation problem considered in an earlier example, see p. 140, in which at
t = 0 s, 1 kmole of N2 exists at P = 100 kP a and T = 6000 K. Take as before the reaction to be
isothermal and isochoric. Consider again the elementary nitrogen dissociation reaction
N2 + N2 2N + N2 ,

(5.285)

which has kinetic rate parameters of


a

7.0 1021

1.6,

cm3 K 1.6
,
mole s

(5.286)
(5.287)

cal
224928.4
.
mole

(5.288)

In SI units, this becomes



3 


m3 K 1.6
1m
1000 mole
cm3 K 1.6
= 7.0 1018
,
a =
7.0 1021
mole s
100 cm
kmole
kmole s





cal
J
kJ
kJ
1000 mole
E =
224928.4
4.186
= 941550
.
mole
cal
1000 J
kmole
kmole

(5.289)
(5.290)

At the initial state, the material is all N2 , so PN2 = P = 100 kP a. The ideal gas law then gives at
t=0

(5.291)
P |t=0 = PN2 |t=0 = N2 t=0 RT,

P |t=0
,
(5.292)
N2 t=0 =
RT
100 kP a

,
(5.293)
=
kJ
8.314 kmole
K (6000 K)
kmole
= 2.00465 103
.
(5.294)
m3
Thus, the volume, constant for all time in the isochoric process, is
V =

nN2 |t=0
1 kmole

=

N2 t=0
2.00465 103

kmole
m3

= 4.9884 102 m3 .

(5.295)

Now the stoichiometry of the reaction is such that


dnN2

(nN2 nN2 |t=0 ) =


| {z }

=
=
=

(5.296)
(5.297)

=0

=1 kmole

nN
nN
V

1
dnN ,
2
1
(nN nN |t=0 ),
2
| {z }

2(1 kmole nN2 ),




1 kmole nN2
2

,
V
V


1 kmole

N2 ,
4.9884 102 m3


kmole

2 2.00465 103
N2 .
m3

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.298)
(5.299)
(5.300)
(5.301)

5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION

179

k(T) (m /kmole/s)
10000
0.0001
1. x 10-12
1. x 10-20
1. x 10-28
1. x 10-36
1000

1500 2000

3000

5000

T (K)
7000 10000

Figure 5.1: k(T ) for Nitrogen dissociation example.


Now the general equation for kinetics of a single reaction, Eq. (5.279), reduces for N2 molar concentration to





dN2

1
E
(5.302)
(N2 )N2 (N )N 1
(N2 )N2 (N )N .
= N2 aT exp
dt
Kc
RT

Realizing that N
= 2, N
= 0, N2 = 1, and N = 2, one gets
2




dN2
1 2N
E
2N2 1
.
= aT exp
dt
K c N 2
RT
{z
}
|

(5.303)

=k(T )

Examine the primary temperature dependency of the reaction




E
k(T ) = aT exp
,
RT
!


kJ
3 1.6
941550 kmole
18 m K
1.6
=
7.0 10
T
exp
,
kJ
kmole s
8.314 kmole
KT


7.0 1018
1.1325 105
.
=
exp
T 1.6
T

(5.304)
(5.305)
(5.306)

Figure 5.1 gives a plot of k(T ) which shows its very strong dependency on temperature. For this
problem, T = 6000 K, so


1.1325 105
7.0 1018
,
(5.307)
exp
k(6000) =
60001.6
6000
m3
= 40071.6
.
(5.308)
kmole s
Now, the equilibrium constant Kc is needed. Recall
PN



i=1 i
Go
Po
.
(5.309)
exp
Kc =
RT
RT
CC BY-NC-ND. 03 August 2012, J. M. Powers.

180

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

f ( N

)
(kmole/m 3 /s)
2
unstable
equilibrium
1

0.0005

0.001

0.0015

0.002

-1

unstable
equilibrium

-2

N2

0.0025

(kmole/m 3 )

stable,
physical
equilibrium

Figure 5.2: Forcing function, f (N2 ), which drives changes of N2 as a function of N2 in


isothermal, isochoric problem.
PN
For this system, since i=1 i = 1, this reduces to




(2goN g oN2 )
Po
exp
,
(5.310)
Kc =
RT
RT
!
o
o


(2(hN T soT,N ) (hN2 T soT,N2 ))
Po
=
exp
,
(5.311)
RT
RT




100
(2(597270 (6000)216.926) (205848 (6000)292.984))
=
exp
,(5.312)
(8.314)(6000)
(8.314)(6000)
kmole
= 0.000112112
.
(5.313)
m3
The differential equation for N2 evolution is then given by
dN2
dt


= 40071.6
|

m3
kmole

2N2

2 2.00465 103
1
1
N 2
0.000112112 kmole
m3
{z

= f (N2 ).

f (N2 )

kmole
m3

N2

2 !

(5.314)
(5.315)

The system is at equilibrium when f (N2 ) = 0. This is an algebraic function of N2 only, and can be
plotted. Figure 5.2 gives a plot of f (N2 ) and shows that it has three potential equilibrium points. It
is seen there are three roots. Solving for the equilibria requires solving
2 !



2 2.00465 103 kmole
N 2
1
m3
2
m3
.
N 2 1
0 = 40071.6
kmole
N 2
0.000112112 kmole
m3
(5.316)
CC BY-NC-ND. 03 August 2012, J. M. Powers.

5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION

181

The three roots are


kmole
N 2 = 0
,
m3 }
| {z
unstable

0.00178121
{z
|

stable

kmole
,
m3 }

0.00225611
{z
|

kmole
.
m3 }

(5.317)

unstable

By inspection of the topology of Fig. 5.2, the only stable root is 0.00178121 kmole
m3 . This root agrees with
the equilibrium value found in an earlier example for the same problem conditions. Small perturbations
from this equilibrium induce the forcing function to supply dynamics which restore the system to its original equilibrium state. Small perturbations from the unstable equilibria induce non-restoring dynamics.
For this root, one can then determine that the stable equilibrium value of N = 0.000446882 kmole
m3 .
One can examine this stability more formally. Define an equilibrium concentration eq
N2 such that
f (eq
N2 ) = 0.

(5.318)

eq
N2 :

Now perform a Taylor series of f (N2 ) about N2 =



df
1 d2 f
eq 2
+
f (N2 ) f (eq
)
(N2 eq
N2
N2 ) +
2 (N2 N2 ) + . . .

d
2
d
eq
| {z }
N2 N =
N2
N2

=0

(5.319)

Now the first term of the Taylor series is zero by construction. Next neglect all higher order terms as
small so that the approximation becomes

df
f (N2 )
( eq
(5.320)
N2 ).
dN2 =eq N2
N2

N2

Thus, near equilibrium, one can write


dN2
df

dt
dN2

eq
N2 =N2

(N2 eq
N2 ).

(5.321)

Since the derivative of a constant is zero, one can also write the equation as

d
df
(N2 eq
( eq
)

N2
N2 ).
dt
dN2 =eq N2
N2

(5.322)

N2

This has a solution, valid near the equilibrium point, of



df

(N2 eq
t ,
N2 ) = C exp
dN2 =eq
N2
N2



df

N2 = eq
t .
N2 + C exp
d
eq
N2 N =
N
2

(5.323)

(5.324)

Here C is some constant whose value is not important for this discussion. If the slope of f is positive,
that is,

df
> 0,
unstable,
(5.325)
dN2 =eq
N2

N2

the equilibrium will be unstable. That is a perturbation will grow without bound as t . If the
slope is zero,

df
= 0,
neutrally stable,
(5.326)
dN2 =eq
N2

N2

CC BY-NC-ND. 03 August 2012, J. M. Powers.

182

concentration (kmole/m 3 )

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

0.002

N2

0.0015
0.001
0.0005

0.001 0.002 0.003 0.004 0.005

t (s)

Figure 5.3: N2 (t) and N (t) in isothermal, isochoric nitrogen dissociation problem.
the solution is stable in that there is no unbounded growth, and moreover is known as neutrally stable.
If the slope is negative,

df
< 0,
asymptotically stable,
(5.327)
dN2 =eq
N2

N2

the solution is stable in that there is no unbounded growth, and moreover is known as asymptotically
stable.
A numerical solution via an explicit technique such as a Runge-Kutta integration is found for
Eq. (5.314). The solution for N2 , along with N is plotted in Fig. 5.3.
Linearization of Eq. (5.314) about the equilibrium state gives rise to the locally linearly valid

d
0.00178121 = 1209.39(N2 0.00178121) + . . .
dt N2

(5.328)

This has local asymptotically stable solution

N2 = 0.00178121 + C exp (1209.39t) .

(5.329)

Here C is some integration constant whose value is irrelevant for this analysis. The time scale of
relaxation is the time when the argument of the exponential is 1, which is
=

1
= 8.27 104 s.
1209.39 s1

(5.330)

One usually finds this time scale to have high sensitivity to temperature, with high temperatures giving
fast time constants and thus fast reactions.
The equilibrium values agree exactly with those found in the earlier example; see Eq. (4.469). Here
the kinetics provide the details of how much time it takes to achieve equilibrium. This is one of the key
questions of non-equilibrium thermodynamics.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION

5.5.2

183

Isobaric systems

The form of the previous section is the most important as it is easily extended to a Cartesian
grid with fixed volume elements in fluid flow problems. However, there is another important
spatially homogeneous problem in which the formulation needs slight modification: isobaric
reaction, with P equal to a constant. Again, in this section only isothermal conditions will
be considered.
In an isobaric problem, there can be volume change. Consider first the problem of isobaric
expansion of an inert mixture. In such a mixture, the total number of moles of each species
must be constant, so one gets
dni
= 0,
dt

inert, isobaric mixture.

(5.331)

Now carry out the sequence of operations, realizing the the total mass m is also constant:
1 d
(ni )
m dt
d  ni 
dt m

d ni V
dt V m
 
d i
dt
d
1 di
2i
dt
dt
di
dt

= 0,

(5.332)

= 0,

(5.333)

= 0,

(5.334)

= 0,

(5.335)

= 0,

(5.336)

i d
.
dt

(5.337)

So a global density decrease of the inert material due to volume increase of a fixed mass
system induces a concentration decrease of each species. Extended to a material with a
single reaction rate r, one could say either
di
d
= i r + i ,
or
dt
dt
 
d i
1
=
i r,
generally valid,
dt

i
=
.

(5.338)
(5.339)
(5.340)

Equation (5.339) is consistent with Eq. (5.275) and is actually valid for general systems with
variable density, temperature, and pressure.
However, in this section, it is required that pressure and temperature be constant. Now
CC BY-NC-ND. 03 August 2012, J. M. Powers.

184

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

differentiate the isobaric, isothermal, ideal gas law to get the density derivative.

P =
0 =

N
X

i=1
N
X

i RT,
RT

i=1
N
X

(5.341)

di
,
dt

(5.342)

di
,
dt
i=1

N 
X
i d
i r +
0 =
,
dt
i=1

0 =

1 d X
0 = r
i +
,
dt i=1 i
i=1
P
r N
i
d
=
PN i=1i .
dt
i=1
P
r N
i
= PNi=1 ,
i=1 i
PN
r i=1 i
,
=
P
RT

(5.344)

N
X

(5.343)

RT r
RT r

PN

i=1

P
PN

k=1 k

(5.345)
(5.346)
(5.347)
(5.348)
(5.349)

(5.350)

PN
Note that if there is no net number change in the reaction,
k=1 k = 0, the isobaric,
isothermal reaction also guarantees there would be no density or volume change. It is
convenient to define the net number change in the elementary reaction as n:

N
X

k .

(5.351)

k=1

Here n is taken to be a dimensionless pure number. It is associated with the number


change in the elementary reaction and not the actual mole change in a physical system; it
is, however, proportional to the actual mole change.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION

185

Now use Eq. (5.350) to eliminate the density derivative in Eq. (5.338) to get
P
di
i RT r N
k=1 k
= i r
,
dt

(5.352)

N
X

RT
i
k
i
= r

,
|{z}
P k=1
reaction effect

|
{z
}

(5.353)

expansion effect

or

di
= r
dt

i
|{z}

reaction effect

yi n
| {z }

expansion effect

(5.354)

There are two terms dictating the rate change of species molar concentration. The first, a
reaction effect, is precisely the same term that drove the isochoric reaction. The second is
due to the fact that the volume can change if the number of moles change, and this induces
an intrinsic change in concentration. Note that the term i RT /P = yi, the mole fraction.
Example 5.14
Study a variant of the nitrogen dissociation problem considered in an earlier example, see p. 145,
in which at t = 0 s, 1 kmole of N2 exists at P = 100 kP a and T = 6000 K. In this case, take the
reaction to be isothermal and isobaric. Consider again the elementary nitrogen dissociation reaction
N2 + N2 2N + N2 ,

(5.355)

which has kinetic rate parameters of


a

7.0 1021

1.6,

cm3 K 1.6
,
mole s

(5.356)
(5.357)

cal
224928.4
.
mole

(5.358)

In SI units, this becomes


a


3 


3
1.6
m3 K 1.6
1m
1000 mole
21 cm K
= 7.0 1018
,
7.0 10
mole s
100 cm
kmole
kmole s





cal
J
kJ
kJ
1000 mole
4.186
= 941550
224928.4
.
mole
cal
1000 J
kmole
kmole

(5.359)
(5.360)

At the initial state, the material is all N2 , so PN2 = P = 100 kP a. The ideal gas law then gives at
CC BY-NC-ND. 03 August 2012, J. M. Powers.

186

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


t=0
= PN2 = N2 RT,
P
,
=
RT
100 kP a

,
=
kJ
8.314 kmole
K (6000 K)
kmole
= 2.00465 103
.
m3

P
N 2

t=0

(5.361)
(5.362)
(5.363)
(5.364)

Thus, the initial volume is


V |t=0 =

nN2 |t=0
1 kmole

=
N2 t=0
2.00465 103

kmole
m3

= 4.9884 102 m3 .

(5.365)

In this isobaric process, one always has P = 100 kP a. Now, in general


P = RT (N2 + N );

(5.366)

therefore, one can write N in terms of N2 :


N

=
=
=

P
N 2 ,
RT
100 kP a

N2 ,
kJ
8.314 kmole
K (6000 K)


kmole
N 2 .
2.00465 103
m3

(5.367)
(5.368)
(5.369)

Then the equations for kinetics of a single isobaric isothermal reaction, Eq. (5.353) in conjunction
with Eq. (5.280), reduce for N2 molar concentration to
!






dN2

RT

E
1
N

( )N2 (N )N
=
aT exp
(N2 + N ) .
N2 2
(N2 ) N2 (N )N 1
dt
Kc N2
P
RT
{z
}
|
=r

(5.370)

Realizing that N
= 2, N
= 0, N2 = 1, and N = 2, one gets
2

dN2
dt

aT exp
{z
|

E
RT

=k(T )

!

2 

RT

1
N
2N2 1
1 N2
.
K c N 2
P

The temperature dependency of the reaction is unchanged from the previous reaction:


E

k(T ) = aT exp
,
RT
!


kJ
3 1.6
941550 kmole
1.6
18 m K
,
T
exp
=
7.0 10
kJ
kmole s
8.314 kmole
KT


7.0 1018
1.1325 105
.
=
exp
T 1.6
T
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.371)

(5.372)
(5.373)
(5.374)

187

5.5. CHEMICAL KINETICS OF A SINGLE ISOTHERMAL REACTION


For this problem, T = 6000 K, so


7.0 1018
1.1325 105
,
exp
60001.6
6000
m3
40130.2
.
kmole s

k(6000) =
=

(5.375)
(5.376)

The equilibrium constant Kc is also unchanged from the previous example. Recall
Kc =

Po
RT

PN
i=1 i

exp

Go
RT

(5.377)

P
For this system, since N
i=1 i = n = 1, this reduces to




(2goN g oN2 )
Po
exp
,
(5.378)
Kc =
RT
RT




(2goN g oN2 )
Po
exp
,
(5.379)
=
RT
RT
!
o
o


(2(hN T soT,N ) (hN2 T soT,N2 ))
Po
=
exp
,
(5.380)
RT
RT




100
(2(597270 (6000)216.926) (205848 (6000)292.984))
=
exp
,(5.381)
(8.314)(6000)
(8.314)(6000)
kmole
= 0.000112112
.
(5.382)
m3
The differential equation for N2 evolution is then given by
dN2
dt


40130.2
1

m3
kmole
N 2

1
1
0.000112112 kmole
m3
!

kJ
8.314 kmole K (6000 K)
,
100 kP a
2N2


2 !
2.00465 103 kmole
N 2
m3
N 2

(5.383)

f (N2 ).

(5.384)

The system is at equilibrium when f (N2 ) = 0. This is an algebraic function of N2 only, and can be
plotted. Figure 5.4 gives a plot of f (N2 ) and shows that it has four potential equilibrium points. It is
seen there are four roots. Solving for the equilibria requires solving

2 !


3 kmole

2.00465

10

m3
1
3
N
m
2
0 =
40130.2
2N2 1
kmole
N 2
0.000112112 kmole
m3
!

kJ
N2 8.314 kmole
K (6000 K)
1
.
100 kP a
(5.385)
The four roots are
N2 = 0.002005
|
{z

kmole
,
m3 }

stable,nonphysical

kmole
0
,
m3 }
| {z
unstable

kmole
0.001583
,
3
|
{z m }
stable,physical

kmole
0.00254
.
3
|
{z m }

(5.386)

unstable

CC BY-NC-ND. 03 August 2012, J. M. Powers.

188

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


f ( N

) (kmole/m 3 /s)
4

unstable
equilibrium

unstable
equilibrium

3
2
1

-0.002

-0.001

0.001

0.002

0.003

N2

(kmole/m 3 )

-1

stable,
physical
equilibrium

stable,
non-physical
equilibrium

Figure 5.4: Forcing function, f (N2 ), which drives changes of N2 as a function of N2 in


isothermal, isobaric problem.
By inspection of the topology of Fig. 5.2, the only stable, physical root is 0.001583 kmole
m3 . Small
perturbations from this equilibrium induce the forcing function to supply dynamics which restore the
system to its original equilibrium state. Small perturbations from the unstable equilibria induce nonrestoring dynamics. For this root, one can then determine that the stable equilibrium value of N =
0.000421 kmole
m3 .
A numerical solution via an explicit technique such as a Runge-Kutta integration is found for
Eq. (5.385). The solution for N2 , along with N is plotted in Fig. 5.5.
Linearization of Eq. (5.385) about the equilibrium state gives rise to the locally linearly valid

d
N2 0.001583 = 967.073(N2 0.001583) + . . .
dt

(5.387)

This has local solution

N2 = 0.001583 + C exp (967.073t) .

(5.388)

Again, C is an irrelevant integration constant. The time scale of relaxation is the time when the
argument of the exponential is 1, which is
=

1
= 1.03 103 s.
967.073 s1

(5.389)

Note that the time constant for the isobaric combustion is about a factor 1.25 greater than for isochoric
combustion under the otherwise identical conditions.
The equilibrium values agree exactly with those found in the earlier example; see Eq. (4.514).
Again, the kinetics provide the details of how much time it takes to achieve equilibrium.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

189

5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS

concentration (kmole/m 3 )

0.002

N2

0.0015

0.001

0.0005

0.001

0.002

0.003

0.004

t (s)
0.005

Figure 5.5: N2 (t) and N (t) in isobaric, isothermal nitrogen dissociation problem.

5.6

Some conservation and evolution equations

Here a few useful global conservation and evolution equations are presented for some key
properties. Only some cases are considered, and one could develop more relations for other
scenarios.

5.6.1

Total mass conservation: isochoric reaction

One can easily show that the isochoric reaction rate model, Eq. (5.279), satisfies the principle
of mixture mass conservation. Begin with Eq. (5.279) in a compact form, using the definition
of the reaction rate r, Eq. (5.281) and perform the following operations:

di
 dt
d ci
dt Mi
d
(ci )
dt
d
(ci )
dt

= i r,

(5.390)

= i r,

(5.391)

= i Mi r,

(5.392)

= i

L
X

Ml li r,

|l=1 {z

=Mi

L
X
d
(ci ) =
Ml lii r,
dt
l=1

(5.393)

(5.394)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

190

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


N X
L
N
X
X
d
Ml li i r,
(ci ) =
dt
i=1 l=1
i=1

L X
N
N

X
d
X
Ml li i r,
ci =

dt i=1
l=1 i=1
| {z }

(5.395)

(5.396)

=1

X
X
d
= r
Ml
li i .
dt
i=1
l=1
| {z }

(5.397)

=0

Therefore, we get

d
= 0.
dt
Note the term

5.6.2

PN

i=1

(5.398)

li i = 0 because of stoichiometry, Eq. (5.24).

Element mass conservation: isochoric reaction

Through a similar series of operations, one can show that the mass of each element, l =
1, . . . , L, in conserved in this reaction, which is chemical, not nuclear. Once again, begin
with Eq. (5.281) and perform a set of operations,
di
= i r,
dt
d
li i = li i r,
dt

(5.399)
l = 1, . . . , L,

d
l = 1, . . . , L,
(li i ) = rli i ,
dt
N
N
X
X
d
(li i ) =
rlii ,
l = 1, . . . , L,
dt
i=1
i=1
!
N
N
X
d X
= r
li i ,
l = 1, . . . , L,
li i
dt i=1
i=1
| {z }
=0
!
N
d X
li i
= 0,
l = 1, . . . , L.
dt i=1
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.400)
(5.401)
(5.402)
(5.403)

(5.404)

5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS

191

P
The term N
i=1 li i represents the number of moles of element l per unit volume, by the
following analysis
N
X
i=1

N
X
moles element l
moles element l moles species i
li i =
=
l e .
moles
species
i
volume
volume
i=1

(5.405)

Here the elemental mole density, l e , for element l has been defined. So the element concentration for each element remains constant in a constant volume reaction process:
dl e
= 0,
dt

l = 1, . . . , L.

(5.406)

One can also multiply by the elemental mass, Ml to get the elemental mass density, el :
el Ml l e ,

l = 1, . . . , L.

(5.407)

Since Ml is a constant, one can incorporate this definition into Eq. (5.406) to get
del
= 0,
dt

l = 1, . . . , L.

(5.408)

The element mass density remains constant in the constant volume reaction. One could also
simply say since the elements density is constant, and the mixture is simply a sum of the
elements, that the mixture density is conserved as well.

5.6.3

Energy conservation: adiabatic, isochoric reaction

Consider a simple application of the first law of thermodynamics to reaction kinetics: that
of a closed, adiabatic, isochoric combustion process in a mixture of ideal gases. One may
be interested in the rate of temperature change. First, because the system is closed, there
can be no mass change, and because the system is isochoric, the total volume is a non-zero
constant; hence,
dm
= 0,
dt
d
(V ) = 0,
dt
d
= 0,
V
dt
d
= 0.
dt

(5.409)
(5.410)
(5.411)
(5.412)

For such a process, the first law of thermodynamics is


dU
.
= Q W
dt

(5.413)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

192

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

But there is no heat transfer or work in the adiabatic isochoric process, so one gets
dU
dt

= 0,

(5.414)

d
(mu) = 0,
dt
du
dm
m +u
= 0,
dt
dt
|{z}

(5.415)
(5.416)

=0

du
= 0.
dt

(5.417)

Thus for the mixture of ideal gases, u(T, 1 , . . . , N ) = uo. One can see how reaction rates
affect temperature changes by expanding the derivative in Eq. (5.417)
!
N
d X
ci u i
= 0,
(5.418)
dt i=1
N
X
d
(ci ui )
dt
i=1


N
X
dci
dui
+ ui
ci
dt
dt
i=1

N 
X
dci
dui dT
+ ui
ci
dT
dt
dt
i=1


N
X
dT
dci
ci cvi
+ ui
dt
dt
i=1

= 0,

(5.419)

= 0,

(5.420)

= 0,

(5.421)

= 0,

(5.422)

N
N
X
dci
dT X
ui ,
ci cvi =
dt i=1
dt
i=1
| {z }
=cv

dT
cv
dt

cv
cv

dT
dt

dT
dt

dT
dt
CC BY-NC-ND. 03 August 2012, J. M. Powers.

N
X
i=1

N
X
i=1
N
X

(5.423)

d
ui
dt

ui Mi

di
,
dt

Mi i

ui Mi i r,

(5.424)
(5.425)
(5.426)

i=1

PN

i=1

cv

i ui

(5.427)

5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS

193

If one defines the net energy change of the reaction as

N
X

i ui ,

(5.428)

rU
dT
=
.
dt
cv

(5.429)

i=1

one then gets

The rate of temperature change is dependent on the absolute energies, not the energy differences. If the reaction is going forward, so r > 0, and that is a direction in which the net
molar energy change is negative, then the temperature will rise.

5.6.4

Energy conservation: adiabatic, isobaric reaction

Solving for the reaction dynamics in an adiabatic isobaric system requires some non-obvious
manipulations. First, the first law of thermodynamics says dU = dQdW . Since the process
is adiabatic, one has dQ = 0, so dU + P dV = 0. Since it is isobaric, one gets d(U + P V ) = 0,
or dH = 0. So the total enthalpy is constant. Then
d
H = 0,
dt

d
dt

d
(mh) = 0,
dt
dh
= 0,
dt
!
N
X
ci hi
= 0,

i=1
N
X
i=1

d
(ci hi ) = 0,
dt

N
X
dci
dhi
+ hi
= 0,
ci
dt
dt
i=1

N
X
dhi dT
dci
ci
+ hi
= 0,
dT dt
dt
i=1

(5.430)
(5.431)
(5.432)
(5.433)
(5.434)
(5.435)
(5.436)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

194

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION


N
X
i=1

dT X dci
+
hi
= 0,
ci cP i
dt
dt
i=1

(5.437)

N
N
X
dT X
dci
hi
ci cP i +
= 0,
dt i=1
dt
i=1
| {z }
=cP

(5.438)


i Mi
= 0,

 
N
d i
dT X
= 0.
+
hi Mi
cP
dt
dt

i=1
N

dT X d
cP
+
hi
dt
dt
i=1

(5.439)
(5.440)

Now use Eq (5.339) to eliminate the term in Eq. (5.440) involving molar concentration
derivatives to get
N

dT X i r
cP
hi
+
= 0,
dt

i=1
dT
dt

(5.441)
r

PN

i=1 hi i

cP

(5.442)

So the temperature derivative is known as an algebraic function. If one defines the net
enthalpy change as
N
X
H
hi i ,
(5.443)
i=1

one gets

dT
rH
.
=
dt
cP

(5.444)

Now differentiate the isobaric ideal gas law to get the density derivative.
P =

N
X

i RT,

i=1
N
X

(5.445)
N

dT X
d
0 =
+
i R
RT i ,
dt
dt
i=1
i=1


N
N
X
dT X
i d
,
+
T i r +
0 =
dt i=1 i i=1
dt

N
N
N
X
1 dT X
1 d X
0 =
i +
+r
.
T dt i=1 i
dt i=1 i
i=1

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(5.446)
(5.447)
(5.448)

5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS


Solving, we get
1 dT
d
= T dt
dt

PN

i=1 i r
PN i
i=1

PN

i=1

195

(5.449)

One takes dT /dt from Eq. (5.442) to get


d
=
dt

1 r
T

PN

i=1

hi i

cP

PN

i=1 i r
PN i
i=1

PN

i=1

(5.450)

Now recall that = /M and cP = cP M, so cP = cP . Then Equation (5.450) can be


reduced slightly:
=1

z }| {
N
X

PN

hi i
cP T

i=1

d
= r
dt
=

i=1
PN
i=1

i
PN

PN

PN

i=1

i
,

hi i
i=1 i
i=1 cP T
,
r
PN
i=1 i


PN
hi
i=1 i cP T 1

= r

(5.452)

P
RT

(5.451)

(5.453)



N
RT X
hi
= r
i
1 ,
P i=1
cP T


N
X
hi
1 ,
= rM
i
cP T
i=1

(5.454)
(5.455)

PN
where M is the mean molecular mass. Note for exothermic reaction
i=1 i hi < 0, so
exothermic reaction induces a density decrease as the increased temperature at constant
pressure causes the volume to increase.
Then using Eq. (5.455) to eliminate the density derivative in Eq. (5.338), and changing
the dummy index from i to k, one gets an explicit expression for concentration evolution:


N
X
di
i
hk
= i r + rM
k
1 ,
(5.456)
dt

cP T
k=1



N

i X
hk

= r i + M
1
(5.457)
k
,

cP T
| {z } k=1
=yi

= r i + yi

N
X
k=1

hk
1
cP T

!

(5.458)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

196

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

Defining the change of enthalpy of the reaction as H


P
number of the reaction as n N
k=1 k , one can also say

PN

k=1

k hk , and the change of




di
H
= r i + yi
n
.
dt
cP T

(5.459)

Exothermic reaction, H < 0, and net number increases, n > 0, both tend to decrease the
molar concentrations of the species in the isobaric reaction.
Lastly, the evolution of the adiabatic, isobaric system, can be described by the simultaneous, coupled ordinary differential equations: Eqs. (5.442, 5.450, 5.458). These require
numerical solution in general. Note also that one could also employ a more fundamental
P
treatment as a differential algebraic system involving H = H1 , P = P1 = RT N
i=1 i and
Eq. (5.338).

5.6.5

Entropy evolution: Clausius-Duhem relation

Now consider whether the kinetics law that has been posed actually satisfies the second law
of thermodynamics. Consider again Eq. (4.387). There is an algebraic relation on the right
side. If it can be shown that this algebraic relation is positive semi-definite, then the second
law is satisfied, and the algebraic relation is known as a Clausius-Duhem relation.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

197

5.6. SOME CONSERVATION AND EVOLUTION EQUATIONS


Now take Eq. (4.387) and perform some straightforward operations on it:

dS|U,V

N
1X

dni
T i=1 i
|
{z
}

0,

(5.460)

irreversible entropy production


dS
dt U,V

N
V X dni 1
0,

T i=1 i dt V

(5.461)

N
V X di

=
0,
T i=1 i dt

(5.462)

!

 Y
N
N
N

1 Y k
E
V X

k
i aT exp
k
k 0,(5.463)
=
1
T i=1 i
K

RT
c
|
{z
} | k=1{z }
| k=1
{z }
k(T )
reverse reaction
forward reaction
{z
}
|
r
!
!
N
N
N
Y
Y

1
V X

1
(5.464)
i k(T )
k k
kk 0,
=
T i=1 i
K
c
k=1
k=1
!
! N
!
N
N
X
Y
Y

1
V

1
= k(T )
k k
kk
i i 0,
(5.465)
T
Kc k=1
i=1
k=1
|
{z
}
=

Change the dummy index from k back to i,


!
!
N
N
Y
Y

V
1

=
k(T )
i i
i i 0,
1
T
Kc i=1
i=1

(5.466)

(5.467)

V
r,
T
d
=
.
T dt

(5.468)

Consider now the affinity term in Eq. (5.466) and expand it so that it has a more useful
form:

N
X
i=1

i i =
=

N
X

g i i ,
i=1
N 
X
g oT,i

i=1

(5.469)
+ RT ln

Pi
Po



i ,

(5.470)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

198

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

N
X

g oT,i i

i=1

{z

=Go

RT

N
X

ln

i=1

N
Go X
= RT

ln
RT
| {z } i=1
=ln KP

= RT

=
=

ln KP ln

Pi
Po

Pi
Po

i

 i


N 
Y
Pi i

(5.471)

(5.472)

,
P
o
i=1
 !
N 
Y
1
Pi i
RT ln
,
+ ln
KP
Po
i=1
 !
N 
1 Y Pi i
,
RT ln
KP i=1 Po

PNi=1 i N 
Po
Y RT i
i
,
RT ln RT
Kc
P
o
i=1
!
N
1 Y i

.
RT ln
Kc i=1 i

(5.473)
(5.474)
(5.475)

(5.476)

(5.477)

Equation (5.477) is the common definition of affinity. Another form can be found by employing the definition of Kc from Eq. (5.270) to get
!

 PNi=1 i
 N

Po
Go Y i
= RT ln
i
,
(5.478)
exp
RT
RT i=1
!!

 PNi=1 i Y
N
Go
Po
i i
= RT
,
(5.479)
+ ln
RT
RT
i=1
!

 PNi=1 i Y
N
P
o
i i .
= Go RT ln
(5.480)
RT
i=1
To see clearly that the entropy production rate is positive semi-definite, substitute
Eq. (5.477) into Eq. (5.466) to get
!
!
!!

N
N
N
Y
dS
V
1 Y i
1 Y i
i
=
k(T )
i
1

RT ln

0,
dt U,V
T
Kc i=1 i
Kc i=1 i
i=1
= RV k(T )

N
Y
i=1

CC BY-NC-ND. 03 August 2012, J. M. Powers.

1
Kc

N
Y
i=1

i i

ln

1
Kc

N
Y
i=1

i i

(5.481)

0.

(5.482)

199

5.7. SIMPLE ONE-STEP KINETICS


Define forward and reverse reaction coefficients, R , and R , respectively, as

R k(T )
R

k(T )
Kc

N
Y

i=1
N
Y

(5.483)

(5.484)

i i ,
i i .

i=1

Both R and R have units of kmole/m3 /s. It is easy to see that


r = R R .

(5.485)

Note that since k(T ) > 0, Kc > 0, and i 0, that both R 0 and R 0. Since
i = i i , one finds that
N
N
1 Y i
1 k(T ) Y i i R
i =

= .
Kc i=1
Kc k(T ) i=1 i
R

Then Eq. (5.482) reduces to



dS
dt U,V

= RV R

R
1
R

ln

R
R

0.

(5.486)

(5.487)

Finally, we get


 
dS
R

= RV (R R ) ln
0.

dt U,V
R

(5.488)

Obviously, if the forward rate is greater than the reverse rate R R > 0, ln(R /R ) > 0,
and the entropy production is positive. If the forward rate is less than the reverse rate,
R R < 0, ln(R /R ) < 0, and the entropy production is still positive. The production
rate is zero when R = R .
Note that the affinity can be written as
 
R
= RT ln
.
(5.489)
R
And so when the forward reaction rate exceeds the reverse, the affinity is positive. It is zero
at equilibrium, when the forward reaction rate equals the reverse.

5.7

Simple one-step kinetics

A common model in theoretical combustion is that of so-called simple one-step kinetics. Such
a model, in which the molecular mass does not change, is quantitatively appropriate only
CC BY-NC-ND. 03 August 2012, J. M. Powers.

200

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

for isomerization reactions. However, as a pedagogical tool as well as a qualitative model for
real chemistry, it can be valuable.
Consider the reversible reaction
A B.
(5.490)
where chemical species A and B have identical molecular masses MA = MB = M. Consider
further the case in which at the initial state, no moles of A only are present. Also take the
reaction to be isochoric and isothermal. These assumptions can easily be relaxed for more
general cases. Specializing then Eq. (5.240) for this case, one has
nA =

A + nAo ,
|{z}
|{z}
=1

nB =

=no

B + nBo .
|{z}
|{z}
=1

Thus

(5.491)
(5.492)

=0

nA = + no ,
nB = .

(5.493)
(5.494)

Now no is constant throughout the reaction. Scale by this and define the dimensionless
reaction progress as /no to get
nA
= + 1,
(5.495)
no
|{z}
=yA

nB

= .
no
|{z}

(5.496)

=yB

In terms of the mole fractions then, one has

yA = 1 ,

yB = .
The reaction kinetics for each species reduce to
no
dA
= r,
o ,
A (0) =
dt
V
dB
= r,
B (0) = 0.
dt
Addition of Eqs. (5.499) and (5.500) gives
d
( + B ) = 0,
dt A
A + B = o ,
A

+ B = 1.
o
o
|{z}
|{z}
=yA

CC BY-NC-ND. 03 August 2012, J. M. Powers.

=yB

(5.497)
(5.498)

(5.499)
(5.500)

(5.501)
(5.502)
(5.503)

201

5.7. SIMPLE ONE-STEP KINETICS


In terms of the mole fractions yi , one then has
yA + yB = 1.

(5.504)

The reaction rate r is then


r =
=
=
=


1 B
kA 1
,
Kc A


A
1 B /o
ko
1
,
o
Kc A /o


1 yB
,
ko yA 1
Kc y A
!

1
ko (1 )
.
Kc 1


(5.505)
(5.506)
(5.507)
(5.508)

Now r = (1/V )d/dt = (1/V )d(no )/dt


= (no /V )d()/dt
= o d/dt.
So the reaction
dynamics can be described by a single ordinary differential equation in a single unknown:
!

1
d
1
= ko (1 )
,
(5.509)
o
dt
Kc 1
!

d
1

1
= k(1 )
.
(5.510)
dt
Kc 1
Equation (5.510) is in equilibrium when
=

1
1
+ ...
1 1
Kc
1 + Kc

(5.511)

As Kc , the equilibrium value of 1. In this limit, the reaction is irreversible. That


is, the species B is preferred over A. Equation (5.510) has exact solution
 
 
1
1 exp k 1 + Kc t
=
.
(5.512)
1 + K1c
For k > 0, Kc > 0, the equilibrium is stable. The time constant of relaxation is
=

k 1+

1
Kc

.

(5.513)

For the isothermal, isochoric system, one should consider the second law in terms of the
Helmholtz free energy. Combine then Eq. (4.394), dA|T,V 0, with Eq. (4.311), dA =
CC BY-NC-ND. 03 August 2012, J. M. Powers.

202

CHAPTER 5. THERMOCHEMISTRY OF A SINGLE REACTION

SdT P dV +

PN

i dni and taking time derivatives, one finds


!
N

X

dA|T,V = SdT P dV +
i dni
0,

i=1
T,V

N
X dni
dA
=
0,
i
dt T,V
dt
i=1

i=1

(5.514)

(5.515)

N
1 dA
V X di

=
0.
T dt
T i=1 i dt

(5.516)

This is exactly the same form as Eq. (5.482), which can be directly substituted into Eq. (5.516)
to give
!
!
!

N
N
N
Y
1 Y i
1 Y i
1 dA
i
i
1

ln

0,

= RV k(T )
T dt T,V
Kc i=1 i
Kc i=1 i
i=1

dA
dt T,V

= RV T k(T )

N
Y

i=1

1
Kc

N
Y
i=1

i i

ln

1
Kc

N
Y
i=1

i i

For the assumptions of this section, Eq. (5.518) reduces to


!
!

1
dA
1
1
= RT ko V (1 )
ln
0,
dt T,V
Kc 1
Kc 1
!
!

1
1
1
= kno RT (1 )
ln
0.
Kc 1
Kc 1

(5.517)

0.
(5.518)

(5.519)
(5.520)

Since the present analysis is nothing more than a special case of the previous section,
Eq. (5.520) certainly holds. One questions however the behavior in the irreversible limit,
1/Kc 0. Evaluating this limit, one finds


 

dA
ln (1 )
ln(1 )
+ . . . 0.
ln 1 +(1 )
(1 )
=
kn
lim
RT
o

| {z }
1/Kc 0 dt T,V
K
| {z c }
>0

(5.521)

Now, performing the distinguished limit as 1; that is the reaction goes to completion,
one notes that all terms are driven to zero for small 1/Kc . Recall that 1 goes to zero
goes to . Note that the entropy inequality is ill-defined for a formally
faster than ln(1 )
irreversible reaction with 1/Kc = 0.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

Chapter 6
Thermochemistry of multiple
reactions
See Turns, Chapter 4, 5, 6
See Kuo, Chapters 1, 2
See Kondepudi and Prigogine, Chapter 16, 19
This chapter will extend notions associated with the thermodynamics of a single chemical
reactions to systems in which many reactions occur simultaneously.

6.1

Summary of multiple reaction extensions

Consider now the reaction of N species, composed of L elements, in J reactions. This section
will focus on the most common case in which J (N L), which is usually the case in large
chemical kinetic systems in use in engineering models. While much of the analysis will only
require J > 0, certain results will depend on J (N L). It is not difficult to study the
complementary case where 0 < J < (N L).
The molecular mass of species i is still given by Eq. (5.1):
Mi =

L
X
l=1

Ml li,

i = 1, . . . , N.

(6.1)

However, each reaction has a stoichiometric coefficient. The j th reaction can be summarized
in the following ways:
N
X

i=1
N
X

i ij

N
X

i ij ,

j = 1, . . . , J,

(6.2)

i=1

i ij

= 0,

j = 1, . . . , J.

i=1

203

(6.3)

204

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

Stoichiometry for the j th reaction and lth element is given by the extension of Eq. (5.24):
N
X

l = 1, . . . , L, j = 1, . . . , J.

liij = 0,

(6.4)

i=1

The net change in Gibbs free energy and equilibrium constants of the j th reaction are defined
by the extensions of Eqs. (5.268, 5.267, 5.270):
Goj

KP,j exp
Kc,j

Po
RT

PNi=1 ij

N
X

j = 1, . . . , J,

(6.5)

j = 1, . . . , J,

(6.6)

j = 1, . . . , J.

(6.7)

i=1

exp

g oT,iij ,

Goj

RT

Goj
RT

The equilibrium of the j th reaction is given by the extension of Eq. (5.253):


N
X

i ij = 0,

j = 1, . . . , J,

(6.8)

i=1

or the extension of Eq. (5.255):


N
X

g i ij = 0,

j = 1, . . . , J.

(6.9)

i=1

The multi-reaction extension of Eq. (4.521) for affinity is


j =

N
X

i ij ,

j = 1, . . . , J.

(6.10)

i=1

In terms of the chemical affinity of each reaction, the equilibrium condition is simply the
extension of Eq. (5.254):
j = 0,
j = 1, . . . , J.
(6.11)
At equilibrium, then the equilibrium constraints can be shown to reduce to the extension
of Eq. (5.266):
KP,j =


N 
Y
Pi ij
i=1

Po

CC BY-NC-ND. 03 August 2012, J. M. Powers.

j = 1, . . . , J,

(6.12)

205

6.1. SUMMARY OF MULTIPLE REACTION EXTENSIONS


or the extension of Eq. (5.274):
Kc,j =

N
Y

ij

j = 1, . . . , J.

(6.13)

i=1

For isochoric reaction, the evolution of species concentration i due to the combined effect
of J reactions is given by the extension of Eq. (5.279):

di
=
dt

J
X
j=1

ij aj T
|
|

exp
{z

E j
RT

kj (T )

}|

1 Y kj

k
k ,
1
Kc,j k=1

} | k=1{z
|
}
{z
}
reverse reaction
forward reaction
{z
}
N
Y

kj

i = 1, . . . , N.

rj =(1/V )dj /dt

(6.14)
The extension to isobaric reactions is straightforward, and follows the same analysis as for
a single reaction. Again, three intermediate variables which are in common usage have been
defined. First one takes the reaction rate of the j th reaction to be the extension of Eq. (5.280)

!

 Y
N
N

1 Y kj
E j
kj

j
rj aj T exp
k
k , j = 1, . . . , J,
(6.15)
1
K

RT
c,j
k=1
|
{z
} | k=1{z
|
}
{z
}
kj (T )

reverse reaction

forward reaction

or the extension of Eq. (5.281)




 Y
N
N


1 Y kj
kj

rj = aj T exp
k
k ,

Kc,j k=1
k=1

|
{z
}
| {z }
|
{z
}
kj (T ), Arrhenius rate
| forward reaction {z reverse reaction }
j

E j
RT

j = 1, . . . , J,

(6.16)

law of mass action

rj =

1 dj
.
V dt

(6.17)

Here j is the reaction progress variable for the j th reaction.


Each reaction has a temperature-dependent rate function kj (T ), which is an extension of
Eq. (5.282):


E j
j
kj (T ) aj T exp
,
j = 1, . . . , J.
(6.18)
RT
CC BY-NC-ND. 03 August 2012, J. M. Powers.

206

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

The evolution rate of each species is given by i , defined now as an extension of Eq. (5.284):
i

J
X

ij rj ,

i = 1, . . . , N.

(6.19)

j=1

The multi-reaction extension of Eq. (5.239) for mole change in terms of progress variables
is
dni =

J
X

ij dj ,

i = 1, . . . , N.

(6.20)

j=1

One also has Eq. (5.243):


dG|T,P =
=

N
X

i=1
N
X

i dni ,
i

i=1

J
X

(6.21)
ik dk ,

k=1
J
X


N
X
G
k
=
,
ik
i

j p
j
i=1
k=1
=

N
X

i=1

N
X

J
X

ik kj ,

(6.22)
(6.23)
(6.24)

j=1

i ij ,

(6.25)

i=1

= j ,

j = 1, . . . , J.

(6.26)

For a set of adiabatic, isochoric reactions, one can show the extension of Eq. (5.429) is
PJ

j=1 rj Uj

dT
=
dt

cv

(6.27)

where the energy change for a reaction Uj is defined as the extension of Eq. (5.428):
Uj =

N
X

ui ij ,

j = 1, . . . , J.

(6.28)

i=1

Similarly for a set of adiabatic, isobaric reactions, one can show the extension of Eq. (5.444):
dT
=
dt

PJ

CC BY-NC-ND. 03 August 2012, J. M. Powers.

j=1 rj Hj

cP

(6.29)

6.1. SUMMARY OF MULTIPLE REACTION EXTENSIONS

207

where the enthalpy change for a reaction Hj is defined as the extension of Eq. (5.443):

Hj =

N
X

hi ij ,

j = 1, . . . , J.

(6.30)

i=1

Moreover, the density and species concentration derivatives for an adiabatic, isobaric set can
be shown to be extensions of Eqs. (5.455, 5.459):



J
N
X
X
hi
d
= M
rj
ij
1 ,
dt
c
T
P
j=1
i=1



J
X
di
Hj
=
nj
,
rj ij + yi
dt
cP T
j=1

(6.31)

(6.32)

where

nj =

N
X

kj .

(6.33)

k=1

In a similar fashion to that shown for a single reaction, one can further sum over all
reactions and prove that mixture mass is conserved, element mass and number are conserved.

Example 6.1
Show that element mass and number are conserved for the multi-reaction formulation.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

208

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS


Start with Eq. (6.14) and expand as follows:

li

J
X

di
dt

di
dt

= li

ij rj ,

(6.34)

j=1

J
X

ij rj ,

(6.35)

j=1

J
X
d
li ij rj ,
(li i ) =
dt
j=1

(6.36)

N X
J
N
X
X
d
li ij rj ,
(li i ) =
dt
i=1 j=1
i=1
!
J X
N
N
X
d X
li ij rj ,
=
li i
dt i=1
j=1 i=1
{z
}
|

(6.37)

(6.38)

=l e

dl e
dt

dl e
dt

J
X
j=1

= 0,

d
(Ml l e ) = 0,
dt
dl e
= 0,
dt

rj

N
X

li ij ,

(6.39)

{z

l = 1, . . . , L,

(6.40)

l = 1, . . . , L,

(6.41)

l = 1, . . . , L.

(6.42)

i=1

=0

It is also straightforward to show that the mixture density is conserved for the multi-reaction, multicomponent mixture:
d
= 0.
(6.43)
dt

The proof of the Clausius-Duhem relationship for the second law is an extension of the
single reaction result. Start with Eq. (5.460) and operate much as for a single reaction model.
dS|U,V

N
1X
dni

T i=1 i
|
{z
}

0,

(6.44)

irreversible entropy production


dS
dt U,V

N
V X dni 1
=
0,

T i=1 i dt V
N
V X di
0,
=

T i=1 i dt

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(6.45)
(6.46)

209

6.1. SUMMARY OF MULTIPLE REACTION EXTENSIONS


N
J
V X X
ij rj 0,
=

T i=1 i j=1

(6.47)

N
J
V XX
=
ij rj 0,
T i=1 j=1 i

(6.48)

J
N
V X X
rj
=
ij 0,
T j=1 i=1 i

(6.49)

J
N
V X Y ij
=
kj

T j=1 i=1 i

N
1 Y ij

1
Kc,j i=1 i

V
=
T

1
1
Kc,j

J
X

kj

j=1

= RV

J
X
j=1

N
Y

i ij

i=1

kj

N
Y
i=1

i ij

N
Y

i ij

i=1

!
!

N
1 Y ij

Kc,j i=1 i

N
X
i=1

i ij 0,

RT ln

ln

(6.50)

N
1 Y ij

Kc,j i=1 i

N
1 Y ij

Kc,j i=1 i

!!

0.

0,
(6.51)
(6.52)

Note that Eq. (6.49) can also be written in terms of the affinities (see Eq. (6.10)) and reaction
progress variables (see Eq. (6.17) as

J
dS
1 X dj
=
0.
(6.53)
j
dt U,V
T j=1
dt

Similar to the argument for a single reaction, if one defines extensions of Eqs. (5.483,
5.484) as

then it is easy to show that

N
Y

ij

Rj

= kj

Rj

N
kj Y ij
=
,
Kc,j i=1 i

(6.54)

i=1

rj = Rj Rj ,

and we get the equivalent of Eq. (5.488):



 
J
X

Rj
dS

RV
R

R
ln
=
0.
j
j
dt U,V
Rj
j=1

(6.55)

(6.56)

(6.57)

Since kj (T ) > 0, R > 0, and V 0, and each term in the summation combines to be positive
semi-definite, one sees that the Clausius-Duhem inequality is guaranteed to be satisfied for
multi-component reactions.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

210

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

6.2

Equilibrium conditions

For multicomponent mixtures undergoing multiple reactions, determining the equilibrium


condition is more difficult. There are two primary approaches, both of which are essentially
equivalent. The most straightforward method requires formal minimization of the Gibbs free
energy of the mixture. It can be shown that this actually finds the equilibrium associated
with all possible reactions.

6.2.1

Minimization of G via Lagrange multipliers

P
Recall
Recall also Eq. (4.397), G = N
i=1 g i ni . Since i = g i =
Eq. (4.395), dG|T,P 0.
P
PN

N
G
,
one
also
has
G
=

n
.
From
Eq.
(4.398),
dG|
=

i
T,P
i
i=1
i=1 i dni . Now
ni
P,T,nj

one must also demand for a system coming to equilibrium that the element numbers are
conserved. This can be achieved by requiring

N
X
i=1

li (nio ni ) = 0,

l = 1, . . . , L.

(6.58)

Here recall nio is the initial number of moles of species i in the mixture, and li is the number
of moles of element l in species i. If one interprets nio ni as ij , the negative of the net
P
mole change, Eq. (6.58) becomes N
i=1 li ij = 0, equivalent to Eq. (6.4).

One can now use the method of constrained optimization given by the method of Lagrange
multipliers to extremize G subject to the constraints of element conservation. The extremum
will be a minimum; this will not be proved, but it will be demonstrated. Define a set of L
Lagrange multipliers l . Next define an augmented Gibbs free energy function G , which is
simply G plus the product of the Lagrange multipliers and the constraints:

G =G+

L
X
l=1

N
X
k=1

lk (nko nk ).

(6.59)

Now when the constraints are satisfied, one has G = G, so assuming the constraints can be
satisfied, extremizing G is equivalent to extremizing G . To extremize G , take its differential
CC BY-NC-ND. 03 August 2012, J. M. Powers.

211

6.2. EQUILIBRIUM CONDITIONS

with respect to ni , with P , T and nj constant and set it to zero for each species:
!



L
N
X
X
G
G

=
l
lk (nko nk ) = 0, i = 1, . . . , N,
+
ni T,P,nj
ni T,P,nj ni T,P,nj l=1 k=1
| {z }
=i

= i

= i
= i

L
X
l=1

L
X
l=1
L
X


nk
lk
l
= 0,
ni T,P,nj
k=1
| {z }
N
X

N
X

(6.60)

i = 1, . . . , N,

(6.61)

ki

lk ki = 0,

i = 1, . . . , N,

(6.62)

k=1

l li = 0,

i = 1, . . . , N.

(6.63)

l=1

Next, for an ideal gas, one can expand the chemical potential so as to get
oT,i
|

+ RT ln
{z

Pi
Po

=i

L
X

l li = 0,

i = 1, . . . , N,

(6.64)

i = 1, . . . , N.

(6.65)

l=1

L
X

n
P
1
i
o

T,i + RT ln PN
l li = 0,

Po

l=1
k=1 nk
|
{z
}
=Pi

Recalling that

PN

k=1 nk

oT,i

= n, in summary then, one has N + L equations

+ RT ln

ni P
n Po

N
X
i=1

L
X

l li = 0,

i = 1, . . . , N,

(6.66)

li (nio ni ) = 0,

l = 1, . . . , L.

(6.67)

l=1

in N + L unknowns: ni , i = 1, . . . , N, l , l = 1, . . . , L.
Example 6.2
Consider a previous example problem, see p. 145, in which
N2 + N2 2N + N2 .

(6.68)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

212

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS


Take the reaction to be isothermal and isobaric with T = 6000 K and P = 100 kP a. Initially one has
1 kmole of N2 and 0 kmole of N . Use the extremization of Gibbs free energy to find the equilibrium
composition.
First find the chemical potentials at the reference pressure of each of the possible constituents.
o

oT,i = goi = hi T soi = h298,i + hi T soi .

(6.69)

For each species, one then finds


oN2
oN

kJ
,
kmole
kJ
.
= 472680 + 124590 (6000)(216.926) = 704286
kmole
= 0 + 205848 (6000)(292.984) = 1552056

To each of these one must add


RT ln

ni P
nPo

(6.70)
(6.71)

to get the full chemical potential. Now P = Po = 100 kP a for this problem, so one only must consider
RT = 8.314(6000) = 49884 kJ/kmole. So, the chemical potentials are


nN 2
,
(6.72)
N2 = 1552056 + 49884 ln
nN + nN 2


nN
N = 704286 + 49884 ln
.
(6.73)
nN + nN 2
Then one adds on the Lagrange multiplier and then considers element conservation to get the
following coupled set of nonlinear algebraic equations:


nN 2
2N = 0,
(6.74)
1552056 + 49884 ln
nN + nN 2


nN
704286 + 49884 ln
N = 0,
(6.75)
nN + nN 2
(6.76)
nN + 2nN2 = 2.
These non-linear equations are solved numerically to get
nN 2

0.88214 kmole,

(6.77)

nN

(6.78)

0.2357 kmole,
kJ
.
781934
kmole

(6.79)

These agree with results found in the earlier example problem, see p. 145.

Example 6.3
Consider a mixture of 2 kmole of H2 and 1 kmole of O2 at T = 3000 K and P = 100 kP a.
Assuming an isobaric and isothermal equilibration process with the products consisting of H2 , O2 ,
H2 O, OH, H, and O, find the equilibrium concentrations. Consider the same mixture at T = 298 K
and T = 1000 K.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

213

6.2. EQUILIBRIUM CONDITIONS

The first task is to find the chemical potentials of each species at the reference pressure and
T = 3000 K. Here one can use the standard tables along with the general equation
o

oT,i = goi = hi T soi = h298,i + hi T soi .

(6.80)

For each species, one then finds

oH2
oO2
oH2 O
oOH
oH
oO

kJ
,
kmole
kJ
,
= 0 + 98013 3000(284.466) = 755385
kmole
= 0 + 88724 3000(202.989) = 520242

(6.81)
(6.82)

kJ
,
kmole
kJ
= 38987 + 89585 3000(256.825) = 641903
,
kmole
kJ
= 217999 + 56161 3000(162.707) = 213961
,
kmole
kJ
.
= 249170 + 56574 3000(209.705) = 323371
kmole
= 241826 + 126548 3000(286.504) = 974790

(6.83)
(6.84)
(6.85)
(6.86)

To each of these one must add

RT ln

ni P
nPo

to get the full chemical potential. Now P = Po = 100 kP a for this problem, so one must only consider
RT = 8.314(3000) = 24942 kJ/kmole. So, the chemical potentials are

H2
O2
H2 O
OH
H
O

=
=
=
=
=
=

520243 + 24942 ln
755385 + 24942 ln
974790 + 24942 ln
641903 + 24942 ln
213961 + 24942 ln
323371 + 24942 ln

nH2 + nO2
nH2 + nO2
nH2 + nO2
nH2 + nO2
nH2 + nO2
nH2 + nO2

nH2
+ nH2 O + nOH
nO2
+ nH2 O + nOH
nH2 O
+ nH2 O + nOH
nOH
+ nH2 O + nOH
nH
+ nH2 O + nOH
nO
+ nH2 O + nOH

+ nH + nO
+ nH + nO
+ nH + nO
+ nH + nO
+ nH + nO
+ nH + nO








(6.87)

(6.88)

(6.89)

(6.90)

(6.91)

(6.92)

Then one adds on the Lagrange multipliers and then considers element conservation to get the following
CC BY-NC-ND. 03 August 2012, J. M. Powers.

214

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS


coupled set of nonlinear equations:


nH2
2H
nH2 + nO2 + nH2 O + nOH + nH + nO


nO2
755385 + 24942 ln
2O
nH2 + nO2 + nH2 O + nOH + nH + nO


nH2 O
974790 + 24942 ln
2H O
nH2 + nO2 + nH2 O + nOH + nH + nO


nOH
H O
641903 + 24942 ln
nH2 + nO2 + nH2 O + nOH + nH + nO


nH
213961 + 24942 ln
H
nH2 + nO2 + nH2 O + nOH + nH + nO


nO
O
323371 + 24942 ln
nH2 + nO2 + nH2 O + nOH + nH + nO
2nH2 + 2nH2 O + nOH + nH
2nO2 + nH2 O + nOH + nO

520243 + 24942 ln

= 0,

(6.93)

= 0,

(6.94)

= 0,

(6.95)

= 0,

(6.96)

= 0,

(6.97)

= 0,

(6.98)

= 4,

(6.99)

= 2.

(6.100)

These non-linear algebraic equations can be solved numerically via a Newton-Raphson technique.
The equations are sensitive to the initial guess, and one can use ones intuition to help guide the
selection. For example, one might expect to have nH2 O somewhere near 2 kmole. Application of the
Newton-Raphson iteration yields

nH2
nO2

=
=

nH2 O
nOH

=
=

nH

nO

3.19 101 kmole,


1.10 101 kmole,

(6.101)
(6.102)

1.50 100 kmole,


2.20 101 kmole,

(6.103)
(6.104)

5.74 102 kmole,


kJ
2.85 105
,
kmole
kJ
.
4.16 105
kmole

(6.106)

1.36 101 kmole,

(6.105)

(6.107)
(6.108)

At this relatively high value of temperature, all species considered have a relatively major presence.
That is, there are no truly minor species.
Unless a very good guess is provided, it may be difficult to find a solution for this set of nonlinear equations. Straightforward algebra allows the equations to be recast in a form which sometimes
CC BY-NC-ND. 03 August 2012, J. M. Powers.

215

6.2. EQUILIBRIUM CONDITIONS


converges more rapidly:

nH2 + nO2

nH2
+ nH2 O + nOH + nH + nO

nH2 + nO2

nO2
+ nH2 O + nOH + nH + nO

nH2 O
nH2 + nO2 + nH2 O + nOH + nH + nO
nOH
nH2 + nO2 + nH2 O + nOH + nH + nO
nH
nH2 + nO2 + nH2 O + nOH + nH + nO
nO
nH2 + nO2 + nH2 O + nOH + nH + nO
2nH2 + 2nH2 O + nOH + nH
2nO2 + nH2 O + nOH + nO

=
=
=
=
=
=



2
H
exp
,
exp
24942



2
755385
O
,
exp
exp
24942
24942




2

O
H
974790
exp
exp
,
exp
24942
24942
24942






O
H
641903
exp
exp
,
exp
24942
24942
24942




H
213961
exp
,
exp
24942
24942




O
323371
exp
,
exp
24942
24942
4,
2.


520243
24942

(6.109)
(6.110)
(6.111)
(6.112)
(6.113)
(6.114)
(6.115)
(6.116)

Then solve these considering ni , exp (O /24942), and exp (H /24942) as unknowns. The same result
is recovered, but a broader range of initial guesses converge to the correct solution.
One can verify that this choice extremizes G by direct computation; moreover, this will show
that the extremum is actually a minimum. In so doing, one must exercise care to see that element
conservation is retained. As an example, perturb the equilibrium solution above for nH2 and nH such
that
nH2
nH

=
=

3.19 101 + ,
1.36 101 2.

(6.117)
(6.118)

Leave all other species mole numbers the same. In this way, when = 0, one has the original equilibrium
solution. For 6= 0, the solution
PN moves off the equilibrium value in such a way that elements are
conserved. Then one has G = i=1 i ni = G().
The difference G() G(0) is plotted in Fig. 6.1. When = 0, there is no deviation from the value
predicted by the Newton-Raphson iteration. Clearly when = 0, G() G(0), takes on a minimum
value, and so then does G(). So the procedure works.
At the lower temperature, T = 298 K, application of the same procedure yields very different
results:
nH2
nO2
nH2 O
nOH
nH
nO
H
O

= 4.88 1027 kmole,


27

= 2.44 10
kmole,
0
= 2.00 10 kmole,

= 2.22 1029 kmole,


49

= 2.29 10
kmole,
54
= 1.67 10
kmole,
kJ
,
= 9.54 104
kmole
kJ
= 1.07 105
.
kmole

(6.119)
(6.120)
(6.121)
(6.122)
(6.123)
(6.124)
(6.125)
(6.126)

At the intermediate temperature, T = 1000 K, application of the same procedure shows the minor
CC BY-NC-ND. 03 August 2012, J. M. Powers.

216

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS


G() - G(0) (kJ)
40

30

20

10

-0.01

-0.005

0.005

0.01
(kmole)

Figure 6.1: Gibbs free energy variation as mixture composition is varied maintaining element
conservation for mixture of H2 , O2 , H2 O, OH, H, and O at T = 3000 K, P = 100 kP a.
species become slightly more prominent:
nH2
nO2
nH2 O
nOH
nH
nO
H
O

6.2.2

= 4.99 107 kmole,


7

= 2.44 10 kmole,
= 2.00 100 kmole,

= 2.09 108 kmole,


12

= 2.26 10
kmole,
13
= 1.10 10
kmole,
kJ
,
= 1.36 105
kmole
kJ
.
= 1.77 105
kmole

(6.127)
(6.128)
(6.129)
(6.130)
(6.131)
(6.132)
(6.133)
(6.134)

Equilibration of all reactions

In another equivalent method, if one commences with a multi-reaction model, one can require
each reaction to be in equilibrium. This leads to a set of algebraic equations for rj = 0, which
CC BY-NC-ND. 03 August 2012, J. M. Powers.

217

6.2. EQUILIBRIUM CONDITIONS


from Eq. (6.16) leads to


Kc,j =

Po
RT

PNi=1 ij

exp

Goj
RT

N
Y

kj

j = 1, . . . , J.

(6.135)

k=1

With some effort it can be shown that not all of the J equations are linearly independent.
Moreover, they do not possess a unique solution. However, for closed systems, only one of
the solutions is physical, as will be shown in the following section. The others typically
involve non-physical, negative concentrations.
Nevertheless, Eqs. (6.135) are entirely consistent with the predictions of the N + L equations which arise from extremization of Gibbs free energy while enforcing element number
constraints. This can be shown by beginning with Eq. (6.65), rewritten in terms of molar
concentrations, and performing the following sequence of operations:
oT,i

+ RT ln

P
PN
Po
k=1 nk /V
ni /V

!
!

L
X

l li = 0,

l=1

L
X

l li
P

o
k
k=1
l=1

 X
L
i P
oT,i + RT ln
l li

Po
l=1

 X
L
RT
o
T,i + RT ln i

l li
Po
l=1


L
X
RT
o
ij T,i + ij RT ln i
l li
ij
Po
l=1
oT,i + RT ln

N
X

ij oT,i +

|i=1 {z

=Goj

Goj

N
X
i=1

+ RT

i
PN

i = 1, . . . , N, (6.136)

= 0,

i = 1, . . . , N, (6.137)

= 0,

i = 1, . . . , N, (6.138)

= 0,

i = 1, . . . , N, (6.139)

= 0,

i = 1, . . . , N,

j = 1, . . . , J,
(6.140)
 X

N
L
X
RT
ij RT ln i

ij
l li = 0,
j = 1, . . . , J, (6.141)
Po
i=1
l=1

N
X
i=1

 X

L
N
X
RT

l
li ij = 0,
ij ln i
Po
l=1
|i=1 {z }

j = 1, . . . , J, (6.142)

=0

Goj + RT

N
X
i=1



RT
= 0,
ij ln i
Po

j = 1, . . . , J. (6.143)

Here, the stoichiometry for each reaction has been employed to remove the Lagrange multiCC BY-NC-ND. 03 August 2012, J. M. Powers.

218

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

pliers. Continue to find


ij
RT
ln i
Po
i=1
ij !

N
X
RT
exp
ln i
Po
i=1
ij
N 
Y
RT
i
Po
i=1
PN

 i=1 ij Y
N
RT
i ij
Po
i=1
N
X

N
Y

=
=
=

i ij =

i=1

Goj

,
j = 1, . . . , J,
(6.144)
RT


Goj
exp
,
j = 1, . . . , J,
(6.145)
RT


Goj
,
j = 1, . . . , J,
(6.146)
exp
RT


Goj
exp
,
j = 1, . . . , J,
(6.147)
RT

PNi=1 ij


Goj
Po
,
j = 1, . . . , J,(6.148)
exp
RT
RT
{z
}
|
=Kc,j

In summary,
N
Y

i ij = Kc,j ,

j = 1, . . . , J,

(6.149)

i=1

which is identical to Eq. (6.135), obtained by equilibrating each of the J reactions. Thus,
extremization of Gibbs free energy is consistent with equilibrating each of the J reactions.

6.3

Concise reaction rate law formulations

One can additional analysis to obtain a more efficient representation of the reaction rate law
for multiple reactions. There are two important cases: 1) J (N L); this is most common
for large chemical kinetic systems, and 2) J < (N L); this is common for simple chemistry
models.
The species production rate is given by Eq. (6.14), which reduces to
J
di
1 X dj
ij
=
,
dt
V j=1
dt

i = 1, . . . , N.

(6.150)

Now recalling Eq. (5.99), one has


dni =

N
L
X
k=1

Dik dk ,

CC BY-NC-ND. 03 August 2012, J. M. Powers.

i = 1, . . . , N.

(6.151)

6.3. CONCISE REACTION RATE LAW FORMULATIONS

219

Comparing then Eq. (6.151) to Eq. (6.20), one sees that


J
X

ij dj =

j=1

1
V

6.3.1

J
X

k=1

ij dj =

j=1

N
L
X

1
V

Dik dk ,

N
L
X
k=1

Dik dk ,

i = 1, . . . , N,

(6.152)

i = 1, . . . , N.

(6.153)

Reaction dominant: J (N L)

Consider first the most common case in which J (N L). One can say the species
production rate is given
N L
J
dk X
1 X
di
Dik
=
=
ij rj ,
dt
V k=1
dt
j=1

i = 1, . . . , N.

(6.154)

One would like to invert


for dk /dt. However, Dik is non-square and has
P and solve directly P
N

D
=
0,
and
no inverse. But since N
i=1 li ij = 0, L of the equations N equations
i=1 li ip
in Eq. (6.154) are redundant.
At this point, it is more convenient to go to a Gibbs vector notation, where there is an
obvious correspondence between the bold vectors and the indicial counterparts:
d
1
d
= D
= r,
dt
V
dt
d
= V DT r,
DT D
dt
d
= V (DT D)1 DT r.
dt

(6.155)
(6.156)
(6.157)

Because of the L linear dependencies, there is no loss of information in this matrix projection.
This system of N L equations is the smallest number of differential equations that can be
solved for a general system in which J > (N L).
Lastly, one recovers the original system when forming
D

d
= V D (DT D)1 DT r.
{z
}
|
dt

(6.158)

=P

Here, the N N projection matrix P is symmetric, has norm of unity, has rank of N L,
has N L eigenvalues of value unity, and L eigenvalues of value zero. And, while application
of a general projection matrix to r filters some of the information in r, because the
N (N L) matrix D spans the same column space as the N J matrix , no information
is lost in Eq. (6.158) relative to the original Eq. (6.155). Mathematically, one can say
P = .

(6.159)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

220

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS

6.3.2

Species dominant: J < (N L)

Next consider the case in which J < (N L). This often arises in models of simple chemistry,
for example one- or two-step kinetics.
The fundamental reaction dynamics are most concisely governed by the J equations
which form
1 d
= r.
V dt

(6.160)

However, r is a function of the concentrations; one must therefore recover as a function


of reaction progress . In vector form, Eq. (6.150) is written as
1
d
d
= .
dt
V
dt

(6.161)

Take as an initial condition that the reaction progress is zero at t = 0 and that there are an
appropriate set of initial conditions on the species concentrations :
= 0,
= o ,

t = 0,
t = 0.

(6.162)
(6.163)

Then, since is a constant, Eq. (6.161) is easily integrated. After applying the initial
conditions, Eq. (6.163), one gets
= o +

1
.
V

(6.164)

Last, if J = (N L), either approach yields the same number of equations, and is equally
concise.

6.4

Adiabatic, isochoric kinetics

Here an example which uses multiple reactions for an adiabatic isochoric system is given.
Example 6.4
Consider the full time-dependency of a problem considered in a previous example in which the
equilibrium state was found; see Sec. 5.3.3.4,. A closed, fixed, adiabatic volume contains at t = 0 s a
stoichiometric mixture of 2 kmole of H2 , 1 kmole of O2 , and 8 kmole of N2 at 100 kP a and 1000 K.
Find the reaction dynamics as the system proceeds from its initial state to its final state.
This problem requires a detailed numerical solution. Such a solution was performed by solving
Eq. (6.14) along with the associated calorically imperfect species state equations for a mixture of
eighteen interacting species: H2 , H, O, O2 , OH, H2 O, HO2 , H2 O2 , N , N H2 , N H3 , N2 H, N O, N O2 ,
N2 O, HN O, and N2 . The equilibrium values were reported in a previous example.
The dynamics of the reaction process are reflected in Fig. 6.2. At early time, t < 4 104 s, the
pressure, temperature, and major reactant species concentrations (H2 , O2 , N2 ) are nearly constant.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

221

6.4. ADIABATIC, ISOCHORIC KINETICS

O2

(kmole/m 3 )
i

HydrogenOxygen-Related
Species
Evolution

10

10

10

10

HO

N2
NO

O2
OH

10

H
O

15

20

Nitrogen
and Nitrogen
Oxides Species
Evolution
b

NO

30

10

H2O2

20

10

N2O

10

H2O2

10

40

10

10

10

t (s)

10

10

3000

10

t (s)

10

250

P (kPa)

2500
T (K)

NO2

10

HO2
HO

H2

(kmole/m 3 )
i

H2

2000

150

induction
time

1500

1000
6

10

200

10

t (s)

10

d
100
6

10

10

t (s)

10

Figure 6.2: Plot of a) H2 (t), H (t), O (t), O2 (t), OH (t), H2 O (t), HO2 (t), H2 O2 (t), b) N (t),
N O (t), N O2 (t), N2 O (t), N2 (t), c) T (t), and d) P (t) for adiabatic, isochoric combustion of
a mixture of 2H2 + O2 + 8N2 initially at T = 1000 K, P = 100 kP a.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

222

CHAPTER 6. THERMOCHEMISTRY OF MULTIPLE REACTIONS


However, the minor species, e.g. OH, N O, HO2 and the major product, H2 O, are undergoing very
rapid growth, albeit concentrations whose value remains small. In this period, the material is in what
is known as the induction period.
After a certain critical mass of minor species has accumulated, exothermic recombination of these
minor species to form the major product H2 O induces the temperature to rise, which accelerates further
the reaction rates. This is manifested in a thermal explosion. A common definition of the end of the
induction period is the induction time, t = tind , the time when dT /dt goes through a maximum. Here
one finds
tind = 4.53 104 s.
(6.165)
At the end of the induction zone, there is a final relaxation to equilibrium.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

Chapter 7
Kinetics in some more detail
Here we give further details of kinetics. These notes are also used to introduce a separate
combustion course and have some overlap with previous chapters.
Let us consider the reaction of N molecular chemical species composed of L elements via
J chemical reactions. Let us assume the gas is an ideal mixture of ideal gases that satisfies
Daltons law of partial pressures. The reaction will be considered to be driven by molecular
collisions. We will not model individual collisions, but instead attempt to capture their
collective effect.
An example of a model of such a reaction is listed in Table 7.1. There we find a N = 9
species, J = 37 step irreversible reaction mechanism for an L = 3 hydrogen-oxygen-argon
mixture from Maas and Warnatz,1 with corrected fH2 from Maas and Pope.2 The model has
also been utilized by Fedkiw, et al.3 We need not worry yet about fH2 , which is known as
a collision efficiency factor. The one-sided arrows indicate that each individual reaction is
considered to be irreversible. Note that for nearly each reaction, a separate reverse reaction is
listed; thus, pairs of irreversible reactions can in some sense be considered to model reversible
reactions.
In this model a set of elementary reactions are hypothesized. For the j th reaction we have
the collision frequency factor aj , the temperature-dependency exponent j and the activation
energy E j . These will be explained in short order. Other common forms exist. Often
reactions systems are described as being composed of reversible reactions. Such reactions are
usually notated by two sided arrows. One such system is reported by Powers and Paolucci4
reported here in Table 7.2. Both overall models are complicated.
1

Maas, U., and Warnatz, J., 1988, Ignition Processes in Hydrogen-Oxygen Mixtures, Combustion and
Flame, 74(1): 53-69.
2
Maas, U., and Pope, S. B., 1992, Simplifying Chemical Kinetics: Intrinsic Low-Dimensional Manifolds
in Composition Space, Combustion and Flame, 88(3-4): 239-264.
3
Fedkiw, R. P., Merriman, B., and Osher, S., 1997, High Accuracy Numerical Methods for Thermally
Perfect Gas Flows with Chemistry, Journal of Computational Physics, 132(2): 175-190.
4
Powers, J. M., and Paolucci, S., 2005, Accurate Spatial Resolution Estimates for Reactive Supersonic
Flow with Detailed Chemistry, AIAA Journal, 43(5): 1088-1099.

223

224

j
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37

CHAPTER 7. KINETICS IN SOME MORE DETAIL

Reaction

aj

O2 + H OH + O
OH + O O2 + H
H2 + O OH + H
OH + H H2 + O
H2 + OH H2 O + H
H2 O + H H2 + OH
OH + OH H2 O + O
H2 O + O OH + OH
H + H + M H2 + M
H2 + M H + H + M
H + OH + M H2 O + M
H2 O + M H + OH + M
O + O + M O2 + M
O2 + M O + O + M
H + O2 + M HO2 + M
HO2 + M H + O2 + M
HO2 + H OH + OH
OH + OH HO2 + H
HO2 + H H2 + O2
H2 + O2 HO2 + H
HO2 + H H2 O + O
H2 O + O HO2 + H
HO2 + O OH + O2
OH + O2 HO2 + O
HO2 + OH H2 O + O2
H2 O + O2 HO2 + OH
HO2 + HO2 H2 O2 + O2
OH + OH + M H2 O2 + M
H2 O2 + M OH + OH + M
H2 O2 + H H2 + HO2
H2 + HO2 H2 O2 + H
H2 O2 + H H2 O + OH
H2 O + OH H2 O2 + H
H2 O2 + O OH + HO2
OH + HO2 H2 O2 + O
H2 O2 + OH H2 O + HO2
H2 O + HO2 H2 O2 + OH

M,j
(mol/cm3 )(
s K j
1

2.00 1014
1.46 1013
5.06 104
2.24 104
1.00 108
4.45 108
1.50 109
1.51 1010
1.80 1018
6.99 1018
2.20 1022
3.80 1023
2.90 1017
6.81 1018
2.30 1018
3.26 1018
1.50 1014
1.33 1013
2.50 1013
6.84 1013
3.00 1013
2.67 1013
1.80 1013
2.18 1013
6.00 1013
7.31 1014
2.50 1011
3.25 1022
2.10 1024
1.70 1012
1.15 1012
1.00 1013
2.67 1012
2.80 1013
8.40 1012
5.40 1012
1.63 1013

PN

i=1 ij

j
0.00
0.00
2.67
2.67
1.60
1.60
1.14
1.14
1.00
1.00
2.00
2.00
1.00
1.00
0.80
0.80
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
2.00
2.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00

Ej

kJ
mole

70.30
2.08
26.30
18.40
13.80
77.13
0.42
71.64
0.00
436.08
0.00
499.41
0.00
496.41
0.00
195.88
4.20
168.30
2.90
243.10
7.20
242.52
1.70
230.61
0.00
303.53
5.20
0.00
206.80
15.70
80.88
15.00
307.51
26.80
84.09
4.20
132.71

Table 7.1: Third body collision efficiencies with M are fH2 = 1.00, fO2 = 0.35, and fH2 O =
6.5.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

225

j
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19

M,j
(mol/cm3 )(
s K j
1

Reaction

aj

H2 + O2 OH + OH
OH + H2 H2 O + H
H + O2 OH + O
O + H2 OH + H
H + O2 + M HO2 + M
H + O2 + O2 HO2 + O2
H + O2 + N2 HO2 + N2
OH + HO2 H2 O + O2
H + HO2 OH + OH
O + HO2 O2 + OH
OH + OH O + H2 O
H2 + M H + H + M
O2 + M O + O + M
H + OH + M H2 O + M
H + HO2 H2 + O2
HO2 + HO2 H2 O2 + O2
H2 O2 + M OH + OH + M
H2 O2 + H HO2 + H2
H2 O2 + OH H2 O + HO2

1.70 1013
1.17 109
5.13 1016
1.80 1010
2.10 1018
6.70 1019
6.70 1019
5.00 1013
2.50 1014
4.80 1013
6.00 108
2.23 1012
1.85 1011
7.50 1023
2.50 1013
2.00 1012
1.30 1017
1.60 1012
1.00 1013

PN

i=1 ij

j
0.00
1.30
0.82
1.00
1.00
1.42
1.42
0.00
0.00
0.00
1.30
0.50
0.50
2.60
0.00
0.00
0.00
0.00
0.00

Ej

cal
mole

47780
3626
16507
8826
0
0
0
1000
1900
1000
0
92600
95560
0
700
0
45500
3800
1800

Table 7.2: Nine species, nineteen step reversible reaction mechanism for an H2 /O2 /N2 mixture. Third body collision efficiencies with M are f5 (H2 O) = 21, f5 (H2 ) = 3.3, f12 (H2 O) = 6,
f12 (H) = 2, f12 (H2 ) = 3, f14 (H2 O) = 20.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

226

CHAPTER 7. KINETICS IN SOME MORE DETAIL

7.1

Isothermal, isochoric kinetics

For simplicity, we will first focus attention on cases in which the temperature T and volume V are both constant. Such assumptions are known as isothermal and isochoric,
respectively. A nice fundamental treatment of elementary reactions of this type is given by
Vincenti and Kruger in their detailed monograph.5

7.1.1

O O2 dissociation

One of the simplest physical examples is provided by the dissociation of O2 into its atomic
component O.
7.1.1.1

Pair of irreversible reactions

To get started, let us focus for now only on reactions 13 and 14 from Table 7.1 in the limiting
case in which temperature T and volume V are constant.
7.1.1.1.1 Mathematical model The reactions describe oxygen dissociation and recombination in a pair of irreversible reactions:
13 : O + O + M O2 + M,
14 : O2 + M O + O + M,

(7.1)
(7.2)

with

a14

2
mole
K
kJ
a13 = 2.90 10
, 13 = 1.00, E 13 = 0
3
cm
s
mole
1

K
kJ
mole
, 14 = 1.00, E 14 = 496.41
.
= 6.81 1018
3
cm
s
mole
17

(7.3)
(7.4)

The irreversibility is indicated by the one-sided arrow. Though they participate in the overall
hydrogen oxidation problem, these two reactions are in fact self-contained as well. So let us
just consider that we have only oxygen in our box with N = 2 species, O2 and O, J = 2
reactions (those being 13 and 14), and L = 1 element, that being O.
Recall that in the cgs system, common in thermochemistry, that 1 erg = 1 dyne cm =
107 J = 1010 kJ. Recall also that the cgs unit of force is the dyne and that 1 dyne =
1 g cm/s2 = 105 N. So for cgs we have
 10

erg
10 erg
erg
kJ
= 4.96 1012
,
.
(7.5)
E 13 = 0
E 14 = 496.41
mole
mole
kJ
mole
5

W. G. Vincenti and C. H. Kruger, 1965, Introduction to Physical Gas Dynamics, Wiley, New York.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

227

7.1. ISOTHERMAL, ISOCHORIC KINETICS

The standard model for chemical reaction induces the following two ordinary differential
equations for the evolution of O and O2 molar concentrations:




E 14
dO
E 13
14
13
= 2 a13 T exp
O O M +2 a14 T exp
O2 M ,
dt
RT
RT
|
{z
}
|
{z
}
dO2
dt

= a13 T
|
|

=k13 (T )

13

exp
{z

{z

=r13

E 13
RT

=k13 (T )

{z

=r13

O O M a14 T
}
|
}

14

=k14 (T )

exp
{z

{z

=r14

E 14
RT

=k14 (T )

{z

=r14

O2 M .

(7.6)

(7.7)

Here we use the notation i as the molar concentration of species i. Also a common usage for
molar concentration is given by square brackets, e.g. O2 = [O2 ]. The symbol M represents
an arbitrary third body and is an inert participant in the reaction. We also use the common
notation of a temperature-dependent portion of the reaction rate for reaction j, kj (T ), where


Ej
j
.
(7.8)
kj (T ) = aj T exp
RT
The reaction rates for reactions 13 and 14 are defined as
r13 = k13 O O M ,
r14 = k14 O2 M .

(7.9)
(7.10)

We will give details of how to generalize this form later. The system Eq. (7.6-7.7) can be
written simply as
dO
= 2r13 + 2r14 ,
dt
dO2
= r13 r14 .
dt
Even more simply, in vector form, Eqs. (7.11-7.12) can be written as
d
= r.
dt
Here we have taken

(7.12)

(7.13)


O
,
=
O2


2 2
=
,
1 1
 
r13
r =
.
r14


(7.11)

(7.14)
(7.15)
(7.16)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

228

CHAPTER 7. KINETICS IN SOME MORE DETAIL

In general, we will have be a column vector of dimension N 1, will be a rectangular


matrix of dimension N J of rank R, and r will be a column vector of length J 1. So
Eqs. (7.11-7.12) take the form
  
 
d O
r13
2 2
.
(7.17)
=
r14
1 1
dt O2
Note here that the rank R of is R = L = 1. Let us also define a stoichiometric matrix of
dimension L N. The component of , li represents the number of element l in species i.
Generally will be full rank, which will vary since we can have L < N, L = N, or L > N.
Here we have L < N and is of dimension 1 2:

= 1 2 .
(7.18)
Element conservation is guaranteed by insisting that be constructed such that
= 0.

(7.19)

So we can say that each of the column vectors of lies in the right null space of .
For our example, we see that Eq. (7.19) holds:


 2 2

= 1 2
= 0 0 .
(7.20)
1 1
The symbol R is the universal gas constant, where
 7

10 erg
erg
J
= 8.31441 107
.
R = 8.31441
mole K
J
mole K

(7.21)

Let us take as initial conditions


O (t = 0) = b
O ,

O2 (t = 0) = b
O2 .

(7.22)

Now M represents an arbitrary third body, so here

M = O2 + O .
Thus, the ordinary differential equations of the reaction dynamics reduce to



dO
E 13
13
O O O2 + O
= 2a13 T exp
dt
RT



E 14
14
+2a14 T exp
O2 O2 + O ,
RT



dO2
E 13
13
= a13 T exp
O O O2 + O
dt
RT



E 14
14
O2 O2 + O .
a14 T exp
RT
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.23)

(7.24)

(7.25)

229

7.1. ISOTHERMAL, ISOCHORIC KINETICS

Equations (7.24-7.25) with Eqs. (7.22) represent two non-linear ordinary differential equations with initial conditions in two unknowns O and O2 . We seek the behavior of these two
species concentrations as a function of time.
Systems of non-linear equations are generally difficult to integrate analytically and generally require numerical solution. Before embarking on a numerical solution, we simplify as
much as we can. Note that
d
dO
+ 2 O2 = 0,
(7.26)
dt
dt

d
(7.27)
O + 2O2 = 0.
dt
We can integrate and apply the initial conditions (7.22) to get
O + 2O2 = b
O + 2b
O2 = constant.

(7.28)

The fact that this algebraic constraint exists for all time is a consequence of the conservation
of mass of each O element. It can also be thought of as the conservation of number of
O atoms. Such notions always hold for chemical reactions. They do not hold for nuclear
reactions.
Standard linear algebra provides a robust way to find the constraint of Eq. (7.28). We
can use elementary row operations to cast Eq. (7.16) into a row-echelon form. Here our goal
is to get a linear combination which on the right side has an upper triangular form. To
achieve this add twice the second equation with the first to form a new equation to replace
the second equation. This gives

 
 
d
O
r13
2 2
.
(7.29)
=
r14
0 0
dt O + 2O2
Obviously the second equation is one we obtained earlier, d/dt(O + 2O2 ) = 0, and this
induces our algebraic constraint. We also note the system can be recast as
 
   

r13
2 2
1 0 d O
.
(7.30)
=
r14
0 0
1 2 dt O2
This is of the matrix form
d
= U r.
(7.31)
dt
Here L and L1 are N N lower triangular matrices of full rank N, and thus invertible.
The matrix U is upper triangular of dimension N J and with the same rank as , R L.
The matrix P is a permutation matrix of dimension N N. It is never singular and thus
always invertable. It is used to effect possible row exchanges to achieve the desired form;
often row exchanges are not necessary, in which case P = I, the N N identity matrix.
Equation (7.31) can be manipulated to form the original equation via
L1 P

d
1
=P
L U} r.
|
{z
dt
=

(7.32)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

230

CHAPTER 7. KINETICS IN SOME MORE DETAIL

What we have done is the standard linear algebra decomposition of = P1 L U.


We can also decompose the algebraic constraint, Eq. (7.28), in a non-obvious way that
is more readily useful for larger systems. We can write

1
(7.33)
O2 = b
O2
O b
O .
2

Defining now O = O b
O , we can say
     

b
O

1
= bO +
O .
1
O2
| {z }
O2
| {z
} | {z
} | {z2 }
=
=D
b
=
=

(7.34)

This gives the dependent variables in terms of a smaller number of transformed dependent
variables in a way which satisfies the linear constraints. In vector form, the equation becomes
b + D .
=

(7.35)

D = 0.

(7.36)

Here D is a full rank matrix which spans the same column space as does . Note that
may or may not be full rank. Since D spans the same column space as does , we must also
have in general

We see here this is true:



1 2

1
12

= (0).

(7.37)

We also note that the term exp(E j /RT ) is a modulating factor to the dynamics. Let
us see how this behaves for high and low temperatures. First for low temperature, we have


E j
lim exp
= 0.
(7.38)
T 0
RT
At high temperature, we have
lim exp

E j
RT

And lastly, at intermediate temperature, we have




E j
exp
O(1)
when
RT

= 1.

T =O

(7.39)

A sketch of this modulating factor is given in Figure 7.1. Note


CC BY-NC-ND. 03 August 2012, J. M. Powers.

Ej
R

(7.40)

231

7.1. ISOTHERMAL, ISOCHORIC KINETICS

exp(- E j /(R T))


1

Ej/ R

Figure 7.1: Plot of exp(E j /R/T ) versus T ; transition occurs at T E j /R.


for small T , the modulation is extreme, and the reaction rate is very small,
for T E j /R, the reaction rate is extremely sensitive to temperature, and
for T , the modulation is unity, and the reaction rate is limited only by molecular
collision frequency.
Now O and O2 represent molar concentrations which have standard units of mole/cm3 .
So the reaction rates
dO2
dO
and
dt
dt
have units of mole/cm3 /s.
Note that after conversion of E j from kJ/mole to erg/mole we find the units of the
argument of the exponential to be unitless. That is


Ej
RT

erg mole K 1

mole erg K

dimensionless.

(7.41)

Here the brackets denote the units of a quantity, and not molar concentration. Let us get
units for the collision frequency factor of reaction 13, a13 . We know the units of the rate
(mole/cm3 /s). Reaction 13 involves three molar species. Since 13 = 1, it also has an
extra temperature dependency. The exponential of a unitless number is unitless, so we need
not worry about that. For units to match, we must have






mole
mole
mole
mole
= [a13 ]
K 1 .
(7.42)
cm3 s
cm3
cm3
cm3
CC BY-NC-ND. 03 August 2012, J. M. Powers.

232
So the units of a13 are

CHAPTER 7. KINETICS IN SOME MORE DETAIL

2
mole
K
[a13 ] =
.
(7.43)
3
cm
s
For a14 we find a different set of units! Following the same procedure, we get





mole
mole
mole
= [a14 ]
K 1 .
(7.44)
cm3 s
cm3
cm3
So the units of a14 are

1
K
mole
[a14 ] =
.
(7.45)
cm3
s
This discrepancy in the units of aj the molecular collision frequency factor is a burden of traditional chemical kinetics, and causes many difficulties when classical non-dimensionalization
is performed. With much effort, a cleaner theory could be formulated; however, this would
require significant work to re-cast the now-standard aj values for literally thousands of reactions which are well established in the literature.
7.1.1.1.2 Example calculation Let us consider an example problem. Let us take T =
5000 K, and initial conditions b
O = 0.001 mole/cm3 and b
O2 = 0.001 mole/cm3 . The
initial temperature is very hot, and is near the temperature of the surface of the sun. This
is also realizable in laboratory conditions, but uncommon in most combustion engineering
environments.
We can solve these in a variety of ways. I chose here to solve both Eqs. (7.24-7.25) without
the reduction provided by Eq. (7.28). However, we can check after numerical solution to see
if Eq. (7.28) is actually satisfied. Substituting numerical values for all the constants to get



2 !
E 13
mole
K
13
7
2a13 T exp
= 2 2.9 10
(5000 K)1 exp(0),
3
cm
s
RT
2

1
mole
,
(7.46)
= 1.16 1014
3
cm
s
1 !



K
E
mole
14
(5000 K)1
2a14 T 14 exp
= 2 6.81 1018
3
cm
s
RT


erg
4.96 1012 mole
exp
,
erg
8.31441 107 mole
(5000 K)
K
1

1
mole
10
,
(7.47)
= 1.77548 10
cm3
s
2



1
E 13
mole
13
13
,
(7.48)
a13 T exp
= 5.80 10
3
cm
s
RT



1
mole
E 14
1
9
14
= 8.8774 10
a14 T exp
.
(7.49)
3
cm
s
RT
CC BY-NC-ND. 03 August 2012, J. M. Powers.

233

7.1. ISOTHERMAL, ISOCHORIC KINETICS


O, O2 (mole/cm3)
0.00150
O2
0.00100

0.00070

0.00050
O

t HsL
10-11

10-10

10-9

10-8

10-7

10-6

Figure 7.2: Molar concentrations versus time for oxygen dissociation problem.
Then the differential equation system becomes
dO
= (1.16 1014 )2O (O + O2 ) + (1.77548 1010 )O2 (O + O2 ),
(7.50)
dt
dO2
(7.51)
= (5.80 1013 )2O (O + O2 ) (8.8774 109 )O2 (O + O2 ),
dt
mole
,
(7.52)
O (0) = 0.001
cm3
mole
O2 (0) = 0.001
.
(7.53)
cm3
These non-linear ordinary differential equations are in a standard form for a wide variety
of numerical software tools. Solution of such equations are not the topic of these notes.
7.1.1.1.2.1 Species concentration versus time A solution was obtained numerically, and a plot of O (t) and O2 (t) is given in Figure 7.2. Note that significant reaction does
not commence until t 1010 s. This can be shown to be very close to the time between
molecular collisions. For 109 s < t < 108 s, there is a vigorous reaction. For t > 107 s,
the reaction appears to be equilibrated. The calculation gives the equilibrium values eO and
eO2 , as
mole
,
(7.54)
t
cm3
mole
,
(7.55)
lim O2 = eO2 = 0.00127
t
cm3
Note that at this high temperature, O2 is preferred over O, but there are definitely O
molecules present at equilibrium.
lim O = eO = 0.0004424

CC BY-NC-ND. 03 August 2012, J. M. Powers.

234

CHAPTER 7. KINETICS IN SOME MORE DETAIL


r
1.5 10-16
1.0 10-16
7.0 10-17
5.0 10-17

3.0 10-17

t HsL
10-10

10-9

10-8

10-7

10-6

Figure 7.3: Dimensionless residual numerical error r in satisfying the element conservation
constraint in the oxygen dissociation example.
We can check how well the numerical solution satisfied the algebraic constraint of element
conservation by plotting the dimensionless residual error r



+ 2 b
b


O2
O
O2
(7.56)
r= O
,
b


O + 2b
O2

as a function of time. If the constraint is exactly satisfied, we will have r = 0. Any non-zero
r will be related to the numerical method we have chosen. It may contain roundoff error
and have a sporadic nature. A plot of r(t) is given in Figure 7.3. Clearly the error is small,
and has the character of a roundoff error. In fact it is possible to drive r to be smaller by
controlling the error tolerance in the numerical method.

7.1.1.1.2.2 Pressure versus time We can use the ideal gas law to calculate the
pressure. Recall that the ideal gas law for molecular species i is
Pi V = ni RT.

(7.57)

Here Pi is the partial pressure of molecular species i, and ni is the number of moles of
molecular species i. Note that we also have
Pi =

ni
RT.
V

(7.58)

Note that by our definition of molecular species concentration that


i =
CC BY-NC-ND. 03 August 2012, J. M. Powers.

ni
.
V

(7.59)

235

7.1. ISOTHERMAL, ISOCHORIC KINETICS


So we also have the ideal gas law as
Pi = i RT.

(7.60)

Now in the Dalton mixture model, all species share the same T and V . So the mixture
temperature and volume are the same for each species Vi = V , Ti = T . But the mixture
pressure is taken to be the sum of the partial pressures:
P =

N
X

Pi .

(7.61)

i=1

Substituting from Eq. (7.60) into Eq. (7.61), we get


P =

N
X

i RT = RT

i=1

N
X

i .

(7.62)

i=1

For our example, we only have two species, so


P = RT (O + O2 ).

(7.63)

The pressure at the initial state t = 0 is


O + b
O2 ),
P (t = 0) = RT (b

= 8.31441 107

(7.64)


erg
mole
mole
(5000 K) 0.001
+
0.001
, (7.65)
mole K
cm3
cm3
dyne
,
(7.66)
= 8.31441 108
cm2
= 8.31441 102 bar.
(7.67)


This pressure is over 800 atmospheres. It is actually a little too high for good experimental
correlation with the underlying data, but we will neglect that for this exercise.
At the equilibrium state we have more O2 and less O. And we have a different number
of molecules, so we expect the pressure to be different. At equilibrium, the pressure is
(7.68)
P (t ) = lim RT (O + O2 ),
t



mole
erg 
mole
7
(5000 K) 0.0004424
= 8.31441 10
,
+ 0.00127
mole K
cm3
cm3
(7.69)
dyne
,
(7.70)
= 7.15 108
cm2
= 7.15 102 bar.
(7.71)
The pressure has dropped because much of the O has recombined to form O2 . Thus there
are fewer molecules at equilibrium. The temperature and volume have remained the same.
A plot of P (t) is given in Figure 7.4.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

236

CHAPTER 7. KINETICS IN SOME MORE DETAIL


P (dyne/cm2)
8.4 108
8.2 108
8. 108
7.8 108
7.6 108
7.4 108
7.2 108
t HsL
10-11

10-10

10-9

10-8

10-7

10-6

Figure 7.4: Pressure versus time for oxygen dissociation example.


7.1.1.1.2.3 Dynamical system form Now Eqs. (7.50-7.51) are of the standard form
for an autonomous dynamical system:
dy
= f(y).
dt

(7.72)

Here y is the vector of state variables (O , O2 )T . And f is an algebraic function of the state
variables. For the isothermal system, the algebraic function is in fact a polynomial.
Equilibrium
The dynamical system is in equilibrium when
f(y) = 0.

(7.73)

This non-linear set of algebraic equations can be difficult to solve for large systems. For
common chemical kinetics systems, such as the one we are dealing with, there is a guarantee
of a unique equilibrium for which all state variables are physical. There are certainly other
equilibria for which at least one of the state variables is non-physical. Such equilibria can
be quite mathematically complicated.
Solving Eq. (7.73) for our oxygen dissociation problem gives us symbolically from Eq. (7.6CC BY-NC-ND. 03 August 2012, J. M. Powers.

237

7.1. ISOTHERMAL, ISOCHORIC KINETICS


7.7)



E 13
E 14
e e e
13
2a13 exp
O O M T + 2a14 exp
eO2 eM T 14 = 0,
RT
RT




E 14
E 13
e e e
14
13
O O M a14 T exp
eO2 eM = 0.
a13 T exp
RT
RT


(7.74)
(7.75)

We notice that eM cancels. This so-called third body will in fact never affect the equilibrium state. It will however influence the dynamics. Removing eM and slightly rearranging
Eqs. (7.74-7.75) gives




E 14
E 13
e e
14
13
O O = a14 T exp
eO2 ,
(7.76)
a13 T exp
RT
RT




E 13
E 14
13
e e
14
a13 T exp
O O = a14 T exp
eO2 .
(7.77)
RT
RT
These are the same equations! So we really have two unknowns for the equilibrium state eO
and eO2 but seemingly only one equation. Note that rearranging either Eq. (7.76) or (7.77)
gives the result


E 14
14
e e
a
T
exp
14
O O
RT
 = K(T ).

(7.78)
=
e
E
O2
13

13
a T exp
13

RT

That is, for the net reaction (excluding the inert third body), O2 O+O, at equilibrium the
product of the concentrations of the products divided by the product of the concentrations
of the reactants is a function of temperature T . And for constant T , this is the so-called
equilibrium constant. This is a famous result from basic chemistry. It is actually not complete
yet, as we have not taken advantage of a connection with thermodynamics. But for now, it
will suffice.
We still have a problem: Eq. (7.78) is still one equation for two unknowns. We solve
this be recalling we have not yet taken advantage of our algebraic constraint of element
conservation, Eq. (7.28). Let us use the equation to eliminate eO2 in favor of eO :

1 b
eO2 =
O eO + b
O2 .
(7.79)
2

So Eq. (7.76) reduces to




 



E 13
1 b
E 14
13
e e
14
e
b
a13 T exp
O O = a14 T exp
O + O2 . (7.80)
2 O
RT
RT
|
{z
}
=eO

Equation (7.80) is one algebraic equation in one unknown. Its solution gives the equilibrium
value eO . It is a quadratic equation for eO . Of its two roots, one will be physical. We note

CC BY-NC-ND. 03 August 2012, J. M. Powers.

238

CHAPTER 7. KINETICS IN SOME MORE DETAIL

f(O) (mole/cm3/s)
300 000
200 000
100 000

-0.004

-0.003

-0.002

3
0.001 O (mole/cm )

-0.001
-100 000
-200 000

Figure 7.5: Equilibria for oxygen dissociation example.


that the equilibrium state will be a function of the initial conditions. Mathematically this
is because our system is really best posed as a system of differential-algebraic equations.
Systems which are purely differential equations will have equilibria which are independent
of their initial conditions. Most of the literature of mathematical physics focuses on such
systems of those. One of the foundational complications of chemical dynamics is the the
equilibria is a function of the initial conditions, and this renders many common mathematical
notions from traditional dynamic system theory to be invalid Fortunately, after one accounts
for the linear constraints of element conservation, one can return to classical notions from
traditional dynamic systems theory.
Consider the dynamics of Eq. (7.24) for the evolution of O . Equilibrating the right hand
side of this equation, gives Eq. (7.74). Eliminating M and then O2 in Eq. (7.74) then
substituting in numerical parameters gives the cubic algebraic equation
33948.3 (1.78439 1011 )(O )2 (5.8 1013 )(O )3 = f (O ) = 0.

(7.81)

This equation is cubic because we did not remove the effect of M . This will not affect
the equilibrium, but will affect the dynamics. We can get an idea of where the roots are
by plotting f (O ) as seen in Figure 7.5. Zero crossings of f (O ) in Figure 7.5 represent
equilibria of the system, eO , f (eO ) = 0. The cubic equation has three roots
eO = 0.003

mole
,
cm3

mole
,
cm3
mole
,
= 0.000442414
cm3

eO = 0.000518944
eO

non-physical,
non-physical,
physical.

(7.82)
(7.83)
(7.84)

Note the physical root found by our algebraic analysis is identical to that which was identified
by our numerical integration of the ordinary differential equations of reaction kinetics.
Stability of equilibria
CC BY-NC-ND. 03 August 2012, J. M. Powers.

239

7.1. ISOTHERMAL, ISOCHORIC KINETICS

We can get a simple estimate of the stability of the equilibria by considering the slope of
f near f = 0. Our dynamic system is of the form
dO
= f (O ).
dt

(7.85)

Near the first non-physical root at eO = 0.003, a positive perturbation from equilibrium induces f < 0, which induces dO /dt < 0, so O returns to its equilibrium.
Similarly, a negative perturbation from equilibrium induces dO /dt > 0, so the system
returns to equilibrium. This non-physical equilibrium point is stable. Note stability
does not imply physicality!
Perform the same exercise for the non-physical root at eO = 0.000518944. We find
this root is unstable.
Perform the same exercise for the physical root at eO = 0.000442414. We find this
root is stable.
In general if f crosses zero with a positive slope, the equilibrium is unstable. Otherwise, it
is stable.
Consider a formal Taylor series expansion of Eq. (7.85) in the neighborhood of an equilibrium point 3O :

d
df
e
e
( O ) = f (O ) +
( eO ) + . . .
(7.86)
| {z } dO =e O
dt O
O
O

=0

We find df /dO by differentiating Eq. (7.81) to get

df
= (3.56877 1011 )O (1.74 1014 )2O .
dO

(7.87)

We evaluate df /dO near the physical equilibrium point at O = 0.0004442414 to get


df
= (3.56877 1011 )(0.0004442414) (1.74 1014 )(0.0004442414)2,
dO
1
= 1.91945 108 .
s

(7.88)

Thus the Taylor series expansion of Eq. (7.24) in the neighborhood of the physical equilibrium gives the local kinetics to be driven by
d
( 0.000442414) = (1.91945 108 ) (O 0.0004442414) + . . . .
dt O

(7.89)

So in the neighborhood of the physical equilibrium we have



O = 0.0004442414 + A exp 1.91945 108 t .

(7.90)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

240

CHAPTER 7. KINETICS IN SOME MORE DETAIL

Here A is an arbitrary constant of integration. The local time constant which governs the
times scales of local evolution is where
=

1
= 5.20983 109 s.
1.91945 108

(7.91)

This nano-second time scale is very fast. It can be shown to be correlated with the mean
time between collisions of molecules.
7.1.1.1.3 Effect of temperature Let us perform four case studies to see the effect of
T on the systems equilibria and it dynamics near equilibrium.
T = 3000 K. Here we have significantly reduced the temperature, but it is still higher
than typically found in ordinary combustion engineering environments. Here we find
mole
,
cm3
= 1.92059 107 s.

eO = 8.9371 106

(7.92)
(7.93)

The equilibrium concentration of O dropped by two orders of magnitude relative to


T = 5000 K, and the time scale of the dynamics near equilibrium slowed by two orders
of magnitude.
T = 1000 K. Here we reduce the temperature more. This temperature is common in
combustion engineering environments. We find
mole
,
cm3
= 2.82331 101 s.

eO = 2.0356 1014

(7.94)
(7.95)

The O concentration at equilibrium is greatly diminished to the point of being difficult


to detect by standard measurement techniques. And the time scale of combustion has
significantly slowed.
T = 300 K. This is obviously near room temperature. We find
mole
,
cm3
s.

eO = 1.14199 1044
= 1.50977 1031

(7.96)
(7.97)

The O concentration is effectively zero at room temperature, and the relaxation time
is effectively infinite. As the oldest star in our galaxy has an age of 4.4 1017 s, we see
that at this temperature, our mathematical model cannot be experimentally validated,
so it loses its meaning. At such a low temperature, the theory becomes qualitatively
correct, but not quantitatively predictive.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

241

7.1. ISOTHERMAL, ISOCHORIC KINETICS

T = 10000 K. Such high temperature could be achieved in an atmospheric re-entry


environment.
mole
,
(7.98)
eO = 2.74807 103
cm3
= 1.69119 1010 s.
(7.99)
At this high temperature, O become preferred over O2 , and the time scales of reaction
become extremely small, under a nanosecond.
7.1.1.2

Single reversible reaction

The two irreversible reactions studied in the previous section are of a class that is common
in combustion modeling. However, the model suffers a defect in that its link to classical
equilibrium thermodynamics is missing. A better way to model essentially the same physics
and guarantee consistency with classical equilibrium thermodynamics is to model the process
as a single reversible reaction, with a suitably modified reaction rate term.
7.1.1.2.1

Mathematical model

7.1.1.2.1.1 Kinetics For the reversible O O2 reaction, let us only consider reaction
13 from Table 7.2 for which
13 : O2 + M O + O + M.

(7.100)

For this system, we have N = 2 molecular species in L = 1 elements reacting in J = 1


reaction. Here

1
cal
mole
11
0.5
E
=
95560
a13 = 1.85 10
. (7.101)
(K)
,

=
0.5,
13
13
cm3
mole
Units of cal are common in chemistry, but we need to convert to erg, which is achieved via

 7


erg
4.186 J
10 erg
cal
= 4.00014 1012
.
(7.102)
E 13 = 95560
mole
cal
J
mole
For this reversible reaction, we slightly modify the kinetics equations to



E 13
dO
1
13
= 2 a13 T exp
O2 M

,
dt
Kc,13 O O M
RT
|
{z
}
dO2
dt

= a13 T
|
|

=k13 (T )

13

exp
{z

{z

{z

=r13

E 13
RT

=k13 (T )



1
O2 M

.
Kc,13 O O M
}
=r13

(7.103)

(7.104)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

242

CHAPTER 7. KINETICS IN SOME MORE DETAIL

Here we have used equivalent definitions for k13 (T ) and r13 , so that Eqs. (7.103-7.104) can
be written compactly as
dO
= 2r13 ,
dt
dO2
= r13 .
dt
In matrix form, we can simplify to
   
d O
2
(r13 ).
=
1
dt O2
| {z }
=
Here the N J or 2 1 matrix is


2
.
1

Performing row operations, the matrix form reduces to



  
d
O
2
(r13 ),
=
0
dt O + 2O2

(7.105)
(7.106)

(7.107)

(7.108)

(7.109)

or

    
2
1 0 d O
=
(r13 ).
0
1 2 dt O2

(7.110)

So here the N N or 2 2 matrix L1 is


1



1 0
.
=
1 2

(7.111)

The N N or 2 2 permutation matrix P is the identity matrix. And the N J or 2 1


upper triangular matrix U is
 
2
.
(7.112)
U=
0
Note that = L U or equivalently L1 = U:
    

2
2
1 0
=
.

0
1
1 2
| {z } | {z } |{z}
=
=U
=L1

(7.113)

Once again the stoichiometric matrix is


= 1 2 .

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.114)

243

7.1. ISOTHERMAL, ISOCHORIC KINETICS


And we see that = 0 is satisfied:


2
1 2
1


= 0 .

(7.115)

As for the irreversible reactions, the reversible reaction rates are constructed to conserve
O atoms. We have

d
O + 2O2 = 0.
dt

Thus, we once again find

O + 2O2 = b
O + 2b
O2 = constant.

As before, we can say

    

b
1
O

= bO +
O .
1
| {z }
O2
O2
} | {z
| {z
} | {z2 }
=
=D
b
=
=


(7.116)

(7.117)

(7.118)

This gives the dependent variables in terms of a smaller number of transformed dependent
variables in a way which satisfies the linear constraints. In vector form, the equation becomes
b + D .
=

Once again D = 0.

(7.119)

7.1.1.2.1.2 Thermodynamics Equations (7.103-7.104) are supplemented by an expression for the thermodynamics-based equilibrium constant Kc,13 which is:


Po
Go13
Kc,13 =
.
(7.120)
exp
RT
RT
Here Po = 1.01326 106 dyne/cm2 = 1 atm is the reference pressure. The net change of
Gibbs free energy at the reference pressure for reaction 13, Go13 is defined as
Go13 = 2g oO g oO2 .

(7.121)

We further recall that the Gibbs free energy for species i at the reference pressure is defined
in terms of the enthalpy and entropy as
o

g oi = hi T soi .

(7.122)

It is common to find hi and soi in thermodynamic tables tabulated as functions of T .


CC BY-NC-ND. 03 August 2012, J. M. Powers.

244

CHAPTER 7. KINETICS IN SOME MORE DETAIL

We further note that both Eqs. (7.103) and (7.104) are in equilibrium when
eO2 eM =

1 e e e
.
Kc,13 O O M

(7.123)

We rearrange Eq. (7.123) to find the familiar


Kc,13

Q
eO eO
[products]
.
= e =Q
[reactants]
O2

(7.124)

If Kc,13 > 1, the products are preferred. If Kc,13 < 1, the reactants are preferred.
Now, Kc,13 is a function of T only, so it is known. But Eq. (7.124) once again is one
equation in two unknowns. We can use the element conservation constraint, Eq. (7.117) to
reduce to one equation and one unknown, valid at equilibrium:
Kc,13 =

eO eO
b
O2 + 12 (b
O eO )

(7.125)

Using the element constraint, Eq. (7.117), we can recast the dynamics of our system by
modifying Eq. (7.103) into one equation in one unknown:


E 13
dO
13
= 2a13 T exp
dt
RT

(O2 +
|

1 b
1
1
1
(O O )) (b
O2 + (b
O O ) + O )
O O (b
O2 + (b
O O ) + O )
,
2 {z
2
K
2
c,13
}|
{z
}
{z
}
|
=O2

=M

=M

7.1.1.2.2 Example calculation Let us consider the same example as the previous section with T = 5000 K. We need numbers for all of the parameters of Eq. (7.126). For O,
we find at T = 5000 K that
erg
o
,
(7.127)
hO = 3.48382 1012
mole
erg
.
(7.128)
soO = 2.20458 109
mole K
So



erg 
9
o
12 erg
(5000 K) 2.20458 10
,
g O = 3.48382 10
mole
mole K
erg
= 7.53908 1012
.
(7.129)
mole
For O2 , we find at T = 5000 K that
erg
o
hO2 = 1.80749 1012
,
(7.130)
mole
erg
soO2 = 3.05406 109
.
(7.131)
mole K
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.126)

245

7.1. ISOTHERMAL, ISOCHORIC KINETICS


So

erg 
erg 
(5000 K) 3.05406 109
mole
mole K
erg
= 1.34628 1013
.
mole

g oO2 =

1.80749 1012

(7.132)

Thus, by Eq. (7.121), we have


Go13 = 2(7.53908 1012 ) (1.34628 1013 ) = 1.61536 1012

erg
.
mole

(7.133)

(7.134)

Thus, by Eq. (7.120) we get for our system


Kc,13 =

1.01326 106 dyne


2
cm
erg
7
8.31441 10 mole K (5000 K)

exp

= 1.187 104

erg
1.61536 1012 mole

erg
8.31441 107 mole
(5000 K)
K

!!

mole
.
cm3

(7.135)

Substitution of all numerical parameters into Eq. (7.126) and expansion yields the following
dO
= 3899.47 (2.23342 1010 )2O (7.3003 1012 )3O = f (O ), O (0) = 0.001.(7.136)
dt
A plot of the time-dependent behavior of O and O2 from solution of Eq. (7.136) is given
in Figure 7.6. The behavior is similar to the predictions given by the pair of irreversible
reactions in Fig. 7.1. Here direct calculation of the equilibrium from time integration reveals
eO = 0.000393328

mole
.
cm3

(7.137)

mole
.
cm3

(7.138)

Using Eq. (7.117) we find this corresponds to


eO2 = 0.00130334

We note the system begins significant reaction for t 109 s and is equilibrated for t
107 s.
The equilibrium is verified by solving the algebraic equation
f (O ) = 3899.47 (2.23342 1010 )2O (7.3003 1012 )3O = 0.

(7.139)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

246

CHAPTER 7. KINETICS IN SOME MORE DETAIL


O, O2 (mole/cm3)
0.00150
O2

0.00100
0.00070
0.00050

O
0.00030

t HsL
10-11

10-9

10-7

10-5

0.001

Figure 7.6: Plot of O (t) and O2 (t) for oxygen dissociation with reversible reaction.
This yields three roots:
eO = 0.003

mole
,
cm3

mole
,
cm3
mole
= 0.000393328
,
cm3

eO = 0.000452678
eO

non-physical,
non-physical,
physical,

(7.140)
(7.141)
(7.142)
(7.143)

is given in Figure 7.6.


Linearizing Eq. (7.136) in the neighborhood of the physical equilibrium yields the equation
d
(7.144)
(O 0.000393328) = (2.09575 107 ) (O 0.000393328) + . . .
dt
This has solution

O = 0.000393328 + A exp 2.09575 107 t .
(7.145)
Again, A is an arbitrary constant. Obviously the equilibrium is stable. Moreover the time
constant of relaxation to equilibrium is
=

1
= 4.77156 108 s.
2.09575 107

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.146)

247

7.1. ISOTHERMAL, ISOCHORIC KINETICS


f(O) (mole/cm3/s)
40 000
30 000
20 000
10 000
-0.004

-0.003

-0.002

-0.001
-10 000

3
0.001 O (mole/cm )

-20 000

Figure 7.7: Plot of f (O ) versus O for oxygen dissociation with reversible reaction.
This is consistent with the time scale to equilibrium which comes from integrating the full
equation.

7.1.2

Zeldovich mechanism of N O production

Let us consider next a more complicated reaction system: that of NO production known
as the Zeldovich6 mechanism. This is an important model for the production of a major
pollutant from combustion processes. It is most important for high temperature applications.

7.1.2.1

Mathematical model

The model has several versions. One is


1:
2:

N + NO N2 + O,
N + O2 NO + O.

(7.147)
(7.148)

similar to our results for O2 dissociation, N2 and O2 are preferred at low temperature. As
the temperature rises N and O begin to appear, and it is possible when they are mixed for
NO to appear as a product.
6

Yakov Borisovich Zeldovich, 1915-1987, prolific Soviet physicist and father of thermonuclear weapons.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

248
7.1.2.1.1
with

CHAPTER 7. KINETICS IN SOME MORE DETAIL


Standard model form Here we have the reaction of N = 5 molecular species

N O
N

=
N2 .
O
O2

(7.149)

We have L = 2 with N and O as the 2 elements. The stoichiometric matrix of dimension


L N = 2 5 is


1 1 2 0 0
.
(7.150)
=
1 0 0 1 2
The first row of is for the N atom; the second row is for the O atom.
And we have J = 2 reactions. The reaction vector of length J = 2 is




 
Ta,1
1
1

a
T
exp

1
r1
 T   N N O Kc,1 N2 O  ,
r =
=
r2
N O2 K1c,2 N O O
a2 T 2 exp TTa,2


k1 N N O K1c,1 N2 O
 .
= 
k2 N O2 K1c,2 N O O

(7.151)

(7.152)

Here, we have

k1
k2



Ta,1
= a1 T exp
,
T


Ta,2
2
.
= a2 T exp
T
1

In matrix form, the model can be written as


N O
1 1
N 1 1  
r1

d
N = 1

0
2
r2 .

dt
O 1
1
O2
0 1
{z
}
|
=

(7.153)
(7.154)

(7.155)

Here the matrix has dimension N J which is 5 2. The model is of our general form
d
= r.
dt
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.156)

249

7.1. ISOTHERMAL, ISOCHORIC KINETICS

Note that our stoichiometric constraint on element conservation for each reaction = 0
holds here:

1 1

 1 1 


0
0
1 1 2 0 0
0
(7.157)
1
=
= 0 0 .
1 0 0 1 2
1
1
0 1

We get 4 zeros because there are 2 reactions each with 2 element constraints.

7.1.2.1.2 Reduced form Here we describe non-traditional, but useful reductions, using
standard techniques from linear algebra to bring the model equations into a reduced form in
which all of the linear constraints have been explicitly removed.
Let us perform a series of row operations to find all of the linear dependencies. Our aim
is to convert the matrix into an upper triangular form. The lower left corner of already
has a zero, so there is no need to worry about it. Let us add the first and fourth equations
to eliminate the 1 in the 4, 1 slot. This gives

1 1
N O
1 1  
N
r1

d
= 1
N
(7.158)
0
2
r2 .

dt

N O + O
0
2
O2
0 1
Next, add the first and third equations to get

N O
1 1
1 1  

N
r1

d
N O + N = 0
1
2
r2 .

dt

0
2
N O + O
0 1
O2

Now multiply the first equation by 1 and add it to the second to get


N O
1 1
N O + N 0 2  
r1

d
N O + N = 0

1
2
r .

dt
N O + O 0
2 2
0 1
O2

Next multiply the fifth equation by 2 and add it to the second to get


1 1
N O

0 2  
N O + N
r1

d

= 0
1
N O + N2

r2 .

dt

0
2
N O + O
0
0
N O + N 2O2

(7.159)

(7.160)

(7.161)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

250

CHAPTER 7. KINETICS IN SOME MORE DETAIL

Next add the second and fourth equations to get


1 1
N O

0 2  
N O + N
r1

d

= 0
1
N O + N2
r2 .

dt

0
0
N + O
0
0
N O + N 2O2

Next multiply the third equation by 2 and add it to the second to get


N O
1 1

0 2  
N O + N
r1

d
N O + N + 2N = 0
0
2

r2 .

dt

0
0
N + O
N O + N 2O2
0
0
Rewritten, this becomes

1 0
1 1

1 1

0 1
1 1
|

0
0
2
0
0
{z

=L1


0 0
1 1
N O

 

0 0
d N 0 2 r1

0 0 N2
0
r2 .
= 0
dt

1 0
O
0
0
0 2
0
0
O2
}
{z
}
|

(7.162)

(7.163)

(7.164)

=U

A way to think of this type of row echelon form is that it defines two free variables, those
associated with the non-zero pivots of U: N O and N . The remain three variables N2 , O
and O2 are bound variables which can be expressed in terms of the free variables.
The last three of the ordinary differential equations are homogeneous and can be integrated to form
N O + N + 2N2 = C1 ,
N + O = C2 ,
N O + N 2O2 = C3 .

(7.165)
(7.166)
(7.167)

The constants C1 , C2 and C3 are determined from the initial conditions on all five state
variables. In matrix form, we can say

N O

C1

1 1 2 0 0
N

0 1 0 1 0 N = C2 .
(7.168)
2

C3
O
1 1 0 0 2
O2
CC BY-NC-ND. 03 August 2012, J. M. Powers.

251

7.1. ISOTHERMAL, ISOCHORIC KINETICS

Considering the free variables, N O and N , to be known, we move them to the right side
to get

2 0 0
N2
C1 N O N
0 1 0 O = C2 N
.
(7.169)
0 0 2
C3 + N O N
O2
Solving, for the bound variables, we find
1

1
1
C

N2

1
N
O
N
2
2
2
O =
.
C2 N
1
1
1
2 C3 2 N O + 2 N
O2

We can rewrite this as

1
1
1


C

N2

1
2
2
2

N
O
O = C2 + 0 1
.
N
1
1
1
2 2
2 C3
O2

(7.170)

(7.171)

We can get a more elegant form by defining N O = N O and N = N . Thus we can say our
state variables have the form

0
N O
1
0

N 0 0

1
1 11 N O

N = C1 +
.
(7.172)
2
2
2 2
O C2 0 1 N
21 C3
21 12
O2
By translating via N O = N O + b
N O and N = N + b
N and choosing the constants C1 , C2
and C3 appropriately, we can arrive at

b
N O
N O
1
0


N
b

N

01 11 N O

N = b
+ 2 2
.
(7.173)

2 N2

N

O b

0 1 | {z }
O
b
12 21
O2
O2
=
{z
}
| {z } | {z } |
=D
=
b
=

This takes the form of

b + D .
=

(7.174)

Here the matrix D is of dimension N R, which here is 5 2. It spans the same column
space as does the N J matrix which is of rank R. Here in fact R = J = 2, so D has the
same dimension as . In general it will not. If c1 and c2 are the column vectors of D, we
CC BY-NC-ND. 03 August 2012, J. M. Powers.

252

CHAPTER 7. KINETICS IN SOME MORE DETAIL

see that c1 c2 forms the first column vector of and c1 c2 forms the second column
vector of . Note that D = 0:

1
0


 0

1

0 0
1 1 2 0 0
1
1

.
(7.175)
2 2 =
D=
0 0
1 0 0 1 2
0 1
21 21
Equations (7.165-7.167) can also be linearly combined in a way which has strong physical relevance. We rewrite the system as three equations in which the first is identical to
Eq. (7.165); the second is the difference of Eqs. (7.166) and (7.167); and the third is half of
Eq. (7.165) minus half of Eq. (7.167) plus Eq. (7.166):

N O + N + 2N2 = C1 ,
(7.176)
O + N O + 2O2 = C2 C3 ,
(7.177)
1
N O + N + N2 + O + O2 =
(C1 C3 ) + C2 .
(7.178)
2
Equation (7.176) insists that the number of nitrogen elements be constant; Eq. (7.177)
demands the number of oxygen elements be constant; and Eq. (7.178) requires the number
of moles of molecular species be constant. For general reactions, including the earlier studied
oxygen dissociation problem, the number of moles of molecular species will not be constant.
Here because each reaction considered has two molecules reacting to form two molecules,
we are guaranteed the number of moles will be constant. Hence, we get an additional
linear constraint beyond the two for element conservation. Note that since our reaction is
isothermal, isochoric and mole-preserving, it will also be isobaric.
7.1.2.1.3

Example calculation Let us consider an isothermal reaction at


T = 6000 K.

(7.179)

The high temperature is useful in generating results which are easily visualized. It insures
that there will be significant concentrations of all molecular species. Let us also take as an
initial condition
mole
b
N O = b
N = b
N2 = b
O = b
O2 = 1 106
.
(7.180)
cm3
For this temperature and concentrations, the pressure, which will remain constant through
the reaction, is P = 2.4942 106 dyne/cm2 . This is a little greater than atmospheric.
Kinetic data for this reaction is adopted from Baulch, et al.7 The data for reaction 1 is
1

1
mole
13
,
1 = 0,
Ta1 = 0 K.
(7.181)
a1 = 2.107 10
3
cm
s
7

Baulch, et al., 2005, Evaluated Kinetic Data for Combustion Modeling: Supplement II, Journal of
Physical and Chemical Reference Data, 34(3): 757-1397.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

253

7.1. ISOTHERMAL, ISOCHORIC KINETICS


For reaction 2, we have
a2 = 5.8394 10

mole
cm3

1

1
K 1.01 s

2 = 1.01,

Ta2 = 3120 K.

(7.182)

Here the so-called activation temperature Ta,j for reaction j is really the activation energy
scaled by the universal gas constant:
Ta,j =

Ej
.
R

(7.183)

Substituting numbers we obtain for the reaction rates





1
0
mole
1
13
0
13
k1 = (2.107 10 )(6000) exp
= 2.107 10
,
(7.184)
3
6000
cm
s


1

mole
1
3120
13
9
1.01
= 2.27231 10
.
k2 = (5.8394 10 )(6000) exp
3
6000
cm
s
(7.185)
We will also need thermodynamic data. The data here will be taken from the Chemkin
database.8 Thermodynamic data for common materials is also found in most thermodynamic
texts. For our system at 6000 K, we find
g oN O = 1.58757 1013
g oN = 7.04286 1012
g oN2 = 1.55206 1013
g oO = 9.77148 1012
g oO2 = 1.65653 1013

erg
,
mole
erg
,
mole
erg
,
mole
erg
,
mole
erg
.
mole

(7.186)
(7.187)
(7.188)
(7.189)
(7.190)

Thus for each reaction, we find Goj :


Go1 = g oN2 + g oO g oN g oN O ,
(7.191)
= 1.55206 1013 9.77148 1012 + 7.04286 1012 + 1.58757 1013 ,(7.192)
erg
,
(7.193)
= 2.37351 1012
mole
Go2 = g oN O + g oO g oN g oO2 ,
(7.194)
13
12
12
13
= 1.58757 10 9.77148 10 + 7.04286 10 + 1.65653 10 ,(7.195)
erg
.
(7.196)
= 2.03897 1012
mole
8

R. J. Kee, et al., 2000, The Chemkin Thermodynamic Data Base, part of the Chemkin Collection
Release 3.6, Reaction Design, San Diego, CA.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

254

CHAPTER 7. KINETICS IN SOME MORE DETAIL

At 6000 K, we find the equilibrium constants for the J = 2 reactions are


Kc,1 =
=
=
Kc,2 =
=
=


Go1
exp
,
RT


2.37351 1012
exp
,
(8.314 107 )(6000)
116.52,


Go2
exp
,
RT


2.03897 1012
,
exp
(8.314 107 )(6000)
59.5861.

(7.197)
(7.198)
(7.199)
(7.200)
(7.201)
(7.202)

Again, omitting details, we find the two differential equations governing the evolution of
the free variables are
dN O
= 0.723 + 2.22 107 N + 1.15 1013 2N 9.44 105 N O 3.20 1013 N N O ,
dt
(7.203)
dN
= 0.723 2.33 107 N 1.13 1013 2N + 5.82 105 N O 1.00 1013 N N O .
dt
(7.204)
Solving numerically, we obtain a solution shown in Fig. 7.8. The numerics show a relaxation
to final concentrations of
lim N O = 7.336 107

lim N = 3.708 108

mole
,
cm3
mole
.
cm3

(7.205)
(7.206)

Equations (7.203-7.204) are of the form


dN O
= fN O (N O , N ),
dt
dN
= fN (N O , N ).
dt

(7.207)
(7.208)

At equilibrium, we must have


fN O (N O , N ) = 0,
fN (N O , N ) = 0.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.209)
(7.210)

255

7.1. ISOTHERMAL, ISOCHORIC KINETICS


N, NO (mole/cm3)
1 10-5
5 10-6

1 10-6
NO

5 10-7

1 10-7
5 10-8
N
10-10

10-9

10-8

10-7

10-6

t s
10-5

Figure 7.8: NO and N concentrations versus time for T = 6000 K, P = 2.4942


106 dyne/cm2 Zeldovich mechanism.
We find three finite roots to this problem:
mole
,
non-physical, (7.211)
cm3
mole
2 : (N O , N ) = (5.173 108 , 2.048 106 )
,
non-physical, (7.212)
cm3
mole
,
physical.
(7.213)
3 : (N O , N ) = (7.336 107 , 3.708 108 )
cm3
1 : (N O , N ) = (1.605 106 , 3.060 108 )

Obviously, because of negative concentrations, roots 1 and 2 are non-physical. Root 3


however is physical; moreover, it agrees with the equilibrium we found by direct numerical
integration of the full non-linear equations.
We can use local linear analysis in the neighborhood of each equilibria to rigorously
ascertain the stability of each root. Taylor series expansion of Eqs. (7.207-7.208) in the
neighborhood of an equilibrium point yields




d
f
f
N
O
N
O
(N O eN O ) +
( eN ) + . . . ,
(N O eN O ) = fN O |e +
| {z } N O e
dt
N e N
=0



fN
fN
d
e
e
( N ) = fN |e +
( N O ) +
( eN ) + . . . .
|{z} N O e N O
dt N
N e N

(7.214)
(7.215)

=0

CC BY-NC-ND. 03 August 2012, J. M. Powers.

256

CHAPTER 7. KINETICS IN SOME MORE DETAIL

Evaluation of Eqs. (7.214-7.215) near the physical root, root 3, yields the system


 

d N O 7.336 107
N O 7.336 107
2.129 106 4.155 105
=
.
2.111 105 3.144 107
N 3.708 108
dt N 3.708 108
{z
}
|
f
=J= |
e

(7.216)

This is of the form


d
f
e
( e ) = J ( e ) .
( ) =
dt
e

(7.217)

It is the eigenvalues of the Jacobian9 matrix J that give the time scales of evolution of the
concentrations as well as determine the stability of the local equilibrium point. Recall that
we can usually decompose square matrices via the diagonalization
J = S S1 .

(7.218)

Here, S is the matrix whose columns are composed of the right eigenvectors of J, and is
the diagonal matrix whose diagonal is populated by the eigenvalues of J. For some matrices
(typically not those encountered after our removal of linear dependencies), diagonalization
is not possible, and one must resort to the so-called near-diagonal Jordan form. This will
not be relevant to our discussion, but could be easily handled if necessary. We also recall the
eigenvector matrix and eigenvalue matrix are defined by the standard eigenvalue problem
J S = S .

(7.219)

We also recall that the components of are found by solving the characteristic polynomial
which arises from the equation
det (J I) = 0,

(7.220)

where I is the identity matrix. Defining z such that


S z e ,

(7.221)

and using the decomposition Eq. (7.218), Eq. (7.217) can be rewritten to form
d
(S z) = |S {z S1} (S z),
| {z }
dt
J
e
dz
= S z,
S
dt
dz
S1 S
= S1 S z,
dt
dz
= z.
dt
9

after Carl Gustav Jacob Jacobi, 1804-1851, German mathematician.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.222)

(7.223)
(7.224)
(7.225)

257

7.1. ISOTHERMAL, ISOCHORIC KINETICS


Eq. (7.225) reduces to the diagonal form
dz
= z.
dt

(7.226)

This has solution for each component of z of


z1 = C1 exp(1 t),
z2 = C2 exp(2 t),
..
.

(7.227)
(7.228)
(7.229)

Here, our matrix J, see Eq. (7.216), has two real, negative eigenvalues in the neighborhood
of the physical root 3:
1
,
s
1
= 2.132 106 .
s

1 = 3.143 107

(7.230)

(7.231)

Thus we can conclude that the physical equilibrium is linearly stable. The local time constants near equilibrium are given by the reciprocal of the magnitude of the eigenvalues.
These are
1 = 1/|1| = 3.181 108 s,
2 = 1/|2| = 4.691 107 s.

(7.232)
(7.233)

Evolution on these two time scales is predicted in Fig. 7.8. This in fact a multiscale problem.
One of the major difficulties in the numerical simulation of combustion problems comes in
the effort to capture the effects at all relevant scales. The problem is made more difficult as
the breadth of the scales expands. In this problem, the breadth of scales is not particularly
challenging. Near equilibrium the ratio of the slowest to the fastest time scale, the stiffness
ratio , is
=

2
4.691 107 s
=
= 14.75.
1
3.181 108 s

(7.234)

Many combustion problems can have stiffness ratios over 106 . This is more prevalent at
lower temperatures.
We can do a similar linearization near the initial state, find the local eigenvalues, and
the the local time scales. At the initial state here, we find those local time scales are
1 = 2.403 108 s,
2 = 2.123 108 s.

(7.235)
(7.236)

So initially the stiffness, = (2.403 108 s)/(2.123 108 s) = 1.13 is much less, but the
time scale itself is small. It is seen from Fig. 7.8 that this initial time scale of 108 s well
CC BY-NC-ND. 03 August 2012, J. M. Powers.

258

CHAPTER 7. KINETICS IN SOME MORE DETAIL

predicts where significant evolution of species concentrations commences. For t < 108 s, the
model predicts essentially no activity. This can be correlated with the mean time between
molecular collisionsthe theory on which estimates of the collision frequency factors aj are
obtained.
We briefly consider the non-physical roots, 1 and 2. A similar eigenvalue analysis of root
1 reveals that the eigenvalues of its local Jacobian matrix are
1 = 1.193 107
2 = 5.434 106

1
,
s

1
.
s

(7.237)
(7.238)

Thus root 1 is a saddle and is unstable.


For root 2, we find
1
,
s
1
= 4.397 107 i7.997 106 .
s

1 = 4.397 107 + i7.997 106

(7.239)

(7.240)

The eigenvalues are complex with a positive real part. This indicates the root is an unstable
spiral source.
A detailed phase portrait is shown in Fig. 7.9. Here we see all three roots. Their local
character of sink, saddle, or spiral source is clearly displayed. We see that trajectories are
attracted to a curve labeled SIM for Slow Invariant Manifold. A part of the SIM is
constructed by the trajectory which originates at root 1 and travels to root 3. The other
part is constructed by connecting an equilibrium point at infinity into root 3. Details are
omitted here.
7.1.2.2

Stiffness, time scales, and numerics

One of the key challenges in computational chemistry is accurately predicting species concentration evolution with time. The problem is made difficult because of the common presence
of physical phenomena which evolve on a widely disparate set of time scales. Systems which
evolve on a wide range of scales are known as stiff, recognizing a motivating example in
mass-spring-damper systems with stiff springs. Here we will examine the effect of temperature and pressure on time scales and stiffness. We shall also look simplistically how different
numerical approximation methods respond to stiffness.
7.1.2.2.1 Effect of temperature Let us see how the same Zeldovich mechanism behaves at lower temperature, T = 1500 K; all other parameters, including the initial species
concentrations are the same as the previous high temperature example. The pressure however, lowers, and here is P = 6.23550105 dyne/cm2 , which is close to atmospheric pressure.
For this case, a plot of species concentrations versus time is given in Figure 7.10.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

259

7.1. ISOTHERMAL, ISOCHORIC KINETICS

x 10

-7

5
0
N (mole/cm 3)

3 sink

saddle

SIM

SIM
-5
-10
-15

spira l
source

-20

2
-4

-3

-2

-1

NO (mole/cm 3)

x 10

-6

Figure 7.9: NO and N phase portraits for T = 6000 K, P = 2.4942 106 dyne/cm2
Zeldovich mechanism.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

260

CHAPTER 7. KINETICS IN SOME MORE DETAIL


N, NO (mole/cm3)
10-6

NO

10-8

10-10

10-12

10-11

10-9

t s

10-7

10-5

0.001

0.1

10

Figure 7.10: N O and N versus time for Zeldovich mechanism at T = 1500 K, P =


6.23550 105 dyne/cm2 .
At T = 1500 K, we notice some dramatic differences relative to the earlier studied
T = 6000 K. First, we see the reaction commences in around the same time, t 108 s. For
t 106 s, there is a temporary cessation of significant reaction. We notice a long plateau
in which species concentrations do not change over several decades of time. This is actually
a pseudo-equilibrium. Significant reaction recommences for t 0.1 s. Only around t 1 s
does the system approach final equilibrium. We can perform an eigenvalue analysis both
at the initial state and at the equilibrium state to estimate the time scales of reaction. For
this dynamical system which is two ordinary differential equations in two unknowns, we will
always find two eigenvalues, and thus two time scales. Let us call them 1 and 2 . Both
these scales will evolve with t.
At the initial state, we find
1 = 2.37 108 s,
2 = 4.25 107 s.

(7.241)
(7.242)

The onset of significant reaction is consistent with the prediction given by 1 at the initial
state. Moreover, initially, the reaction is not very stiff; the stiffness ratio is = 17.9.
At equilibrium, we find
mole
,
cm3
mole
= 4.2 1014
,
cm3

lim N O = 4.6 109

lim N

CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.243)
(7.244)

261

7.1. ISOTHERMAL, ISOCHORIC KINETICS


and
1 = 7.86 107 s,
2 = 3.02 101 s.

(7.245)
(7.246)

The slowest time scale near equilibrium is an excellent indicator of how long the system
takes to relax to its final state. Note also that near equilibrium, the stiffness ratio is large,
= 2 /1 3.8 105 . This is known as the stiffness ratio. When it is large, the scales in
the problem are widely disparate and accurate numerical solution becomes challenging.
In summary, we find the effect of lowering temperature while leaving initial concentrations
constant
lowers the pressure somewhat, slightly slowing down the collision time, and slightly
slowing the fastest time scales, and
slows the slowest time scales many orders of magnitude, stiffening the system significantly, since collisions may not induce reaction with their lower collision speed.
7.1.2.2.2 Effect of initial pressure Let us maintain the initial temperature at T =
1500 K, but drop the initial concentration of each species to
mole
b
.
N O = b
N = b
N2 = b
O2 = b
O = 108
cm3

(7.247)

With this decrease in number of moles, the pressure now is


P = 6.23550 103

dyne
.
cm2

(7.248)

This pressure is two orders of magnitude lower than atmospheric. We solve for the species
concentration profiles and show the results of numerical prediction in Figure 7.11 Relative to
the high pressure P = 6.2355 105 dyne/cm2 , T = 1500 K case, we notice some similarities
and dramatic differences. The overall shape of the time-profiles of concentration variation
is similar. But, we see the reaction commences at a much later time, t 106 s. For
t 104 s, there is a temporary cessation of significant reaction. We notice a long plateau
in which species concentrations do not change over several decades of time. This is again
actually a pseudo-equilibrium. Significant reaction recommences for t 10 s. Only around
t 100 s does the system approach final equilibrium. We can perform an eigenvalue analysis
both at the initial state and at the equilibrium state to estimate the time scales of reaction.
At the initial state, we find
1 = 2.37 106 s,
2 = 4.25 105 s.

(7.249)
(7.250)

The onset of significant reaction is consistent with the prediction given by 1 at the initial
state. Moreover, initially, the reaction is not very stiff; the stiffness ratio is = 17.9.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

262

CHAPTER 7. KINETICS IN SOME MORE DETAIL


N, NO (mole/cm3)
10-8
NO
10-10

10-12

10-14
N
10-16

10-8

10-6

10-4

t s
0.01

100

Figure 7.11: N O and N versus time for Zeldovich mechanism at T = 1500 K, P =


6.2355 103 dyne/cm2 .
Interestingly, by decreasing the initial pressure by a factor of 102 , we increased the initial
time scales by a complementary factor of 102 ; moreover, we did not alter the stiffness.
At equilibrium, we find
lim N O = 4.6 1011

lim N = 4.2 1016

mole
,
cm3
mole
,
cm3

(7.251)
(7.252)
(7.253)

and
1 = 7.86 105 s,
2 = 3.02 101 s.

(7.254)
(7.255)

By decreasing the initial pressure by a factor of 102 , we decreased the equilibrium concentrations by a factor of 102 and increased the time scales by a factor of 102 , leaving the stiffness
ratio unchanged.
In summary, we find the effect of lowering the initial concentrations significantly while
leaving temperature constant
lowers the pressure significantly, proportionally slowing down the collision time, as well
as the fastest and slowest time scales, and
does not affect the stiffness of the system.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

263

7.1. ISOTHERMAL, ISOCHORIC KINETICS

7.1.2.2.3 Stiffness and numerics The issue of how to simulate stiff systems of ordinary
differential equations, such as presented by our Zeldovich mechanism, is challenging. Here a
brief summary of some of the issues will be presented. The interested reader should consult
the numerical literature for a full discussion. See for example the excellent text of Iserles.10
We have seen throughout this section that there are two time scales at work, and they are
often disparate. The species evolution is generally characterized by an initial fast transient,
followed by a long plateau, then a final relaxation to equilibrium. We noted from the phase
plane of Fig. 7.9 that the final relaxation to equilibrium (shown along the green line labeled
SIM) is an attracting manifold for a wide variety of initial conditions. The relaxation onto
the SIM is fast, and the motion on the SIM to equilibrium is relatively slow.
Use of common numerical techniques can often mask or obscure the actual dynamics.
Numerical methods to solve systems of ordinary differential equations can be broadly categorized as explicit or implicit. We give a brief synopsis of each class of method. We cast
each as a method to solve a system of the form
d
= f().
dt

(7.256)

Explicit: The simplest of these, the forward Euler method, discretizes Eq. (7.256) as
follows:
n+1 n
= f(n ),
(7.257)
t
so that
n+1 = n + t f(n ).

(7.258)

Explicit methods are summarized as


easy to program, since Eq. (7.258) can be solved explicitly to predict the new
value n+1 in terms of the old values at step n.
need to have t < f astest in order to remain numerically stable,
able to capture all physics and all time scales at great computational expense for
stiff problems,
requiring much computational effort for little payoff in the SIM region of the phase
plane, and thus
inefficient for some portions of stiff calculations.
Implicit: The simplest of these methods, the backward Euler method, discretizes
Eq. (7.256) as follows:
n+1 n
= f(n+1 ),
t

(7.259)

10

A. Iserles, 2008, A First Course in the Numerical Analysis of Differential Equations, Cambridge University Press, Cambridge, UK.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

264

CHAPTER 7. KINETICS IN SOME MORE DETAIL


so that
n+1 = n + t f(n+1 ).

(7.260)

Implicit methods are summarized as


more difficult to program since a non-linear set of algebraic equations, Eq. (7.260),
must be solved at every time step with no guarantee of solution,
requiring potentially significant computational time to advance each time step,
capable of using very large time steps and remaining numerically stable,
suspect to missing physics that occur on small time scales < t,
in general better performers than explicit methods.
A wide variety of software tools exist to solve systems of ordinary differential equations.
Most of them use more sophisticated techniques than simple forward and backward Euler
methods. One of the most powerful techniques is the use of error control. Here the user
specifies how far in time to advance and the error that is able to be tolerated. The algorithm,
which is complicated, selects then internal time steps, for either explicit or implicit methods,
to achieve a solution within the error tolerance at the specified output time. A well known
public domain algorithm with error control is provided by lsode.f, which can be found in
the netlib repository.11
Let us exercise the Zeldovich mechanism under the conditions simulated in Fig. 7.11,
T = 1500 K, P = 6.2355 103 dyne/cm2 . Recall in this case the fastest time scale near
equilibrium is 1 = 7.86 105 s 104 s at the initial state, and the slowest time scale is
= 3.02 101 s at the final state. Let us solve for these conditions using dlsode.f, which
uses internal time stepping for error control, in both an explicit and implicit mode. We
specify a variety of values of t and report typical values of number of internal time steps
selected by dlsode.f, and the corresponding effective time step tef f used for the problem,
for both explicit and implicit methods, as reported in Table 7.3.
Obviously if output is requested using t > 104 s, the early time dynamics near t
104 s will be missed. For physically stable systems, codes such as dlsode.f will still provide
a correct solution at the later times. For physically unstable systems, such as might occur in
turbulent flames, it is not clear that one can use large time steps and expect to have fidelity
to the underlying equations. The reason is the physical instabilities may evolve on the same
time scale as the fine scales which are overlooked by large t.

7.2

Adiabatic, isochoric kinetics

It is more practical to allow for temperature variation within a combustor. The best model
for this is adiabatic kinetics. Here we will restrict our attention to isochoric problems.
11

Hindmarsh, A. C., 1983,ODEPACK, a Systematized Collection of ODE Solvers, Scientific Computing, edited by R. S. Stepleman, et al., North-Holland, Amsterdam, pp. 55-64. Source code at
http://www.netlib.org/alliant/ode/prog/lsode.f
CC BY-NC-ND. 03 August 2012, J. M. Powers.

265

7.2. ADIABATIC, ISOCHORIC KINETICS

t (s)
102
101
100
101
102
103
104
105
106

Explicit
Ninternal
106
105
104
103
102
101
100
100
100

Explicit
tef f (s)
104
104
104
104
104
104
104
105
106

Implicit
Ninternal
100
100
100
100
100
100
100
100
100

Implicit
tef f (s)
102
101
100
101
102
103
104
105
106

Table 7.3: Results from computing Zeldovich NO production using implicit and explicit
methods with error control in dlsode.f.

7.2.1

Thermal explosion theory

There is a simple description known as thermal explosion theory which provides a good
explanation for how initially slow exothermic reaction induces a sudden temperature rise
accompanied by a final relaxation to equilibrium.
Let us consider a simple isomerization reaction in a closed volume
A B.

(7.261)

Let us take A and B to both be calorically perfect ideal gases with identical molecular masses
MA = MB = M and identical specific heats, cvA = cvB = cv ; cP A = cP B = cP . We can
consider A and B to be isomers of an identical molecular species. So we have N = 2 species
reacting in J = 1 reactions. The number of elements L here is irrelevant.
7.2.1.1

One-step reversible kinetics

Let us insist our reaction process be isochoric and adiabatic, and commence with only A
present. The reaction kinetics with = 0 are



1
dA
E
A
,
(7.262)
= a exp
dt
Kc B
RT
|
{z
}
=k
|
{z
}
=r



dB
E
1
A
,
(7.263)
= a exp
dt
Kc B
RT
|
{z
}
=k
|
{z
}
=r

A (0) = b
A ,
B (0) = 0.

(7.264)
(7.265)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

266

CHAPTER 7. KINETICS IN SOME MORE DETAIL

For our alternate compact linear algebra based form, we note that



1
E
,
A

r = a exp
Kc B
RT
and that
d
dt

A
B

 
1
(r).
=
1

Performing the decomposition yields


  

d
1
A
=
(r).
0
dt A + B

(7.266)

(7.267)

(7.268)

Expanded, this is
    

1 0 d A
1
(r).
=
1 1 dt B
0

(7.269)

Combining Eqs. (7.262-7.263) and integrating yields

Thus, Eq. (7.262) reduces to

d
( + B ) = 0,
dt A
A + B = b
A ,
B = b
A A .

dA
= a exp
dt

E
RT




1 b
.
A
A
Kc A

Scaling, Eq. (7.273) can be rewritten as


 




d
1
A
E
A
A
1
1
= exp

.
b
b
d(at) b
RTo T /To
A
A Kc
A
7.2.1.2

(7.270)
(7.271)
(7.272)

(7.273)

(7.274)

First law of thermodynamics

Recall the first law of thermodynamics and neglecting potential and kinetic energy changes:
dU
.
= Q W
dt

(7.275)

Here U is the total internal energy. Because we insist the problem is adiabatic Q = 0.
= 0. Thus we have
Because we insist the problem is isochoric, there is no work done, so W
dU
= 0.
dt
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.276)

267

7.2. ADIABATIC, ISOCHORIC KINETICS


Thus, we find
U = Uo .

(7.277)

Recall the total internal energy for a mixture of two calorically perfect ideal gases is
U = nA uA + nB uB ,
(7.278)
n

n
A
B
= V
uA +
uB ,
(7.279)
V
V
(7.280)
= V (A uA + B uB ) ,


 

PA
PB
+ B hB
,
(7.281)
= V A hA
A
B


= V A hA RT + B hB RT ,
(7.282)
 



o
o
= V A cP (T To ) + hTo ,A RT + B cP (T To ) + hTo ,B RT ,(7.283)


o
o
(7.284)
= V (A + B )(cP (T To ) RT ) + A hTo ,A + B hTo ,B ,


o
o
= V (A + B )((cP R)T cP To ) + A hTo ,A + B hTo ,B ,
(7.285)


o
o
= V (A + B )((cP R)T (cP R + R)To ) + A hTo ,A + B hTo ,B , (7.286)


o
o
= V (A + B )(cv T (cv + R)To ) + A hTo ,A + B hTo ,B ,
(7.287)


o
o
(7.288)
= V (A + B )cv (T To ) + A (hTo ,A RTo ) + B (hTo ,B RTo ) ,

= V (A + B )cv (T To ) + A uoTo ,A + B uoTo ,B .
(7.289)

Now at the initial state, we have T = To , so




A uoTo ,A + b
B uoTo ,B .
Uo = V b

(7.290)

So, we can say our caloric equation of state is




o
o
b
b
U Uo = V (A + B )cv (T To ) + (A A )uTo ,A + (B B )uTo ,B , (7.291)


A + b
B )cv (T To ) + (A b
A )uoTo ,A + (B b
B )uoTo ,B . (7.292)
= V (b

As an aside, on a molar basis, we scale Eq. (7.292) to get

u uo = cv (T To ) + (yA yAo )uoTo ,A + (yB yBo )uoTo ,B .

(7.293)

And because we have assumed the molecular masses are the same, MA = MB , the mole
fractions are the mass fractions, and we can write on a mass basis
u uo = cv (T To ) + (cA cAo )uoTo ,A + (cB cBo )uoTo ,B .

(7.294)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

268

CHAPTER 7. KINETICS IN SOME MORE DETAIL

Returning to Eq. (7.292), our energy conservation relation, Eq. (7.277), becomes


0 = V (b
A + b
B )cv (T To ) + (A b
A )uoTo ,A + (B b
B )uoTo ,B .
(7.295)

Now we solve for T

A + b
B )cv (T To ) + (A b
A )uoTo ,A + (B b
B )uoTo ,B ,
0 = (b
b
A o
b
B o
0 = cv (T To ) + A
uTo ,A + B
uT ,B ,
b
b
A + b
B
A + b
B o
b
A uoTo ,A b
B uoTo ,B
T = To + A
+ B
.
b
b
A + b
B cv
A + b
B cv

Now we impose our assumption that b


B = 0, giving also B = b
A A ,
b
A A uoTo ,A B uoTo ,B

,
b
b
cv
A
A cv
b
A uoTo ,A uoTo ,B
= To + A
.
b
cv
A

T = To +

A :

(7.296)
(7.297)
(7.298)

(7.299)
(7.300)

In summary, realizing that hTo ,A hTo ,B = uoTo ,A uoTo ,B we can write T as a function of
T = To +

A A ) o
(b
o
(hTo ,A hTo ,B ).
b
A cv

(7.301)

o
o
A that T
We see then that if hTo ,A > hTo ,B , that as A decreases from its initial value of b
will increase. We can scale Eq. (7.301) to form
!

 

o
o
hTo ,A hTo ,B
T
A
=1+ 1
.
(7.302)
b
To
cv To
A

We also note that our caloric state equation, Eq. (7.293) can, for yAo = 1, yBo = 0 as
u uo = cv (T To ) + (yA 1)uoTo ,A + yB uoTo ,B ,
= cv (T To ) + ((1 yB ) 1)uoTo ,A + yB uoTo ,B ,
= cv (T To ) yB (uoTo ,A uoTo ,B ).

(7.303)
(7.304)
(7.305)

Similarly, on a mass basis, we can say,


u uo = cv (T To ) cB (uoTo ,A uoTo ,B ).

(7.306)

For this problem, we also have


Kc = exp
CC BY-NC-ND. 03 August 2012, J. M. Powers.

Go
RT

(7.307)

269

7.2. ADIABATIC, ISOCHORIC KINETICS


with
Go =
=
=
=

g oB g oA ,
o
o
hB T soB (hA T soA ),
o
o
(hB hA ) T (soB soA ),
o
o
(hTo ,B hTo ,A ) T (soTo ,B soTo ,A ).

So
o

Kc = exp
= exp
= exp

hTo ,A hTo ,B T (soTo ,A soTo ,B )


RT

cv To
RT

hTo ,A hTo ,B T (soTo ,A soTo ,B )


cv To

1 1
k 1 TTo

(7.308)
(7.309)
(7.310)
(7.311)

(7.312)
!!

hTo ,A hTo ,B
T (soTo ,A soTo ,B )

cv To
To
cv

(7.313)
!!

(7.314)

Here we have used the definition of the ratio of specific heats, k = cP /cv along with R =
cP cv . So we can solve Eq. (7.273) by first using Eq. (7.314) to eliminate Kc and then
Eq. (7.301) to eliminate T .
7.2.1.3

Dimensionless form

Let us try writing dimensionless variables so that our system can be written in a compact
dimensionless form. First lets take dimensionless time to be
= at.

(7.315)

Let us take dimensionless species concentration to be z with


z=

A
.
b
A

(7.316)

Let us take dimensionless temperature to be with


=

T
.
To

(7.317)

Let us take dimensionless heat release to be q with


o

hT ,A hTo ,B
q= o
.
cv To

(7.318)

Let us take dimensionless activation energy to be with


=

E
.
RTo

(7.319)
CC BY-NC-ND. 03 August 2012, J. M. Powers.

270

CHAPTER 7. KINETICS IN SOME MORE DETAIL

And let us take the dimensionless entropy change to be with


=

(soTo ,A soTo ,B )
.
cv

(7.320)

So our equations become





dz
1

z
= exp
(1 z) ,
d

Kc
= 1 + (1 z)q,


1 1
Kc = exp
(q ) .
k1

(7.321)
(7.322)
(7.323)

It is more common to consider the products. Let us define for general problems
=

B
B
=
.
b
A + B
A + b
B

(7.324)

Thus is the mass fraction of product. For our problem, b


B = 0 so
=

Thus,

B
c A
= A
.
b
b
A
A

(7.325)

= 1 z.

(7.326)

We can think of as a reaction progress variable as well. When = 0, we have = 0, and


the reaction has not begun. Thus, we get



d
1

(1 )
= exp
,
(7.327)
d

Kc
= 1 + q,
(7.328)


1 1
Kc = exp
(q ) .
(7.329)
k1
7.2.1.4

Example calculation

Let us choose some values for the dimensionless parameters:


= 20,

= 0,

q = 10,

7
k= .
5

With these choices, our kinetics equations reduce to






25
20
d
(1 ) exp
,
= exp
d
1 + 10
1 + 10
CC BY-NC-ND. 03 August 2012, J. M. Powers.

(7.330)

(0) = 0.

(7.331)

271

7.2. ADIABATIC, ISOCHORIC KINETICS

1.0

0.8

0.6

0.4

0.2

1. 106

2. 106

3. 106

4. 106

5. 106

Figure 7.12: Dimensionless plot of reaction product concentration versus time for adiabatic isochoric combustion with simple reversible kinetics.
The right side of Eq. (7.331) is at equilibrium for values of which drive it to zero.
Numerical root finding methods show this to occur at 0.920539. Near this root, Taylor
series expansion shows the dynamics are approximated by
d
( 0.920539) = 0.17993( 0.920539) + . . .
d

(7.332)

Thus the local behavior near equilibrium is given by


= 0.920539 + C exp (0.17993 ) .

(7.333)

Here C is some arbitrary constant. Clearly the equilibrium is stable, with a time constant
of 1/0.17993 = 5.55773.
Numerical solution shows the full behavior of the dimensionless species concentration
( ); see Figure 7.12. Clearly the product concentration is small for some long period of
time. At a critical time near = 2.7 106 , there is a so-called thermal explosion with a
rapid increase in . Note that the estimate of the time constant near equilibrium is orders
of magnitude less than the explosion time, 5.55773 << 2.7 106 . Thus, linear analysis
here is a poor tool to estimate an important physical quantity, the ignition time. Once the
ignition period is over, there is a rapid equilibration to the final state. The dimensionless
temperature plot is shown in Figure 7.13. The temperature plot is similar in behavior to
the species concentration plot. At early time, the temperature is cool. At a critical time,
the thermal explosion time, the temperature rapidly rises. This rapid rise, coupled with the
exponential sensitivity of reaction rate to temperature, accelerates the formation of product.
This process continues until the reverse reaction is activated to the extent it prevents further
creation of product.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

272

CHAPTER 7. KINETICS IN SOME MORE DETAIL

10

1. 106

2. 106

3. 106

4. 106

5. 106

Figure 7.13: Dimensionless plot of temperature versus time for adiabatic, isochoric
combustion with simple reversible kinetics.
7.2.1.5

High activation energy asymptotics

Let us see if we can get an analytic prediction of the thermal explosion time, 2.7 106 .
Such a prediction would be valuable to see how long a slowing reacting material might take
to ignite. Our analysis is similar to that given by Buckmaster and Ludford in their Chapter
1.12
For convenience let us restrict ourselves to = 0. In this limit, Eqs. (7.327-7.329) reduce
to




q

d
(1 ) exp
,
(7.334)
= exp
d
1 + q
(k 1)(1 + q)
with (0) = 0. The key trouble in getting an analytic solution to Eq. (7.334) is the presence
of in the denominator of an exponential term. We need to find a way to move it to the
numerator. Asymptotic methods provide one such way.
Now we recall for early time << 1. Let us assume takes the form
= 1 + 2 2 + 3 3 + . . .

(7.335)

Here we will assume 0 < << 1 and that 1 ( ) O(1), 2 ( ) O(1), . . . , and will define
in terms of physical parameters shortly. Now with this assumption, we have
1
1
=
.
2
1 + q
1 + q1 + q2 + 3 q3 + . . .
12

(7.336)

J. D. Buckmaster and G. S. S. Ludford, 1983, Lectures on Mathematical Combustion, SIAM, Philadelphia.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

273

7.2. ADIABATIC, ISOCHORIC KINETICS


Long division of the term on the right side yields the approximation
1
= = 1 q1 + 2 (q 2 21 q2 ) + . . . ,
1 + q
= 1 q1 + O(2 ).
So


exp

1 + q


exp (1 q1 + O(2 )) ,

e exp q1 + O(2 ) ,

(7.337)
(7.338)

(7.339)
(7.340)

We have moved from the denominator to the numerator of the most important exponential
term.
Now, let us take the limit of high activation energy by defining to be

1
.

(7.341)

Let us let the assume the remaining parameters, q and k are both O(1) constants. When
is large, will be small. With this definition, Eq. (7.340) becomes




e1/ exp q1 + O(2 ) .


(7.342)
exp
1 + q
With these assumptions and approximations, Eq. (7.334) can be written as


d
(1 + . . . ) = e1/ exp q1 + O(2 )
d


(1 1 . . . ) (1 + . . . ) exp

q
(k 1)(1 + q1 + . . . )



(7.343)

Now let us rescale time via


1
= e1/ .

With this transformation, the chain rule shows how derivatives transform:
d d
1 d
d
=
= 1/
.
d
d d
e d

(7.344)

(7.345)

With this transformation, Eq. (7.343) becomes



1
1 d
(1 + . . . ) = 1/ exp q1 + O(2 )
1/
e d
e


(1 1 . . . ) (1 + . . . ) exp


q
.
(k 1)(1 + q1 + . . . )
(7.346)

CC BY-NC-ND. 03 August 2012, J. M. Powers.

274

CHAPTER 7. KINETICS IN SOME MORE DETAIL

This simplifies to

d
(1 + . . . ) = exp q1 + O(2 )
d


(1 1 . . . ) (1 + . . . ) exp

q
(k 1)(1 + q1 + . . . )



(7.347)

Retaining only O(1) terms in Eq. (7.347), we get


d1
= exp (q1 ) .
d

(7.348)

This is supplemented by the initial condition 1 (0) = 0. Separating variables and solving,
we get
exp(q1 )d1 = d ,
1
exp(q1 ) = + C.
q

(7.349)
(7.350)

Applying the initial condition gives

1
exp(q(0)) = C,
q
1
= C.

(7.351)
(7.352)

So

1
1
exp(q1 ) = ,
q
q
exp(q1 ) = q + 1,


1
,
exp(q1 ) = q
q

 
1
,
q1 = ln q
q

 
1
1
1 = ln q
.
q
q

(7.353)
(7.354)
(7.355)
(7.356)
(7.357)

For q = 10, a plot of 1 ( ) is shown in Fig. 7.14. We note at a finite that 1 begins to
exhibit unbounded growth. In fact, it is obvious from Eq. (7.348) that as
1
,
q
CC BY-NC-ND. 03 August 2012, J. M. Powers.

275

7.2. ADIABATIC, ISOCHORIC KINETICS

1
1.4
1.2
1.0
0.8
0.6
0.4
0.2

0.00

0.02

0.04

0.06

0.08

0.10

0.12

Figure 7.14: 1 versus for ignition problem.


that
1 .

That is there exists a finite time for which 1 violates the assumptions of our asymptotic
theory which assumes 1 = O(1). We associate this time with the ignition time, i :
1
i = .
q

(7.358)

Let us return this to more primitive variables:


 
1
1
1
i =
exp
,

q

exp 1
i =
,
q
exp
i =
.
q

(7.359)
(7.360)
(7.361)

For our system with = 20 and q = 10, we estimate the dimensionless ignition time as
i =

exp 20
= 2.42583 106 .
(20)(10)

(7.362)

This is a surprisingly good estimate, given the complexity of the problem. Recall the numerical solution showed ignition for 2.7 106 .
CC BY-NC-ND. 03 August 2012, J. M. Powers.

276

CHAPTER 7. KINETICS IN SOME MORE DETAIL

In terms of dimensional time, ignition time prediction becomes


exp
,
aq


1 RTo
=
a
E

ti =

(7.363)
cv To
o
o
hTo ,A hTo ,B

exp

E
RTo

(7.364)

Note the ignition is suppressed if the ignition time is lengthened, which happens when
he activation energy E is increased, since the exponential sensitivity is stronger than
the algebraic sensitivity,
o

the energy of combustion (hTo ,A hTo ,B ) is decreased because it takes longer to react
to drive the temperature to a critical value to induce ignition, and
the collision frequency factor a is decreased, which suppresses reaction.

7.2.2

Detailed H2 O2 N2 kinetics

Here is an example which uses multiple reactions for an adiabatic isothermal system is given.
Consider the full time-dependency of a problem similar to the thermal explosion problem
just considered.
A closed, fixed, adiabatic volume, V = 0.3061251 cm3 , contains at t = 0 s a stoichiometric
hydrogen-air mixture of 2 105 mole of H2 , 1 105 mole of O2 , and 3.76 105 mole of
N2 at Po = 2.83230 106 P a and To = 1542.7 K.13 Thus the initial molar concentrations
are
H2 = 6.533 105 mole/cm3 ,
O2 = 3.267 105 mole/cm3 ,

H2 = 1.228 104 mole/cm3 .


The initial mass fractions are calculated via ci = Mi i /. They are
cH2 = 0.0285,
cO2 = 0.226,
cN2 = 0.745.
To avoid issues associated with numerical roundoff errors at very early time for species
with very small compositions, the minor species were initialized at a small non-zero value
near machine precision; each was assigned a value of 1015 mole. The minor species all have
i = 1.803 1016 mole/cm3 . They have correspondingly small initial mass fractions.
13

This temperature and pressure correspond to that of the same ambient mixture of H2 , O2 and N2 which
was shocked from 1.01325 105 P a, 298 K, to a value associated with a freely propagating detonation.
CC BY-NC-ND. 03 August 2012, J. M. Powers.

277

7.2. ADIABATIC, ISOCHORIC KINETICS


0

H
O
H2
O2
OH
H 2O
H 2O2
HO 2
N2

10

10
ci

10

10

15

10

10

10

10

10

t (s)

Figure 7.15: Plot of cH2 (t), cH (t), cO (t), cO2 (t), cOH (t), cH2 O (t), cHO2 (t), cH2 O2 (t), cN2 (t), for
adiabatic, isochoric combustion of a mixture of 2H2 + O2 + 3.76N2 initially at To = 1542.7 K,
Po = 2.8323 106 P a.
We seek the reaction dynamics as the system proceeds from its initial state to its final
state. We use the reversible detailed kinetics mechanism of Table 7.2. This problem requires
a detailed numerical solution. Such a solution was performed by solving the appropriate
equations for a mixture of nine interacting species: H2 , H, O, O2 , OH, H2 O, HO2 , H2 O2 ,
and N2 . The dynamics of the reaction process are reflected in Figs. 7.15-7.17.
At early time, t < 107 s, the pressure, temperature, and major reactant species concentrations (H2 , O2 , N2 ) are nearly constant. However, the minor species, e.g. OH, HO2 ,
and the major product, H2 O, are undergoing very rapid growth, albeit with math fractions
whose value remains small. In this period, the material is in what is known as the induction
period.
After a certain critical mass of minor species has accumulated, exothermic recombination
of these minor species to form the major product H2 O induces the temperature to rise, which
accelerates further the reaction rates. This is manifested in a thermal explosion. A common
definition of the end of the induction period is the induction time, t = tind , the time when
dT /dt goes through a maximum. Here one finds
tind = 6.6 107 s.

(7.365)

A close-up view of the species concentration profiles is given in Fig. 7.18.


At the end of the induction zone, there is a final relaxation to equilibrium. The equilibCC BY-NC-ND. 03 August 2012, J. M. Powers.

278

CHAPTER 7. KINETICS IN SOME MORE DETAIL

4000
3500
3000

T (K)

2500
2000
1500
1000
500
0

10

10

10
t (s)

Figure 7.16: Plot of T (t), for adiabatic, isochoric combustion of a mixture of 2H2 +O2 +3.76N2
initially at To = 1542.7 K, Po = 2.8323 106 P a.

P (Pa)

10

10

10

10

10
t (s)

Figure 7.17: Plot of P (t), for adiabatic, isochoric combustion of a mixture of 2H2 + O2 +
3.76N2 initially at To = 1542.7 K, Po = 2.8323 106 P a.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

279

7.2. ADIABATIC, ISOCHORIC KINETICS

10

ci 10

H
O
H2
O2
OH
H 2O
H 2O2

10

10

HO 2
N2
2

6
t (s)

10
7

x 10

Figure 7.18: Plot near thermal explosion time of cH2 (t), cH (t), cO (t), cO2 (t), cOH (t), cH2 O (t),
cHO2 (t), cH2 O2 (t), cN2 (t), for adiabatic, isochoric combustion of a mixture of 2H2 +O2 +3.76N2
initially at at To = 1542.7 K, Po = 2.8323 106 P a.
rium mass fractions of each species are
cO 2
cH
cOH
cO
c H2
c H2 O
cHO2
c H2 O 2
c N2

=
=
=
=
=
=
=
=
=

1.85 102 ,
5.41 104 ,
2.45 102 ,
3.88 103 ,
3.75 103 ,
2.04 101 ,
6.84 105 ,
1.04 105 ,
7.45 101 .

(7.366)
(7.367)
(7.368)
(7.369)
(7.370)
(7.371)
(7.372)
(7.373)
(7.374)

We note that because our model takes N2 to be inert that its value remains unchanged.
Other than N2 , the final products are dominated by H2 O. The equilibrium temperature is
3382.3 K and 5.53 106 P a.

CC BY-NC-ND. 03 August 2012, J. M. Powers.

280

CHAPTER 7. KINETICS IN SOME MORE DETAIL

CC BY-NC-ND. 03 August 2012, J. M. Powers.

Bibliography
M. M. Abbott and H. C. van Ness, 1989, Thermodynamics with Chemical Applications,
Second Edition, Schaums Outline Series in Engineering, McGraw-Hill, New York.
First published in 1972, this is written in the style of all the Schaums series. It has extensive solved
problems and a crisp rigorous style that is readable by undergraduate engineers. It has a chemical
engineering emphasis, but is also useful for all engineers.

P. Atkins, 2010, The Laws of Thermodynamics: a Very Short Introduction, Oxford University Press, Oxford.
This is a short book by an eminent chemist summarizing the foundations of thermodynamics for an
interested general reader.

A. Bejan, 2006, Advanced Engineering Thermodynamics, Third Edition, John Wiley, Hoboken, New Jersey.
This is an advanced undergraduate text first published in 1988. It gives a modern treatment of
the science of classical thermodynamics. It does not confine its attention to traditional engineering
problems, and considers applications across biology and earth sciences as well; some readers will
find parts of the discussion to be provocative, as received wisdom is occasionally challenged. The
thermodynamics of irreversible processes are discussed in detail.

R. S. Berry, S. A. Rice, and J. Ross, 2000, Physical Chemistry, Second Edition, Oxford
University Press, Oxford.
This is a rigorous general text in physical chemistry at a senior or first year graduate level. First
appearing in 1980, it has a full treatment of classical and statistical thermodynamics as well as
quantum mechanics.

L. Boltzmann, 1995, Lectures on Gas Theory, Dover, New York.


This is a detailed monograph by the founding father of statistical thermodynamics. This edition is a
translation of the original German version, Gastheorie, from 1896-1898.

C. Borgnakke and R. E. Sonntag, 2009, Fundamentals of Thermodynamics, Seventh Edition,


John Wiley, New York.
281

This classic and popular undergraduate mechanical engineering text, which in earlier editions was
authored by J. G. van Wylen and Sonntag, has stood the test of time and has a full treatment of
most classical problems. Its first edition appeared in 1965.

M. Born, 1949, Natural Philosophy of Cause and Chance, Clarendon Press, Oxford.
This monograph is a summary of the Waynflete Lectures delivered at Oxford University in 1948 by
the author, the winner of the 1954 Nobel Prize in physics. The lectures consider various topics, and
include an important chapter on thermodynamics with the authors earlier defense of the approach of
Carath`eodory playing a prominent role.

R. Boyle, 2003, The Sceptical Chymist, Dover, New York.


This is a reprint of the original 1661 work of the famous early figure of the scientific revolution.

P. W. Bridgman, 1943, The Nature of Thermodynamics, Harvard, Cambridge.


This short monograph by the winner of the 1946 Nobel Prize in physics gives a prosaic introduction
to thermodynamics which is directed at a scientifically literate audience who are not interested in
detailed mathematical exposition.

H. B. Callen, 1985, Thermodynamics and an Introduction to Thermostatistics, Second Edition, John Wiley, New York.
This advanced undergraduate text, an update of the 1960 original, has an emphasis on classical physics
applied to thermodynamics, with a few chapters devoted to quantum and statistical foundations.

S. Carnot, 2005, Reflections on the Motive Power of Fire, Dover, New York.
This is a translation of the authors foundational 1824 work on the heat engines, Reflexions sur la
Puissance Motrice du Feu et sur les Machines propres a
` developper cette Puissance (Reflections on
the Motive Power of Fire and on Machines Fitted to Develop that Power). Also included is a paper
of Clausius.

Y. A. C
engel and M. A. Boles, 2010, Thermodynamics: An Engineering Approach, Seventh
Edition, McGraw-Hill, Boston.
This popular undergraduate mechanical engineering text which first appeared in 1989 has most of the
features expected in a modern book intended for a large and varied audience.

S. Chandrasekhar, 2010, An Introduction to the Study of Stellar Structure, Dover, New


York.
This monograph, first published in 1939, on astrophysics is by the winner of the 1983 Nobel Prize
in physics. It has large sections devoted to rigorous axiomatic classical thermodynamics in the style
of Carath`eodory, highly accessible to engineering students, which show how thermodynamics plays a
critical role in understanding the physics of the heavens.

282

R. J. E. Clausius, 2008, The Mechanical Theory of Heat, Kessinger, Whitefish, Montana.


This is a reprint of the 1879 translation of the 1850 German publication of the great German scientist
who in many ways founded classical thermodynamics.

S. R. de Groot and P. Mazur, 1984, Non-Equilibrium Thermodynamics, Dover, New York.


This is an influential monograph, first published in 1962, which summarizes much of the work of the
famous Belgian school of thermodynamics. It is written at a graduate level and has a strong link to
fluid mechanics and chemical reactions.

E. Fermi, 1936, Thermodynamics, Dover, New York.


This short 160 page classic clearly and efficiently summarizes the fundamentals of thermodynamics.
It is based on a series of lectures given by this winner of the 1938 Nobel Prize in physics. The book
is highly recommended, and the reader can benefit from multiple readings.

R. P. Feynman, R. B. Leighton, and M. Sands, 1963, The Feynman Lectures on Physics,


Volume 1, Addison-Wesley, Reading, Massachusetts.
This famous series documents the introductory undergraduate physics lectures given at the California
Institute of Technology by the lead author, the 1965 Nobel laureate in physics. Famous for their clarity,
depth, and notable forays into challenging material, they treat a wide range of topics from classical
to modern physics. Volume 1 contains chapters relevant to classical and modern thermodynamics.

J. B. J. Fourier, 2009, The Analytical Theory of Heat, Cambridge, Cambridge.


This reprint of the 1878 English translation of the 1822 French original, Theorie Analytique de la
Chaleur is a tour de force of science, engineering, and mathematics. It predates Carnot and the
development of the first and second laws of thermodynamics, but nevertheless successfully develops
a theory of non-equilibrium thermodynamics fully consistent with the first and second laws. In so
doing, the author makes key advances in the formulation of partial differential equations and their
solution by what are now known as Fourier series.

J. W. Gibbs, 1957, The Collected Works of J. Willard Gibbs, Yale U. Press, New Haven.
This compendium gives a complete reproduction of the published work of the monumental American
engineer and scientist of the nineteenth century, including his seminal work on classical and statistical
thermodynamics.

W. T. Grandy, 2008, Entropy and the Time Evolution of Macroscopic Systems, Oxford,
Oxford.
This monograph gives an enlightening description, fully informed by both historical and modern
interpretations, of entropy and its evolution in the context of both continuum and statistical theories.

E. A. Guggenheim, 1933, Modern Thermodynamics by the Methods of Willard Gibbs, Methuen,


London.
This graduate-level monograph was important in bringing the work of Gibbs to a wider audience.

283

E. P. Gyftopoulos and G. P. Beretta, 1991, Thermodynamics: Foundations and Applications, Macmillan, New York.
This beginning graduate text has a rigorous development of classical thermodynamics.

C. S. Helrich, 2009, Modern Thermodynamics with Statistical Mechanics, Springer, Berlin.


This is an advanced undergraduate text aimed mainly at the physics community. The author includes a
full treatment of classical thermodynamics and moves easily into statistical mechanics. The authors
undergraduate training in engineering is evident in some of the style of the text which should be
readable by most undergraduate engineers after a first class in thermodynamics.

J. O. Hirschfelder, C. F. Curtis, and R. B. Bird, 1954, Molecular Theory of Gases and


Liquids, Wiley, New York.
This comprehensive tome is a valuable addition to any library of thermal science. Its wide ranging
text covers equations of state, molecular collision theory, reactive hydrodynamics, reaction kinetics,
and many other topics all from the point of view of careful physical chemistry. Much of the work
remains original.

J. R. Howell and R. O. Buckius, 1992, Fundamentals of Engineering Thermodynamics,


Second Edition, McGraw-Hill, New York.
This is a good undergraduate text for mechanical engineers; it first appeared in 1987.

W. F. Hughes and J. A. Brighton, 1999, Fluid Dynamics, Third Edition, Schaums Outline
Series, McGraw-Hill, New York.
Written in the standard student-friendly style of the Schaums series, this discussion of fluid mechanics
includes two chapters on compressible flow which bring together fluid mechanics and thermodynamics.
It first appeared in 1967.

F. P. Incropera, D. P. DeWitt, T. L. Bergman, and A. S. Lavine, 2006, Fundamentals of


Heat and Mass Transfer, Sixth Edition, John Wiley, New York.
This has evolved since its introduction in 1981 into the standard undergraduate text in heat transfer.
The interested thermodynamics student will find the subject of heat transfer builds in many ways
on a classical thermodynamics foundation. Especially relevant to thermodynamics are chapters on
boiling and condensation, heat exchangers, as well as extensive tables of thermal properties of real
materials.

J. H. Keenan, F. G. Keyes, P. G. Hill, and J. G. Moore, 1985, Steam Tables: Thermodynamic


Properties of Water Including Vapor, Liquid, and Solid Phases, John Wiley, New York.
This is a version of the original 1936 data set which is the standard for waters thermodynamic
properties.

J. Kestin, 1966, A Course in Thermodynamics, Blaisdell, Waltham, Massachusetts.


284

This is a foundational textbook that can be read on many levels. All first principles are reported in a
readable fashion. In addition the author makes a great effort to expose the underlying mathematical
foundations of thermodynamics.

R. J. Kee, M. E. Coltrin, and P. Glarborg, 2003, Chemically Reacting Flow: Theory and
Practice, John Wiley, New York.
This comprehensive text gives an introductory graduate level discussion of fluid mechanics, thermochemistry, and finite rate chemical kinetics. The focus is on low Mach number reacting flows, and
there is significant discussion of how to achieve computational solutions.

C. Kittel and H. Kroemer, 1980, Thermal Physics, Second Edition, Freeman, San Francisco.
This is a classic undergraduate text, first introduced in 1970; it is mainly aimed at physics students.
It has a good introduction to statistical thermodynamics and a short effective chapter on classical
thermodynamics. Engineers seeking to broaden their skill set for new technologies relying on microscale thermal phenomena can use this text as a starting point.

D. Kondepudi and I. Prigogine, 1998, Modern Thermodynamics: From Heat Engines to


Dissipative Structures, John Wiley, New York.
This is a detailed modern exposition which exploits the authors unique vision of thermodynamics with
both a science and engineering flavor. The authors, the second of whom is one of the few engineers
who was awarded the Nobel Prize (chemistry 1977, for the work summarized in this text), often
challenge the standard approach to teaching thermodynamics, and make the case that the approach
they advocate, with an emphasis on non-equilibrium thermodynamics, is better suited to describe
natural phenomena and practical devices than the present approach, which is generally restricted to
equilibrium states.

K. K. Kuo, 2005, Principles of Combustion, Second Edition, John Wiley, New York.
This is a readable graduate level engineering text for combustion fundamentals. First published in
1986, it includes a full treatment of reacting thermodynamics as well as discussion of links to fluid
mechanics.

K. J. Laidler, 1987, Chemical Kinetics, Third Edition, Prentice-Hall, Upper Saddle River,
NJ.
This is a standard advanced undergraduate chemistry text on the dynamics of chemical reactions. It
first appeared in 1965.

L. D. Landau and E. M. Lifshitz, 2000, Statistical Physics, Part 1, Volume 5 of the Course
of Theoretical Physics, Third Edition, Butterworth-Heinemann, Oxford.
This book, part of the monumental series of graduate level Russian physics texts, first published in
English in 1951 from Statisticheskaya fizika, gives a fine introduction to classical thermodynamics as
a prelude to its main topic, statistical thermodynamics in the spirit of Gibbs.

285

B. H. Lavenda, 1978, Thermodynamics of Irreversible Processes, John Wiley, New York.


This is a lively and opinionated monograph describing and commenting on irreversible thermodynamics. The author is especially critical of the Prigogine school of thought on entropy production rate
minimization.

A. Lavoisier, 1984, Elements of Chemistry, Dover, New York.


This is the classic treatise by the man known as the father of modern chemistry, translated from the

1789 Traite Elementaire


de Chimie, which gives the first explicit statement of mass conservation in
chemical reactions.

G. N. Lewis and M. Randall, 1961, Thermodynamics and the Free Energy of Chemical
Substances, Second Edition, McGraw-Hill, New York.
This book, first published in 1923, was for many years a standard reference text of physical chemistry.

H. W. Liepmann and A. Roshko, 2002, Elements of Gasdynamics, Dover, New York.


This is an influential text in compressible aerodynamics that is appropriate for seniors or beginning
graduate students. First published in 1957, it has a strong treatment of the physics and thermodynamics of compressible flow along with elegant and efficient text. Its treatment of both experiment
and the underlying theory is outstanding, and in many ways is representative of the approach to
engineering sciences fostered at the California Institute of Technology, the authors home institution.

J. C. Maxwell, 2001, Theory of Heat, Dover, New York.


This is a short readable book by the nineteenth century master, first published in 1871. Here the
mathematics is minimized in favor of more words of explanation.

M. J. Moran and H. N. Shapiro, 2007, Fundamentals of Engineering Thermodynamics,


Sixth Edition, John Wiley, New York.
This is a standard undergraduate engineering thermodynamics text, and one of the more popular.
First published in 1988, it has much to recommend it including good example problems, attention to
detail, good graphics, and a level of rigor appropriate for good undergraduate students.

P. M. Morse, 1969, Thermal Physics, Second Edition, Benjamin, New York.


This is a good undergraduate book on thermodynamics from a physics perspective. It covers classical
theory well in its first sections, then goes on to treat kinetic theory and statistical mechanics. It first
appeared in 1964.

I. M
uller and T. Ruggeri, 1998, Rational Extended Thermodynamics, Springer-Verlag, New
York.
This modern, erudite monograph gives a rigorous treatment of some of the key issues at the frontier
of modern continuum thermodynamics.

286

I. M
uller and W. Weiss, 2005, Entropy and Energy, Springer-Verlag, New York.
This is a unique treatise on fundamental concepts in thermodynamics. The author provide mathematical rigor, historical perspective, and examples from a diverse set of scientific fields.

I. M
uller, 2007, A History of Thermodynamics: the Doctrine of Energy and Entropy,
Springer-Verlag, Berlin.
The author gives a readable text at an advanced undergraduate level which highlights some of the
many controversies of thermodynamics, both ancient and modern.

I. M
uller and W. H. M
uller, 2009, Fundamentals of Thermodynamics and Applications:
with Historical Annotations and Many Citations from Avogadro to Zermelo, Springer,
Berlin.
The author presents an eclectic view of classical thermodynamics with much discussion of its history.
The text is aimed at a curious undergraduate who is unsatisfied with industrial-strength yet narrow
and intellectually vapid textbooks.

W. Nernst, 1969, The New Heat Theorem: Its Foundation in Theory and Experiment,
Dover, New York.
This monograph gives the authors exposition of the development of the third law of thermodynamics.
It first appeared in English translation in 1917 and was originally published in German.

W. Pauli, 2000, Thermodynamics and the Kinetic Theory of Gases, Dover, New York.
This is a monograph on statistical thermodynamics by the winner of the 1945 Nobel Prize in physics.
It is actually derived from his lecture course notes given at ETH Zurich, as compiled by a student in
his class, E. Jucker, published in 1952.

M. Planck, 1990, Treatise on Thermodynamics, Dover, New York.


This brief book, which originally appeared in German in 1897, gives many unique insights from the
great scientist who was the winner of the 1918 Nobel Prize in physics. It is rigorous, but readable by
an interested undergraduate student.

H. Poincare, 1892, Thermodynamique: Lecons Profess`ees Pendant le Premier Semestre


1888-89, Georges Carre, Paris.
This text of classical undergraduate thermodynamics has been prepared by one of the premier mathematicians of the nineteenth century.

I. Prigogine, 1967, Introduction to Thermodynamics of Irreversible Processes, Third Edition,


Interscience, New York.
This is a famous book that summarizes the essence of the work of the Belgian school for which the
author was awarded the 1977 Nobel Prize in chemistry. This book first appeared in 1955.

287

W. J. M. Rankine, 1908, A Manual of the Steam Engine and Other Prime Movers, Seventeenth Edition, Griffin, London.
This in an accessible undergraduate text for mechanical engineers of the nineteenth century. It first
appeared in 1859. It contains much practical information on a variety of devices, including steam
engines.

L. E. Reichl, 1998, A Modern Course in Statistical Physics, John Wiley, New York.
This full service graduate text has a good summary of key concepts of classical thermodynamics and
a strong development of modern statistical thermodynamics.

O. Reynolds, 1903, Papers on Mechanical and Physical Subjects, Volume III, The SubMechanics of the Universe, Cambridge, Cambridge.
This volume compiles various otherwise unpublished notes of Reynolds and includes his detailed derivations of general equations of conservation of mass, momentum, and energy employing his transport
theorem.

W. C. Reynolds, 1968, Thermodynamics, Second Edition, McGraw-Hill, New York.


This is an unusually good undergraduate text written for mechanical engineers. The author has
wonderful qualitative problems in addition to the usual topics in such texts. A good introduction
to statistical mechanics is included as well. This particular edition is highly recommended; the first
edition appeared in 1965.

S. I. Sandler, 1998, Chemical and Engineering Thermodynamics, Third Edition, John Wiley,
New York.
This is an advanced undergraduate text in thermodynamics from a chemical engineering perspective
with a good mathematical treatment. It first appeared in 1977.

E. Schrodinger, 1989, Statistical Thermodynamics, Dover, New York.


This is a short monograph written by the one of the pioneers of quantum physics, the co-winner of
the 1933 Nobel Prize in Physics. It is based on a set of lectures delivered to the Dublin Institute for
Advanced Studies in 1944, and was first published in 1946.

A. H. Shapiro, 1953, The Dynamics and Thermodynamics of Compressible Fluid Flow,


Vols. I and II, John Wiley, New York.
This classic two volume set has a comprehensive treatment of the subject of its title. It has numerous
worked example problems, and is written from a careful engineers perspective.

J. M. Smith, H. C. van Ness, and M. Abbott, 2004, Introduction to Chemical Engineering


Thermodynamics, Seventh Edition, McGraw-Hill, New York.
This is probably the most common undergraduate text in thermodynamics in chemical engineering.
It first appeared in 1959. It is rigorous and has went through many revisions.

288

A. Sommerfeld, 1956, Thermodynamics and Statistical Mechanics, Lectures on Theoretical


Physics, Vol. V, Academic Press, New York.
This is a compilation of the authors lecture notes on this subject. The book reflects the authors
stature of a leader of theoretical physics of the twentieth century who trained a generation of students
(e.g. Nobel laureates Heisenberg, Pauli, Debye and Bethe). The book gives a fine description of
classical thermodynamics with a seamless transition into quantum and statistical mechanics.

J. W. Tester and M. Modell, 1997, Thermodynamics and Its Applications, Third Edition,
Prentice Hall, Upper Saddle River, New Jersey.
First appearing in 1974, this entry level graduate text in thermodynamics is written from a chemical
engineers perspective. It has a strong mathematical development for both classical and statistical
thermodynamics.

C. A. Truesdell, 1980, The Tragicomic History of Thermodynamics, 1822-1854, Springer


Verlag, New York.
This idiosyncratic monograph has a lucid description of the history of nineteenth century thermal
science. It is written in an erudite fashion, and the reader who is willing to dive into a difficult subject
will be rewarded for diligence by gain of many new insights.

C. A. Truesdell, 1984, Rational Thermodynamics, Second Edition, Springer-Verlag, New


York.
This is an update on the evolution of classical thermodynamics in the twentieth century. The book
itself first appeared in 1969. The second edition includes additional contributions by some contemporaneous leaders of the field.

S. R. Turns, 2000, An Introduction to Combustion, Second Edition, McGraw-Hill, Boston.


This is a senior-level undergraduate text on combustion which uses many notions from thermodynamics of mixtures. It first appeared in 1996.

H. C. van Ness, 1983, Understanding Thermodynamics, Dover, New York.


This is a short readable monograph from a chemical engineering perspective. It first appeared in 1969.

W. G. Vincenti and C. H. Kruger, 1976, Introduction to Physical Gas Dynamics, Krieger,


Malabar, Florida.
This graduate text on high speed non-equilibrium flows contains a good description of the interplay
of classical and statistical mechanics. There is an emphasis on aerospace science and fundamental
engineering applications. It first appeared in 1965.

F. M. White, 2010, Fluid Mechanics, Seventh Edition, McGraw-Hill, Boston.


This standard undergraduate fluid text draws on thermodynamics in its presentation of the first law
and in its treatment of compressible flows. It first appeared in 1979.

289

B. Woodcraft, 1851, The Pneumatics of Hero of Alexandria, London, Taylor Walton and
Maberly.
This is a translation and compilation from the ancient Greek of the work of Hero (10 A.D.-70 A.D.).
The discussion contains descriptions of the engineering of a variety of technological devices including
a primitive steam engine. Other devices which convert heat into work are described as well.

L. C. Woods, 1975, The Thermodynamics of Fluid Systems, Clarendon, Oxford.


This graduate text gives a good, detailed survey of the thermodynamics of irreversible processes,
especially related to fluid systems in which convection and diffusion play important roles.

290

You might also like