You are on page 1of 111

1

s
t

Y
e
a
r

T
h
e
r
m
o
d
y
n
a
m
i
c
s

:

M
e
c
h
a
n
i
c
a
l

E
n
g
i
n
e
e
r
i
n
g


P
a
g
e

1

o
f

2

F

B
e
y
r
a
u

&

R

M
a
r
t
i
n
e
z
-
B
o
t
a
s


R
e
v
i
s
i
o
n

D
a
t
e
:

2
8
/
0
9
/
1
0


1
1
M
E

T
H
E
R
M
O
D
Y
N
A
M
I
C
S

2
0
1
0
-
1
1

b
y

F
r
a
n
k

B
e
y
r
a
u

a
n
d

R
i
c
a
r
d
o

M
a
r
t
i
n
e
z
-
B
o
t
a
s


M
e
c
h
a
n
i
c
a
l

E
n
g
i
n
e
e
r
i
n
g

D
e
p
a
r
t
m
e
n
t


I
m
p
e
r
i
a
l

C
o
l
l
e
g
e

L
o
n
d
o
n




S
e
c
t
i
o
n



C
o
n
t
e
n
t
s



L
e
c
t
u
r
e
s
,

T
u
t
o
r
i
a
l
s

a
n
d

T
e
s
t
s

S
e
c
t
i
o
n

1
:


I
n
t
r
o
d
u
c
t
i
o
n

t
o

T
h
e
r
m
o
d
y
n
a
m
i
c
s

S
e
c
t
i
o
n

2
:


B
a
s
i
c

C
o
n
c
e
p
t
s


H
i
s
t
o
r
y

l
i
n
k
e
d

t
o

p
o
w
e
r

p
l
a
n
t

d
e
v
e
l
o
p
m
e
n
t


R
e
l
e
v
a
n
c
e

i
n

t
h
e

c
o
n
t
e
x
t

o
f

e
n
e
r
g
y

r
e
s
e
r
v
e
s

a
n
d

t
h
e

e
n
v
i
r
o
n
m
e
n
t


M
a
c
r
o
s
c
o
p
i
c

v
e
r
s
u
s

m
i
c
r
o
s
c
o
p
i
c

d
e
s
c
r
i
p
t
i
o
n
s

o
f

m
a
t
t
e
r


T
h
e

m
a
c
r
o
s
c
o
p
i
c

v
i
e
w
p
o
i
n
t


S
y
s
t
e
m
s
,

s
y
s
t
e
m

s
t
a
t
e
s
,

a
n
d

c
h
a
n
g
e
s

o
f

s
t
a
t
e

v
i
a

e
n
e
r
g
e
t
i
c

i
n
t
e
r
a
c
t
i
o
n
s


S
y
s
t
e
m
s

a
n
d

c
o
n
t
r
o
l

v
o
l
u
m
e
s


P
r
o
p
e
r
t
i
e
s

a
n
d

s
t
a
t
e


E
q
u
i
l
i
b
r
i
u
m


P
r
o
c
e
s
s
e
s

a
n
d

c
y
c
l
e
s

L
e
c
t
u
r
e
:

w
e
e
k

2

T
u
t
o
r
i
a
l

n
o
.

1

&

2

S
e
c
t
i
o
n

3
:

E
n
e
r
g
y
,

H
e
a
t
,

W
o
r
k

a
n
d

t
h
e

F
i
r
s
t

L
a
w


F
o
r
m
s

o
f

e
n
e
r
g
y
:

k
i
n
e
t
i
c
,

p
o
t
e
n
t
i
a
l

a
n
d

i
n
t
e
r
n
a
l


I
n
t
e
r
a
c
t
i
o
n

w
i
t
h

t
h
e

s
u
r
r
o
u
n
d
i
n
g
s
:

H
e
a
t


I
n
t
e
r
a
c
t
i
o
n

w
i
t
h

t
h
e

s
u
r
r
o
u
n
d
i
n
g
s
:

W
o
r
k


D
i
s
p
l
a
c
e
m
e
n
t

a
n
d

s
h
a
f
t

w
o
r
k


F
i
r
s
t

L
a
w

f
o
r

a

s
y
s
t
e
m

(
c
y
c
l
i
c

a
n
d

n
o
n
-
c
y
c
l
i
c
)

L
e
c
t
u
r
e
s
:

w
e
e
k
s


4
,

5

&

6


T
u
t
o
r
i
a
l

n
o
.

3

S
t
r
u
c
t
u
r
e
d

T
u
t
o
r
i
a
l

1

o

T
i
t
l
e
:

E
n
e
r
g
y
,

h
e
a
t
,

d
i
s
p
l
a
c
e
m
e
n
t

w
o
r
k

a
n
d

t
h
e

1
s
t

L
a
w

o
f

T
h
e
r
m
o
d
y
n
a
m
i
c
s

o

D
a
t
e
:

w
e
e
k


6

S
e
c
t
i
o
n

4
:

P
r
o
p
e
r
t
i
e
s

o
f

S
u
b
s
t
a
n
c
e
s


P
u
r
e

s
u
b
s
t
a
n
c
e
s


T
w
o
-
p
r
o
p
e
r
t
y

r
u
l
e
,

s
t
a
t
e

d
i
a
g
r
a
m
s


I
n
t
e
n
s
i
v
e

a
n
d

e
x
t
e
n
s
i
v
e

p
r
o
p
e
r
t
i
e
s


I
n
t
e
r
n
a
l

e
n
e
r
g
y
,

e
n
t
h
a
l
p
y

a
n
d

s
p
e
c
i
f
i
c

h
e
a
t
s


I
d
e
a
l

a
n
d

p
e
r
f
e
c
t

g
a
s
e
s


P
h
a
s
e

c
h
a
n
g
e
,

v
a
p
o
u
r

a
n
d

l
i
q
u
i
d

p
r
o
p
e
r
t
i
e
s

L
e
c
t
u
r
e
s
:

w
e
e
k
s


7
,

8

&

9

T
u
t
o
r
i
a
l

n
o
.

4

S
t
r
u
c
t
u
r
e
d

T
u
t
o
r
i
a
l


o

T
i
t
l
e
:

P
r
o
p
e
r
t
i
e
s

o
f

s
u
b
s
t
a
n
c
e
s


1
s
t

Y
e
a
r

T
h
e
r
m
o
d
y
n
a
m
i
c
s

:

M
e
c
h
a
n
i
c
a
l

E
n
g
i
n
e
e
r
i
n
g


P
a
g
e

2

o
f

2

F

B
e
y
r
a
u

&

R

M
a
r
t
i
n
e
z
-
B
o
t
a
s


R
e
v
i
s
i
o
n

D
a
t
e
:

2
8
/
0
9
/
1
0


2
S
e
c
t
i
o
n

5
:

T
h
e

F
i
r
s
t

L
a
w

f
o
r

F
l
o
w

P
r
o
c
e
s
s
e
s


C
o
n
t
r
o
l

v
o
l
u
m
e

a
n
a
l
y
s
i
s


T
h
e

d
e
r
i
v
a
t
i
o
n

o
f

t
h
e

f
i
r
s
t

l
a
w

f
o
r

a

f
l
o
w

p
r
o
c
e
s
s
:

g
e
n
e
r
a
l


F
i
r
s
t

l
a
w

a
p
p
l
i
e
d

t
o

a

s
t
e
a
d
y

f
l
o
w

p
r
o
c
e
s
s


T
h
r
o
t
t
l
i
n
g

p
r
o
c
e
s
s


N
o
z
z
l
e
s
,

p
u
m
p
s
,

c
o
m
p
r
e
s
s
o
r
s

a
n
d

t
u
r
b
i
n
e
s


H
e
a
t

e
x
c
h
a
n
g
e
r
s

a
n
d

m
i
x
i
n
g

c
h
a
m
b
e
r
s


U
n
s
t
e
a
d
y

e
n
e
r
g
y

p
r
o
c
e
s
s
e
s

L
e
c
t
u
r
e
s
:

w
e
e
k
s

1
0
,

1
1
,

1
6

&

1
7

T
u
t
o
r
i
a
l

n
o
.

5

S
t
r
u
c
t
u
r
e
d

T
u
t
o
r
i
a
l

2

o

T
i
t
l
e
:


1
s
t

L
a
w

o
f

T
h
e
r
m
o
d
y
n
a
m
i
c
s

f
o
r

f
l
o
w

p
r
o
c
e
s
s
e
s
:

t
h
e

S
F
E
E

P
r
o
g
r
e
s
s

T
e
s
t
:

(
b
o
t
h

T
h
e
r
m
o
d
y
n
a
m
i
c
s

a
n
d

F
l
u
i
d

M
e
c
h
a
n
i
c
s

i
n

o
n
e

t
e
s
t
)

S
e
c
t
i
o
n

6
:

T
h
e

S
e
c
o
n
d

L
a
w

o
f

T
h
e
r
m
o
d
y
n
a
m
i
c
s


H
e
a
t

e
n
g
i
n
e
s
:

g
e
n
e
r
a
l

a
p
p
r
o
a
c
h


R
e
v
e
r
s
i
b
i
l
i
t
y

i
n

t
h
e
r
m
o
d
y
n
a
m
i
c
s


C
l
a
u
s
i
u
s

s
t
a
t
e
m
e
n
t

o
f

t
h
e

S
e
c
o
n
d

L
a
w


T
e
m
p
e
r
a
t
u
r
e

s
c
a
l
e


E
f
f
i
c
i
e
n
c
y

o
f

a

h
e
a
t

e
n
g
i
n
e


H
e
a
t

p
u
m
p
s

a
n
d

c
o
e
f
f
i
c
i
e
n
t

o
f

p
e
r
f
o
r
m
a
n
c
e

L
e
c
t
u
r
e
s
:

w
e
e
k
s

1
8
,

1
9
,

2
0

&

2
1

T
u
t
o
r
i
a
l

n
o
.

6

S
t
r
u
c
t
u
r
e
d

T
u
t
o
r
i
a
l

3

o

T
i
t
l
e
:

2
n
d

L
a
w

o
f

T
h
e
r
m
o
d
y
n
a
m
i
c
s
:

E
f
f
i
c
i
e
n
c
y

o
f

r
e
v
e
r
s
i
b
l
e

e
n
g
i
n
e
s


S
e
c
t
i
o
n

7
:

C
o
n
s
e
q
u
e
n
c
e
s

o
f

t
h
e

S
e
c
o
n
d

L
a
w


C
l
a
u
s
i
u
s

i
n
e
q
u
a
l
i
t
y


D
e
f
i
n
i
t
i
o
n

o
f

e
n
t
r
o
p
y


S
t
a
t
e

d
i
a
g
r
a
m
s

u
s
i
n
g

e
n
t
r
o
p
y


T
d
s

r
e
l
a
t
i
o
n
s
h
i
p
s


I
s
e
n
t
r
o
p
i
c

p
r
o
c
e
s
s
e
s

f
o
r

p
e
r
f
e
c
t

g
a
s

a
n
d

s
t
e
a
m


I
s
e
n
t
r
o
p
i
c

e
f
f
i
c
i
e
n
c
y
:

c
o
m
p
r
e
s
s
o
r
s

a
n
d

t
u
r
b
i
n
e
s


A
p
p
l
i
c
a
t
i
o
n

t
o

a

c
y
c
l
e

L
e
c
t
u
r
e
s
:

w
e
e
k
s

2
2
,

2
3
,

2
4

&

2
5

(
R
e
v
i
s
i
o
n
)

T
u
t
o
r
i
a
l

n
o
.

7

S
t
r
u
c
t
u
r
e
d

T
u
t
o
r
i
a
l

4

o

T
i
t
l
e
:

U
s
e

o
f

i
s
e
n
t
r
o
p
i
c

e
f
f
i
c
i
e
n
c
y

a
n
d

2
n
d

L
a
w

r
e
l
a
t
i
o
n
s
h
i
p
s


E
x
a
m
p
l
e

C
l
a
s
s
e
s



W
e
e
k
s

3
0
,

3
1

&

3
2
,

c
h
e
c
k

1
M

b
o
a
r
d

T
o
p
i
c
s
:

o

S
t
e
a
d
y

f
l
o
w

e
n
e
r
g
y

e
q
u
a
t
i
o
n

a
p
p
l
i
e
d

t
o

a

s
i
m
p
l
e

p
o
w
e
r

p
l
a
n
t

o

S
u
c
c
e
s
s
i
o
n

o
f

p
r
o
c
e
s
s
e
s

i
n

a

s
y
s
t
e
m

o

I
s
e
n
t
r
o
p
i
c

e
f
f
i
c
i
e
n
c
y

i
n

a

r
e
h
e
a
t

t
u
r
b
i
n
e
s

a
n
d

i
n

a
n

i
n
t
e
r
c
o
o
l
e
d

c
o
m
p
r
e
s
s
o
r

s
y
s
t
e
m

o

S
e
c
o
n
d

l
a
w

u
s
i
n
g

h
e
a
t

e
n
g
i
n
e
s

S
u
m
m
e
r

C
l
i
n
i
c

T
u
t
o
r
i
a
l
s


W
e
e
k
s

3
0
,

3
1

&

3
2
,

c
h
e
c
k

1
M

b
o
a
r
d

T
h
e
r
m
o
f
l
u
i
d
s

E
x
a
m


C
h
e
c
k

1
M

b
o
a
r
d


1ME Thermodynamics Sec. 1
1
1. INTRODUCTION TO THERMODYNAMICS

You will have some idea of the subject, as mechanical engineers see it, from the
Introduction to the 1M Thermo-Fluids course. If you have already learned some
Thermodynamics, perhaps during Chemistry courses at school, it would be wise to put
that knowledge to one side for a while and to start again as an engineer, since the
emphasis is likely to be rather different (especially when we reach the topic of entropy).

1.1 A VERY BRIEF HISTORY

If we ignore the elementary steam turbine attributed to the 2nd-century Greek
mathematician Hero of Alexandria, the beginnings of engineering thermodynamics are
to be found in the attempts of 18th century pioneers such as the English inventor
Thomas Newcomen to build improved steam engines, then used mainly for pumping
water out of mines. The aim was to consume less steam, therefore to burn less fuel, for
a given quantity of work done in raising water. James Watt, Scottish engineer (and
instrument maker to Glasgow University) is well known for his engine of 1765.
Developments such as Watt's were essentially practical in nature, with little underlying
science. The science really began with the 1824 theory of French engineer Sadi
Carnot, who tried with reasonable success to draw an analogy between the
performance of water wheels (in terms of the water reservoir levels) and of "heat
engines" (in terms of the temperature of the heat source etc.). Notable contributors to
thermodynamics theory in the 19th century include the physicist (and brewery owner)
James Joule, in demonstrating the 1st Law, and Lord Kelvin and William Rankine, both
of Glasgow University, from whose work the 2nd Law was developed. It was not until
the 20th century that the last (we believe!) misconceptions were eradicated.

1.2 THE RELEVANCE OF THERMODYNAMCS IN THE 21ST CENTURY

High standards of living have been synonymous with high per capita energy
consumption and high emissions of pollutants (particularly those responsible for acid
rain and photochemical smogs and for increasing the greenhouse effect). Virtually all
domestic activity involves the consumption of energy, directly or otherwise, while the
success of industries is often critically dependent on energy supplies and their cost.
(Note the rise of the Energy Manager over the last two decades.) The economy of an
entire nation can be heavily influenced by its primary energy sources (the fuels
available to it), the ways in which that energy is converted into useful forms (particularly
electricity) and the ways in which it is finally used. Two major problems face society
today:
(i) How can we maintain adequate and affordable supplies of energy in useful form,
for the developing and the developed world, from dwindling reserves of coal, oil
and gas ("fossil" fuels) or from alternative sources (nuclear materials, the sun,
wind, waves, tides, ...)?
(ii) How can we avoid, or learn to live with, the resulting adverse effects on our
surroundings, locally and globally?
Some recent political leaders have argued that time needs to be taken to work out
possible solutions to these problems, but your generation will not have that luxury! You
will have to solve the problems! The answers to global-scale questions such as how to
limit carbon dioxide emissions, or more local ones such as how should the U.K. replace
North Sea gas when it runs out, will depend to a large extent on the knowledge and
skills of engineers with a sound grasp of the basic principles of Thermodynamics.
1ME Thermodynamics Sec. 2
1
2. BASIC CONCEPTS


2.1 MACROSCOPIC VIEWPOINT: "CLASSICAL THERMODYNAMICS"
(&B 4
th
Ed., Sect. 1-1)

We know that the temperature of a substance is determined by the energy of the
molecules, and that the pressure in a gas or liquid is related to the number of
molecules in a given space, their mass and their mean-square speed. It would be
impossible to use such "microscopic" detail to analyse the behaviour of a quantity of
material of interest to engineers (e.g. the gas inside a cylinder of an internal
combustion engine), although there is a body of knowledge called Statistical
Thermodynamics which is based on "average" behaviour of sufficiently large numbers
of molecules; it is of more interest to chemists and physicists.

We would prefer to regard temperature as something measurable by a calibrated
instrument called a thermometer (or thermocouple) and pressure as the local force per
unit area on a surface, measurable by a pressure gauge (or transducer). At the scales
(in terms of lengths, areas, volumes) we are concerned with, the empty space between
and inside molecules is irrelevant; we prefer to regard a liquid or a gas as a
"continuum", i.e. as homogeneous matter. This "macroscopic" viewpoint forms the
basis of the only branch of Thermodynamics we shall cover, known as Classical
Thermodynamics. The continuum idea is also the foundation for your study of Fluid
Mechanics.


2.2 FRAMEWORKS FOR STUDYING ENERGY TRANSFERS:
CONTROL VOLUMES AND SYSTEMS (&B Sect. 1-3)

Just as we focus on a "particle" or a solid body when we apply Newton's laws of motion
and calculate changes in the energy, velocity, position, etc. of the body under the
influence of forces, we need an equivalent "object" on which to use the laws of
Thermodynamics to calculate energy changes, changes in temperature, pressure, etc.
Likening energy transfers to flows of money, we need to define something analogous to
a bank account, on which to draw up a balance.
If we look at the steam turbine in the
power plant diagram of the 1M
Thermodynamics & Fluid Mechanics
Introduction, we might choose the
turbine itself as our "object". Steam
flows in at one place and (with altered
properties) out at another. Energy
"flows" out as work (to drive the
electrical generator) and as heat to the
atmosphere (the casing is hot).



The turbine is a fixed region of space which, for the purpose of our "energy
accounting", we call a CONTROL VOLUME. Its surface, across which both ENERGY
(in the form of heat or work) and MATERIAL (here steam) may flow, is called a
CONTROL SURFACE. (It is possible to choose control surfaces which expand or
contract as a process takes place, but we shall not do so in this course.) You will also
use control volumes when applying the laws of Fluid Mechanics.
1ME Thermodynamics Sec. 2
2

In Thermodynamics, we are not often interested in changes in the properties of the
solids from which items of engineering plant are constructed, so here we might equally
well consider our control surface as the inside surface of the turbine casing.
In some situations, it makes more sense to focus not on a fixed region of space but on
a fixed, identifiable quantity of material. An example is the air + petrol vapour mixture
in the cylinder of a reciprocating engine, while the inlet and exhaust valves are shut and
the mixture is being compressed prior to ignition.
For energy balance purposes, we call
such a fixed quantity of material a
SYSTEM. The SYSTEM BOUNDARY is
the real or imaginary surface which
separates the system from its
SURROUNDINGS. Material clearly
cannot cross a system boundary (here
the cylinder head and walls plus the
piston face), but energy (heat or work)
may do so. In this example, the piston is
moving upwards and the system (air +
petrol vapour) is changing its shape and
volume but not its material contents or
its mass.



Note that some textbooks, including &B, also use the term "open system" to
mean control volume, and the term "closed system" to mean system.

From now until Sec. 5, we shall concentrate on the system as a framework for applying
the laws of thermodynamics, at this stage just the law of conservation of energy. A
convenient feature of a system is that the principle of conservation of mass is
automatically satisfied; we are dealing with a fixed mass. Before considering energy
and energy transfers between a system and it surroundings, we must think about
precise descriptions of the system at any point in a thermodynamic process: the ideas
of properties and state, and the equilibrium between the system and its surroundings.
This is all part of the "language" of thermodynamics; if we proceed too quickly to the
more interesting parts of the course without an adequate grasp of this language, we
shall come unstuck!


2.3 PROPERTIES AND STATE (&B Sects. 1-4, 1-5 & 1-6)

2.3.1 Describing a system and its state
To describe a system adequately, we must know (a) relevant facts about its material
nature, and (b) its STATE, i.e. its condition at a given instant. Characteristics which
determine the state are called PROPERTIES.

1ME Thermodynamics Sec. 2
3
Examples:
(i) Motion of a smooth sphere on a smooth horizontal plane (hardly
Thermodynamics!) : System: the sphere. Material nature: mass and shape
are sufficient; colour, moment of inertia etc. are not relevant. State: two
coordinates and two components of velocity form a sufficient definition;
temperature, angular velocity etc. are irrelevant.

(ii) The air in the cylinder of a diesel engine after it has been compressed and
just before the fuel is injected : System: the air (not any of the engine
parts). Material nature: its chemical composition, "air" is good enough, and
its mass, perhaps 0.001 kg. State: its volume, perhaps 50 x 10
-6

m
3
, and
its temperature, perhaps 1000 K, are sufficient. Pressure could be
calculated from these, so does not also have to be specified; we shall see
shortly that any two of the properties volume, temperature and pressure
are sufficient to fix the state.

Note that the properties of a system in a given state do not depend on the process by
which the system arrived in that state. If you were climbing Mount Everest and had
reached the summit, the only relevant property defining your state might be the altitude
you had reached (or your potential energy relative to what it was at sea level, which
depends only on altitude unless you had lost weight in the attempt). The time taken to
reach the summit and the amount of food consumed on the way would depend on the
route taken, the weather, etc.; they are not properties, but your altitude at the summit
would be the same regardless of how you got there, so it is a property.

2.3.2 Classifying properties

Thermodynamic properties (those relevant to studies of energy) can be classifed into
two types.
Those whose magnitudes are proportional to the system's mass are called
EXTENSIVE properties. Examples:
potential energy, calculated as mgz and measured in J, where m is the mass of
the system, g is the gravitational acceleration and z is the height above some
datum level;
volume, V, measured in m
3
(at fixed pressure and temperature, the volume of a
system comprising 2 kg of a certain gas is twice that of one comprising 1 kg of
the same gas)

Those whose magnitudes do not not depend on the system's mass are called
INTENSIVE properties. Examples:
pressure, P; temperature, T; density, ; the reciprocal of density, i.e. volume
per unit mass, which is called specific volume, v, and measured in m
3
/kg;
potential energy per unit mass, i.e. mgz m = gz, measured in J/kg

Extensive properties can be used in fixing the state of a system only if we know the
system's mass. In some thermodynamic analysis, particularly when we base it on a
control volume rather than a system, it is not necessary to know masses. (For
example, in an "ideal" diesel engine, where we can ignore friction and heat loss, a
1ME Thermodynamics Sec. 2
4
knowledge of the air temperatures and pressures at the beginning and at the end of the
compression stroke is sufficient to calculate the work done on the air by the piston per
unit mass of air; the actual work done is then the product of this quantity, called the
specific work, and the actual mass of air in the cylinder, which depends on the size of
the cylinder.) It is therefore safer to say that the state of a system is determined by the
values of its intensive properties. Two systems have the same state if the values of all
the relevant intensive properties in one system are the same as the corresponding
properties in the other; the two systems need not have the same mass.

2.3.3 The "Two-property" rule (called the "state postulate" in &B)

Experience shows that only a limited number of properties is necessary to fix the state
of a system. For most situations relevant to engineers, the effects of electric and
magnetic fields and of surface tension are negligible. If the effects of gravity and of
motion (of the system as a whole) can also be neglected, the system is described as a
SIMPLE COMPRESSIBLE SYSTEM, and the "two-property" rule states that:


The state of a simple compressible system is fixed completely by the
values of any two of its intensive properties, provided those two
properties are independent of each other.


"Independent" means that the value of one of those two properties is not determined by
the value of the other; so, for example, density and specific volume will not fix a state.
A common case of two properties not being independent is when the system consists
of a single substance, e.g. water, in both the liquid and vapour (steam) phases;
pressure is then determined by temperature and vice versa. (If water is boiling at
atmospheric pressure, you know its temperature is 100C.) If you know the pressure of
such a two-phase system, the other property needed to determine the state must be
something other than temperature.

The small number of properties needed to describe such a system is a feature of
Classical Thermodynamics. In a gravitational field, pressure varies with height, but in
most systems of interest to us in Thermodynamics the variation is small and we can
assign a single value of pressure to the entire system. If motion of the whole system
(not molecular motion) is important, only a few extra properties are needed: velocity
components.


2.4 EQUILIBRIUM (&B Sect. 1-5)

2.4.1 Internal equilibrium

If a system is completely isolated from its surroundings, i.e. no energy can cross the
system boundary and it cannot be influenced by any external effects, and if none of its
properties changes with time, it is said to be in INTERNAL EQUILIBRIUM.




1ME Thermodynamics Sec. 2
5
Example: gas (0.2 kg of CO
2
, say) in
one side of a rigid, thermally insulated
box, separated from an evacuated part
by a diaphragm through which nothing
can pass. ("Rigid" implies that the gas
cannot be compressed or expanded by
movement of the walls, so no work can
be done, in the absence of any stirring
device inside the box. The insulation
prevents any heat transfer across the
system boundary. If the

gas properties are measured at intervals, they will eventually become invariant with
time; the system will be in internal equilibrium. Now suppose that the diaphragm
breaks. Immediately afterwards, temperature T and pressure P will vary in time and
space throughout the box (a small number of property values will no longer suffice to
describe the instantaneous state); the system is not now in internal equilibrium.
However, a new equilibrium state (T and P uniform and constant) will be reached when
the gas fills the entire box and all motion has ceased.

2.4.2 Equilibrium between a system and its surroundings

The system (gas in a cylinder) in the left-hand diagram below is at a pressure
P
0
and and temperature T
0
greater than those of its surroundings, P
1
and T
1
.

It is in
equilibrium with the surroundings only because of the thermal insulation (the piston is
assumed to be a non-conductor) and the peg preventing outward motion of the piston.
If the insulation is removed to allow a transfer of heat and the peg is also withdrawn,
allowing the gas to expand and do work on the surrounding atmosphere, the system is
no longer in equilibrium with its surroundings.

The interaction between the system and the surroundings will continue until the causes
of the heat transfer (the temperature difference) and the work transfer (the pressure
difference) have been removed. When the system is at P
1
and T
1
, it once again in
equilibrium with the surroundings. The term "mechanical equilibrium" is used when
there is no tendency for any part of the system boundary to move, because the
pressure of the system balances the external pressure (and any forces caused by
weights, springs etc.). The term "thermal equilibrium" is used when there is no
temperature difference across the system boundary hence no reason for any heat
transfer. A combination of mechanical and thermal equilibrium (and some other types
of equilibrium which need not concern us) is referred to as "thermodynamic
equilibrium".




1ME Thermodynamics Sec. 2
6
2.5 PROCESSES (&B Sect. 1-6)

2.5.1 Definition

When a system changes from one state of equilibrium (internally and with its
surroundings) to another, we say that it has undergone a PROCESS. The succession
of states through which the system passes during the change from the INITIAL STATE
to the FINAL STATE is called the process PATH. These intermediate states are not
equilibrium states, otherwise there would be no reason for the process to continue; a
process can only take place when the equilibrium between system and surroundings is
removed. To define a process completely, the initial and final states, the path and the
energy transfers between system and surroundings should be specified.

2.5.2 Quasi-equilibrium processes

The Thermodynamics of systems which are not in equilibrium is a difficult subject, not
appropriate to an undergraduate course. How, then, can we deal with processes in
engineering equipment, e.g. the changes in state of air inside the cylinder of a
reciprocating engine, or of steam as it flows through a turbine?

Consider whan would happen, during the air compression process in a diesel engine
cylinder, if the piston were suddenly stopped. Just before that happened. the air would
be in a non-equilibrium state, with temperature, pressure and velocity varying
throughout the cylinder. How long would it take, from the instant the piston stopped, for
the system (the air) to settle down to a state of equilibrium, with all properties uniform


and constant*? If that "settling time" were small compared with the time taken for any
property to change noticeably as a result of piston movement (before we stopped it),
we could say that the system had been nearly in equilibrium, at that point in the
process. How nearly would depend on the piston speed, i.e. on the nature of the
process, not of the system. In a real diesel engine (even a high-speed one running at,
say, 4000 rev/min) "information" about the pressure increase, caused by the piston
advancing into the air, is transmitted to all parts of the air much faster than the speed of
the piston itself, and the air would indeed be nearly in equilibrium throughout this
compression process.

A process which takes place slowly enough for the system to deviate only
infinitesimally from equilibrium (internally and with its surroundings) is called a QUASI-
EQUILIBRIUM (or quasi-static) process. Fortunately, real processes in much
engineering equipment of interest to us can be considered to be quasi-equilibrium
processes.


2.5.3 Diagrams of process paths

The Two-Property Rule allows us to plot process paths on a two-dimensional diagram,
as a graph of one property against any other independent property. For a system,
pressure and specific volume are easily-measured properties and are commonly used
to illustrate quasi-equilibrium process paths. Example for the compression process in a
diesel engine cylinder.

Remember: "uniform" means that a property does not vary in space (at a given time);
"constant" and "steady" mean that it does not vary in time (at a given position).
1ME Thermodynamics Sec. 2
7

The initial state, or state 1, is the point marked by the number 1 (circled to remind us
that it does not represent a quantity, such as a pressure of 1 bar). The specific volume
in state 1 is v
1
and the pressure is P
1
. The final state is state 2. The left-hand P - v
diagram is for a non-equilibrium process in which the intermediate states are
undefined; only the initial and final states, where the system is in equilibrium, can be
plotted. The right-hand P - v diagram shows a quasi-equilibrium process, in which all
the intermediate states can be considered to be equilibrium states, and can therefore
be plotted as points; the continuous line joining these states represents the process
path.



2.5.4 Cyclic Processes, or Cycles

The sequence of processes taking place in a steadily-operating reciprocating engine
repeats itself hundreds or thousands of times a minute. An idealised model of the
processes in a diesel engine cylinder, for example, is as follows, where we take the
system as air alone (neglecting the small additional mass of the injected fuel).

Process 1 - 2 (meaning from state 1 to state 2): compression of the air as the piston
moves from its outermost (or "bottom dead centre") position to its innermost (or "top
dead centre") position.

Process 2 - 3: expansion of the air while there is a heat transfer to it (modelling the
effect of the fuel combustion in a real engine), for part of the piston's return stroke,
while the pressure stays constant. (In practice, the rise in pressure which would occur if
combustion took place in a fixed-volume vessel is approximately counterbalanced by
the reduction in pressure which would occur due to piston movement in the absence of
combustion.)

Process 3 - 4: continued expansion of the air to the maximum volume (piston's
outermost position) while the pressure falls, modelling that part of the return stroke
after combustion has been completed

Process 4 - 1: instantaneous* reduction of pressure to its value in state 1 (* no piston
movement therefore no change in volume). This models the removal of the combustion
products from the cylinder and their replacement by a fresh charge of air; from a
thermodynamic point of view, it is irrelevant that this occurs in a four-stroke engine by
1ME Thermodynamics Sec. 2
8
an inward stroke of the piston while the exhaust valve is open, followed by an outward
stroke while the inlet valve is open.

A sequence of processes like this,
where the final state of the last process
is identical to the initial state of the first
process, is called a CYCLIC PROCESS
or simply a CYCLE.

P - v diagrams of cycles are particularly
useful for evaluating the net power
output of an engine and its efficiency,
meaning the net power divided by the
rate of energy supply via the fuel. (With
the idealised models in this course, we
consider not the energy available from
burning the fuel, a 2nd-year topic, but
the equivalent rate of heat transfer to
our system, a fixed mass of air alone.)

























1ME Thermodynamics Ex. 1&2
1
EXERCISES 1 & 2
Introduction & Basic Concepts

1. In power plant terminology, "station efficiency" means the net electrical power output
of a power station divided by the theoretical rate of input of energy in the fuel. This
energy input rate is the product of the mass of fuel consumed in unit time and the
calorific value, or heating value, of the fuel. Calorific value is the energy theoretically
converted from one form, chemical energy, to another, internal energy (which will be
defined in Sec. 2) of the gaseous products, when unit mass of the fuel is completely
burned.

(a) A large coal-fired power station with a maximum electrical output of 1000 MW
may have a station efficiency of 35%. If the coal has a calorific value of 22000 kJ/kg,
how many tonnes of coal are consumed per day when such a plant is running
continuously (i.e. for 24 h) at its maximum rating?

(b) Typical U.K. coal has a sulphur content of 1.6% (by mass). Assuming that all of
this is converted to sulphur dioxide (SO
2
) during combustion, what mass of SO
2
will
be emitted to the atmosphere per day by the power station in part (a)?
(Approximate relative atomic masses S: 32, O: 16)

To comply with the European Community's 1988 Directive on emission of sulphur
dioxide and other pollutants from large power plant, the U.K. electricity generating
industry has opted for a combination of fitting expensive flue gas desulphurisation
(FGD) equipment to the newest of its coal-fired plants and replacing much of the
home-produced coal with alternative fuels. At present, this means either imported
coal with lower sulphur content (posing transport and handling problems, also
"cheating" because the SO
2
emission limits applied to the U.K. took account of our
high-sulphur coal!) or natural gas (virtually no sulphur). Gas is favoured by the
privatised generation industry. New gas-fired combined-cycle plant is much cheaper
and quicker to build than new coal-fired plant (although the running costs are not
necessarily lower than those of existing coal-fired plant without FGD).

(Do not spend too long on this. "Don't know" is a valid answer at present, but
discuss with tutors.)


2. If a balloon is inflated by connecting it to a bottle of compressed gas, what would be
the most appropriate "framework", system or control volume, for thermodynamic
calculations on (a) the bottle, (b) the bottle's valve, (c) the bottle plus the balloon?
State the reason in each case.









1ME Thermodynamics Ex. 1&2
2
3 Two mountaineers set out from base camp, A, to reach the same higher-altitude
camp, B. (This is analogous to a substance undergoing processes.) One takes a
short but steep route, the other a longer route with a more gradual slope. When
either of them has reached B , their final state, which of the following can be
regarded as properties defining that state?

(a) the latitude and longitude
(b) the altitude
(c) the distance travelled (along the mountain surface) during the climb

One climber now returns to A. What word could be used to describe the process of
ascending to B then descending to A,
(d) using the same route in each direction?
(e) descending by the route which the other climber used to reach B?
(f) If an avalanche had moved camp A down the mountain while the climbers were
at camp B, would that word still be a valid description of the sequence of processes
old A-B + B-new A?


4 A tank, with a leak-proof partition dividing it into two sections, has one section initially
filled with a liquid (the system), as shown on the left.



(a) The partition is suddenly removed (so that a process takes place). After
some time, the liquid settles down to a new level, as shown on the right.

i. Is the system in equilibrium in its initial state and in its final state?
ii. On the centre diagram, sketch (very roughly) the liquid surface in some
intermediate state.
iii. Can this process be described as a quasi-equilibrium process? (Give
the reason.)

(b) From the same initial state, a different process occurs if , instead of removing
the partition, it is left in place and a small hole is drilled in it so that liquid can
flow slowly through to the other section. Sketch the liquid surface at an
intermediate state in this new process, and state, with a reason, whether this
is a quasi-equilibrium process.









1ME Thermodynamics Ex. 1&2
3

5 The four sketches below represent different ways in which a gas (the system,
shown shaded) in a cylinder can undergo an expansion process from an initial
equilibrium state. In a, a diaphragm, which initially separated the gas from an
evacuated section of the cylinder, has burst, allowing the gas to expand rapidly to fill
the entire cylinder. In b, the gas is separated from the evacuated section by a
lightweight piston attached to the end of the cylinder by a compression spring;
initially, the spring was not compressed, and the piston was held in place by a catch;
the expansion was caused by releasing the catch. In c, a heavy weight was placed
on the piston, whose upper surface is now open to the atmosphere; initially the
upward force on the piston due to the gas pressure just balanced the downward
force due to atmospheric pressure plus the weight, and the expansion was caused
by heating the gas. In d, a connecting rod links the piston to a rotating crankshaft, as
in a reciprocating engine, and the expansion was again caused by heating the gas.


For each case, state, with reasons, whether the process is, is not, or could be a
quasi-equilibrium process. If the answer is "could be", say what conditions would
have to be satisfied for it to be a quasi-equilibrium process.


6 The valve on a compresed air bottle, containing air at an initial pressure of 130 bar,
is opened slightly to discharge air slowly to the atmosphere (where the pressure is
approximately 1 bar) until the pressure in the bottle has fallen to 80 bar. (Numbers
are for illustration; no calculations are required.)

(a) Could any of the following usefully be defined as a system when attempting
to describe this process, and if not, why? (i) the air initially in the bottle (ii) the air
which escapes (iii) the air which remains in the bottle.

(b) Could the air which is still in the bottle at the end of the process be said to
have undergone a quasi-equilibrium process? Why, and under what conditions?

(c) Could the same be said for the air which escapes?









1ME Thermodynamics Ex. 1&2
4
_____________________________________________________________________

ANSWERS (Complete solutions in some cases)

1. (a) 11200 tonnes (b) 359 tonnes (each to 3 significant figures)

2. (a) Control volume, because gas crosses the boundary (or "control surface"), at one
place (or "port")
(b) Control volume, because gas crosses the control surface, at two ports (one
inflow, one outflow)
(c) System, because no gas crosses the boundary

3. (a) are both properties (at camp B the climbers have only one longitude and one
latitude), so is (b) (assuming they remain on the surface, their altitude is uniquely
determined by their longitude and latitude, rather like the value of any property of a
simple compressible substance being fixed by the values of any other two
independent properties). (c) is not a property, since it depends on the route between
A and B (analogous to energy transfers, heat or work, between a system and its
surroundings).
(d) and (e) Cycle, because the sequence of processes begins and ends at A in
both cases, regardless of the process paths (the routes). (f) No, because if at least
one of the properties at camp A changes, e.g. latitude, A no longer represents the
same state as before.

4. (a) (i) Yes, in both states. No property changes with time and only a single value of
the most relevant property, the height of the liquid surface, is needed to describe the
system.
(ii) No. In an intermediate state, surface height changes in a complicated way
with position across the tank; a single value cannot be assigned to the system as a
whole. Also, if the process were stopped by replacing the partition, the heights on
each side would continue to oscillate for some finite time before reaching
equilibrium.

(b) Yes. At any instant during the process, only two properties would be needed to
describe the state (the virtually uniform height on each side of the partition). Also, if
the process were stopped by blocking the hole, the two heights would stop changing
almost immediately; intermediate states would be almost equilibrium states).

5. (a) Not q-e. Unlike a sudden expansion of gas into the atmosphere, the system
boundary can always be defined, as the cylinder's inside surface (since "adding a
vacuum" to the system does not change its definition as a a fixed, identifiable
quantity of matter). However, in intermediate states the pressure (and other
properties) varies throughout the gas. This is called an unresisted expansion.

(b) Not q-e. Immediately after the catch is released to start the process, the spring
exerts no force on the piston, so the expansion begins rapidly. As the expansion
proceeds and the spring is compressed, the spring force increases from zero; this is
a partially-resisted expansion.

(c) Could be q-e if the heating is sufficiently slow, so that the gas pressure does not
vary in space; it would then be a fully-resisted expansion. (In this particular
1ME Thermodynamics Ex. 1&2
5
example, the pressure would also be constant throughout the process because the
forces on the piston would balance at all times:- gas pressure x piston area =
atmospheric pressure x piston area + weight of the "weight")

(d) Could be q-e; similar to (c), with the effect of the weight replaced by the force
exerted on the piston by the connecting rod. (Unlike (c), no reason why this should
be a constant-pressure process.)

6. (a)
(i) No. The boundary (except for that part formed by the bottle's inside
surface) cannot be defined in the final state.
(ii) No. No part of the boundary can be defined in the final state.
(iii) Yes. The boundary is always the bottle's inside surface. (If
unconvinced, take a microscopic view; at any time during the
process, molecules of this gas can be found anywhere in the bottle;
initially, molecules which will eventually escape can also be found
anywhere in the bottle, so in the initial state only, we might regard
the gas as two systems which share the same volume.)

(b) Yes, provided the process takes place slowly so that the air remaining
in the bottle has time to redistribute itself, keeping the pressure (and other
properties) uniform throughout the bottle.

1ME Thermodynamics Sec. 3
1
3. ENERGY, HEAT, WORK AND THE FIRST LAW

3.1 FORMS OF ENERGY (&B 4
th
Ed, Sect. 1-7)

You should be familiar with the concepts of kinetic and potential energy for solid
bodies. If an entire thermodynamic system of mass m has a velocity magnitude C in
some frame of reference, it too has a property KINETIC ENERGY (K.E.) of mC
2
/2 ; e.g. if
we could identify and keep track of a particular 4 kg of steam as it flowed out of a
steam turbine at 300 m/s, that system would have a K.E. of 4 300
2
/ 2 = 180000 J =
180 kJ. (The kilojoule is the most conveniently-sized energy unit for most
thermodynamics applications, but you are advised for the time being to carry out
calculations using the fundamental unit, the joule.) A system has a POTENTIAL ENERGY
(P.E.) of mgz if its centre of mass is at a height z above some datum level (see Sec.
2.3.2); e.g. 20 kg of feed water entering the boiler of a steam plant at 4 m above floor
level has a P.E. of 20 9.81 4 = 785 J relative to that datum. (Identifying the systems
in these examples and defining their system boundaries might be difficult; we shall see
later that the control volume is a more appropriate framework for handling such cases.)

The energy associated with the molecular motion and inter-molecular forces in a
substance (i.e. on the microscopic scale) is important in thermodynamics, since large
changes can occur when a system changes state. This form of energy is related more
to temperature than to the other properties of a system, and is the property called
INTERNAL ENERGY (I.E.), with symbol U.

Neglecting the energies associated with magnetic, electric and surface tension effects,
the TOTAL ENERGY E of a system is then given by

E = K.E. + P.E. + I.E. = mC
2
/2 + mgz + U . ... (3-1)

The specific total energy e (an intensive property) is the sum of the specific K.E, P.E.
and I.E. :

e = E/m = C
2
/2 + gz + U/m = C
2
/2 + gz + u .

Note that we use small (lower-case) letters as the symbols for specific properties,
expressed per unit mass, and capital (upper-case) letters for the extensive properties,
proportional to the mass of the system. (Specific K.E. and P.E. do not have single
symbols to represent them.) You must make sure that your handwriting distinguishes
between u and U, for example!

You will soon see that thermodynamics is only concerned with changes in the various
forms of energy, not with their absolute values. The datum for any form of energy, the
state at which we define the energy to be zero, is arbitrary.

For situations where the system is an appropriate framework for handling problems,
e.g. gas in a piston-cylinder arrangement, changes in K.E. and P.E. of the system are
very often zero or insignificant compared with changes in I.E. In such cases, where a
system undergoes a process from state 1 to state 2, we can write:

change in total energy = E = E
2
- E
1
U
2
- U
1
= m (u
2
- u
1
) .

1ME Thermodynamics Sec. 3
2
We next have to consider how a system's energy can change, i.e. how energy can
cross the system boundary. Such energy transfers are classified in Sec. 3.2.

3.2 HEAT TRANSFER (&B Sect. 3-1)

If a system changes state as a result of a temperature difference between system and
surroundings, the transfer of energy across the system boundary is called a heat
transfer.

Note that in thermodynamic usage,
"heat" is not a property of a system. A
system has a certain energy E
1
at the
start of a process, the energy Q which
then crosses the boundary may be
called heat during the process, and at
the end of the process the system has a
new value, E
2
, of the property energy.
The sign convention is that an energy
transfer from the surroundings to the
system is positive, and vice versa.

A process in which no heat transfer
occurs is called an adiabatic process.
("Adiabatic" has no other meaning.)
In the 1st year course, we shall not concern ourselves with the detailed mechanisms of
heat transfer (conduction, convection and radiation); a 2nd year course is devoted to
these.


3.3 WORK TRANSFER (&B Sects. 3-2, 3-3 & 3-4)

In Mechanics, work is usually defined as the product of a force and the distance moved
in the direction of the force. In Thermodynamics, we need a wider definition; two
roughly equivalent versions are as follows.

(a) A work transfer is any transfer of energy between a system and its surroundings
which is not the result of a temperature difference.

(b) An energy transfer from a system to its surroundings is a work transfer if
the only effect on the surroundings could have been the rise of a weight.
Example: a process where a gas expands in a vertical cylinder

The word "only" is used because although a heat transfer can also cause the
rise of a weight, this can never be the only effect on the surroundings; this will
1ME Thermodynamics Sec. 3
3
become clear when we deal with the 2nd Law of Thermodynamics. The words
"could have been" are necessary to cover a case such as a process taking place
in which an electric current crosses the system boundary; the current could have
powered a "perfect" electric motor (no "I
2
R" heating from the windings) which
rotated a drum which wound up a rope to which a weight was attached. The
case of a rotating shaft crossing the system boundary is also covered by this
wording.

The symbol for a work transfer is W and we shall use the mechanical engineer's
traditional sign convention that W is positive for a transfer from system to surroundings
(or work done by the system on the surroundings). This is the opposite of the
convention for heat transfer, arising from the subject's origins in steam engines; the
early engineers wanted work from their engines and paid for fuel to provide heat to the
water/steam in the engine, so it was natural to regard both these quantities as positive.
Joule's famous experiments in Manchester in the 1840s showed that work and heat are
equivalent forms of energy transfer. We now have to consider two important types of
work transfer: displacement and shaft work.


3.3.1 Displacement Work

Suppose that a gas in a vertical cylinder is at a pressure P greater than the pressure
P
a
of the surrounding atmosphere, and is maintained in equilibrium by a large number
of small weights placed on top of the piston (assuming the piston itself to be
weightless).



Now remove one of the weights. The piston will rise by a small amount and the gas
pressure will fall by a small amount, until system (the gas) and surroundings reach a
new equilibrium state. During this process, the difference between the upward force on
the piston, due to gas pressure, and the downward force, due to the weights, will only
differ very slightly from zero; the expansion of the gas is then said to be fully-resisted.
We can consider this to be a quasi-equilibrium process. Note that the work done by the
system all occurs at that part of the system boundary which moves, hence the terms
"boundary displacement work" or "moving boundary work", as &B call it, or just
"displacement work".
The amount of work done in this process is the product of the force exerted on the
moving boundary (the piston face) by the gas, P A, and the distance moved by the
boundary, z. In the limit of infinitesimally small weights, z becomes an infinitesimally
small distance z and the work done is dW = P A dz. Now A dz is the volume dV
through which the boundary moves, so

dW = P dV . ... (3-2)

1ME Thermodynamics Sec. 3
4
If we now continue to remove the small weights one by one, the gas will undergo a
finite quasi-equilibrium process, and the work done will be obtained by integrating eqn
3-2:

W P V
V
V
d =

1
2
for a quasi-equilibrium process ... (3-3)

If the expansion were not fully resisted, e.g. if we removed several of the small weights
at once, the process would not be a quasi-equilibrium (q-e) one, and W P dV .

Interpreting eqn 3-3 graphically, displacement work W is the area under the process
path on a P V diagram. To evaluate it numerically, we have to know the relationship
between P and V.



The left-hand diagram might represent the case above, where we remove weights from
the piston. The right-hand diagram might represent the air compression process in a
diesel engine.




3.3.2 Shaft work

Displacement work is concerned with part of a system boundary moving in the direction
of a normal (perpendicular) force, the pressure force. Work can also be done if a
boundary moves in the direction of a tangential or shearing force; this is called shear
work. Simple example: a vessel is filled with liquid, and its lid, in contact with the liquid,
is pushed so that it slides across the top of the vessel; shear work is done on the liquid,
related to the shear stress in the liquid at the boundary (the lid), the contact area
between lid and liquid, and the distance moved by the lid. This idea will be covered in
more depth in Fluid Mechanics.

1ME Thermodynamics Sec. 3
5


Shear work transfer to a fluid commonly occurs during stirring processes, as in the left-
hand diagram above. If the system is defined as the fluid alone, with boundary A
(centre diagram), the paddle surfaces form part of the boundary, where both normal
and shear forces are exerted on the fluid; displacement and shear work transfer occur
together, in a way which is difficult to evaluate. However, if we redefine the system to
include the paddle and the immersed part of its driving shaft, with boundary B (right-
hand diagram), the only part of this boundary which is moving relative to the fluid is the
shaft cross-section. Only shear forces are now involved. Work transfer as a result of
shear forces in a rotating shaft is known as shaft work, with the symbol W
sh
. This is
the most common means of specifying work transfers in engineering plant. In the
combined-cycle plant shown in the Introduction to Thermo. & Fluid Mech. notes, shaft
work occurs to the air in the gas turbine engine's compressor, and from the combustion
products or steam in the two turbines. Shaft power, i.e. rate of shaft work transfer, can
be measured as

&
W
sh
= torque x angular velocity

e.g. in the case of a single shaft crossing a boundary, transmitting a torque of 2.1 x 10
6

N m while rotating at a steady 3000 rev/min (note that "r.p.m." is not an approved
abbreviation), the shaft power = 2.1 x 10
6
N m x 2 x 3000/60 rad/s =
660 x 10
6
Nm/s ( J/s W) = 660 MW, the output of a typical steam turbine in a
large power station. We usually have to consider shaft work in control volume analyses
(Sec 5), rather than system analyses.


3.3.3 Work done in a cycle

We can evaluate the net work done by a system in a cycle by summing (with the
correct signs) the values of W for each of the processes making up the cycle, i.e.

W
net
= W = W
12
+ W
23
+ W
34
+ . . . .

where W
23
(pronounced "W two three", not "W twenty-three") means the work done in
process 2 3 (i.e. from state 2 to state 3), etc. Because a system returns to its initial
state on completion of a cycle, and therefore to its initial volume V
1
(and specific
volume v
1
), the volume must decrease during some of the processes in the cycle if it
increases during others. W in some processes is therefore negative and in others
positive, so the net work in the cycle may be negative, zero or positive.
1ME Thermodynamics Sec. 3
6

In the simple 2-process cycle sketched above, W
12
is positive (V increases) and W
21

is negative and of smaller magnitude. W
net
= W
12
+ W
21
is therefore positive, and
we can see that the magnitude of W
net
is the difference between the shaded areas in
the left-hand and the centre P V diagrams, i.e.


| W
net
| for a cycle is the area enclosed in the P V diagram.

Considering just 1 kg of the substance forming a system of total mass m, the net
specific work |W
net
/m| is the area enclosed in the P v diagram. (Remember specific
volume v = V/m.)

It should be obvious (by thinking of the areas) that if the process paths proceed in a
clockwise direction round a cycle, W
net
will be positive. This is true for the idealised
model of a diesel engine cycle, the P V diagram sketched in Sec. 2.5.4. (It would not
be an engine if we did not get positive net work from it !) Cycles where the process
paths go anti-clockwise involve net work transfer to the system; they occur in
refrigerators and heat pumps, devices intended to produce heat transfers; we shall not
consider them until Sec. 6, and then only briefly.

3.4 FIRST LAW OF THERMODYNAMICS (&B Sects. 4-1 & 4-2)

On the macroscopic scale, it is a fundamental law of nature (proposed by Joule in the
1840s) that energy is conserved; it cannot be created or destroyed. If this is applied to
a thermodynamic system undergoing a process from state 1 to state 2, we can write


Q
12
W
12
= E
2
E
1
for process 1 2 ... (3-4)


Q
12
represents the algebraic sum of all the heat transfers during the process (+
ve
to
the system and -
ve
from it) and W
12
represents the algebraic sum of all the work
transfers during the process (-
ve
to the system and +
ve
from it).
1ME Thermodynamics Sec. 3
7
Q
12
is the net energy added to the system by heat transfer(s) and W
12
is the net
energy added by work transfer(s). These together increase the system's total energy
from E
1
to E
2
. Usually, we only have to deal with a single Q and/or a single W. For a
system defined only as a quantity of fluid (excluding any structure or machinery), it does
not matter where on the boundary these energy transfers occur, or when they occur. It
also makes no difference whether or not it is a q-e process.

In a cycle, the system's state, and therefore its total energy E, is the same at the start
and end of the cycle. The change of energy of the system must then be zero, i.e.


Q W = 0 for a cycle ... (3-5)


where Q and W mean the algebraic sums (correct signs required!) of all the heat
transfers and work transfers, respectively, round the complete cycle. In other words,

(Q
12
+ Q
23
+ . . .) (W
12
+ W
23
+ . . .) = 0.

Later, in Sec. 5, we shall see how the First Law applies to control volumes, and derive
an equation to use in place of eqn 3-4.


Some very simple examples:

(a) Concerning a single process:
Gas (the system) in a cylinder is compressed by a piston, the magnitude of the work
transfer being 30 kJ. If the internal energy of the gas increases by 20 kJ during this
process, what heat transfer takes place?

Eqn 3-4 applies. We are not concerned with changes of kinetic or potential
energy of the system as a whole, so eqn 3-4 simplifies to
Q
12
W
12
= U
2
U
1

where U is the system's internal energy. Compression means reduction of
volume, so the displacement work (= PdV) is negative, i.e. W = -30 kJ.
Q
12
= U
2
U
1
+ W = 20 + (30) = 10 kJ, the minus sign indicating
heat transfer from the gas to the surroundings.

(b) Concerning a cycle:
Water, filling a closed, rigid, insulated tank, is stirred for 100 s by a paddle wheel
which is driven by a 1 kW motor. Then the insulation is removed, allowing a heat
transfer from the water until it returns to its initial state. What is the magnitude of the
heat transfer Q
21
in the second of the two processes?

Eqn 3-5 applies. In the first process, there is no heat transfer (assuming the
insulation to be perfect), i.e. Q
12
= 0, while W
12
= 1 x 100 = 100 kJ. In
the second process, there is no work transfer (no shaft work because the
paddle is not operating, and no displacement work because there is no change
of volume), i.e. W
21
= 0. For a cycle of two processes, eqn 3-5 becomes
(Q
12
+ Q
21
) - (W
12
+ W
21
) = 0,
1ME Thermodynamics Sec. 3
8
Q
21
= W
12
+ W
21
Q
12
= 100 + 0 0 = 100 kJ,
i.e. the magnitude of Q
21
is 100 kJ.

These two problems could be done in your head, using simple ideas of energy
conservation, but for more complex problems it is wise to write down the appropriate
version of the First Law equation, using the correct symbols for heat, work and internal
energy with subscripts which identify the particular process.

1ME Thermodynamics Sec. 3
9
Example
Diesel engine performance calculation using idealised model of the cycle
The model:

System: taken as air throughout
Processes:
12 Compression of air
23 Heat addition to air at const
pressure (modelling fuel
injection and combustion)
34 Expansion of air
41 Heat rejection from air at
const. volume (modelling
expulsion of combustion
gases then intake of fresh air)




Given data:
Air inlet conditions: pressure P
1
= 1 bar, temperature T
1
= 15C
(From the two-property rule, state 1 is fixed by these property values)
Compression ratio V
1
/ V
2
= v
1
/ v
2
= 17
Maximum temperature T
3
= 2500 K

Assumptions:
Air behaves as a perfect gas, meaning that
P V = R T and u = C
v
T where R and C
v
are constants
with the following values: R = 287 J/kg K, C
v
= 717 J/kg K
(This will be covered in detail in Sec. 4)
The compression and expansion processes are adiabatic and are
described by the process-path relation P V
1.4
= constant
(The reason for this choice of process-path relation will become clear in Sec.
7.4)

Find:
the air properties P, v and T in each of the states 1, 2, 3 and 4
the heat transfer Q and displacement work transfer W in each of the
processes 12, 23, 34 and 41
the net work in the cycle
the thermal efficiency of the cycle, defined as (net work) / (heat input)
(Note carefully that net work means the sum of all the work transfers,
each with its correct sign, but heat input means the sum of those heat
transfers which have positive values)
1ME Thermodynamics Ex. 3
1
1M Thermodynamic EXERCISES 3
Energy, Heat, Work and the 1st Law

1. Consider a system comprising 2 kg of water in a lake above a waterfall; in this state,
state 1, it is flowing sufficiently slowly that its kinetic energy can be neglected. It falls
down the waterfall, a vertical distance of 50 m; immediately before it enters the pool at
the bottom of the waterfall, it is in state 2. A little way downstream from the bottom of
the waterfall, after all the turbulent motion has died down (and assuming we can still
identify it), it is in state 3. Assume that there are no energy transfers
(heat or work) between the system and its surroundings.
(a) Evaluate the changes in potential energy PE, kinetic energy KE, internal
energy U and total energy E between state 1 and state 2. (Remember that "change"
always means final value minus initial value, not the other way round.)

(b) Repeat part (a) for the changes between state 1 and state 3.

(c) Assume that the specific internal energy u of water is a function only of its
absolute temperature T and is given by u = 4190 T where u is the change in u,
in J/kg, and T is the change in T, in K. (This is quite a good approximation for
temperatures in the relevant range.) Evaluate the change in temperature of the
system between state 1 and state 3.


2. (a) A gas inside a rigid, insulated vessel is stirred by a paddle wheel (or fan). The
paddle wheel is driven by allowing a weight to fall, unwinding a rope from a drum
attached to the shaft. State what form of energy transfer, to or from the system, takes
place
(i) if the system is defined as the gas plus the paddle wheel and that part of the
shaft whichis inside the vessel.

(ii) if the system is defined as the gas plus the paddle wheel, shaft, drum, rope
and weight.


(b) A gas inside a rigid, insulated vessel is heated by an electric resistance heater.
Ignoring the wires, state what form of energy transfer, to or from the system, takes
place
(i) if the system is defined as the gas.

(ii) if the system is defined as the gas plus the resistance heater.

(iii) if the system is defined as the gas plus the resistance heater plus the battery.



1ME Thermodynamics Ex. 3
2
3. A balloon made from light, flexible material, initially collapsed, is connected via a valve,
initially closed, to a rigid storage bottle containing a gas at pressure 60 bar. The valve
is partly opened and gas flows slowly into the balloon until its volume is 0.25 m
3
.
Atmospheric pressure is 1.013 bar (1 bar = 10
5
N/m
2
).
For a system defined as the gas initially in the bottle, and taking note of the words in
italics above:

(a) sketch the position of the system boundary at the start and end of this
process;

(b) calculate the work transfer across that part of the system boundary consisting
of the inner surface of the balloon (N.B. not all the data above are relevant to this
calculation);

(c) repeat part (b) for that part of the system boundary consisting of the inner
surface of the bottle.


4. A gas in a cylinder, with a leak-proof, frictionless piston, expands slowly from state 1 to
state 2. (Leak-proof means that we can always define the system; frictionless and
slowly mean a quasi-equilibrium process.) The cause of the process does not need to
be known for this question.

For each of the following types of process, (a) to (f), sketch the process path on a P
V diagram, shade the area representing the magnitude of the work transfer, then
derive an expresion (simplified as far as possible) for the displacement work done by
the gas.

(a) A constant-pressure process.

(b) A constant-volume process.

(c) A process where gas pressure varies linearly with piston displacement.

(d) A process obeying the relationship PV
n
= constant, known as a "polytropic"
process, where n is a constant (determined by the nature of the process) greater
than 1, say 1.2.

(e) Another process as in part (d), but where n = 1.

(f) An adiabatic process in which the temperature changes from T
1
to T
2
, the
mass of gas is m and its specific internal energy is related to its temperature by u
= C
v
T, where C
v
is a constant and x denotes any change in property x. (Is it
possible in this case to evaluate W in the same way as in parts (a) to (e)? If not, try
using the 1st Law.)







1ME Thermodynamics Ex. 3
3
5. An idealised model of the cycle on which a spark-ignition (petrol or gasoline) four-
stroke reciprocating engine works is as follows, based on a system comprising air only.

Process Real SI engine Idealised model
1 2
Compression
stroke
Compression of
air + petrol vapour mixture
Adiabatic compression of
air
P
1
V
1
n
= P
2
V
2
n

2 3 Rapid pressure rise after
spark ignites mixture
Heat transfer to air in a
constant-volume process
3 4
Power stroke
Expansion of
combustion products
Adiabatic expansion of air
P
3
V
3
n
= P
4
V
4
n

4 1 Exhaust stroke :
exhaust valve opens, gases
expelled at P = approx.
P
atmos.
followed by
Induction stroke :-
exhaust valve closes, inlet
valve opens, cylinder fills
with new air + petrol vapour
mixture at P=approx.
P
atmos.

Heat transfer from air in a
constant-volume process

(If the exhaust and
induction strokes are
modelled as constant-
pressure processes, at
atmos. pressure, they
have no net effect;
one is the reverse of the
other.)




The diagrams above represent the real four-stroke spark-ignition engine, the arrows
on the left-hand sketch showing the directions of piston motion and crankshaft
rotation.
(The P V diagram could be plotted on an oscilloscope if the engine was equipped
with a pressure transducer in one cylinder and a displacement transducer giving a
signal proportional to instantaneous volume in the cylinder; it is easier in practice to
provide a signal proportional to angle of rotation of the crankshaft. Such diagrams are
known as "indicator diagrams".)
(a) Sketch the idealised model of the cycle on a P V diagram.
1ME Thermodynamics Ex. 3
4

(b) Derive expressions for the work transfer in each of the four processes making
up the idealised cycle, i.e. W
12
, W
23
, W
34
and W
41
, and hence for the net work
W
net
.

(c) The relevant properties of air can be assumed to be related by the "equation
of state" PV = mRT, where m is the mass of air in the cylinder, T is its absolute
temperature and R, the "gas constant" for air, equals 287 J/kg K. (More about this in
Sec. 4.) For both the compression and the expansion process, n may be taken as
1.4. To model a particular engine, take the maximum volume V
1
as 1.6 litre, the
compression ratio V
1
/V
2
as 9 and the maximum temperature T
3
as 3000C. P
1
may
be taken as 1 bar and T
1
as 15C.
For the idealised model of the cycle, calculate
(i) the mass of air in the cylinder,
(ii) the values of V, then P, then T, in each of states 2, 3 and 4,
(iii) the work transfer in each of the four processes and the net work in one
cycle.

(d) If the net work per thermodynamic cycle, calculated in part (c)(iii), applied to a
real four-stroke engine running at a speed of 5000 rev/min, what would be the engine's
power output (in kW)? Why would a real engine with the same cylinder capacity,
compression ratio and maximum temperature as in part (c) have a power output
rather less than this?

(e) A simple, commonly-used measure of engine performance is the "mean
effective pressure" P
me
, defined as the net work per cycle divided by the displaced or
swept volume, i.e. W
net
/ V
s
or W
net
/(V
1
- V
2
) (see the P V diagram above).
Interpret P
me
graphically and evaluate it for the conditions in part (c).

6. A vertical cylinder of cross-sectional area 0.1 m
2
, with its lower end closed and upper
end open, is fitted with a heavy piston, of mass m
p
= 203.9 kg. Initially, the piston is
supported (e.g. on a ring projecting inwards a small distance from the cylindrical wall)
such that there is a volume V
1
of 0.05 m
3
between its lower face and the closed end
of the cylinder. This volume is occupied by a mass m of a gas, initially at a pressure
P
1
of 1 bar and temperature T
1
of 26.8C. Atmospheric pressure P
at
, acting on the
piston's upper face, is also 1 bar.

The gas can be modelled as a "perfect gas". You will see in Sec. 4 of the course that
this means its properties are related by Pv = RT (or PV = mRT ) (equation of state)
and u = C
v
T (or U = m C
v
T ) (internal energy vs temperature is a straight
line). For this gas, R = 0.290 kJ/kg K and C
v
= 0.725 kJ/kg K. (Note that this R is not
the same as the universal gas constant, R . If the system mass m represents n
kilomoles of the substance, then nR is the same as mR. You are probably familiar with
the equation of state in the form PV = nR T, but engineers prefer PV = mRT, since
mass is usually a much more useful quantity than number of kilomoles.)

1ME Thermodynamics Ex. 3
5
A heat transfer to the gas takes place, until the piston has risen through a distance of
200 mm.The process occurs sufficiently slowly and the effect of friction is sufficiently
small that the system comprising the gas undergoes a quasi-equilibrium process.

(a) Do three sketches of the cylinder and piston:- at the beginning of the process
(gas in state 1), at the point where the piston is just about to rise off its support
(gas in state 2), and at the end of the process (gas in state 3). Write on, or beside,
each sketch the known values of the gas properties. (This may seem like a waste
of time, but it is good practice and can later save time in searching through the
wording of a question to look for given data.)

(b) Sketch the process path on a P V (or P v ) diagram.

(c) Use the equation of state in state 1 to show that m = 0.0575 kg (to 3 sig. fig.)

(d) Considering the forces on the piston, show that P
2
= 1.20 bar. What is P
3
?

(e) Write down (in symbols) the 1st Law equation for the gas system in process
12, with the aim of finding Q
12
, the heat transfer necessary to begin lifting the
piston. Satisfy yourself that the unknowns are W
12
and T
2
, and that W
12
= 0.
Use the equation of state in states 1 and 2 (or just in state 2) to show that T
2
=
360 K, and hence that Q
12
= 2.50 kJ.

(f) Write down (in symbols) the 1st Law equation for the gas system in process 2
3, with the aim of finding Q
23
, the heat transfer while the piston is moving. Satisfy
yourself that the unknowns are W
23
and T
3
. Show that the displacement work
W
23
= 2.40 kJ. Use the equation of state in states 2 and 3 (or just in state 3) to
show that T
3
= 504 K, and hence that Q
23
= 8.50 kJ.


7. A system undergoes the following cycle of quasi-equilibrium processes.

Process 1 2: Constant volume, with heat transfer of 170 kJ to the system

Process 2 3: Constant pressure, with heat transfer of 100 kJ from the
system and work transfer of 42 kJ to the system

Process 3 1: Adiabatic

(a) Sketch the cycle on a P V diagram. (If you cannot decide on the direction of
process path 1 2 because the properties of the substance forming the system are
not given, it is safe to assume that internal energy and pressure increase with
temperature at constant volume.)

(b) Evaluate the heat and work transfers in process 3 1.







1ME Thermodynamics Ex. 3
6
______________________________________________________________________


ANSWERS

1. (a) PE
2
PE
1
= 981 J; KE
2
KE
1
= +981 J; U
2
U
1
= 0; E
2
E
1
= 0

(b) PE
3
PE
1
= 981 J; KE
3
KE
1
= 0; U
3
U
1
= +981 J; E
3
E
1
= 0

2. (a) (i) Shaft work, negative (ii) None (Within the system, the weight loses
potential energy and the gas gains internal energy, but no energy crosses the system
boundary.)
(b) (i) Heat, positive
(ii) Electrical work (see notes, Sec. 3.3), negative
(iii) None (Within the system, the battery loses chemical energy and the gas gains
internal energy, but again no energy crosses the system boundary.)

3. Light, flexible material implies that the gas pressure inside the expanded balloon
needs to be only very slightly greater than atmospheric pressure in order to stretch the
material and support its weight. Hence the process can be approximated as a
constant-pressure process with gas pressure equal to atmospheric pressure, the word
slowly implying quasi-equilibrium. The initial pressure of the gas in the bottle is
irrelevant here; it falls almost to atmospheric pressure as it flows through the valve. So
the displacement work W = P dV which in this case becomes P dV = P (V
2
- V
1
) =
1.013 x 10
5
x 0.25 J = 25.3 kJ (positive; work done by gas on surrounding
atmosphere).

4. (a) P
1
(V
2
- V
1
)
(b) 0 (No piston movement, therefore no displacement work)
(c)

(P
1
+ P
2
) (V
2
- V
1
)
(d) (P
2
V
2
- P
1
V
1
) / (1 - n )
(e) P
1
V
1
ln (V
2
/ V
1
)

5. (c) (i) m = 1.94 x 10
-3
kg
(ii) V
2
= 1.778 x 10
-4
m
3
, P
2
= 21.7 bar, T
2
= 693.6 K
V
3
= 1.778 x 10
-4
m
3
, P
3
= 102.3 bar, T
3
= 3273.2 K
V
4
= 1.600 x 10
-3
m
3
, P
4
= 4.72 bar, T
4
= 1359.2 K
(iii) W
12
= -563.2 J, W
23
= 0, W
34
= +2659.2 J, W
41
= 0. W
net
= W = 2096.0
J

(d) 87.3 kW

(e) 14.74 bar

6. (b) 1 2 is a constant volume process, 2 3 is a constant-pressure process
(d) P
3
= 1.20 bar

7. Q
31
= 0 (given as adiabatic), W
31
= 112 kJ


1ME Thermodynamics Struc. Tut. 1
1
1M THERMODYNAMICS
PROBLEM FOR STRUCTURED TUTORIAL
Energy, heat, displacement work and the 1st Law of Thermodynamics
______________________________________________________________________

This problem does not rely on material from Sec. 4 (Properties of Substances), although
some basic ideas on gas properties are introduced.

Consider a rigid cylinder of cross-sectional area A = 0.05 m
2
, open to the atmosphere at
one end and fitted with a lightweight (i.e. negligible mass), leakproof, frictionless piston.
The piston is linked by a spring to a fixed point beyond the open end of the cylinder.
The space between the piston and the closed end is occupied by a gas, initially at a
pressure P
1
of 8 bar (absolute). Atmospheric pressure P
atm
can be taken as 1 bar.
The spring, which has a stiffness or spring constant k of 25 kN/m, is initially in
compression such that gas is in equilibrium with its surroundings.
Suppose the gas is now heated slowly so that it expands, doing displacement work on
the surroundings, until the piston's displacement x is 0.2 m.



What must be the spring force F
1
in the initial equilibrium state?

Force balance on piston:
so F
1
=

Does the system defined as the gas undergo a quasi-equilibrium (q-e) process?



Can we calculate the displacement work transfer from the gas using W = P dV ?




To do this integral, we need the relationship between P and gas volume V. What is it?

Consider the equilibrium of forces acting on the piston at any instant:
PA (on left face) = (on right face)
So P =
Now V =
1ME Thermodynamics Struc. Tut. 1
2






Suppose V
1
= 0.01 m
3
. What are B, C, and the final volume V
2
of the gas?

B =
C =

V
2
=


What is the final pressure, P
2
?

P
2
=


Now sketch the P V diagram for this process.











Now calculate W.













1ME Thermodynamics Struc. Tut. 1
3
How can we find the heat transfer Q in this process?

Use





(Warning: don't confuse U and u, etc.; take care with handwriting)


For simple compressible substances in general, u could be found if we knew the values
of any two other independent intensive properties, such as P and v (v = V/m).
Suppose that u for this gas depends on the value of only one other property,
temperature T. This is a very good approximation for gases under a wide range of
conditions. Let us assume that u = C
v
T where C
v
is a constant, still a good
approximation provided that T is not too large. (More of this in lectures, Sec. 4; C
v
is
called the "specific heat at constant volume"; its use is not restricted to processes where
the volume of the gas is constant.)
So Q = W + m C
v
(T
2
T
1
) (for this substance in this process).
Suppose that C
v
for this gas is 625 J/kgK, and that T
1
is 400 K. All we need now are
values for m and T
2
. For these, we need to know a relationship between the properties,
which applies in any given state, called an "equation of state".
A very good approximation for gases under a wide range of conditions is P v = R T,
where R is another constant for the particular gas (not what what you may know as the
universal gas constant; more of this in lectures, Sec. 4.)
Suppose R is 250 J/kgK for this gas. Writing v = V/m, the equation of state becomes
P V = m R T.
What is the mass of gas, m ?

Use
m =

What is the final temperature of the gas, T
2
?

Use
T
2
=
(or,



1ME Thermodynamics Struc. Tut. 1
4
So what is Q ?

Q =



Check that the sign of Q is correct: the calculation gave a positive result, which is OK
because the heat transfer was from the surroundings to the gas.

Of the energy which was transferred to the gas by heating, about three quarters has
gone to increase the internal energy of the gas and about one quarter has been
transferred from the gas to become energy stored in the compressed spring.
1ME Thermodynamics Struc. Tut. 1
5
A supplementary problem, which you will probably not have time to go
through in detail during the tutorial, but should read through afterwards:

Suppose the cylinder of the previous problem is closed at both ends, with the gas again
initially at P
1
= 8 bar (absolute) and the space on the other side of the piston completely
evacuated. The piston is initially prevented from moving by a peg, and a spring is
chosen such that there is initially no tension or compression.



The peg is now removed. Suppose the piston moves through a distance of 0.2 m before
it is stopped and held, e.g. by a ratchet mechanism. The gas then settles to a new
equilibrium state.
Is this a quasi-equilibrium (q-e) process, and if not, why?
No. As soon as the peg is removed, a finite pressure difference exists across the
piston, so the expansion is rapid; it is likely that a single value of pressure (or
other property) is inadequate to describe the state of the gas at any instant. The
expansion is initially unresisted.
Would it still be a non-q-e process if the spring were initially compressed so that it
exerted a force of 10 kN on the piston?
In state 1, gas would exert force P
1
A = 8 x 10
5
x 0.05 N = 40 kN on piston, so
piston would still have a large force imbalance; expansion would initially be partly-
resisted. Answer is therefore Yes.
What about an initial spring force of 40 kN?
Would now be a q-e process, because fully - resisted
(but no process would occur without some cause other than removing the peg.)
What about a case with no spring but a piston having significant mass ?
Resistance would now be caused by the inertia of the piston;
how closely the process approached q-e would depend on the piston's mass.
Going back to the case of the initially uncompressed spring and the lightweight piston,
can we calculate the work transfer to or from the gas using W = P dV ?
No, because it is not a q-e process;
at any instant during the process, P is not uniform throughout the gas.
Is there a work transfer?
Yes. Defining our system as the gas, there is energy in the compressed spring,
part of the system's surroundings, at the end of the process.
Would there have been a work transfer in the absence of the spring?
No, because there would never have been any resistance.
(You don't have to do work to "push on a vacuum"!)
How can W be found?
By concentrating on the surroundings; need to find the work done on the spring.
The spring force F increases linearly with piston displacement x. Here, F is initially
zero, so it reaches 5 kN (i.e. 0 + kx ) after 0.2 m where the piston is stopped. To
find the work W
spring
done on the spring: Sketch, or imagine, the graph of F
against x.
1ME Thermodynamics Struc. Tut. 1
6
|W
spring
| = F dx = ( 0 + 5000 ) x 0.2 = 500 J or 0.5 kJ
So W
spring
= 0.5 kJ (note sign convention; work transfer to spring, therefore
ve
value)
So what is W for the gas?
|W | = |W
spring
| = 0.5 kJ, so W = 0.5 kJ (intuitive way; work transfer from
gas, must be +
ve
) or W = W
spring
= ( 0.5 kJ ) = 0.5 kJ (strict algebraic
way; + sign appears automatically)
(Choose which way suits you, but don't mix them.)

1ME Thermodynamics Sec. 4
1
4. PROPERTIES OF SUBSTANCES

4.1 INTRODUCTION
To predict the heat and work transfers in processes and cycles, we need to know the
relevant properties of the system in any equilibrium state, or, for properties like internal
energy which have an arbitrary datum, the change during the process. For example,
when applying the 1st Law to the air in a diesel engine cylinder during the compression
stroke, the change in specific internal energy u
2
u
1
needs to expressed in terms of
other properties which are known in advance (P
1
, v
1
and v
2
) or found from a
knowledge of the process path (P
2
or T
2
).
Detailed knowledge of properties is also needed in order to select a suitable working
fluid for a plant based on a particular thermodynamic cycle (although considerations of
availability, cost, safety etc. are also very important). For example, what makes
water/steam the best choice of working fluid for a conventional power plant consisting of
boiler (heat in), turbine (work out), condenser (heat out) and pump (work in)? Why do
refrigerators use substances such as ammonia or hydrofluorocarbon compounds as their
working fluid, rather than water?
In 1M, attention will be confined to pure substances. By the end of Section 4, we shall be
able to relate the properties of gases such as air with sufficient accuracy to model their
behaviour in some important types of process occurring in power plant and in other
engineering applications. For liquids/vapours such as water/steam, tables and charts are
needed in general. The 1M treatment will avoid table reading/interpolating and detailed
analysis of steam power plant; these will be left until the start of the 2M course.

4.2 PURE SUBSTANCES (&B 4th ed., Sect. 2-2)
A substance whose chemical composition is the same throughout, and does not change
with time, is called a pure substance. Examples:
o nitrogen gas
o liquid nitrogen
o air (several gases, but uniformly mixed)
o air/petrol vapour mixture
o products from burning (provided chemical reaction has ceased and all constituents
are a fuel in air uniformly mixed)
o water/steam mixture (chemically the same throughout, whatever the physical
condition, e.g. water droplets in steam, vapour bubbles in water, water at the
bottom of a vessel with steam above it, ...)

The following examples are not pure substances:
o mixture of gaseous (constituents have different boiling points; liquid is and liquid
air richer in nitrogen)
o air/petrol droplets (distinct regions with different chemical composition)

4.3 DIAGRAMS OF STATE (&B Sects, 2-4 & 2-5)
A phase of a substance is identified by a particular configuration of the molecules: solid
(molecules held in fixed positions by attractive forces), liquid (molecules still quite closely
spaced but no longer fixed), vapour or gas (widely spaced molecules in random motion).
Processes where a substance changes phase are of great importance in engineering
thermodynamics, e.g. the refrigerant evaporating to extract "latent heat" from the food
1ME Thermodynamics Sec. 4
2
compartment of a freezer, the steam partially condensing in the low-pressure turbine of a
steam power plant.
Phase changes are conveniently depicted on property diagrams such as the following
pressure-temperature diagram (for a substance like water which expands on freezing).
The triple point (0.00612 bar, 0.01C for
ice/water/steam) is the only state where
solid, liquid and vapour phases can co-
exist in equilibrium. Below the triple-point
pressure, solids change phase directly to
vapour (sublimation) when heated, e.g.
solid carbon dioxide, "dry ice", for which
the triple-point pressure is above
atmospheric. The critical point is a state
where liquid which is just about to
evaporate and vapour which is just about
to condense become identical, and
supercritical fluid is the high-pressure
continuation of the liquid and vapour
phases where the two phases are
indistinguishable; they will be discussed further in Sec. 4.6.
From now on, diagrams will be simplified since we shall not be concerned with the
thermodynamics of the solid phase or with solid-liquid or solid-vapour phase changes.
All equilibrium states of a simple compressible system comprising a pure substance are
represented by a point on a surface plotted in pressure specific volume temperature
coordinates.












Below the critical pressure (P at the critical point), there is a region where liquid and
vapour are in equilibrium. For water/steam, the state of the contents of a kettle when it is
boiling lies somewhere in this region. Sec. 4.6 will introduce more terminology to
describe this region, with important applications to the boiler and condenser of a steam
power plant, for example.
Some substances, e.g. air, are normally encountered in the vapour phase in states far
away from the liquid+vapour (or two-phase) region, at temperatures well above the
PvT
equilibrium
surface for a
pure
substance

(excluding
solid region)
1ME Thermodynamics Sec. 4
3
critical temperature. They are then normally called gases (or "permanent" gases) rather
than vapours, and relationships between their properties can be greatly simplified.
We shall next consider the properties of gases, and return to vapours and liquids in
Sec. 4.6.

4.4 MODELLING THE BEHAVIOUR OF GASES (&B Sect. 2-6)

There is a variety of empirically-based equations which relate P, v and T in a given state.
They are known as equations of state. The simplest of these, the ideal-gas equation of
state, can be derived theoretically (without recourse to measurements) if the following
assumptions are made about the behaviour of the molecules of a gas:
o no momentum loss during collisions with container walls (molecules ~ rigid, elastic
spheres)
reasonable since pressure in a sealed container does not fall with time;
o molecules have negligible volume
reasonable since gases are highly compressible;
o attractive force between adjacent molecules is negligible
reasonable since gases expand to fill the entire volume available.
The kinetic theory of gases then yields the equation with which you are probably familiar:
P V = n R T where n is the number of moles occupying volume V at pressure P
and temperature T, and R is a constant, the same for all gases,
called the universal gas constant and equal to 8.314 kJ/kmol K. (A kilomole, kmol, is the
molar mass expressed in kg, not grams SI, not cgs, units.)
This is in accordance with the relationships derived experimentally in England by Robert
Boyle in 1662 (PV = const. for fixed T) and in France by J. Charles in the early 1800s
(T/V = const. for fixed P ). Engineers are normally more interested in the mass of a
system than in the number of moles, so we will now derive two equivalent but more
useful forms of this equation of state.
Multiplying and dividing by the molar mass M (the mass in kg of one kmol of the gas,
which equals the mass in g of one gram-mole),
P V = (M n) ( R / M) T = m R T since system mass m = M n.
R = R / M is known as the gas constant for the particular gas. Dont confuse it with R .
e.g. R for nitrogen is 8.314 / 28.01 = 0.297 kJ/kg K.
Dividing both sides by m to remove the extensive properties so that the equation applies
to one kg rather than the whole system,
P (V / m) = P v = R T .
The assumptions and theory leading to this equation also imply that specific internal
energy u depends on only one other property, temperature, as Joule demonstrated
experimentally in 1843.




1ME Thermodynamics Sec. 4
4
It is sufficient to specify this model of gas properties as follows (remember it !) :


IDEAL GAS MODEL P v = R T or P V = m R T ... (4-1)
and
u = some function of T only ... (4-2)


It can be applied with good accuracy to model the behaviour of air and other common
gases over the ranges of pressure and temperature relevant to most engineering
equipment of interest to us.
In 1M at least, we shall not need any greater sophistication, although &B describe
several more-complex equations of state starting with that of van der Waals:
( ) T R b v
v
a
P
2
=

+ where non-zero a accounts for inter-molecular attraction and non-


zero b for molecular volume. Expressions for u as a function of P as well as T would also
be needed.
If the processes of interest to us cover a sufficiently small range of temperature that we
can approximate the ideal-gas u(T) relationship by a linear equation in T, i.e. a straight
line fit, we have the perfect gas version of the ideal gas model. The now-constant slope
of the u(T) curve allows us to write
u
2
u
1
= const. x (T
2
T
1
)
for the change in specific internal energy in process 12, needed to obtain Q W from
the 1st Law equation. We shall use this simplification for almost all calculations on gas
systems in 1M. More details of the perfect gas model will be given after defining specific
heats in Sec. 4.5. Note that a perfect gas is an ideal gas; u is still a function of T only,
but a particular function.

4.5 ENTHALPY AND SPECIFIC HEATS (&B Sects. 2-11)
Consider a process in which there is heat transfer to a gas (the system) at constant
pressure, e.g. in a vertical cylinder closed by a piston, with a weight on the piston to
maintain the pressure at the initial value P
1
.

The 1st Law (for any process with negligible K.E. and P.E. change) is
Q W = U
2
U
1
.
In a q-e process, W = P dV and if P is constant, this becomes
1ME Thermodynamics Sec. 4
5
W

1
2
d
V
V
V P = P
1
(V
2
V
1
)

= P
2
V
2
P
1
V
1
since P
2
= P
1
.

Q = U
2
U
1
+ P
2
V
2
P
1
V
1

= (U
2
+ P
2
V
2
) (U
1
+ P
1
V
1
) .
So the heat transfer in a constant-pressure process equals the change in U + PV. Now
U, P and V are all properties of the system, so U + PV is also a property. It frequently
appears in thermodynamics (especially when, in Sec. 5, we write the 1st Law for a
control volume rather than a system), so to save space and time it has its own name
ENTHALPY and symbol H.


The enthalpy of a system of mass m is defined as H = U + P V . Dividing each
term by m, the specific enthalpy of the substance is h = u + P v . ... (4-3)


Since P v = R T for an ideal gas, h for an ideal gas also equals u + R T , a function
of T only.

The two properties C
v
, specific heat at constant volume, and C
p
, specific heat at
constant pressure, can now be defined formally. The definitions seem to have little to do
with heat or pressure or volume; they are confusingly named, but we shall see shortly
how those names arose.
For any substance, where in general u and h are functions of two other properties,

p v
T
h
C
T
u
C

p v
, .
You probably don't yet recognize the mathematical notation (partial derivatives) but don't
worry. We shall only make use of specific heats when using the ideal gas model, in
which u and h depend on T only. C
v
and C
p
are then simply the slopes of the u(T) and
h(T) curves, respectively:


For an ideal gas, C
u
T
C
h
T
v p

d
d

d
d
= = , ... (4-4)

In a process between states 1 and 2, the changes in specific internal energy and specific
enthalpy of an ideal gas are therefore

=
2
1
d
v 1 2
T
T
T C u u ,

=
2
1
d
p 1 2
T
T
T C h h , ... (4-5)
C
v
and C
p
being functions of T which need to be known before the integrals can be
evaluated.

1ME Thermodynamics Sec. 4
6
In the perfect gas model, u(T) is a straight-line approximation so
d
d
u
T
= C
v
= constant.
The expression for u
2
u
1
in eqns 4-5 then becomes very simple, as follows.


If an ideal gas is also perfect (constant C
v
), u
2
u
1
= C
v
(T
2
T
1
) ... (4-6)




In 1M, we shall almost always model gases as perfect, so that eqn 4-6 applies. Suitable
constant values for C
v
and C
p
are given for common gases in Table E5, together with M
and R.
N.B. &B do not use the term "perfect gas". Instead, they describe this approximation as
"using average specific heats" (e.g. C
v,av
).

Since h u + P v , we can write for an ideal gas:

d
d

d
d
h u R T
h
T
u
T
R = =


For an ideal gas, C
p
C
v
= R ... (4-7)


In the perfect gas model, where C
v
is constant, it follows from eqn 4-7 that C
p
is also
constant.
Hence the expression for h
2
h
1
in eqns 4-5 can also be simplified for perfect gases:


If an ideal gas is also perfect (constant C
p
), h
2
h
1
= C
p
(T
2
T
1
) ... (4-8)


1ME Thermodynamics Sec. 4
7
The ratio of the two specific heats,
v
p
C
C
, has usually been given the symbol although in
C&B call it k.

By combining the definition of with eqn 4-7, we can express C
v
and C
p
in terms of R
and , for any ideal gas (perfect or not).
If any two of C
v
, C
p
, R and are known, the other two can be found from C
p
C
v
= R
and
C
p
/ C
v
= . However, do not expect the values in Table E5 to be exactly consistent with
these relationships; the perfect gas values listed there are approximations for some
chosen temperature range and are deliberately rounded.

How C
v
and C
p
got their names and how not to be fooled
by them
If a system undergoes a constant-volume process from state 1 to
state 2 (e.g. heating the substance in a rigid vessel), the
displacement work P dV is zero. If there is no shaft work (e.g.
no stirring), the 1st law
Q W = U
2
U
1
= m ( u
2
u
1
)
becomes
Q / m = u
2
u
1
.
If the substance is a gas which can be modelled as a perfect gas, then
u
2
u
1
becomes C
v
( T
2
T
1
) (eqn 4-6).
So C
v
= Q / [ m ( T
2
T
1
) ], the heat transfer per unit mass per unit temperature
change in a constant-volume process, assuming constant C
v
and no shaft work, hence
the name specific heat at constant volume.
Enthalpy was introduced at the start of Sec. 4.5 by considering a constant-pressure
process, for which Q / m = (u
2
+ P
2
v
2
) (u
1
+ P
1
v
1
) = h
2
h
1
.
For a perfect gas, h
2
h
1
becomes C
p
( T
2
T
1
) (eqn 4-8).
So C
p
= Q / [ m ( T
2
T
1
) ], the heat transfer per unit mass per unit temperature
change in a constant-pressure process, assuming constant C
p
and no shaft work, hence
the name specific heat at constant pressure.

What is important to remember is that eqn 4-6 and 4-8 are valid for any process with a
perfect gas. Use them to relate the change in internal energy or enthalpy to the change
in temperature, regardless of whether v and P change during the process.

1ME Thermodynamics Sec. 4
8
EXAMPLE

In a diesel engine cylinder, the air at the beginning of the compression stroke has
pressure P
1
= 1 bar, temperature T
1
= 26.8C and volume V
1
= 0.001 m
3
.

The compression process, from state 1 to state 2, may be assumed to be fully resisted
and adiabatic (slow enough for intermediate states to be considered as equilibrium
states yet fast enough for heat transfer to the cylinder walls to be negligible). The
engine's compression ratio V
1
/ V
2
= 20. Air may be taken to behave as a perfect gas
with the properties given in Table E5.

(a) By considering an infinitesimally small part of the process, during which the air
volume changes from V to V + dV (dV having a negative value) and the air
temperature changes from T to T + dT, show that the equation of the process path
is
T V
1
= constant or P V

= constant .

(b) Calculate the mass of air in the cylinder, the temperature T
2
and pressure P
2
at the
end of the compression stroke, and the work transfer W in the process.




SOLUTION

(a) Next page

(b) To be done in lecture






1ME Thermodynamics Sec. 4
9
4. Solution to Example on page 8



1ME Thermodynamics Sec. 4
10
4.6 THE BEHAVIOUR OF VAPOURS AND LIQUIDS (&B Sects. 2-3 &2-4)
To describe the relevant regions of the PvT equilibrium surface and introduce some
new terminology, we will consider a constant-pressure, quasi-equilibrium, heating
process for a system comprising a pure substance, initially in the liquid phase (so the
pressure is below the critical value). Such a process will occur if the liquid completely fills
a vertical cylinder closed by a piston, with a weight on the piston to maintain the
pressure at the initial value P
1
,
and we arrange a heat transfer to
the liquid so that the temperature
and specific volume increase.
Another instance of a (nearly)
constant-pressure heating process
is when water flows through the
tubes of a power station boiler.
A two-dimensional slice through
the
PvT diagram gives the Pv
diagram shown on the left. C is
the critical point and states under
the curve BCD are those where
liquid and vapour are both
present. (The low-v end of the
diagram has been stretched in
order to see the liquid region ; if
drawn to scale, BC would almost
touch the P axis.) Note that the
isotherms (lines of constant
temperature) have a horizontal
portion under BCD, but as T
increases to values well above
the critical temperature, their
shape tends toward the hyperbola
Pv = RT = const. which corresponds to an ideal gas.

The constant-pressure process is plotted on the Pv diagram as 12345. In state 1,
the fluid is known as subcooled liquid. Heating it until it is just about to begin evaporating
(or boiling) brings us to state 2, on the line BC, where the fluid is described as saturated
liquid and its temperature is called the saturation temperature, T
sat
. Higher pressures
have correspondingly higher saturation temperatures, as can be seen from the
isotherms; e.g. at 1.013 bar ("standard" atmospheric presure), T
sat
= 100C, but at 7 bar
T
sat
= 165C. The line BC is called the saturated liquid line; it is the locus of saturated
liquid state points as P varies.
Further heating to state 3 causes some of the liquid to evaporate, and the two-phase
fluid here is described as wet vapour. Its temperature remains at T
sat
until we reach
state 4, on the line CD, where the last of the liquid has just evaporated and the fluid is
called saturated vapour.The line CD, the locus of saturated vapour state points as P
varies, is called the saturated vapour line, and the entire line BCD is referred to as the
saturation line. (Note that there is another school of thought which uses the term
'saturated' to refer also to wet vapour, and distinguishes states on CD by calling them
'dry saturated'. To avoid confusion, these notes will restrict use of the term 'saturated
1ME Thermodynamics Sec. 4
11
vapour' to states on CD.) Continued heating takes us to state 5, where the fluid is called
superheated vapour.
Another way of expressing the one-to-one link between P and T for saturated liquid, wet
vapour and saturated vapour states is to say, for example, that at 165C the saturation
pressure P
sat
is 7 bar.
The term 'subcooled liquid' is used for state 1 because T
1
< T
sat
at P
1
. Another name
for the fluid in state 1 is 'compressed liquid', because P
1
> P
sat
at T
1
. The term
'superheated vapour' is used for state 5 because T
5
> T
sat
at P
5
.


Summary of terminology Temperature specific volume
diagram

Another two-dimensional slice through the PvT diagram gives the Tv diagram shown
above on the right ; the liquid, two-phase and vapour regions are in the same places,
relative to the saturation line, as on a Pv diagram.


Specific properties in the saturated liquid
and vapour states are denoted by
subscripts f and g, respectively, i.e.
v
f
, u
f
, h
f
on the saturated liquid line,
v
g
, u
g
, h
g
on the saturated
vapour line.
The change in a specific property, at
constant P or T, between the saturated
liquid and the saturated vapour state is
denoted by the subscript fg. This is
normally used only for specific enthalpy,
where
h
fg
means h
g
h
f
.
Returning to the constant-pressure process

1ME Thermodynamics Sec. 4
12
1-2-3-4-5 used for illustration at the start of Sec. 4.6, that part of the process from state 2
to state 4 involves the evaporation of all the liquid to vapour. We know that in a constant-
pressure process for any substance, Q = H = m h (Sec. 4.5), so
Q
24
/ m = h
4
h
2
= h
g
h
f
= h
fg
.
h
fg
is thus the latent heat of evaporation of the substance. It varies with P or T.
Dryness fraction
Since isobars and isotherms coincide in the wet vapour region, P and T are not
independent properties in this region. Although a wet vapour state is completely fixed if
we know P and v, or T and v, or P and u, etc., it is convenient to use a "pseudo property"
called dryness fraction (or, in older literature, quality). This expresses how much of a wet
vapour consists of vapour, and has the symbol x. It is defined as


Dryness fraction x =
mass of vapour
mass of vapour + mass of liquid

vap
vap liq
vap
=
+
=
m
m m
m
m
... (4-9)

Thus x = 0 for saturated liquid, 0 < x < 1 for wet vapour, x = 1 for saturated vapour. For
subcooled liquid, superheated vapour and supercritical fluid, x has no meaning.
A system comprising m kg of a wet vapour is a mixture of x m kg of saturated vapour
with specific volume v
g
and (1 x) m kg of saturated liquid with specific volume v
f
.
Physically, it can be vapour bubbles in liquid (typical of low x) or liquid drops and films in
vapour (typical of high x) or liquid and vapour occupying separate parts of a container.
The arrangement does not matter and the word 'mixture' is used here to mean any
arrangement.
Volume of system = volume occupied by vapour + volume occupied by liquid
V = m v = x m v
g
+ (1 x) m v
f

The effective specific volume v of the mixture (i.e. what we would plot, below the
saturation line, on a Pv or Tv diagram) is therefore given by
v = x v
g
+ (1 x) v
f
, which can be re-written as x
v v
v v
f
g f




=

.
Thus for any wet vapour state on a Pv or Tv diagram, x can be pictured as the ratio
(distance from sat. liq. line to state point) / (distance between sat. liq. and sat. vap.
lines).
It won't often look right because we deliberately stretch the v scale at the low-v end of
the diagram.
By repeating the above analysis for the internal energy U and enthalpy H of a system in
a wet vapour state, similar equations are obtained for the mixture u and h.
The way in which dryness fraction is used in problem solving is usually by means of
eqns (4-10) below, not via its definition (eqn 4-9):


v = x v
g
+ (1 x) v
f

For a wet vapour state, u = x u
g
+ (1 x) u
f
... (4-10)
h = x h
g
+ (1 x) h
f

1ME Thermodynamics Sec. 4
13
Except at high pressure, where v
f
is not << v
g
, the equation for v can be simplified to
v x v
g
since the (1 x) v
f
term is usually negligible. However, the equations for u and
h cannot be simplified in the same way. Since values of h
fg
are often listed in property
tables, another way of writing the equation for h is h = h
f
+ x h
fg
.
Steam Tables
There are clearly no simple relationships connecting vapour properties, as there are for
gases. However, lengthy equations (based on experimental data and on thermodynamic
theory which is beyond the scope of the present 1M and 2M TF courses) allow computer
calculation of any property, given two other independent properties.
Properties of water and steam v, u, h and specific entropy s, which we shall meet in
Sec. 7 are included in the Data & Formulae book, for use in 2M and later years. For
now, values will be provided within 1M problems requiring steam properties for their
solution, so no table reading will be necessary. Section 1.7 of the 2M course will deal
with the layout of the steam tables and how to use them in problem solving.


1M E Thermodynamics Sec. 4

14

EXAMPLE (Solution started below, to be completed in lecture)
A vertical cylinder is fitted with a leak-proof, frictionless piston which supports a
mass m
p
of 60 kg. The cylinders cross-sectional area A is 0.001963 m
2
. The
piston is initially at rest, with its lower face a distance L
1
= 50 mm above the
closed end of the cylinder. Atmospheric pressure, P
a
= 1 bar, acts on the upper
face of the piston.
A heat transfer Q to the contents of the cylinder causes the piston to rise slowly
through a distance of 22.9 mm.
Determine the magnitude of this heat transfer when the cylinder contains:
(a) a gas which can be modelled as a perfect gas, To be done now
with specific heats ratio = 1.4 ;
(b) steam, initially with dryness fraction x
1
= 0.9.
SOLUTION
P
1
P
2
P
a
Q
L
1
L
2
initially finally
V
2 V
1


Assume quasi-equilibrium process. Force balance on piston at any point gives
P P
m g
A
a
p
= + = +

1 10
60 9 81
0 001963
5
.
.
= 3.998 x 10
5
N/m
2

4.00 bar
i.e. constant-pressure process, P = P
1
= P
2
= 4 bar

1st law : Q W = m (u
2
u
1
)
Now W =

V P d P
1
(V
2
V
1
) = P
2
V
2
P
1
V
1
= m (P
2
v
2
P
1
v
1
)
1M E Thermodynamics Sec. 4

15

so Q = m (u
2
+ P
2
v
2
[u
1
+ P
1
v
1
] ) = m (h
2
h
1
) ,
for either substance.
Now need a different approach for each substance.
(a) Gas, modelled as a perfect gas with = 1.4 (What gas? Table E5 no use)
h
2
h
1
= C
p
(T
2
T
1
) for a perfect gas
Q = m C
p
(T
2
T
1
) . Don't know C
p
, T
1
or T
2
and can't
evaluate m from P
1
V
1
/ (R T
1
) since R and T
1
unknown.
However, can write C
p
as R
1

(using C
p
C
v
= R and C
p
/ C
v
= )
so Q = m R (T
2
T
1
) / ( 1) .
Since m R T
2
= P
2
V
2
and m R T
1
= P
1
V
1
(eqn of state),
Q =
1

(P
2
V
2
P
1
V
1
)
=
1

P
1
(V
2
V
1
) =
1

P
1
A (L
2
L
1
)
=
4 . 0
4 . 1
x 4 x 10
5
x 0.001963 x 0.0229
= 62.9 J

1ME Thermodynamics Ex. 4
1
EXERCISES 4
Properties of Substances


1. A system consisting of a gas, and some of the properties defining its state, are
shown below.

To model the behaviour of this system in any process, we would need:
equations describing the type of process;
equations which relate the gas properties to each other in any particular state
during the process.
In this question, we are concerned only with the property modelling equations, for a
given state.
For gases, however complicated or simple a model we choose to use, there are two
basic property modelling equations:
an equation of state relating P, v and T (the properties which are easiest to
measure)
a relationship for the specific internal energy in terms of other properties.
(a) Write down the equation of state when the chosen model is an ideal gas, first in a
form involving n and the universal gas constant R , then in the two, more useful,
engineering forms involving the gas constant R for the particular gas.
(b) Write down the ideal gas relationship for change in specific internal energy u,
then express it in terms of C
v
, the specific heat at constant volume

.
(b) Other property relationships, involving properties such as specific enthalpy and
the specific heats, follow from the two basic equations. Making use of the
definition of specific enthalpy, h, show that for the ideal gas model, specific
enthalpy and the specific heat at constant pressure, C
p
, both depend on
temperature only, then derive the relationship between C
p
,C
v
and R.
Hence show that C
v
= R / ( 1) and C
p
= R / ( 1) .








1ME Thermodynamics Ex. 4
2
2. The perfect gas model is a simplified version of the ideal gas model. All the
relationships in question 1 still apply but some of them can be further simplified.
(a) Is the equation of state the same for the ideal and perfect gas models?
(b) Write down the simplified perfect gas relationship for change in u.
If it is required to assign a numerical value to u in a given state, an arbitrary
datum state must be chosen where u is taken to be zero. If T
0
is the temperature
in this datum state, write down the perfect gas expression for u at any other
temperature T, first for any datum state and then for a datum at absolute zero
temperature.
(c) Repeat part (b) for h.
(d) From the universal gas constant R = 8.314 kJ/kmol K and the gas molar masses
M given in Table E5, determine perfect-gas-model values for the following
properties, then compare them with the values in Table E5 and account for any
differences.
(i) For air: R, C
p
and C
v
, given that C
p
/

C
v
= 1.4
(ii) For carbon dioxide: R, C
v
and , given that C
p
= 0.84 kJ/kg K
(iii) For hydrogen: R, C
p
and , given that C
v
= 10.18 kJ/kg K


3. (a) A rigid vessel contains 0.1 kg of air, initially at 1 bar and 20C. A heat transfer to
the gas increases its temperature to 1000C.
What type of process does the gas undergo?
Assuming that the air can be modelled as a perfect gas with the properties given
in Table E5, calculate:
(i) the initial and final volume of the air;
(ii) the final pressure of the air;
(iii) the change of internal energy and enthalpy of the air;
(iv) the work and heat transfers.
(b) A vertical cylinder, closed at the top by an unconstrained, frictionless and leak-
proof piston, contains 0.1 kg of air, initially at 1 bar and 20C. A heat transfer to
the gas increases its temperature to 1000C.
What type of process does the gas undergo?
Assuming that the air behaves as a perfect gas with the properties given in Table
E5, calculate (i) to (iv) as in part (a).








1ME Thermodynamics Ex. 4
3
4. A certain gas with a molar mass of 29 kg/kmol is to be modelled as an ideal but not
perfect gas. Its specific heat at constant pressure is therefore not taken as constant,
but can be approximated over the temperature range 100 K to 2000 K by:
C
p
(in kJ/kg K) = a + b T
where T is the absolute temperature, in K, and a & b are constants with values a =
0.9912 kJ/kg K and b = 0.1076 x 10
-3
kJ/kg K
2
.
A piston-cylinder arrangement containing 2 kg of this gas, initially at a pressure of
1 bar, is heated from 15C to 1000C while the gas pressure remains constant.
(a) Starting from the 1st law equation for any process between states 1 and 2
undergone by a system of mass m (neglecting changes in the system's kinetic
and potential energy), show that in a constant-pressure quasi-equilibrium process
only, the heat transfer is given by
Q = m ( h
2
h
1
)
and hence calculate the heat transfer in this particular process.
(b) Calculate the change U
2
U
1
in the internal energy of the system.
(c) Calculate the work transfer W and state its direction, i.e. from system to
surroundings (work done by the gas) or from surroundings to system (work done
on the gas).
(Reminder : If this had been an examination question, some marks would probably
have been allocated to a simple, labelled sketch of the system in its initial and
then in its final state, showing known property values.)


















1ME Thermodynamics Ex. 4
4

5. The Diesel cycle is the simplest idealised model of the processes taking place in one
of the cylinders of a diesel, or compression-ignition (CI), engine. In the model, the
working fluid is taken to be air throughout the cycle, hence the more complete name
"air-standard Diesel cycle". (Question 5 of Exercises 3 concerned an idealised model
for the processes in one of the cylinders of a petrol, or spark-ignition, engine, known
as the air-standard Otto cycle.)
A comparison of the processes in a real diesel engine and their equivalents in the
Diesel cycle model is given in the following table.

Process Real CI engine Idealised model
1 2
Compression
stroke
Compression of air Adiabatic compression of
air
Process path equation:
P
1
V
1
n
= P
2
V
2
n

2 3
1st part of
power stroke
Diesel fuel sprayed into
cylinder, at about top-dead-
centre, and ignites due to high
temperature after compression
process.
Injection + combustion takes a
considerable time, during which
the gas (now air + other gases)
expands at roughly constant
pressure
Heat transfer to air in a
constant-pressure process
3 4
2nd part of
power stroke
Continued expansion of gas Adiabatic expansion of air
Process path equation:
P
3
V
3
n
= P
4
V
4
n

4 1 Exhaust stroke:-
exhaust valve opens, gases
expelled at P = approx.
P
atmos.

followed by
Induction stroke:-
exhaust valve closes,
inlet valve opens,
cylinder fills with fresh air
at P = approx. P
atmos.

Heat transfer from air in a
constant-volume process
(If the exhaust and
induction strokes are
modelled as constant-
pressure processes, at
atmos. pressure, they
have no net effect; one
is the reverse of the
other.)

(a) Sketch the idealised cycle on a pressure volume (PV ) diagram.
(b) Derive expressions for the heat transfer in each process of the idealised cycle
and hence an expression for the net work in the cycle.
1ME Thermodynamics Ex. 4
5
(N.B. In question 5 of Exercises 3, the net work was found by the obvious
method of summing the work transfers in all processes of the cycle. Since for any
cycle, net work transfer equals net heat transfer, an alternative method of finding
net work is to sum the heat transfers, as asked here; it is often faster).
(c) The efficiency
c
of a cycle is defined as the net work (what we want) divided by
the heat supplied (essentially what we pay for). ("Heat supplied" means the sum
of those process heat transfers which have positive values, from the
surroundings to the working fluid, modelling real real processes in which fuel is
burned.) Show that the efficiency of the Diesel cycle may be expressed as
( )
2 3
1 4
c

1
T T
T T

= .
(d) In a particular Diesel cycle, T
1
= 15C, T
3
= 2400C, the compression ratio
V
1
/

V
2
= 16 and the index n of the compression and expansion process
equations is 1.4. If the working fluid, air, can be assumed to behave as a perfect
gas (and it is a big "if", considering the temperature range !), calculate the
specific net work (net work divided by system mass) and the efficiency of the
cycle.
(e) What is the main difference between the idealised cycles for the spark-ignition
engine (question 5 of Exercises 3) and the compression-ignition, or diesel,
engine (this question)?
What is the main difference between the operating principles of these two types
of engine?

6. On a pressure specific volume (P v) diagram for water/steam, sketch the
saturated liquid and saturated vapour lines and mark the position of the critical point.
On the same diagram, sketch the paths of the four separate processes described
below. Label the paths a, b, c & d and indicate their directions with arrows.
(a) Superheated vapour is compressed isothermally until it reaches the saturated
liquid state.
(b) Superheated vapour in a rigid vessel is heated.
(c) Subcooled (or compressed) liquid is expanded isothermally until it
reaches the saturated liquid state.
(d) Wet steam with dryness fraction 0.8 is cooled at constant pressure
until the dryness fraction is 0.2.


7. A sealed, rigid tube contains a mixture of liquid water and water vapour / steam (i.e.
wet steam) at an initial pressure of 1 bar. When this fluid is heated, the process path
passes through the critical state.
(a) Sketch the process on a Pv diagram, not forgetting the saturation line.
(b) Find the fluid's initial dryness fraction.
(c) For a tube of volume 20 cm
3
, calculate the heat transfer to the tube contents
during the part of the process between the initial state and the critical state.


1ME Thermodynamics Ex. 4
6
8. This problem will be done as a group activity during a structured tutorial
(unless you are informed otherwise)

A rigid vessel with a volume of 0.35 m
3
is filled with saturated steam at a pressure of
1 atm. The vessel is then sealed and cooled to 20C.
(a) Sketch the process on a Pv diagram, including the saturation line and
isotherms for 100C and 20C.
(b) Tables show that at 20C, the saturation pressure of water/steam is 0.02339 bar.
Is it true to say that the final pressure in this process is 0.02339 bar? If so, why?
(c) Calculate the heat transfer in this process.

_____________________________________________________________________
ANSWERS
1. (a) PV = n R T PV = mRT , Pv = RT
(b) u = function of T only u = C
v
dT
2. (a) Yes (b) u = C
v
T u = C
v
(T T
0
), u = C
v
T
(c) h = C
p
T h = C
p
(T T
0
), h = C
p
T
(d) (i) R = 0.287 kJ/kg K C
v
= 0.651 kJ/kg K C
p
= 1.005 kJ/kg K
(ii) R = 0.189 kJ/kg K C
v
= 0.718 kJ/kg K = 1.29
(iii) R = 4.12 kJ/kg K C
p
= 14.30 kJ/kg K = 1.41
3. (a) (i) V
2
= V
1
= 0.0841 m
3
(ii) P
2
= 4.34 bar
(iii) U
2
U
1
= 70.6 kJ, H
2
H
1
= 99.0 kJ (iv) W = 0, Q = 70.6 kJ
(b) (i) V
1
= 0.0841 m
3
, V
2
=0.365 m
3
(ii) P
2
= 1 bar
(iii) U
2
- U
1
= 70.6 kJ, H
2
- H
1
= 99.0 kJ
(iv) W = 28.1 kJ, Q = 98.7 kJ or W = 28.4 kJ, Q = 99.0 kJ
4. (a) Q = 2118 kJ (b) U
2
- U
1
= 1553 kJ (c) W = 565 kJ, from system
5. (b) W
net
= m C
p
(T
3
T
2
) m C
v
(T
4
T
1
)
(d) W
net
= 1032 kJ/kg
c
= 56.7% (T
2
= 873.5 K, T
4
= 1379.4 K)
(e) Cycles The heat addition process occurs at constant pressure in the Diesel
cycle, compared with constant volume in the Otto cycle.
Engines In the diesel engine, only air is compressed, compared with air and
fuel vapour in the spark ignition engine; this enables higher compression ratios to
be used which, as will be seen later, results in a higher efficiency.
6.
7. (b) 0.00122 (c) 10.3 kJ
8. (b) Yes, because steam is wet steam (liquid/vapour mixture) in final state
(c) 492.5 kJ
1ME Thermodynamics Sec.5
5. THE FIRST LAW FOR FLOW PROCESSES
5.1 INTRODUCTION
In Sec. 3, we dealt with the First Law of Thermodynamics as applied to a system, a xed,
identiable mass of a substance. Eqn 3-4 expressed the fact that in a process from state
1 to state 2, the net energy addition to the system, i.e. Q
12
W
12
, must equal the increase
E
2
E
1
in the systems energy. The only way in which a systems total energy E (kinetic +
potential + internal) can change during the process is by means of energy transfers across
the system boundary, either as heat (caused by a temperature difference between system
and surroundings) or as work (any other cause). We noted that in common applications
of this equation, especially to gases in piston-cylinder arrangements, e.g. reciprocating
internal combustion engines, E
2
E
1
can be approximated closely by the change U
2
U
1
in internal energy only; also that work transfers are usually in the form of displacement
work only.
However, in Sec. 2.2 we saw that the system is not always the most convenient frame-
work for energy accounting, and we dened an alternative framework called a control
volume, a xed region of space rather than a xed mass of substance.
Suppose you wished to determine what work transfer was necessary to compress the
intake air from a pressure of about 1 bar to perhaps 20 bar in the compressor section of
this gas turbine engine (a Rolls-Royce RB211-24G).
It would not be easy to identify and keep track of a particular 1 kg, say, of air as it passed
through 14 rows of rotating blades, having some complicated combination of displace-
ment and shear work done on it in each blade row. The control volume is the obvious
framework here. If we draw a control surface round the whole compressor, as shown,
we need not be too concerned with the detail of the process inside the compressor. En-
ergy transfers across this control surface are: shaft work (see Sec. 3.2.2) where the
rotating shaft crosses the surface; heat transfer (since the air temperature rises as it is
compressed), likely in this case to be very small compared with the shaft work; plus a
form of energy transfer which did not arise with systems, namely the energy (kinetic +
potential + internal) in the air which ows into the control volume at the upstream end and
out again at the downstream end, where it is in a different state.
1
1ME Thermodynamics Sec.5
All the components of the combined-cycle power plant illustrated in the Introduction to the
1M Thermo-Fluids course are best modeled as control volumes.
We now need a version of the First Law equation which applies to ow processes, pro-
cesses in which the working uid ows through some piece of equipment which we wish
to dene as a control volume. Eqn 3-4, Q
12
W
12
= E
2
E
1
, is not adequate as it stands
because it does not account for energy transported into and out of a control volume with
the owing uid.
In Sec. 5.2, a suitable equation will be obtained by dening a system, looking at how its
boundary is displaced as it passes through a control volume, and applying the 1st Law
to it via eqn 3-4, which is valid for this system. In principle, we could do this every time
we met a control volume, but the whole idea is to avoid having to do so. Once we have
derived a 1st Law equation which is valid for a control volume, we can use it for all control
volumes without ever again having to think about systems owing through them.
5.2 THE 1ST LAW EQUATION FOR CONTROL VOLUMES
(C &B, 4th 4-3 & 4-4)
Consider a control volume, shown below, with two ports. Ports are places on the control
surface where uid can ow into or out of the control volume. Suppose that there is an
inow through the left-hand port and an outow through the top port. The pieces of pipe,
leading to and from the ports, are outside the control volume.
It is possible to regard the continuous process as a series of nonow processes under-
gone by an imaginary system. Let us dene the system, shown by the shading in the
above diagram. The left-hand diagram represents the initial state (subscript in), in which
the system occupies the whole control volume plus a small part of the inow pipe. After
a certain time, t, the system occupies the control volume plus a small part of the outow
pipe. It now becomes possible to apply the energy equation which has been developed
from a study of systems.
First consider the state of the imaginary system in the conguration shown in (a) above.
An enlarged view of the part of the system containing the element of uid that is about
to enter is shown in the gure below. The thermodynamic and mechanical states of the
element of mass m may be assumed to be known. The internal energy of the element
of mass is U
in
, and if the element is moving it also possesses kinetic energy m
in
C
2
in
/2,
with reference to a datum height the potential energy of the element would be m
in
gZ
in
.
The total energy of the mass entering the control volume is therefore,

mu +
1
2
mC
2
+mgZ

in
2
1ME Thermodynamics Sec.5
At the time when the element of mass is about to enter the control system, the total
energy of the uid elements within the control volume is E
i
. Therefore the total energy of
the imaginary system at the initial state is,
E
t
....
system energy
= E
i
....
c.v. energy
+

internal
....
mu +
kinetic
. .. .
1
2
mC
2
+
potential
. .. .
mgZ

in
. .. .
energy of the small mass at inlet
In a similar way the energy of the system in the nal state (time t +t) is,
E
t+t
= E
f
+

mu +
1
2
mC
2
+mgZ

out
During the process the imaginary system changes from its initial state to its nal state,
how has this change come about? In this period of time (t) there has been a transfer
of energy to the system in the form of Q, and shaft work has been done by the sys-
tem as W
sh
. In the process the system boundary has been displaced at inlet and exit,
therefore W
sh
cannot be the only work done by the system. For an element of mass
m
in
to enter the control volume, the system must be compressed by a certain amount,
the volume has decreased by m
in
v
in
. This is accompanied by a force P
in
A acting over
a length L = m
in
v
in
/A. The work done by the surroundings on the system is therefore
m
in
P
in
v
in
. In a similar way, the displacement work done by the system on the surround-
ings is m
out
P
out
v
out
. The total work done by the system becomes,
W
sh
+ (m
out
P
out
v
out
m
in
P
in
v
in
)
1
It is now possible to apply the rst law to the imaginary system and write,
Q[W
sh
+ (mP v)
out
(mP v)
in
] =

E
f
+m
out

u +
1
2
C
2
+gZ

out

E
i
+m
in

u +
1
2
C
2
+gZ

in

1
The Pv terms are often referred as ow work, ow energy or pressure energy by various authors.
There is no need to give this term a special name.
3
1ME Thermodynamics Sec.5
By introducing the denition of specic enthalpy h = u + Pv, it is possible to simplify the
above equation. Because specic enthalpy is dened as a combination of properties u,
P and v, it is itself a property. That is, a change in enthalpy between two states depends
only upon the end states, and it is independent of the process. Hence,
QW
sh
=

E
f
+m
out

h +
1
2
C
2
+gZ

out

E
i
+m
in

h +
1
2
C
2
+gZ

in

By rearranging the above equation and dividing by t, we can write,


Q
t

W
sh
t
=
(E
f
E
i
)
t
+

m
out
t

h +
1
2
C
2
+gZ

out

m
in
t

h +
1
2
C
2
+gZ

in

?
rate of heat transfer,

Q
?
rate of work transfer,

W
?
change of total en-
ergy of the c.v.,
dE
dt
?
outlet mass ow rate,
m
out
?
inlet mass ow rate,
m
in
4
1ME Thermodynamics Sec.5
The rst law becomes,

Q

W
sh
=
dE
dt
+

m
out

h +
1
2
C
2
+gZ

out

m
in

h +
1
2
C
2
+gZ

in

(5.1)
It is not necessary to remember the details of this derivation. You only need to be aware
of the reason why a uid owing into a c.v. brings with it a specic energy h +C
2
/2 +gZ,
not u+C
2
/2+gZ. Essentially, it is because the surroundings have had to do displacement
work, equal to the value of Pv at the inlet port, on the uid in order to push it into the c.v.;
this increases the uids specic energy from u+C
2
/2+gZ when it is just outside the c.v.,
to u + Pv +C
2
/2 + gZ when it is just inside. A similar argument applies to uid leaving a
c.v.; work Pv now has to be done by the uid inside the c.v., so the c.v. loses more energy
than just the I.E. + K.E. + P.E. of the uid.
5.3 STEADY-FLOW ENERGY EQUATION
If we wanted to examine ow processes such as the lling of a compressed gas bottle
by connecting it to a high-pressure supply line, the control volume dened as the bottle
would have only one port (inow through its valve). Such a process would not be steady,
because the conditions inside the control volume would change continuously; E
f
would
differ from E
i
. Unsteady ow processes are beyond the scope the the current 1M course,
but eqn (5-1) is a useful tool for analyzing them. We shall now concentrate on steady
ow processes, in which the conditions inside the control volume and at its ports do not
change with time. control volume will abbreviated to c.v. from now on.
Steady ow implies
1. that the mass ow rate m through each port is constant and m
in
= m
out
; also that
there are a constant heat transfer rate

Q and shaft work transfer rate

W
sh
across the
control surface; as usual.
2. that the energy of the c.v. contents does not change with time; different material
occupies the c.v. at different times, but at every point within the c.v. the local proper-
ties do not change with time, so the total energy of material in the c.v. at one instant
equals that at any other instant; hence dE/dt = 0, however long the process lasts.
With these conditions the above equation can be written for the steady ow as follows,

Q

W
sh
= m

h +
1
2
C
2
+gZ

out
m

h +
1
2
C
2
+gZ

in
(5.2)
Eqn (5.2) is a version of one of the most important and widely used equations in thermo-
dynamics; it is called the steady ow energy equation, often abbreviated to SFEE. You
must remember this one!
5
1ME Thermodynamics Sec.5
The terms in eqn 5-2 should be interpreted as follows.

Q

W
sh
Rates of heat and shaft work transfer across the control
surface. The c.v. is of xed and shape so there is no
displacement work due to control surface movement.
m

h +
1
2
C
2
+gZ

in
Rate at which energy is brought into the c.v. with the
inowing uid; the reason why specic enthalpy h ap-
pears instead of specic internal energy u has been
explained in Sec. 5.2. Remember that h, C and Z are
evaluated at the inlet port.
m

h +
1
2
C
2
+gZ

out
Rate at which energy is removed from the c.v. with
the outowing uid. Remember that h, C and Z are
evaluated at the outlet port.
For a c.v with three or more ports, eqn 5-2 has to be expanded so that there is a term
+ m(h +C
2
/2 +gZ)
in
for each inow port and a term m(h +C
2
/2 +gZ)
out
for each out-
ow port. The SFEE can be written for a c.v. with any number of ports as follows, where

denotes the sum of all such terms:

Q

W
sh
=

outow

h +
1
2
C
2
+gZ

inow

h +
1
2
C
2
+gZ

(5.3)
With three or more ports, we also have to use the mass continuity equation explicitly, to
express the fact that the mass in the c.v. stays constant in steady ow. To match eqn 5-3,
it is:

outow
m =

inow
m (5.4)
We can apply eqn 5-2, or 5-3 plus 5-4, to the steady operation of pumps, compressors, tur-
bines, heat exchangers including boilers and condensers, valves in pipelines, processes
in which two or more uid streams are mixed, ... etc. The c.v. can be drawn round just
one of the components, or round any two or more components through which the uid
ows successively.
Before we look at the application of the SFEE to some common examples of engineering
plant, it is worth noting that as part of the design process, we often decide on appropriate
thermodynamic states at the inow and outow ports before considering what the mass
6
1ME Thermodynamics Sec.5
ow rate(s) should be. The SFEE can then be used in a form where all the terms are
divided by mass ow rate; for example, eqn 5-2 would be written as:

Q
m

W
sh
m
=

h +
1
2
C
2
+gZ

out

h +
1
2
C
2
+gZ

in
This would enable us to nd the values of

Q/ m, the heat transfer per unit mass of working
uid and

W
sh
/ m, the shaft work transfer rate per unit mass, called the specic work.
(

Q/ m is not called the specic heat transfer, to avoid confusion with the property specic
heat.) The actual heat transfer rate

Q and shaft power

W
sh
would then be found by
multiplying

Q/ m and

W
sh
/ m by the desired value of m. A small m would result in a
small version of that component or plant, and a large m would produce a large version,
since the size is related to the ow rate through it, from the mass ow rate expression
m = AC (density x cross-sectional area x velocity) = AC/v. Velocity magnitude C
and specic volume v are xed by the thermodynamic calculations, so A is proportional
to m.
5.4 USING THE SFEE: SOME EXAMPLES
(C &B Section 4-4)
When we were dealing with systems, we used the subscripts 1 and 2 (or 2 and 3, etc.)
to denote the initial and nal states in a process. Now, with a steady ow process in a
c.v., there is no real beginning or end to the process in terms of time, but if we follow the
working uid through the c.v. the process does have a beginning at the inow port and an
end at the outow port. We shall therefore use numerical subscripts to denote the ports
of the c.v.
In Fluid Mechanics, you have been calculating mass ow rates through pipes etc. from
a given velocity prole over the cross-section. We will not go into such detail here; local
uid velocity will be assumed to be uniform over the ow cross-section at each port.
5.4.1 Flow through a boiler
If the c.v. is drawn to enclose only the uid which is being heated to saturation temperature
then evaporated (and possibly then superheated), there will be a heat transfer to the c.v.
contents from the hot gases produced by burning the fuel.
Facts: 2 ports.

W
sh
= 0.
Assumptions: K.E. & P.E changes are negligible.
SFEE reduces to
7
1ME Thermodynamics Sec.5
In the laboratory steam plant experiment, you will need to use this equation to nd the
heat transfer rate to the working uid, having measured the mass ow rate and derived
the inlet and outlet enthalpies by applying the SFEE to the feed water pump and to the
throttling calorimeter.
5.4.2 Flow through a partly-closed valve a throttling process
Facts: 2 ports,

W
sh
= 0.
Assumptions:

Q is negligible
(valve well lagged).
K.E. & P.E changes are negligible.
SFEE reduces to
i.e. throttling is a constant-enthalpy process.
This is the basis of the throttling calorimeter, which you will meet in the laboratory steam
plant experiment. For the laboratory boiler, the exit steam is wet, i.e. the exit state in is
in the liquid+vapour region of the Pv diagram. Pressure and temperature in this state are
therefore not independent properties, so the exit enthalpy cannot be found from these
measurements. A sample of the boiler exit steam is therefore throttled by passing it
through a well-lagged, partly-closed valve. The valve exit state is in the superheated
steam region, so a measurement of now-independent pressure and temperature after the
valve gives the valve-exit enthalpy which equals the valve-inlet enthalpy, and this equals
the boiler exit enthalpy.
5.4.3 Flow through a pump or compressor
Devices for increasing the pressure of a uid stream are generally called pumps for liquids
and compressors for gases.
Facts: 2 ports.
Assumptions:

Q is negligible.
K.E. & P.E changes are negligible.
8
1ME Thermodynamics Sec.5
(For the type of pump which is designed to move a liquid without a signicant pressure
rise, and for some compressors, the assumption of negligible K.E. change may not be
valid.)
SFEE reduces to
The specic work

W
sh
/ m , or work per unit mass, required to compress the uid is there-
fore h
1
h
2
, which will have a negative numerical value; this is in accordance with our
sign convention, because the work transfer is from the surroundings to the c.v. contents.
For liquids, which are almost incompressible so that v constant, we can write
u is almost independent of P for liquids, and can be regarded as a function of T only,
so that if, as usual, the liquid undergoes only a small temperature change in the pump,
u
1
u
2
0 and
|

W
sh
/ m| v(P
2
P
1
) .
This is a very useful relationship for liquid pumps. It does not apply to gas compressors,
where the uid is not incompressible and there is a signicant temperature rise during the
process. (Do not be fooled into thinking that the temperature cannot rise because there
is no heat transfer! Your knowledge of the 1st Law should tell you by now that u or h can
increase as a result of a negative (to a system or c.v.) work transfer as well as a positive
heat transfer.)
5.4.4 Flow through a nozzle
Nozzles are devices for increasing the kinetic energy, hence the velocity, of a uid stream,
at the expense of internal energy. Until you deal in the 2nd-year course with supersonic
ows, think of nozzles just as converging ducts, i.e. cross-sectional area decreasing in
the ow direction. They range from simple cones, e.g. at the end of a water hosepipe, to
the more-carefully-shaped nozzles which form the tail pipe (or jet pipe) of an aircraft jet
engine (where a high-speed jet is needed to produce thrust), or the passages between
the xed blades of a turbine (where the uid has to be accelerated and swirled in order to
impart angular momentum to the rotating blades).
Facts: 2 ports.

W
sh
= 0.
Assumptions:

Q is negligible.
P.E change is negligible.
SFEE reduces to
Sometimes, the inlet kinetic energy can be ignored compared with the exit value, so the
exit velocity can be calculated from
C
2
2
/2 = h
1
h
2
.
9
1ME Thermodynamics Sec.5
Remember that h
1
h
2
becomes C
p
(T
1
T
2
) for a perfect gas (although it not a constant-
pressure process!), but for steam the values of h must be looked up in tables.
5.4.5 Flow through a turbine
The uid expands and the objective is to obtain shaft work at the expense of internal
energy.
Facts: 2 ports.
Assumptions:

Q is negligible.
K.E & P.E changes are negligible.
SFEE reduces to
(same as for a compressor).
This time, h
2
< h
1
and, as the sign convention requires, the numerical value of

W
sh
is
positive.
In some cases, such as the low-pressure section of a steam turbine, the exit K.E. is
signicant and the SFEE has to be written as:

W
sh
= m[h
1
(h
2
+C
2
2
/2)] .
5.4.6 Flow through a heat exchanger
The purpose of a heat exchanger is to transfer energy from a hot uid stream to a cold
uid stream, usually without mixing the two streams. (Cases where the streams mix will
be considered in Sec. 5.4.7.) An example is the condenser of the steam plant in the
1MTF laboratory experiment, where the steam condenses to water as it gives up energy
to a separate ow of cooling water. Simple heat exchangers are often of shell-and-tube
construction; one stream ows through a number of parallel tubes or perhaps a single
tube with U-bends that makes more than one pass through the outer casing known as
the shell. The other stream ows through the space between the shell and the outside of
the tubes. The choice of c.v. depends on what we are trying to calculate.
For the c.v. in the rst diagram, drawn round the entire heat exchanger:
Facts: 4 ports.

W
sh
= 0.
Assumptions:

Q to or from outer casing
is negligible (the outside
is well lagged). K.E. & P.E.
changes are negligible.
10
1ME Thermodynamics Sec.5
Note that the heat transfer between the two streams occurs entirely within the c.v., not
across the control surface, so it does not appear in the SFEE.
SFEE reduces to
used with the mass continuity equations m
1
= m
2
and m
3
= m
4
.
For the c.v. in the second diagram, drawn round just one of the uid streams (in at port 3
and out at port 4) and using the standard block-diagram symbol for a heat exchanger:
Facts: 2 ports.

W
sh
= 0.
There is now a

Q.
Assumptions: K.E. & P.E. changes
are negligible.
SFEE reduces to
with m
3
= m
4
.
At the same time,

Q
12
= m
1
(h
2
h
1
) , where m
1
= m
2
.

Q
12
=

Q
34
, and if the SFEEs for
the two individual streams are added, we get the same SFEE as for the c.v. in the rst
diagram, drawn round the complete heat exchanger.
5.4.7 Flow through a mixing chamber
A simple example is the mixing of hot water and cold water streams in a shower unit.
Another example is a spray condenser, where a superheated vapour is condensed by
spraying cold liquid into it. We are usually dealing with a 3-port c.v., having two inlets and
one outlet. It does not matter whether the chamber is a vessel or just a junction between
two pipes.
Facts: 3 ports.

W
sh
= 0.
Assumptions:

Q is negligible; if the two
inlet streams are at different
temperatures, there will be a
heat transfer inside the
chamber but not across the
control surface.
SFEE reduces to m
1
h
1
+ m
2
h
2
m
3
h
3
= 0
used with the mass continuity equation m
1
+ m
2
= m
3
.
11
1ME Thermodynamics Sec.5
5.5 Mass and volume ow rates a reminder
You should be familiar with calculating mass and volume ow rates, from Fluid Mechanics.
You will often need to make similar calculations as part of a 1st Law analysis or design
calculation using the SFEE.
For example, suppose you had used the SFEE to nd out what mass ow rate m of
air+combustion products was needed to produce a required power from a land-based gas
turbine. You would also have found the specic volume v and velocity C at the turbine exit.
In order to nd what size of duct would be needed to convey these gases to the exhaust
stack (or chimney)
2
, you would use the equations
volume ow rate

V = m/ = mv where density = 1/v and
volume ow rate

V = AC where A = duct cross sectional area.
So, A = (/4)D
2
= mv/C where D = duct diameter.
In Thermodynamics, the mass ow rate relationship m = CA is more often written m =
CA/v.
2
Gas turbine exhaust gases, at around 500C, are too good to throw away! Use them in a heat exchanger
to produce steam from a ow of water, and use that steam for heating purposes or to drive a steam tur-
bine and thus augment the power (as in the combined cycle power plant illustrated in the Thermo-Fluids
course Introduction). What mass ow rate of steam can we generate in the heat exchanger? More SFEE
calculations!
12
1M E Thermodynamics Sec. 5
13


1ME Thermodynamics Ex. 5
1
EXERCISES 5
The 1st Law for Flow Processes

1. A boiler operates at a pressure of 4 bar (absolute) and produces wet steam with a
dryness fraction of 0.90. The feed water enters the boiler at a temperature of 40C.
The fuel burners provide a heat transfer rate of 100 kW to the water/steam.
(a) Apply the steady flow energy equation (SFEE) to a control volume, drawn round
the water and steam only, to calculate the mass flow rate of water/steam through the
boiler.

(b) Find the mean velocity of the wet steam as it leaves the boiler through a pipe of
diameter 50 mm.

Take the conditions of the wet steam as: P = 4 bar, v
f
= 0.001084 m
3
/kg, v
g
= 0.4625
m
3
/kg, h
f
= 604.9 kJ/kg and h
g
= 2738.5 kJ/kg


2. The diagram on the right represents
an industrial gas turbine engine,
driving an electrical generator. In the
simplest mechanical arrangement, as
shown here, the turbine, compressor
and generator are directly coupled, on
the same shaft. The shaft power from
the turbine has to drive the
compressor as well as the electrical
generator. (We shall see later that the
compressor consumes a large
fraction of the turbine's gross power
output.)


(a) Atmospheric air at a temperature of 15C flows with negligible kinetic energy into
the compressor. It leaves the compressor at a rate of 316 kg/s, at a pressure of
12.56 bar (absolute) and a temperature of 395C, through a duct of cross-
sectional area 0.4 m
2
.

(i) Assuming that air behaves as a perfect gas, with properties as given on the
data sheet, use the equation of state and the mass flow rate relationship in state
2 to find the mean velocity of the air at the compressor exit.
(ii) Use the SFEE to find the power required to drive the compressor,
noting that the exit kinetic energy cannot be neglected.
(b) The mass flow rate of fuel injected into the combustion chamber is only about 2%
of the mass flow rate of air, so it reasonable to model the flow process in the
combustor and the turbine by assuming that the working fluid is air alone.
Consistent with this approximation, we can model the combustion process as
heat addition to the air. If the air temperature at the combustor exit is to be
1162C (limited by the ability of the highly stressed parts of the engine to
withstand the hot gas over a long period of time; creep resistance is the principal
problem), find the heat transfer rate which is equivalent to the burning of the fuel.
1M TF Thermodynamics Ex. 5
2
(Use the SFEE on a control volume drawn round the combustor, with two ports
only since we are neglecting the fuel flow. Neglect combustor exit K.E.)
(c) If the turbine exit temperature is 522C, calculate the shaft power produced by the
turbine, neglecting kinetic energies at entry and exit.
(d) Calculate:
(i) the shaft power available to drive the electrical generator, which equals the
electrical power if we neglect losses due to friction in the shaft bearings and
losses in the generator itself (which will total only a few per cent);
(ii) the efficiency of the plant, defined as the electricity output divided by the
heat input.

3. Air enters a propulsion nozzle at 900 K and 0.6 bar through an inlet pipe of area 1.3
m
2
. At the nozzle exit the state is 0.265 bar and 710 K. The nozzle is well lagged so
the process may be assumed to be adiabatic. The mass flow rate is 80 kg/s. Find the
velocity at the nozzle exit. (Pressures are absolute unless stated otherwise, here and
in subsequent questions).

4. An electric hair dryer (which can be regarded as a fan and a resistance heater inside
a duct) has a fan motor rated at 100 W and a 600 W heater. The air inlet is fairly
large, and the exit has a diameter of 30 mm. The dryer heats air from 18C, 1 bar to
50C, 1 bar, with a negligible heat loss from its casing.
(a) Taking air as a perfect gas, find the inlet and exit densities and the change in
specific enthalpy.

(b) Assuming that the kinetic energies are negligible, calculate the air mass flow rate
and hence the exit mean velocity.

(c) Recalculate the mass flow rate, taking into account the exit velocity found in (b).
Was the assumption in (b) valid?

5. An air compressor takes in air at 1 bar, 20C and delivers it at 300C, at a flow rate of
0.1 kg/s. The compressor is cooled by a flow of water, also at a rate of 0.1 kg/s; the
cooling water temperature rises from 15C to 70C as it flows round the compressor
casing. All K.E.s can be assumed negligible.
(a) Write down the SFEE, using symbols which you have defined on a block diagram,
(i) for a control volume which excludes the water cooling "jacket", and (ii) for a
control volume containing only the water cooling jacket. Show that if these two
equations are added, the result is the same as if the SFEE had been written for a
control volume surrounding the entire water-cooled compressor.

(b) What is the power required to drive the compressor? (N.B. h for a liquid can be
closely approximated by h
f
at the same temperature.


6. Air flows through a turbine at a rate of 45.2 kg/s with inlet conditions of 10 bar and
1500 K. The inlet area is considered to be large. At the outlet, the pressure is 2 bar,
1M TF Thermodynamics Ex. 5
3
the temperature is 350 K and the area is 0.047 m
2
. stating your assumptions,
calculate work output.

7. A mixing process is to be used to produce nitrogen (N
2
) in the saturated liquid state at
80K (where its saturation pressure is 1.37 bar), by mixing streams of subcooled (or
compressed) liquid nitrogen and gaseous nitrogen. The subcooled liquid is supplied at
70K, 1.37 bar, at a flow rate of 0.1 kg/s. The gas is supplied at 300K, 10 bar, and is
throttled to the same pressure as the liquid N
2
before entering the mixing chamber.
The equipment is well insulated but the heat transfer rate from the surroundings to the
mixing chamber is estimated to be 50 W. It is required to find the necessary flow rate
of N
2
gas.

(a) Sketch a block diagram, define a suitable control volume and number the ports.

(b) Using the following extracts from
N
2
property tables and making an
appropriate approximation for the
enthalpy of subcooled liquids, find the
mass flow rate of N
2
gas.

T h
f
h at 10 bar
70K 136.6 kJ/kg
80K 115.9 kJ/kg
300K 309.2 kJ/kg



_____________________________________________________________________
ANSWERS

1. (a) 0.0424 kg/s (b) 9.00 m/s
2. (a) (i) 120.6 m/s (ii) 123.6 MW (b) 242.5 MW (c) 204.3 MW
(d) (i) 80.7 MW (ii) 33.3%
3. 674 m/s
4. (a)
1
= 1.197 kg/m
3
,
2
= 1.078 kg/m
3
, h
2
h
1
= 32.3 kJ/kg
(b) 0.0217 kg/s, 28.4 m/s
(c) 0.0214 kg/s; result in (b) was in error by less than 2%, hence the assumption
was reasonably valid.
5. [ In (a), there should be heat transfer rate terms in the separate SFEEs but not in the
added equation.] (b) 51.3 kW
6. (a) 47.23 MW
7. 0.00475 kg/s
1ME Thermodynamics Struc. Tut. 2
1

PROBLEM FOR STRUCTURED TUTORIAL

1st Law of Thermodynamics for Flow Processes:
the Steady-Flow Energy Equation


Nitrogen flows steadily into a compressor at 1 bar, 20C, with a mean velocity of 50 m/s,
and is compressed adiabatically to 10 bar, 320C. It then flows through a cooler where
its temperature is reduced to 40C, with a negligible change in pressure. Kinetic energies
are negligible except at the compressor inlet.
The compressor is driven by a small steam turbine. Steam enters the turbine at 20 bar,
400C, with a mass flow rate of 0.15 kg/s, and leaves as wet steam at 0.2 bar with
dryness fraction 0.95. Kinetic energies and heat loss from the turbine casing are
negligible.
The steam is then condensed to the saturated liquid state in a water-cooled condenser;
the cooling water enters the condenser at 20C with a mass flow rate of 16 kg/s. Energy
losses in the transmission between steam turbine and nitrogen compressor (friction in
the bearings, and in a gearbox if one is needed because the design shaft speeds differ)
leave 95% of the turbine shaft power available for driving the compressor.


(a) Sketch a block diagram of the plant. This has been done for you; note that it is
good practice to write the known data on the diagram, both to define symbols and
to save searching through the text as you tackle each step of the solution.)

(b) Find the shaft power produced by the steam turbine.

(c) Find the mass flow rate of nitrogen.

(d) Find the heat transfer rate from the nitrogen in the cooler.

(e) To minimise the pressure loss due to friction in the pipe, it is required to keep the
mean velocity of the nitrogen below 10 m/s after it leaves the cooler. Pipes with
internal diameters of 50 mm and (at higher cost) 75 mm are available. Which size
should be used?

(f) Find the cooling water temperature as it leaves the condenser.

Turn over

1ME Thermodynamics Struc. Tut. 2
2


1ME Thermodynamics Sec. 6
1
6. THE SECOND LAW OF THERMODYNAMICS

6.1 INTRODUCTION

You have recently been tackling problems involving various power plant components
(reciprocating engines, boilers, turbines, heat exchangers, pumps, compressors, etc.)
using the 1st Law for either a system or a control volume. If you were asked to find,
for example, the shaft power produced by a turbine, the inlet and outlet states of the
working fluid had to be specified. So far, you have not met any techniques which
would allow you to predict the work output of that turbine, given only the inlet state
and perhaps the outlet pressure (which by itself does not fix the outlet state). This is
clearly something which designers of such plant need to be able to do. At this point in
the 1M course, you are in much the same position as steam engine designers in the
first quarter of the 19th century, before the 2nd Law of Thermodynamics was
recognised.
The 2nd Law provides us with such techniques, by telling us what would be the
performance of a perfect example of the plant component (e.g. a turbine with no
energy losses to the surroundings and no frictional effects). The property entropy is a
consequence of the 2nd Law. We shall see in Sec. 7 that entropy provides a measure
of how imperfect a real component is, and allows us to predict the performance of a
real component if we assume that it is imperfect to the same extent as another
component of which we have some experience.
The 1st Law is not concerned with the directions in which energy transfers take place,
but the 2nd Law places some restrictions on the direction, and also tells us that work
is a more valuable form of energy transfer than heat (you can do more things with n kJ
of work than you can with n kJ of heat). Like the 1st Law, the 2nd Law is a law of
nature which cannot be proved directly, but its many consequences all make sense
and we can be confident that it is true.
An idea of the directional implications can be gained by considering a fluid in a vessel
with thermally-conducting walls. If the fluid is stirred by a paddle wheel, the work input
will increase its internal energy, hence its temperature, so there will be a heat transfer
from the fluid to the surroundings through the walls. However, if you started again with
the paddle wheel stationary and supplied heat through the wall to the fluid, would you
expect the paddle wheel to begin rotating? The process cannot be completely
reversed.
The 2nd Law has been stated in a number of different ways, at first sight unrelated to
each other, but if any one of them is accepted as true, the others all follow from it. The
most easily understood version is the one first stated in 1850 by the German physicist
R. Clausius, along the lines that
"Heat cannot be transferred from a body to a hotter body without a work input".
Have you ever observed a hot drink on your desk getting hotter while the desk below it
cools down?! On the other hand, a refrigerator causes a heat transfer from the food in
the cold storage compartment to the warmer kitchen, but it stops working when there
is no longer an electrical work input.

The "Clausius statement" and another equally important statement of the 2nd Law
due to Kelvin and Planck, are written more formally in the language of systems and
cycles. Before we go much further, we need to grasp the idea of a heat engine,
defined here as a system which goes through a cycle of processes and produces net
1ME Thermodynamics Sec. 6
2
positive work (i.e. W > 0). The steam and water in a steam power plant therefore
constitutes a heat engine; it goes through a cycle of processes in the boiler, turbine,
condenser and pump, and the shaft work transfer from it in the turbine is greater than
the magnitude of the work transfer to it in the pump. In our idealized model of the
processes in a spark-ignition reciprocating engine, the working fluid (air only) is a heat
engine, although the real petrol engine does not match this definition because the
fluids (air, petrol vapour, combustion products) do not actually go through a cycle. It
may help to think of a heat engine as a generalized machine for converting heat into
work, as long as you remember that the working fluid inside it, going through a
thermodynamic cycle, is what really matters.


6.2 HEAT ENGINES, THERMAL RESERVOIRS AND THERMAL EFFICIENCY
(&B 4
th
Ed., Sects. 5-1 & 5-2)
In this rather abstract but nevertheless useful
description of heat engines, the heat transfer
from the surroundings to the working fluid is
assumed to come from a single source of
energy at a high temperature. This heat
source is called a thermal reservoir, and can
be imagined as a mass of material sufficiently
large that when energy is removed from it in
the form of a heat transfer to the working fluid,
its temperature does not drop significantly.
Similarly, the heat transfer from the working
fluid (during another part of the cycle) is
assumed to be to another thermal reservoir
(sometimes called a heat sink), at a low
temperature. We shall call these reservoirs
the hot reservoir, at temperature T
h
, and the
cold reservoir, at temperature T
c
, as shown on
the right. The corresponding heat transfers
are Q
h
and Q
c
.
Q
h
, Q
c
and the net work W
net
can be thought of as the amounts of energy transferred
to or from the working fluid in one cycle, or as the amounts in unit time during steady
operation, i.e. rates of heat transfer and power.

In everyday life, the efficiency of any operation is thought of as "what you want out of
it divided by what you have to pay to get it". As far as engines are concerned, what
you want is the net work (positive work minus the magnitude of the negative work),
and what you have to pay for is the heat input (in practice, the fuel which has to be
burned to provide the heat input). Note that the heat transfer from an engine, or the
heat output, is here regarded as having little value. (If, in a real power plant, this
"waste heat" can be used, "efficiency" has to be re-defined.)


The thermal efficiency
th
of a heat engine is defined as
th
= W
net
/ Q
h
. ... (6-1)


Note carefully that the numerator is net work, i.e. the algebraic sum of all the work
transfers in the cycle, but the denominator is the heat input, or heat addition, only, i.e.
1ME Thermodynamics Sec. 6
3
the sum of those heat transfers which are positive. The heat output, or heat rejection,
does not feature in this definition unless we re-write it as follows.
Since Q W = 0 for a cycle, W
net
= Q
h
+ Q
c
= Q
h
|Q
c
| .


For the rest of Sec. 6, let us forget our sign convention for Q and W, and take all Qs
and Ws to represent only the magnitudes of the heat and work transfers. Directions
of the energy transfers will be made clear by arrows on the diagrams showing thermal
reservoirs and heat engines.

With this new meaning for the symbols Q and W, we can now write
W
net
= Q
h
Q
c
.
Therefore
th
can also be written as

th
= ( Q
h
Q
c
) / Q
h
= 1 ( Q
c
/ Q
h
) ... (6-2).



We can represent the performance of a 660
MW steam plant, e.g. one of the six such units
at Drax Power Station in Yorkshire, by the
diagram on the right, where the numbers for
the Qs are reasonable approximations.


th
= 660 / 1650 or 1 (990 / 1650)
= 0.40 or 40%.

The 990 MW is rejected to the surroundings,
first to the cooling water in the condenser then
to the plumes of air-plus-water-vapour
emerging from the cooling towers. The
purpose of the cooling towers is to cool the
condenser cooling water so that it can be
returned to the condenser to repeat its cooling
job. (Don't confuse this cooling water with the
working fluid, the water/steam which goes
round the circuit of feed pump, boiler, turbine
and condenser.)


Before we look at the profound implications of the 2nd Law for heat engines and their
efficiency, in Sec. 6.4, we must introduce the concept of a reversible process.


6.3 REVERSIBLE PROCESSES (&B Sect. 5-7)

If a system undergoes a process (e.g. from state 1 to state 2) followed by the reverse
of that process (from state 2 to state 1, with Q
21
and W
21
equal in magnitude but
1ME Thermodynamics Sec. 6
4
opposite in direction to Q
12
and W
12
), and as a result the surroundings as well as the
system are restored to their original states, the system's original process (1 to 2) is
called a reversible process. However, if the surroundings are not in their original state
when the system has returned to its original state (state 1), the system's original
process (1 to 2) is called an irreversible process.


6.3.1 Examples of reversible and irreversible processes

(a) A ball, initially at rest, slides down a
slope, with no friction. Its potential
energy is converted into kinetic energy.
We can imagine that the process can be
reversed by arranging for the ball to roll
up a similar slope. Because we assume
no friction, all the kinetic energy will be
converted back to potential energy. The
ball will have returned to its initial state,
and the surroundings (the slopes) will
not have been affected at all. This is a
reversible process.
If there is friction, the ball will not reach its original height during the reverse
process. To get it there, work would have to be done on it by the surroundings.
In addition, the slopes will gain internal energy (their temperature will rise
slightly). Hence the surroundings will not return to their initial state, and the ball
rolling down the slope will have undergone an irreversible process.

(b) A gas is trapped in a vertical cylinder by
a frictionless piston. Initially, it is
maintained in equilibrium by a large
number of small weights on the top face
of the piston, and a slow process occurs
when the weights are removed one at a
time. This is the situation described in
Sec. 3.3.1, when we derived the
expression for displacement work in a q-
e process. If the weights are next put
back onto the piston, one by one, the
gas (the system) will retrace its process
path and return to its initial state while
the surroundings also return to their
initial state. On the other hand, if there is
friction between the piston and the
cylinder wall, the internal energy of the
surroundings will increase during both
the forward and the reverse process, so
the surroundings will not return to their
initial state.

Another irreversible process would occur if the weights were all removed at
once. In this sudden expansion, there would also be friction between adjacent
1ME Thermodynamics Sec. 6
5
streams of the gas; viscosity would cause a transfer of energy from kinetic to
internal. So an expansion process has to be fully-resisted (see Sec. 3.3.1) for it
to be reversible.
We can begin to see that that friction is a sign of irreversibility, and that reversible
means almost the same thing as quasi-equilibrium.

(c) An unresisted, adiabatic expansion
of a gas takes place, in a cylinder
with a heavy, frictionless piston. The
gas does some work in raising the
piston; much of this is converted to
internal energy of the cylinder wall
when the piston hits the stops. To
reverse this process, apply a
pressure > atmospheric above the
piston; work done on the gas in the
reverse process is > work done by
the gas in the original process, so
when the piston has returned to its



initial position, the gas has more internal energy than at the start of the original
process. To restore the gas to its initial state, remove the insulation and allow a
heat transfer to surroundings. The surroundings are thus not restored to their
initial state, so the original process must have been irreversible.


6.3.2 Reversible work and heat transfers

Work transfer to or from a system occurs only in compression or expansion
processes. For these to be reversible, the compression or expansion must be fully
resisted. The system and its surroundings depart from equilibrium only by an
infinitesimal amount, throughout the process.


W = P dV for a reversible process. W P dV for an irreversible process.


The 2nd Law (Clausius version) implies that heat transfer from a system at
temperature T to surroundings at T
0
can only occur if T > T
0
. The reverse of this
process would involve heat transfer from a cold body (at T
0
) to a hotter one (at T), so
is impossible. The only way that this heat transfer can be reversible is if T = T
0
. The
process would take an infinite time (if you believed it would happen at all!) If we allow
T to be only slightly greater than T
0
, we have a nearly reversible heat transfer
process.


Heat transfer across an infinitesimally small temperature difference is reversible.
Heat transfer across a finite temperature difference is irreversible.





1ME Thermodynamics Sec. 6
6
6.3.3 Effects which make a process irreversible

Common causes of irreversibility (which you must recognise but need not know how
to prove):
Friction, involving solids and/or fluids. Flow processes where mechanical forms of
energy are converted to internal energy, those which require an
extended Bernoulli equation (with terms for unrecoverable pressure
losses), are irreversible.
Unresisted expansion Heat transfer across finite temperature differences
Combustion Mixing between different fluids



6.4 CONSEQUENCES OF THE 2
ND
LAW FOR HEAT ENGINES
(&B Sect. 5-9 & 5-10)

A reversible heat engine is one in which all
the processes of its cycle are reversible,
as defined in Sec. 6.3. There is no such
thing in practice, because if the heat
transfers Q
h
and Q
c
were reversible, the
working fluid would be at a temperature
only infinitesimally lower than T
h
during the
process in which the heat transfer Q
h
took
place, and only infinitesimally higher than
T
c
during the process in which the heat
transfer Q
c
took place. Either it would run
infinitely slowly or the heat exchangers
(transferring energy between the
reservoirs and the working fluid) would
need infinitely large surface areas.


A reversed heat engine is one operating in
the reverse direction, so that during each
cycle there is a net work input to the
working fluid, a heat transfer from the cold
reservoir to the working fluid and a heat
transfer from the working fluid to the hot
reservoir. Reversed heat engines serve
either as refrigerators, if their purpose is to
remove energy from the cold reservoir, or
as heat pumps, if their purpose is to add
energy to the hot reservoir. Practical
refrigerators and heat pumps are not
reversible, for the same reasons as
described above for engines, but we shall
make use of some theory concerning
reversed reversible heat engines, i.e.
reversible refrigerators or heat pumps.



1ME Thermodynamics Sec. 6
7
The Clausius statement of the 2nd Law is
formally expressed as:


"It is impossible to construct a heat
engine which operates in reverse without
a net work input from the surroundings."


We are now in a position to find out what
is the most efficient type of heat engine
possible, a theoretical ideal against which
to judge the performance of any real heat
engine (or of an idealized model of a real
engine, which matches our strict
definition of a heat engine).


In the left-hand diagram below, two heat engines are operating between the same
thermal reservoirs. Engine 1 (shown as "E1") is irreversible, but engine 2 (shown as
"E2") is reversible. The thermal efficiencies, by definition, are
1
= W
1
/Q
h1
, for engine
1, and
2
= W
2
/Q
h2
for engine 2. The relative sizes of the engines are chosen so that
W
1
= W
2
= W; these are the net work transfers per cycle.

Suppose that engine 1, the irreversible one, is more efficient than engine 2, i.e.

1
>
2
. Then
W / Q
h1
> W / Q
h2
.
Because engine 2 is reversible, we can reverse it and be sure that Q
h2
, Q
c2
and W
have the same magnitude (but opposite direction) as when it was operating as an
engine. Now arrange to drive the reversed engine directly from the output of engine 1
(since the Ws are equal in magnitude). This is the arrangement in the right-hand
diagram.



The inequality W / Q
h1
> W / Q
h2
is still true (remember that the symbols now stand only for the magnitudes of the Qs
and Ws).
Therefore Q
h2
> Q
h1
. ... (i)

1ME Thermodynamics Sec. 6
8
Now apply the 1st Law to a system consisting of E1 and the reversed E2 , whose
boundary is shown shaded; note that no work crosses this boundary.
Q
h1
Q
c1
+ Q
c2
Q
h2
= 0.
Re-arranging,
Q
c2
Q
c1
= Q
h2
Q
h1
, which is > 0 ( from (i) ).
This means shows that more heat is extracted from the cold reservoir than is rejected
to it, while more heat is added to the hot reservoir than is taken from it. Thus E1 plus
the reversed E2 form a system which works in a cycle, without any work input (across
its shaded boundary), while causing a heat transfer from the cold reservoir to the hot
one. It violates the 2nd Law! What is wrong? Everything follows logically from our
assumption that the irreversible engine E1 is more efficient than the reversible engine
E2. Therefore


A reversible engine is the most efficient heat engine possible.

This is not an unexpected result, since reversible engines, whose cycles consist only
of reversible processes, must not involve friction, fluid mixing, unresisted expansions,
etc. In other words, they are loss-free.

A similar argument can be used to show that


All reversible heat engines operating between the same
two thermal reservoirs have the same thermal efficiency


We can now identify the "perfect" heat engine, but we have yet to find out what its
thermal efficiency is; this is very important because it will define an upper limit for the
efficiency of all real engines.

This maximum possible thermal efficiency cannot be more than 100%, because the
1st Law says that W = Q
h
Q
c
, so W cannot exceed Q
h
. Could it equal 100%, i.e.
could Q
c
be zero?

In the diagram on the right, E1 is an
engine which is supposed to have an
efficiency of 100%, so that Q
c1
= 0. Its size
is chosen so that it produces exactly
enough net work per cycle to drive the
reversed reversible engine E2.
As in the previous argument, the system
comprising E1 plus the reversed E2 is
seen to be removing energy from the cold
reservoir and adding it to the hot reservoir.
(Q
h2
> Q
h1
because Q
h1
= W but
Q
h2
= W + Q
c2
.)
So the 2nd Law is violated, and the
assumption that Q
c
could be zero must be
false.


1ME Thermodynamics Sec. 6
9

This, at last, leads to the alternative statement of the 2nd Law, the Kelvin-Planck
statement, which will be of more direct use to us than the Clausius statement. Thus:


It is impossible for any heat engine to receive heat from a single
thermal reservoir and to produce an equivalent quantity of work.

In other words:

A cyclically-operating engine must reject heat to a low-temperature
"sink" as well as receive heat from a high-temperature "source".

This means the same as:

No heat engine, even a reversible one, can have an efficiency of 100%.




6.5 THE EFFICIENCY OF REVERSIBLE HEAT ENGINES
(&B Sect. 5-11)
How can we evaluate it?
We showed that all reversible engines operating between the same two thermal
reservoirs (one, the heat source, at a temperature T
h
and the other, the heat sink, at
a lower temperature T
c
), have the same thermal efficiency, which is greater than that
of any irreversible engine working between the same temperatures. This means that
the efficiency must depend only on those two temperatures. Since efficiency is
defined as
th
= W
net
/ Q
h
, which equals 1 ( Q
c
/ Q
h
) (eqn 6-1 & 6-2),

Q
c
/ Q
h
= f ( T
c
, T
h
) (where f stands for "function of").

The absolute temperature scale was actually defined by Kelvin in terms of
reversible heat engines, such that T
c
/ T
h
= Q
c
/ Q
h
. (It was also necessary to
define the size of one degree on this scale by giving a numerical value to the
temperature of some reproducible physical phenomenon. The most recent
international agreement was that the one and only state where the solid, liquid and
vapour phases of water are all in equilibrium, the "triple point", should be assigned
the value 273.16 K, and that T in degrees Celsius should equal T in degrees Kelvin
minus 273.15.)

It was not the only possible definition; other equally valid forms of the function f
would have yielded different temperature scales, including some where T could tend
to minus infinity as a system's internal energy decreased. However, the merit of the
Kelvin scale is that it can be shown to be identical with a scale based on the
properties of ideal gases. This "ideal gas temperature scale" essentially says that if a
fixed mass of an ideal gas is held at constant volume, its temperature is a number
proportional to a measurement of its pressure; this is why we can write Pv = RT. In
this course, you do not need to know the details of how Q
c
/ Q
h
came to equal T
c
/
T
h
. It involves another "thought experiment", with three reversible heat engines,
1ME Thermodynamics Sec. 6
10
outlined in &B. You should, however, appreciate that the concept of absolute zero
temperature on the Kelvin scale, to and below which it is impossible to go, arises
because Q
c
can never be zero (Sec. 6.3); if T Q, then neither can T be zero.

What matters to us is that we can now state the efficiency of a reversible heat
engine as



th, rev
= 1 T
c
/ T
h
. ... (6-3)


Remember that
th
= 1 Q
c
/ Q
h
for any heat engine, but = 1 T
c
/ T
h
only for
a reversible heat engine.

If the heat transfers Q
c
and Q
h
are to be reversible, there must only be an
infinitesimally small temperature difference between the working fluid and the
thermal reservoir with which it is exchanging heat. We can therefore regard
T
c
and T
h
as the minimum and maximum temperatures, respectively, of the engine's
working fluid. Going back to the example of Drax power station, in Sec. 6.2, imagine
that the entire steam plant cycle could have been reversible, with the same
maximum (565C at the boiler outlet) and minimum (say 30C in the condenser if the
cooling water is at 20C) steam temperatures as in the actual plant. Its efficiency
would then have been 1 T
c
/ T
h
= 1 (30+273.2) / (565+273.2) = 63.8%.
Anyone who thinks that its actual cycle efficiency of around 40% is disgustingly low
should be told:

(a) that it actually achieves 40/63.8 = about 63% of the theoretically possible
efficiency;

(b) that if they wished to see more work (electrical power output) produced from
the 990 MW of heat rejected to the cooling water (for each of the six
boiler/turbine/condenser/feed pump units), they would have to find a suitable
low-temperature heat sink so that an additional cycle (a "bottoming" cycle) like
this could be set up:

There is nothing available at a lower temperature than the atmosphere, lakes,
rivers, and seas, so getting extra work out of the plant by using the rejected
heat is out of the question. We could use the rejected energy as heat, but the
opportunities to make use of a large mass flow rate of fluid at a temperature
only a few degrees above ambient are very limited. In fact, some of the
condenser cooling water at Drax is used to heat local glasshouses growing
salad crops, but the cost of transporting this "low-grade" heat further away
from the power station would be prohibitive.
1ME Thermodynamics Sec. 6
11


6.5 THE "QUALITY" OF ENERGY (&B Sect. 5-11, 2
nd
half)

The discussion above should give you the idea that a supply of heat at a high
temperature is more valuable than the same quantity of heat at a low
temperature. You can use high-temperature (or "high-grade" heat) as the heat
source for a heat engine, to produce work (as shaft work to power machinery or
generate electricity), but "low-grade" heat, at temperatures not much above ambient,
cannot sensibly be used for this purpose.

A similar idea is that work is a more valuable form of energy transfer than heat.
The left-hand diagram opposite represents
a quantity of work W being converted
completely into heat, such that Q
h
= W.
The 2nd Law does not prohibit a reversed
heat engine from exchanging heat with
only one reservoir. (Cut out the reversed
engine and this is just a representation of
friction, mechanical energy being
converted to internal energy). In the right-
hand diagram, a normally-operating heat
engine, only part of Q
h
can be converted
to W, the rest becoming the compulsory
Q
c
.





6.6 THE CARNOT ENGINE AND CYCLE
(&B Sects. 5-8 & 5-11)

How do real engines differ from reversible ones? For a start, real processes are not
reversible. For example, the boiler in a steam plant involves heat transfer across a
large temperature difference, between the combustion products (the hot gases
produced by burning the fuel) and the water/steam; the process in the steam turbine
involves friction, between adjacent layers of the viscous steam and between the
steam and the solid surfaces (blades, casing and hub). The other main difference is
that the working fluid does not take in all its heat input at a constant-temperature heat
source, and reject all its heat at another constant temperature. In a steam plant's
boiler, the water begins its heat addition process at the comparatively low temperature
at which it enters from the feed pump; as the heat addition proceeds, the water
temperature rises until it reaches saturation value. This temperature variation during
the heat transfer processes is common to most real engines (or idealized models of
them); for example, in the model of a spark-ignition (petrol) engine which you
analysed in Qu.5 of Exercises Sheet 3, the temperature rose from 420C to 3000C
during the constant-volume heat addition process. While the heat rejection in the
condenser of a steam plant is at constant temperature (the saturation temperature),
the constant-volume heat rejection in the spark-ignition engine model is obviously not.

The first idea of how, in theory, an ideal, fully reversible engine might operate is due
to the French engineer Sadi Carnot, who in 1824 proposed the following four-process
cycle, a remarkable piece of thinking considering that the 2nd Law had not then been
1ME Thermodynamics Sec. 6
12
formally stated. The system (i.e. the heat engine) undergoing this cycle may be
thought of as a gas in a piston-cylinder arrangement, but it does not have to be so.
We shall describe it in terms of a frictionless piston-cylinder arrangement where the
piston and the cylindrical walls are always thermally insulated, but the closed end of
the cylinder, the cylinder head, has removable insulation, allowing it to be brought into
contact with a heat source or sink. In state 1, the gas has its maximum volume and its
temperature T
1
equals T
c
, the temperature of a heat sink (a thermal reservoir).

Process 12: Reversible, isothermal compression. The cylinder head insulation is
removed and the heat sink is brought into contact with the head. The piston begins its
slow, inwards motion from the bottom-dead-centre position (at the crankshaft end of the
cylinder), doing work on the gas. During every small interval of time where the work
transfer is W, the heat transfer Q to to the sink is equal in magnitude to W, so the
internal energy change U is zero and the gas temperature remains unchanged at T
c
,
the reservoir temperature. The heat transfer is reversible because there is no
temperature difference between the gas and the reservoir.

Process 23: Reversible, adiabatic compression. The cylinder head insulation is replaced
and the piston completes its inwards motion, to the top-dead-centre position (at the
head end of the cylinder), doing further work on the gas without any heat transfer. In
each small time interval, W (negative value, original sign convention) now equals -U,
and the gas temperature rises. At the top-dead-centre position (where the gas has its
minimum volume), the gas temperature T
3
= T
h
, the temperature of a heat source
(another thermal reservoir).

Process 34 Reversible isothermal expansion. The cylinder head insulation is removed
again and the heat source at T
h
is brought into contact with the head. The piston begins
its slow outwards motion. During every small interval of time where the heat transfer
from the source to the gas is Q, the work transfer W from the gas is equal in
magnitude to Q, so the internal energy change U is zero and the gas temperature
remains unchanged at T
h
, the reservoir temperature. This heat transfer is also
reversible.

Process 41: Reversible, adiabatic expansion. The cylinder head insulation is replaced and
the piston completes its outwards motion, to the bottom-dead-centre position, as the
gas does further work without any heat transfer. In each small time interval, -W equals
U (original sign convention), and the gas temperature falls. At the bottom-dead-centre
position, the gas temperature has returned to T
c
.

This cycle, known as the Carnot cycle, is
shown here on a Pv diagram. A
hypothetical engine working on the cycle
is called a Carnot engine; it serves as the
unattainable target which power plant
designers strive to approach, subject
always to economic realities.

Since the cycle is composed entirely of
reversible processes and has all its heat
addition at one temperature, T
h
, and all
its heat rejection at one temperature, T
c
,
its efficiency is

th, rev
= 1 T
c
/ T
h
(eqn. 6-3).
This is often called the Carnot efficiency
and written as
Carnot
.


1ME Thermodynamics Sec. 6
13


6.7 REFRIGERATORS AND HEAT PUMPS (REVERSED HEAT ENGINES)
(&B Sects. 5-5 & 5-12)

Any cycle where the sequence of process paths goes round the Pv diagram in an
anti-clockwise (or counter-clockwise) direction has a negative value of P dv, the net
work, and is therefore a refrigerator or heat pump cycle.

"Efficiency" is not an appropriate word to
describe the performance of reversed
engines. "What we want" is Q
c
, for
refrigerators, or Q
h
, for heat pumps; "what
we have to pay" is W. From the 1st Law,
Q
h
/ W is certainly 1, because
Q
h
= W + Q
c
. Q
c
/ W may be 1.
Numbers > 1 (or 100%) can hardly be
called "efficiencies". Therefore Q
c
/ W and
Q
h
/ W are called "coefficient of
performance" (COP).




For a refrigerator, COP = Q
c
/ W = Q
c
/ (Q
h
Q
c
).
If it is reversible, so that Q
h
/ Q
c
= T
h
/ T
c
, COP
rev
= T
c
/ (T
h
T
c
).

For a heat pump, COP = Q
h
/ W = Q
h
/ (Q
h
Q
c
).
If it is reversible, so that Q
h
/ Q
c
= T
h
/ T
c
, COP
rev
= T
h
/ (T
h
T
c
).


The simplest way to handle the algebra of calculations on refrigerators and heat
pumps is often to make direct use of the 1st Law, as W = Q
h
Q
c
, and, in the
reversible case, the 2nd Law as Q
h
/ Q
c
= T
h
/ T
c
.
1ME Thermodynamics Ex. 6
1
THERMODYNAMICS EXERCISES 6
The Second Law of Thermodynamics


1. An idealized model of a diesel engine uses a fixed mass m of air as the working
fluid, in a cycle where combustion is represented by a heat transfer Q
in
to the air and
the exhaust and induction processes are represented by a heat rejection Q
out
from
the air. If Q
in
/ m = 1800 kJ/kg and Q
out
/ m = 800 kJ/kg (i.e. |Q
out
| / m = 800 kJ/kg),
what is the thermal efficiency predicted by the model?


2. The four diagrams (a) to (d) represent systems undergoing cyclic processes while
exchanging heat with one or two thermal reservoirs. The symbols Q and W here
refer to the magnitudes of the heat and work transfers during one cycle; the
directions of the energy "flow" are indicated by the arrows.
For each case, answer the following questions: Is the 1st law obeyed? Is the 2nd
Law obeyed? If both are obeyed, what kind of device or process does it represent?





3. What is the theoretical maximum thermal efficiency obtainable from each of the
following power plants, in which Tmax and Tmin are the maximum and minimum
temperatures of the working fluid?
(a) A gas turbine engine with T
max
= 1500K and T
min
= 750K
(b) A steam plant with T
max
= 565C and T
min
= 30C
(c) A combined-cycle plant (gas turbine plus steam plant, with a "waste heat" boiler
using the energy in the gas turbine's exhaust to raise steam, as illustrated in the
Introduction to the 1M Thermo-Fluids course), with T
max
from (a) and T
min
from
(b).
1ME Thermodynamics Ex. 6
2
(d) The steam plant in this Department's Thermo-Fluids Teaching Laboratory, with
boiler pressure 7 bar gauge and condensate storage tank temperature 60C.
This boiler produces wet steam so T
max
= T
sat
at boiler pressure, and T
min
for the
cycle occurs in the condensate storage tank.

If the combined-cycle plant in (c) used the gas turbine in (a), could the maximum
steam temperature in the waste heat boiler equal that of the steam plant in (b)?


4. The temperature in the food compartment of a domestic refrigerator is to be kept at
5C when the kitchen where it is situated is at 18C. The average rate of heat
transfer from the kitchen into the food compartment is 1 kW (caused more by
opening the door than by poor insulation), so this is also the average rate of heat
transfer from the food compartment to the working fluid (1st Law applied to the food
compartment).
(a) Calculate the coefficient of performance (COP) of a reversible refrigerator (a
reversed reversible heat engine) under these conditions, and hence the
minimum possible power input from the motor.
(b) If the actual refrigerator has a COP of 15% of the ideal, reversible one, calculate
the actual power input required, the rate of heat rejection from the working fluid
to the kitchen (via the heat exchanger at the rear of the refrigerator), and hence
the kitchen's net rate of energy gain.


5. (a) In the early 1990s, advertising by the electricity distribution companies in the
U.K. attempted to persuade us to "Heat Electric". What is wrong, thermodynamically
if not commercially, with using electricity to heat houses?
(b) A house heated by electric heaters requires an electricity (i.e. electrical work)
input of 10 kW in order to maintain the inside temperature at 20C when the outside
temperature is 5C. It is proposed to replace the electric heaters with a heat pump
(thermodynamically a good idea, since Q
h
= W + Q
c
instead of just W ). A suitable
electrically-driven heat pump has a coefficient of performance 20% of that of a
reversible one. The price of electricity is 7.5 pence per kWh.
(i) What is the running cost of the original electric heating, in pence per hour?
(ii) What is the COP of the proposed heat pump and the power input it
requires?
(iii) What is the running cost of the proposed heating scheme, in pence per
hour?
(iv) How would you decide whether to replace the electric heaters with the heat
pump? (Think what contributes to the total cost: purchase (probably with a
loan), running, maintenance, depreciation)


6. A building requires a heat supply of 20 kW at 85C for its heating system. A nearby
factory has available for sale, at a price of 3 pence/kWh, the waste heat from a
chemical process, at a temperature of 200C. Two schemes for using this waste
heat are being evaluated, as follows.
(a) Use the waste heat directly.
1ME Thermodynamics Ex. 6
3
(b) Use the waste heat as the high-temperature energy source for a heat engine,
which will reject heat to a river at a temperature of 10C. The engine will drive a heat
pump which will take heat from the river and reject heat at 85C to the building.

(c) Sketch and label a block diagram of the proposed arangement (b) (in terms of
engines and thermal reservoirs). Assuming that the heat engine and heat pump are
both reversible, calculate the running costs of the two schemes. Hint : it is simpler
not to introduce efficiency and coefficient of performance into the calculation; just
use the 1st Law and the 2nd Law (or rather its consequence Q
h
/T
h
= Q
c
/T
c
for a
reversible engine or heat pump).

______________________________________________________________________
ANSWERS
1. 55.6%
2. (a) Yes, Yes, Engine (b) Yes, Yes, Refrigerator or Heat Pump (c) Yes, No (d) Yes,
Yes, Friction
3. (a) 50.0% (b) 63.8% (c) 79.8% (d) 24.9% No; cannot transfer any
more heat from the
exhaust gases at 477C to the water/steam once the steam has reached this
temperature (2nd Law)
4. (a) COPrev = 21.4, min. power input = 0.0467 kW (b) Actual power input =
0.312 kW, rate of heat rejection from working fluid = 1.312 kW, kitchen's net rate of
energy gain = 0.312 kW
5. (b) (i) 75 p/h (ii) COP = 2.35, power input = 4.26 kW (iii) 32.0 p/h
6. Direct use: 60 p/h Using engine + heat pump: 31.3 p/h
1ME Thermodynamics Struc. Tut. 3
1

PROBLEM FOR STRUCTURED TUTORIAL

2nd Law of Thermodynamics: Efficiency of Reversible Engines
_______________________________________________________________________

Thermal efficiency
th
is defined for all heat engines as W
net
/ Q
h
, = 1 Q
c
/ Q
h
.
In Sec. 6.5 of the Thermodynamics lecture notes, the thermal efficiency of any reversible
heat engine operating between a heat source at temperature T
h
and a heat sink at
temperature T
c
is shown to be

th
= 1 T
c
/ T
h
,
simply by accepting that Kelvin's absolute temperature scale is defined by the reversible
engine relationship
Q
c
/Q
h
= T
c
/ T
h
.
The solution to the problem below may provide a more convincing demonstration,
although it is restricted to a system consisting of a perfect gas.

By using the 1st Law and relationships for perfect gas properties (including the process
path relationship for slow, frictionless, fully-resisted, adiabatic compression or expansion
the example on p.7 of the Sec. 4 notes), show that if 1 kg of a perfect gas in a piston-
cylinder arrangement executes a Carnot cycle, the thermal efficiency is 1 T
c
/T
h
.


Q.5 on Exercises Sheet 3 concerned the idealized model for a spark-ignition (petrol or
gasoline) reciprocating engine. The net work per cycle, W
net
, was 2094 J and if the heat
addition had been asked for, it would have been found as Q
23
= m C
v
(T
3

T
2
) = 3591 J. The cycle thermal efficiency was therefore W
net
/ Q
23
= 0.583 (or 58.3%).
What would be the cycle efficiency for a Carnot engine having the same maximum and
minimum temperatures, 3000C and 15C?
Why is the efficiency of the idealized spark-ignition engine cycle (known as the Otto cycle)
much less than the Carnot efficiency, although all the processes in the Otto cycle are what
we would now call reversible?
1ME Thermodynamics Sec. 7
1
7. CONSEQUENCES OF THE 2ND LAW OF THERMODYNAMICS


In order to analyse the performance of a real power plant, whose efficiency is less than
that of a reversible heat engine, or to design one to achieve a given performance, we
shall make use of a property called entropy, due to Prof. R. Clausius of Berlin who
provided the first formal statement of the 2nd Law (Sec. 6.1 & 6.4). Entropy allows a
real, irreversible process to be compared with its ideal, reversible equivalent.
Before defining entropy and showing how it is related to the other properties of a
substance, we need to look at a relationship called the Clausius inequality, involving the
heat transfers to or from a heat engine and the temperatures of the working fluid while
those heat transfers are taking place.

7.1 THE CLAUSIUS INEQUALITY (&B Sect. 6-1)

In Sec. 6, we concentrated on heat engines exchanging heat with only two thermal
reservoirs. If they were to be reversible, the working fluid had to be at a temperature
only infinitesimally different from that of the reservoir during the heat transfer process.
Real engines have heat addition or heat rejection processes where the working fluid
temperature changes as the process occurs, e.g. in a steam plant's boiler while the
water is being heated to saturation temperature. To be reversible, such a process would
need an infinitely large number of thermal reservoirs, one at temperature T, the next at
T + dT, the next at T + 2 dT, and so on. In a heat addition process, the system would
have to make contact with the first reservoir while the working fluid temperature rose
from T to T + dT, with the next reservoir while the temperature rose from T + dT to
T + 2 dT, etc. The total Q
h
would then be the sum of an infinitely large number of heat
transfers dQ
h
, each occurring at a different value of T. Total Q
c
would be a similar
sum.
(It is important to remember that Q is not a continuous function of any property of the system. The system
does not have a "value of Q" in every state, whereas it does have a unique value of properties like v, u, h
etc. Therefore "dQ" must only be regarded as a very small part of a total heat transfer Q. Mathematically,
it is not like du, for example, which does mean a very small change in the property u.)
We could write the sum of all these small parts of the heat addition, round the whole
cycle, as

h
dQ , and the sum of both the heat addition and heat rejection as

dQ =

h
dQ

c
dQ . (The circle on the integral sign denotes the summation over one
complete cycle, and we are still using Q to represent only a magnitude, hence the minus
sign in front of

c
dQ .)

The Clausius inequality concerns another "cyclic integral",

(dQ/T) . Let us evaluate

(dQ/T) for a reversible heat engine using only two reservoirs.



(dQ/T) = (dQ
h
/T
h
) (dQ
c
/T
c
)

= (1/T
h
) (dQ
h
) (1/T
c
) (dQ
c
)

= Q
h
/T
h
Q
c
/T
c
= 0

1ME Thermodynamics Sec. 7
2
Now let us repeat this for an irreversible engine which has the same Q
h
as the
reversible one. The irreversible engine has a lower efficiency (Sec. 6.4), so W will be
smaller and therefore Q
c
will be larger than for the reversible engine.

(dQ/T) (1/T
h
) (dQ
h
) (1/T
c
) (dQ
c, irrev
)

The last term has a larger magnitude than in the reversible case, so for the irreversible
engine

(dQ/T) < 0.

The combination of these two results is the Clausius inequality:


(dQ/T) 0 , "=" for a reversible cycle and "<" for an irreversible cycle ... (7-1)


It states that when a system undergoes a cycle, the sum of {the heat transfers (positive
to the system and negative from it) divided by the system boundary temperature at
which they take place} is zero if all the processes in the cycle are reversible, or less than
zero if they are not.
Note that T refers to the working fluid temperature rather than the thermal reservoir
temperature; these are the same for a completely reversible engine, but the Clausius
inequality holds also for an engine where all the processes inside the system boundary
are reversible but the heat transfers with the reservoirs take place across finite
temperature differences.

We have demonstrated its validity for a heat engine operating between two thermal
reservoirs only, but it is true also for cases where heat transfer occurs over a range of
temperatures. The formal proof is rather tedious, and you do not need to know it. &B
do not give it.


7.2 THE PROPERTY ENTROPY (&B Sects. 6-2 & 6-3)

Important: from now on, we will use our original sign convention for heat transfer
(+
ve
to the system,
ve
from it) and for work transfer (+
ve
from the system,
ve
to it).

7.2.1 Definition

Consider a system undergoing a reversible
process 'A' (see Pv diagram) from state 1
to state 2, then completing a cycle by
returning to state 1 in another reversible
process 'B'. From the Clausius inequality
2 1
(dQ/T) + (dQ/T) = 0
1 2
via A via B



If process 'B' is replaced by a different reversible process 'C', (dQ/T) via A + C still
equals zero, so the value of (dQ/T) from state 2 to state1 depends only on the initial
1ME Thermodynamics Sec. 7
3
and final states, not on the particular reversible process path between these states.
(dQ/T)
rev
must therefore represent the change in value of a property of the system
(just as U
1
U
2
, the change in internal energy between states 2 and 1, does not
depend on the process path taken).

This property is called entropy, with the symbol S, and is defined by

2
dS = dQ
rev
/ T , so that S
2
S
1
= (dQ
rev
/

T) ... (7-2)
1

Note that the process 1 2 has to be reversible, unlike any relationships for property
change that you have met before. If you are keen to relate this concept of entropy to
those you may have met before, concerning "disorder", "randomness", "chaos" etc.,
start by reading the relevant sections in &B. We shall not attempt to cover this in the
present course.

The entropy S of a system is an extensive property, with units J/K or kJ/K. The specific
entropy s, = S/m where m is the system's mass, is an intensive property with units
J/kg K or kJ/kg K.

Since real processes are not reversible, how is the entropy change of a system, or of a
fluid flowing through a control volume, actually calculated? In principle, one can
imagine a reversible process which ends in the same state as the actual process, then
use eqn 7-2 for the imaginary process. &B give an example of this (Sect. 6-1) and
older textbooks often devoted several pages to ingenious imaginary processes which
allowed eqn 7-2 to be used for finding S
2
S
1
. In practice, we do not often evaluate
entropy changes in this way. Instead, we relate entropy to other properties. For steam,
this is too complex to be done every time, just as it is for enthalpy, etc., so we use
tables; you will have noted the columns for specific entropy. For perfect gases,
algebraic relationships are used, and these will be derived in Sec. 7.4. Before that, we
shall introduce a fundamental relationship for entropy change in a general process
(reversible or not), which has philosophical as well as engineering implications; &B
call it the "increase of entropy" principle.

7.2.2 Entropy change in reversible and irreversible processes

Consider three processes between states 1
& 2. 'A' and 'B' are reversible but 'C' is
irreversible. 12 via A followed by 21 via B
is a reversible cycle, so the Clausius
inequality gives
2 1
(dQ/T) + (dQ/T) = 0 ... (i)
1 2
via A via B

12 via A followed by 21 via C is an
irreversible cycle. So





1ME Thermodynamics Sec. 7
4
2 1
(dQ/T) + (dQ/T) < 0 ... (ii)
1 2
via A via C

Subtracting (i) from (ii) and re-arranging,

1 1
(dQ/T) > (dQ/T)
2 2
via B via C

1
But B is reversible, so (dQ/T) = S
1
S
2
. Therefore in a general process,
2
via B


dS (or S
final
S
initial
) dQ / T ("=" if reversible, ">" if irreversible) ... (7-3)


For an adiabatic process, Q = 0, so eqn 7-3
says
S
final
S
initial
= 0 if reversible,
S
final
S
initial
> 0 if irreversible.

If we had written eqn 7-3 for any infinitesimally small
part of the process, i.e. without the integral signs, the
result above for a reversible process would have
been that S is constant throughout a reversible
adiabatic process. A process where entropy is
constant is called "isentropic". This is illustrated here
on a state diagram, plotting S on the horizontal axis.
Such diagrams, with T or h as the ordinate, will be
covered in more detail in Sec. 7.5.
An important point to remember is:


A reversible and adiabatic process is isentropic. If any two of these descriptions
apply to a process: "reversible", "adiabatic", "isentropic", then the third also
applies.


Even when heat transfer does occur across a system boundary, it is possible to
imagine a larger system which includes what we had previously regarded as the
surroundings, so that no heat transfer occurs across the boundary of the larger system,
sometimes called an isolated system.


Thus the total entropy of a system plus its surroundings increases whenever
an irreversible process occurs. The entropy of the system itself may decrease,
but in that case the entropy of the surroundings will increase by a larger amount.

1ME Thermodynamics Sec. 7
5
For a process with heat transfer Q, eqn 7-
3 says that the system entropy will increase
if Q is +ve, and will increase by more in an
irreversible process than in a reversible
one. If Q is ve, the system entropy will
decrease if the process is reversible, but
for an irreversible process, eqn 7-3 does
not predict whether S
final
is > or < S
initial
; it
depends which effect "wins", ve Q trying
to decrease S or irreversibility (friction,
etc.) trying to increase S.


Whatever the magnitude and sign of Q, the results above can be summarised by
noting that:


In an irreversible process, the final state is further to the right on a T s
(or h s) diagram than in a reversible process with the same Q.



From eqn 7-2, if we know what a reversible process path looks like on a T S diagram
we can evaluate Q as the area under the process path, T dS, which also equals
m T ds, where m is the mass of the system This corresponds with evaluating W as the
area under a reversible process path on a P V diagram.
For a reversible cycle, the area enclosed on a T S diagram is Q
net
= Q, just as the
area enclosed on a P V diagram is W
net
= W. Because W
net
= Q
net
for a cycle,
the area enclosed on the T S diagram equals that enclosed on the P V diagram.


Q = T dS Q
net
= Q = T dS


1ME Thermodynamics Sec. 7
6
The Carnot cycle consists of two
isothermal processes with heat transfer
and two reversible adiabatic (and therefore
isentropic) processes. It is more instructive
to plot this on a T s diagram
(temperature vs specific entropy) than on
the P v diagram shown in Sec. 6.6.
Any Carnot cycle is a rectangle on a T
s diagram.




7.3 TWO IMPORTANT THERMODYNAMIC RELATIONSHIPS:
THE "T ds EQUATIONS" (&B Sect. 6-7)

Consider a system of unit mass, going through an infinitesimally small part of a
reversible process, during which the heat transfer is dQ and the work transfer dW. The
1st Law is then

dQ dW = du .

Because the process is reversible, dQ = T ds (eqn 7-2) and
dW = P dv (Sec. 6.3.2). Substituting these in the 1st Law equation above, and taking
P dv to the r.h.s.,


T ds = du + P dv ... (7-4)


Eqn 7-4, sometimes called "the 1st T ds equation", can be divided by T then integrated
over the whole of the reversible process to give

s
2
s
1
= ds = [ du / T ] + [ P dv / T ] ... (7-5)
The only quantities appearing in eqn 7-5 are properties, i.e. quantities whose values
depend only on the state, not (as with Q and W) on the process by which the system
arrived in the state. Therefore eqn 7-5 applies to a reversible or an irreversible
process, with any substance. This is important because it can be used to relate s to
other properties.
Using the definition of enthalpy, h = u + Pv , dh = du + d(Pv) = du + P dv + v dP ,
eqn 7-4 can also be written in the following form, sometimes called "the 2nd T ds
equation":


T ds = dh v dP ... (7-6)


The equivalent of eqn 7-5 is
s
2
s
1
= ds = [ dh / T ] [ v dP / T ] ... (7-7)
which also applies to both reversible and irreversible processes with any substance.


1ME Thermodynamics Sec. 7
7
7.4 ENTROPY CHANGE IN PROCESSES WITH PERFECT GASES
(&B Sect. 6-9)

For an ideal gas, du = C
v
dT (definition of C
v
) and P/T = R/v (eqn of state),
so the integrated T ds eqn, eqn 7-5, becomes
s
2
s
1
= [ C
v
dT / T ] + [ R dv / v ]
For a perfect gas, C
v
does not depend on T so it can come outside the integral sign,
giving


s
2
s
1
= C
v
ln (T
2
/T
1
) + R ln (v
2
/v
1
) ... (7-8)


This gives the specific entropy change in any process where a perfect gas goes from a
state specified by (T
1
, v
1
) to a state specified by (T
2
, v
2
). Two alternative forms of
eqn 7-8 can be found using the eqn of state:
Pv = RT , so (P
2
/P
1
) (v
2
/v
1
) = (T
2
/T
1
)
ln (P
2
/P
1
) + ln (v
2
/v
1
) = ln (T
2
/T
1
) .
Also, C
p
C
v
= R . Substituting in eqn (7-8) gives versions for use when T and P or v
and P are known instead of T and v :


s
2
s
1
= C
p
ln (T
2
/T
1
) R ln (P
2
/P
1
) ... (7-9)
s
2
s
1
= C
p
ln (v
2
/v
1
) + C
v
ln (P
2
/P
1
) ... (7-10)


Eqn (7-10) is probably easier to remember than (7-8) and (7-9), which can be derived
from (7-10) if they have been forgotten.

By setting s
2
s
1
= 0 in eqn 7-10, we can find the equation describing an isentropic
process path on a P v diagram. Dividing by C
v
,

0 = ln (v
2
/v
1
) + ln (P
2
/P
1
) = ln [(v
2
/v
1
)

(P
2
/P
1
)]
(v
2
/v
1
)

(P
2
/P
1
) = 1 ,
P
2
v
2

= P
1
v
1



The same will apply to any small part of the process, so the process path is given by
Pv

= const. This is the result obtained in Sec. 4.9, for a perfect gas undergoing a slow,
frictionless, fully-resisted, adiabatic expansion or compression. We should now
recognise that "slow, frictionless, fully-resisted" means "reversible". To describe a
reversible adiabatic process path for a perfect gas in terms of T and v, or of T and P,
just substitute from the equation of state Pv = RT into the process equation Pv

=
constant (or alternatively set s
2
s
1
= 0 in eqn 7-8 or 7-9). The three versions of this
result are as follows (where the constants are different in each case).
1ME Thermodynamics Sec. 7
8


for a perfect gas in a

P v

= const,

T v
- 1
= const,

T / P
( - 1)/
= const reversible adiabatic


process ... (7-11)


Eqns 7-8 to 7-11 do not apply to steam! As for other properties, use tables to find s for
steam.


7.5 T s DIAGRAMS FOR STEAM AND FOR PERFECT GASES
(&B Sect. 6-5)

On a T s diagram for steam, the
saturation line has a "bell" shape, unlike
the heavily skewed shape on a P v
diagram.

Lines of constant pressure resemble
those on a T v diagram.

L is the subcooled (or compressed)
liquid region, V is the superheated
vapour region and L+V (liquid + vapour)
is the wet steam region.

On a T s diagram for a perfect gas
(with the saturation line somewhere in
the bottom left-hand corner, too small to
be seen), lines of constant pressure are
exponential curves. This follows from
eqn 7-6, by putting dP = 0 and writing dh
as C
p
dT. The important thing to
remember is that the isobars diverge
as T and s increase; it is this effect
which allows us to get net positive work
out of a gas turbine engine, as we shall
see in Sec. 8. Constant-volume lines are
similar, but with a larger gradient, i.e.
they are steeper.
1ME Thermodynamics Sec. 7
9
7.6 ISENTROPIC PROCESSES AND ISENTROPIC EFFICIENCY

7.6.1 Isentropic Processes and their use (&B Sect. 6-9, 2
nd
half))

Isentropic (i.e. reversible adiabatic) processes serve as ideal models of the real
processes taking place in pumps, compressors, nozzles, turbines, etc. They are useful
models because these real processes are often nearly adiabatic (any heat transfer is
small compared with the enthalpy change, for example) and irreversible effects such as
friction are not too large. If the initial state is known (two independent properties), but
only the final pressure, we can model the process as isentropic and say that the final
specific entropy is the same as the initial value. Then the final state is fixed since
pressure and specific entropy are two independent properties. With the initial and final
states known, we can apply the 1st Law (for a system or control volume) to find the
work transfer, kinetic energy change, etc. This is what was hinted at in the first two
paragraphs of Sec. 6.1.
The exact procedure depends on whether the working fluid is steam or a perfect gas.
Consider a case where the initial pressure and temperature are known, for example at
the entry to a turbine, but only the pressure is known in the final state, at the turbine
exit.
For steam, we would look up the initial specific entropy in the tables, and equate the
final specific entropy to this value. If this value were greater than the saturated vapour
value s
g
at the final pressure, the final state would be superheated, and we would have
to look in Table 3 to find the temperature at which this entropy value appeared in the
column for final pressure. If the s value were less than s
g
, we would use the equation
s = x s
g
+ (1 x) s
f
to find the final dryness fraction x.
For a perfect gas, e.g. air modelling the gases in the turbine of a gas turbine engine,
we would find the final temperature directly from the third of equations 7-11 (or eqn 7-9
with s
2
s
1
= 0).

7.6.2 Isentropic Efficiency
(&B Sect. 6-12, where it is called "adiabatic efficency")

This is a measure of how irreversible is the flow process in a real compressor or pump
or turbine. It defines the actual work transfer rate (power) relative to what it would have
been in an equivalent isentropic process. It assumes that both the real and the
isentropic processes are adiabatic, and that kinetic energy changes are negligible, so
that the SFEE gives W/m as h
1
h
2
. The equivalent isentropic process is taken to be
one starting at the same initial state, state 1, as the real process and ending at the
same final pressure, P
2
.
The symbol for isentropic efficiency is
s
and the definitions (for this course) are as
follows.


For a turbine,
s
= (h
1
h
2
) / (h
1
h
2s
) ... (7-12 a)
For a compressor or pump,
s
= (h
2s
h
1
) / (h
2
h
1
) ... (7-12 b)

s
for a compressor is the reciprocal of that for a turbine in order that the numerical
value of something called an "efficiency" should be less than 1. (Specific work, = h
1

1ME Thermodynamics Sec. 7
10
h
2
, is less for the real turbine than for the isentropic one, but the magnitude of the
specific work for the real compressor is greater than that for the isentropic one. With
irreversibilty, you always lose !). In each definition, state 2 is the final state in the real
process and state 2s is the final state in the isentropic process. P
2s
is the same as P
2
(see above) and we never need to write it as P
2s
. The idea of isentropic efficiency is
most easily understood on an h s diagram (which will be covered in more detail in
Sec. 7.7).

Turbine :




Compressor :





Typical values of
s
: large steam turbine 80 90%
turbine of large gas turbine engine 85 95%
compressor of large gas turbine engine 85 90%
1ME Thermodynamics Sec. 7
11

7.7 ENTHALPY-ENTROPY DIAGRAMS (&B Sects. 6-3 & 6-5)

In Sec. 7.6, enthalpy-entropy (hs) diagrams were seen to provide a convenient
method of visualizing the isentropic efficiency of an adiabatic flow process with shaft
work transfer. They are commonly used because the vertical dimension of a process
path ( | h
2
h
1
| ) represents the magnitude of the specific work transfer (neglecting
changes in kinetic and potential energy) and the horizontal dimension ( s
2
s
1
)
represents the extent of the irreversibility.

7.7.1 hs diagram for a perfect gas

For a perfect gas, h
2
h
1
= C
p
(T
2
T
1
) , or h = C
p
T since the arbitrary datum for
internal energy can be taken as absolute zero temperature, and h = u + Pv = u + RT so
the datum for h is then the same. An hs diagram for a perfect gas is therefore just a
Ts diagram with a different vertical scale; there is no difference when you sketch
them.

7.7.1 hs diagram for water/steam
(Scale chart provided to accompany steam tables)

Because h is not proportional to T for water/steam (as well as being non-linear, it
depends on P), the hs diagram looks like a distorted Ts diagram. The main
differences are as follows.
(a) The saturation line "bell" is pushed over to the left and the critical point is no longer
at the top of the bell (look at Table 2 to see that h
g
has its maximum value at about 30
bar, not at the critical pressure); the subcooled liquid region is not shown on most hs
diagrams.
(b) The lines of constant pressure in the wet steam region (under the bell) are no
longer horizontal (it is only temperature, not enthalpy, which stays constant during
constant-pressure evaporation), and they diverge as s increases. Feature (b) follows
from the 2nd T ds equation:

T ds = dh v dP. For an isobar, dP = 0, and for wet steam, T = constant = T
sat

T
sat
ds = dh and the gradient
dh
/
ds
= T
sat
.

So the isobars under the bell are straight lines, with steeper gradient as T
sat
or P
increase. However, for superheated steam the isobars are curved, similar to those for
perfect gases. Isotherms obviously coincide with isobars in the wet steam region, but
far into the superheated vapour region they become nearly horizontal, as the vapour
approaches ideal gas behaviour.
1ME Thermodynamics Sec. 7
12


1ME Thermodynamics Ex. 7
1
1M Thermo-Fluids THERMODYNAMICS EXERCISES 7
Entropy and its uses



1. The features of a Carnot engine which ensure that it has the maximum possible
thermal efficiency, 1 (T
min
/T
max
), are: (a) all the processes are reversible; (b) all
the heat addition takes place at T
max
and all the heat rejection at T
min
, where T
max

and T
min
are the maximum and minimum temperatures of the working fluid in the
cycle. In many real power plant cycles, the irreversibilities are not very large and yet
the thermal efficiency is much less than the Carnot efficiency.

To demonstrate that this is partly a
result of heat transfer occurring over
a range of working fluid temperature,
consider the reversible cycle plotted
on this T s diagram, which differs
from a Carnot cycle only in process
34. Calculate its efficiency and
compare it with that of a Carnot
engine having the same maximum
and minimum temperatures.
(Hint: evaluate the heat transfers as
shown near the end of Sec. 7.2 of the
lecture notes.)



2. Carbon dioxide in a piston-cylinder arrangement has an initial volume of 0.05 m
3

and an initial temperature of 300 K. It is compressed isothermally (i.e. heat is
transferred from the gas such that its temperature remains constant) until the
volume is 0.01 m
3
.

(a) Starting from the relationship T ds = du + P dv and taking carbon dioxide
as an ideal gas,show that the change in specific entropy in this process is
given by

s
2
s
1
= R ln (v
2
/ v
1
) .

(b) Evaluate s
2
s
1
.

(c) If the magnitude of the heat transfer from the gas to the surroundings is
100 kJ/kg, is the process reversible?
(Hint: look at eqn 7-3 in the lecture notes, and remember that ">" means
"closer to zero" for quantities which have negative values.)

(d) What is the specific work done by the piston on the gas?





1M TF Thermodynamics Ex. 7
2
3. Air, which can be treated as a perfect gas, is compressed in the compressor of a
gas turbine engine, from 1 bar, 15C to 12 bar.

(a) Assuming that the compression is isentropic, sketch the process path on a
T s diagram, including the isobars for 1 bar and 12 bar.

(b) Calculate the final temperature in the isentropic process.
(From the lecture notes, either eqn 7-11 or eqn 7-9 is applicable.)

(c) If the compression were not isentropic but could still be considered
adiabatic, with a final temperature of 350C, state whether the process
would be reversible or irreversible, sketch it on the same T s diagram,
and calculate the specific entropy change of the air.


4. Saturated steam at a pressure P1 of 20 bar (absolute) expands reversibly and
adiabatically in a turbine, to a pressure P2 of 0.6 bar (absolute).

(a) Find the turbine exit specific entropy s
2
, dryness fraction x
2
and specific
enthalpy h
2
.

(b) What is the turbine power output if the steam mass flow rate is 2 kg/s and
the kinetic energy change is negligible?

The steam leaving the turbine is condensed, with a negligible pressure drop.
The condenser exit state, state 3, is on the saturated liquid line.

(c) From the "2nd T ds equation", T ds = dh v dP , show that the entropy
change in the condenser is given by
s
3
s
2
= (h
3
h
2
) / T
sat


(d) Calculate the heat transfer from the steam in the condenser, for a 2 kg/s
mass flow rate,
(i) using the SFEE (1st Law), asuming negligible kinetic energy change,
(ii) using the 2nd Law equation Q = T dS , or Q/m = T ds, with s
3
taken
from Tables.

(Note that the results from (i) and (ii) are the same because the assumption of
negligible pressure drop indicates the absence of friction, and the constant
temperature during condensation implies that the temperature difference necessary
for the heat transfer to take place is all outside the control surface (in the pipe walls
and the cooling water rather than in the condensing steam); the process is therefore
assumed to be internally reversible, strictly all that is necesary for Q = T dS to be
true.)

(e) Sketch a T s diagram showing the saturation lines, the 20 bar and 0.6 bar
isobars and the processes 12 and 23.


1M TF Thermodynamics Ex. 7
3
5. In Question 3(c), air was compressed adiabatically but non-isentropically in the
compressor of a gas turbine engine, from 1 bar, 15C to 12 bar, 350C.
What was the isentropic efficiency of the compressor? (Take air to behave as a
perfect gas.)


6.

The purpose of a turbo-supercharger, or
just a turbocharger, is to boost the air
pressure at the intake of a reciprocating
engine so that, as a result of the
increased density, a greater mass of air is
taken into the cylinders during each
induction stroke. This means that more
fuel can be burned, thus increasing the
engine's power. The work required by the
compressor is provided by a turbine which
is driven by the engine's exhaust gases.
(The turbocharger consists of the
compressor, the turbine and the shaft
connecting them; it is rather like a gas
turbine engine, but has no combustion
chamber of its own and the turbine has to
produce only enough power to drive the
compressor.)

(a) What is the advantage of a turbocharger over its predecessor, the
mechanically-driven supercharger consisting only of a compressor driven from the
engine's crankshaft?
(b) In a particular turbocharger, the conditions at the engine's exhaust (the
turbocharger turbine inlet) are P
1
= 1.7 bar, T
1
= 650C, and the turbine exit
pressure P
2
is atmospheric, which can be assumed here to be 1 bar. If the turbine
has an isentropic efficiency of 75%, the process in the turbine can be assumed to
be adiabatic and the exhaust gases behave as a perfect gas with specific heats
ratio = 1.40, what is the temperature T
2
of the gases at the turbine outlet?

7. Air, which may be modelled as a perfect gas with the properties given in Table E5,
flows steadily and adiabatically through a compressor which has an isentropic
efficiency
s
of 0.80. (Reminder: the words "flows steadily" mean that a control
volume analysis is appropriate, and the correct version of the 1st Law is the SFEE.)
At the compressor inlet, the air is at 1 bar, 280 K and has a mean flow speed of 250
m/s. At the compressor exit, the pressure is 8 bar and the mean speed is 150
m/s.The air then enters a cooler (a heat exchanger) where it is cooled at constant
pressure to a temperature of 280 K; the cooler inlet and outlet have the same cross-
sectional area as the compressor outlet.

(a) Sketch a block diagram of the two components, using conventional symbols,
and label it with state numbers; add the appropriate symbols and values for all the
known air properties.
(Reminder: you should do this even without being asked.)
1M TF Thermodynamics Ex. 7
4
(b) Sketch the compression and cooling processes on a T s diagram. Add the
imaginary isentropic process 1 2s which is used in the definition of isentropic
efficiency.
(c) Find the air temperature T
2s
at the end of the imaginary isentropic process, then
use the given isentropic efficiency to find the actual compressor exit temperature T
2
.
(d) Calculate the specific work transfer W/m to the air in the compressor.
(Hint: the 2nd Law was used implicitly in (c) to find T
2
; it will be of no
further use, so back to the 1st Law!)
(e) Calculate the mean flow speed of the air at the cooler outlet.
(Hint: which fundamental conservation law, and which feature of a perfect
gas, have not yet been used?)
(f) Find the heat transfer from the air in the cooler, per unit mass of air.


8. In Question 4(a), steam expanded adiabatically in a turbine from 20 bar, saturated,
to 0.6 bar. Suppose now that that turbine had an isentropic efficiency of 0.80
instead of 1.
(a) What would be the exit specific enthalpy h
2
and exit dryness fraction x
2
?
(b) What would be the change in specific entropy, s
2
s
1
?
(c) Find the power output, again for a mass flow rate of 2 kg/s and negligible
kinetic energy change.
(d) Verify that the result in (c) is 80% of the isentropic turbine's power, Q.4(b).
Provided that the K.E. and P.E. changes are negligible, turbine isentropic
efficiency can be interpreted as the ratio of the actual power (or specific work)
to the power (or specific work) from an imaginary isentropic turbine having the
same initial state and same final pressure. In some texts, that is how isentropic
efficiency is formally defined, but difficulties arise with that definition when we
have to deal with significant K.E. changes. You should use only the definition
given in Sec. 7.6.2 of the lecture notes.


1M TF Thermodynamics Ex. 7
5
ANSWERS

1. 0.667, compared with Carnot efficiency of 0.750

2. (b) 0.3042 kJ/kg K (c) No (d) 100 kJ/kg (For the gas, W / m = 100
kJ/kg)

3. (b) 313C (c) Irreversible; +0.0658 kJ/kg K

4. (a) s 2 = 6.340 kJ/kg K, x 2 = 0.8135, h 2 = 2225.3 kJ/kg
(b) 1147 kW (d) () 3731 kW

5. 0.890

6. (b) 553C (If your answer was 520C, you forgot that the process wasn't
isentropic.)

7. (c) T 2s = 507.2 K, T 2 = 564.0 K (d) (-)266.8 kJ/kg (e) 74.5 m/s
(f) (-)295.3 kJ/kg

8. (a) h 2 = 2340 kJ/kg, x 2 = 0.864 (b) +0.319 kJ/kg K (c) 917 kW

1ME Thermodynamics Struc. Tut. 4
1

PROBLEM FOR STRUCTURED TUTORIAL

Use of isentropic efficiency and 2nd Law relationships
_______________________________________________________________________

An aircraft jet engine consists of a compressor, combustion chamber, turbine and nozzle,
as shown in Fig. 1. Unlike a land-based gas turbine engine which provides a net shaft
power output, the aim here is to produce thrust, by accelerating the turbine exhaust
through the nozzle (often called the jet-pipe). The turbine's only purpose is to drive the
compressor.



In a stationary test of a particular engine, the flow conditions are:

Compressor inlet: P
1
= 1 bar (absolute), T
1
= 20C
Compressor outlet: P
2
= 10 bar (absolute), T
2
= 344C
Turbine inlet: T
3
= 1100 K
Nozzle exit: P
5
= 1 bar (absolute)

The pressure drop in the combustion chamber is negligible. All kinetic energies are
small except at the nozzle exit. The isentropic efficiency of the process in the turbine is
0.90. The process in the nozzle can be assumed to be reversible. The fuel mass flow rate
is small compared with the air mass flow rate. The combustion process can be modelled
as heat addition to the air. (It is actually quite a good approx-imation to take the working
fluid as air throughout, because gas turbines must use much more air than is necessary to
supply oxygen for fuel combustion, in order to keep the maximum temperature T
3
below
the limit which can be tolerated by the combustion chamber and turbine blade materials.)

(a) Sketch the processes on a T s diagram.
Making appropriate assumptions, calculate
(b) the isentropic efficiency of the compression process,
(c) the turbine exit temperature T
4
and pressure P
4
,
(d) the jet velocity C
5
at the nozzle exit.
Fig. 1
Components
of a jet engine

You might also like