You are on page 1of 49

THERMODYNAMICS

“A theory is the more impressive the greater the simplicity of its premises, the more
varied the kinds of things that it relates and the more extended the area of its
applicability. Therefore classical thermodynamics has made a deep impression
upon me. It is the only physical theory of universal content which I am convinced,
within the areas of the applicability of its basic concepts, will never be overthrown.”

Albert Einstein

1
LIST OF SYMBOLS

A area, m 2
C velocity, m / s
CV constant volume specific heat, J / kgK kJ / kgK 
CP constant pressure specific heat, J / kgK kJ / kgK 
g gravitational acceleration, m / s 2 ; moisture content, kg / kg
H enthalpy, J kJ 
h specific enthalpy J / kg kJ / kg 
m mass, kg
n polytropic index; amount of substance, kmol
p pressure, Pa kPa , bar 
R specific gas constant, J / kgK
Ro universal gas constant, J / kmol K
S entropy, J / K kJ / K 
s specific entropy, J / kgK kJ / kgK 
T temperature, K
Q heat transfer, J kJ 
q heat transfer per unit mass, J / kg kJ / kg 
U internal energy, J kJ 
u specific internal energy, J / kg kJ / kg 
V volume, m3
v specific volume, m 3 / kg
W work transfer, J kJ 
w work transfer per unit mass, J / kg kJ / kg 
x dryness fraction

 density, kg / m 3
 thermal efficiency
 relative humidity
 temperature, C

2
1. BASIC CONCEPTS AND DEFINITIONS

1.1 Thermodynamic System

A thermodynamic system can be defined as a finite portion of matter or region of


space containing matter whose behaviour is being investigated. The system is
separated from the surroundings by a boundary.

Thermodynamics involves the study of the energy of a system, its forms and
transformations, and the relationships among the various physical quantities which
are affected by, or cause, these transformations.

There are two types of system:

(i) Closed: There is no flow of mass across the system boundary which may be
deformable (Fig 1.1a). The same matter remains within the
boundary. Only energy in the form of heat or work can cross the
boundary.

(ii) Open: Mass may flow across the system boundary in addition to heat and
work (Fig 1.1b). In this case the system is referred to as the control
volume and the boundary as the control surface.

Mass
flow
Gas (system)
System
( Control volume)

Boundary

( a ) Closed system ( b ) Open system

FIG 1.1

1.2 Classical Thermodynamics

Classical thermodynamics seeks to describe the behaviour of a system and the


energy interactions at the system boundary in terms of the large scale bulk
characteristics of the system, such as its pressure p Pa  , volume V m 3 and  
temperature T K  . This is a macroscopic description: the quantities used apply to
the system as a whole and no special assumptions are made regarding the structure

3
of matter. The laws of thermodynamics are based solely on macroscopic
observation and rest on three simple, but far reaching, concepts:

(i) the existence of relationships (equations of state) between the properties of a


system in equilibrium,

(ii) the conservation of energy which leads to the formulation of the First Law of
Thermodynamics,

(iii) reversibility and irreversibility, which takes account of the fact that there is a
preferred direction for natural processes and leads to the formulation of the
Second Law of Thermodynamics.

1.3 Thermodynamic Properties and State

The quantities used in the macroscopic description of a system are called


thermodynamic properties (or thermodynamic coordinates). An extensive property is
one which depends on the mass of the system e.g. volume. An intensive property is
one which is independent of the system mass e.g. pressure, temperature. If the
value of an extensive property is divided by the mass, the resulting property is
intensive and referred to as a specific property e.g. specific volume v m 3 / kg (the
reciprocal of specific volume is density  kg / m 3 ).

The condition of a system, as identified by a particular set of values of its properties,


is called the thermodynamic state of the system.

Relationships exist between the properties of any given system and the values of
only a few properties will generally be sufficient to identify a state completely. The
number of independent properties needed to specify the state of a system is
sometimes referred to as the number of degrees of freedom of the system. A simple
compressible substance has two degrees of freedom. For example, for a fixed mass
of gas in a cylinder, the thermodynamic state is specified completely when, say, the
pressure and volume are known. The state can be represented by a point on a
diagram using pressure and volume as coordinates (Fig 1.2). With this pressure and
volume, all the other thermodynamic properties are fixed. Any third property, say
temperature, must be a single-valued function of these two:

T  f  p ,V  (1.1)

Such a relation is called an equation of state.

4
Pressure Property A

Defined state
Defined state
p A

V Volume B Property B

FIG 1.2

1.4 Thermodynamic Equilibrium

An important assumption implicit in the classical thermodynamic description of the


state of a system is that the system is in thermodynamic equilibrium i.e. the
thermodynamic properties are well-defined and, where appropriate, are uniform
throughout the system, such that each has a single value which does not change
with time.

Thermodynamic equilibrium requires that the system be in thermal, mechanical and


chemical equilibrium. Thermal equilibrium requires that the temperature of the
system be uniform and that there is no difference in temperature between the system
and the surroundings. Mechanical equilibrium requires that the pressure of the
system be uniform and that there are no unbalanced forces acting between the
system and the surroundings. Chemical equilibrium requires that the system be
incapable of a spontaneous change in composition.

1.5 Changes in State

A system undergoes a process when it changes from one equilibrium state to


another. The states through which the system passes during a process can be
described by the thermodynamic properties of the system only if, at every stage of
the process, the system is infinitesimally close to thermodynamic equilibrium. Such
a process is effectively a succession of equilibrium states and is called a quasistatic
process. To be quasistatic a process must be carried out sufficiently slowly, such
that all properties are well defined and gradients in the intensive coordinates are no
more than infinitesimal. There can never be more than an infinitesimal unbalanced
force acting on the system or an infinitesimal temperature difference between the
system and the surroundings. An equation of state is valid for all states passed
through by the system and since equilibrium states may be represented on a
diagram of properties, the complete path of the process can be plotted (Fig 1.3).

5
p

V
FIG 1.3

A system can, of course, pass from an initial equilibrium state to a final equilibrium
state by means of a process which is not a succession of equilibrium states. Then,
at each instant, the thermodynamic properties will not have definite values applying
to the system as a whole and the process cannot be described in terms of the
properties. No equation of state exists for such nonequilibrium states. The path of
the process is indetermined and cannot be shown on a property diagram.

A series of processes can be put together such that the system returns to its initial
state. Such a series of processes is called a thermodynamic cycle (Fig 1.4).

p
1 2

4
3

V
FIG 1.4

Thermodynamic properties are functions of state (or point functions). This gives
properties the characteristics of exact differentials i.e. the change in the value of a

6
property depends only on the initial and final states of the system and is independent
of the process undergone by the system during the change of state. When the state
of a system can be defined by two independent properties, say x and y , any third
property z can be expressed as a function of these two

z  f x , y  (1.2)

When x and y change by dx and dy , the infinitesimal change in z is given by

 z   z 
dz    dx    dy (1.3)
 x  y  y  x

When x changes from x1 to x 2 and y changes from y1 to y 2 the finite change in z


is

dz  z 2  z1  z x2 y 2   z x1 y1 
2
z  1
(1.4)

As values of z at the point  x1 y1  and  x 2 y 2  are fixed, so  z is fixed. It does not


matter how x and y are varied during the integration between the given limits – the
integral is path-independent.

1.6 Heat and Work

A system can exchange energy with the surroundings in two ways, either as heat or
work. If the driving potential is anything other than temperature, the energy transfer
is called work.

Both heat and work are energy in transition i.e. energy which crosses the system
boundary during a change in the state of the system. This energy is not stored in the
system as heat or work but is converted to some other form of energy after crossing
the system boundary. Heat and work are therefore not thermodynamic properties
but depend upon the path followed by the system during the change of state. Heat
and work are path functions and infinitesimals of heat and work, dQ and dW , are
inexact differentials. The integration of these inexact differentials for a system
undergoing a process from state 1 to state 2 will be written as

 dQ  Q12 or simply Q 


1
2

 dW  W12 or simply W 


1

where Q12 and W12 represent the amount of heat and work transferred during the
process.

7
Note that
2

 dQ  Q2  Q1
1
2

 dW  W2  W1
1

as Q1 , Q 2 , W1 and W 2 have no meaning.

A sign convention is necessary to indicate the direction of the heat and work flow
(Fig 1.5). Work done on the surroundings by the system and heat flowing into the
system will be taken as positive; work done on the system by the surroundings and
heat flowing from the system will be taken as negative.

Q + ve Q - ve
System
W - ve W + ve

FIG 1.5

Piston c.s.a.A

pA
Pressure F
p

FIG 1.6

8
1.7 Work in a Quasistatic Process

Consider a simple compressible fluid contained in a cylinder fitted with a frictionless


piston. Suppose that the system is undergoing a quasistatic process from an initial
equilibrium state  p1 V1  to a final equilibrium state  p 2 V 2  . Consider an intermediate
equilibrium state at which the system pressure is p (Fig 1.6). The force exerted by
the system on the movable part of the boundary is p A and this is balanced by the
opposing force F  p A . If F is changed infinitesimally so that the piston moves a
distance dx (positive for an expansion, negative for compression), the work done is

d W  pAdx (1.5)

as Adx  dV , then

dW  pdV (1.6)

and for the whole process

2
W   1
pdV (1.7)

If the path between the initial and final states is plotted on a p  V diagram, it is clear
that the work done is represented by the area under the curve (Fig 1.7). If two such
curves are combined to constitute a cycle, the net work for the cycle is represented
by the area within the closed figure:

a i
Wnet   pdV   pdV (1.8)
i a

  pdV

p p
2
i Wnet
W
p

1
a

dv V V

FIG 1.7

9
In terms of unit mass of substance, equations (1.6) and (1.7) become

dw  pdv (1.9)

2
and w  1
pdv (1.10)

Example 1.1

Unit mass of a fluid at a pressure of 3 bar and a specific volume of 0.18 m 3 / kg


contained in a cylinder behind a piston expands quasistatically to a pressure of
0.6 bar according to a law p  c / v 2 where c is a constant. Calculate the work
done.

2 2 dv 1 1
w   1
pdv  c 
1 v 2
 c 
 v1

v2 

Now c  p1 v12  3  10 5  0.18 2  9.72  10 3

c 9.72 x 10 3
and v2    0.402m3 /kg
p2 0.6 x 10 5

therefore
 1 1 
w  9.72 x 10 3   
 0.18 0.402 

 9.72  10 3  3.07

 29840 J / kg
i.e. work is done by the fluid.

10
2. THE FIRST LAW OF THERMODYNAMICS

2.1 Closed Systems

The First Law of Thermodynamics is a statement of the principle of the conservation


of energy applied to a system. This principle requires that the algebraic sum of all
the energy transfers across the system boundary must be equal to the change in the
energy of the system. Since heat and work are the only forms of energy that may
cross a boundary, the First Law can be written as

 E  E 2  E1  Q12  W12 (2.1)

The energy term E consists of three terms

E  U  KE  PE (2.2)

where

(i) the internal energy, U , represents the energy possessed by the elementary
particles of the substance by virtue of their microscopic kinetic and potential
energy,

(ii) the macroscopic kinetic energy, KE , represents the kinetic energy of the
system due to its motion

1
KE  mC 2 (2.3)
2

(iii) the macroscopic potential energy, PE , represents the potential energy of the
system due to its position in the gravitational field

PE  mg z (2.4)

where z is the elevation of the system above some datum.

Generally the closed systems studied in thermodynamics are stationary and there is
no change in KE and PE . The first law may therefore be expressed as

U  U 2  U 1  Q12  W12 (2.5)

which is known as the non-flow energy equation. Internal energy is a


thermodynamic property and is a function of as many properties as are necessary to
specify the state of the system. dU is an exact differential. For an infinitesimal
process

dU  dQ  dW (2.6)

11
where regarding U as a function of, say, p and T

 U   U 
dU    dT    dp (2.7)
 T  p  p T

For a cyclic process  dU  0

and  dQ   dW (2.8)

In terms of a unit mass of substance, equations (2.5), (2.6) and (2.8) are written as

u 2  u1  q12  w12 (2.9)

du  dq  dw (2.10)

 dq   dw (2.11)

2.2 Enthalpy and Specific Heats

The enthalpy H is defined as

H  U  pV (2.12)

As a function of thermodynamic properties, H is itself a thermodynamic property. In


specific terms

h  u  pv (2.13)

The constant – volume specific heat CV is defined as

 u 
CV    (2.14)
 T  v

The constant – pressure specific heat C P is defined as

 h 
CP    (2.15)
 T  P

Both CV and C P are thermodynamic properties. Each is the rate of change of a


property with temperature while some other property is held constant. For example,
C P gives the slope of a constant pressure line on an h  T diagram for a substance.

12
Originally, CV and C P were defined in relation to the amount of heat added to a
system in a constant volume and constant pressure process and the resulting
change in temperature:

dq  CV dT ) const . volume (2.16)

dq  C P dT ) const . pressure (2.17)

The equivalence of these alternative definitions of specific heats requires that the
heat addition processes are quasistatic.

2.3 Open Systems

In most practical engineering problems the fluid flow rate through open systems is
constant. This type of flow is called steady flow. Under steady-flow conditions the
energy of a system is constant and an ‘energy balance’ applied to a stationary
control volume gives

heat added energy of energy of work done


to the + matter = matter + by
system entering leaving system (2.18)
system system

Consider a fluid flowing in a steady flow with a mass flow rate m  kg / s through the
open system shown in Fig 2.1. The fluid enters at section 1 and leaves at section 2.
A steady rate of flow of heat Q is supplied and W is the rate of work output.

1
C1
p1 .
Q
.
W
FIG 2.1

z1 2
p2
C2
z2

Datum

13
When a fluid flows across a system boundary work is always done on or by the
system. Therefore the work term in the energy balance equation (2.18) has two
components (i) the work required to push the fluid into and out of the system (flow
work) and (ii) W .

At entry the flow work per unit mass (on system)  p1v1

At exit the flow work per unit mass (by system)  p 2 v 2

Rewriting equation (2.18):

 C12   C 22 
 
Q  m  u1   
 z1 g   m  u 2   z2 g   W  m  p 2 v 2  p1 v1 
 2   2 

 C2   C2 
or Q  W  m  u 2  2  z 2 g  p 2 v 2   m  u1  1  z1 g  p1 v1 
 2   2 

Using the definition of enthalpy this becomes

 C2   C2 
Q  W  m  h2  2  z 2 g   m  h1  1  z1 g  (2.19)
 2   2 

This is known as the steady-flow energy equation. In terms of a unit of mass


passing through the system, it can be written as

C2 C2 
q  w  h2  h1    2  1   g z 2  z 1  (2.20)
 2 2 

In applying equation (2.20) to the various open systems commonly encountered in


engineering, various simplifications can be made. For example, in a throttling
process, differences in kinetic and potential energy are negligible, no work crosses
the boundary and because the throttling takes place in a short length of pipe, any
heat transfer can be ignored. Thus the equation reduces to

h1  h2 (2.21)

Example 2.1

In the turbine of a gas turbine unit the gases flow through the turbine at 17 kg / s and
the power developed by the turbine is 14000 kW . The specific enthalpies of the
gases at inlet and outlet are 1200 kJ / kg and 360 kJ / kg respectively, and the
velocities of the gases at inlet and outlet are 60 m / s and 150 m / s respectively.

14
Calculate the rate of heat transfer from the turbine. Find also the area of the inlet
pipe given that the specific volume of the gases at inlet is 0.5 m 3 / kg .

Neglecting changes in height, the steady-flow energy equation becomes

 C2   C2 
Q  W  m  h2  2   m  h1  1 
 2   2 

C12 60 2
Kinetic energy at inlet    1800 J / kg  1.8 kJ / kg
2 2

C 22 150 2
Kinetic energy at outlet    11250 J / kg  11.25 kJ / kg
2 2

Also W  14000 kW

Substituting in the above equation

Q  14000  17 360  11.25  17 1200  1.8

Therefore
Q  119 .3 kW

i.e. Heat rejected  119.3 kW

To find the inlet area

C1 A1 m v1
m  and A1 
v1 C1
therefore

17 x 0.5
A1   0.142m 2
60

15
4. THE SECOND LAW OF THERMODYNAMICS

4.1 Functions of the 2nd Law

(i) The 1st Law emphasises the equivalence of heat and work as energy transfer
mechanisms, but tells us nothing about the conversion from one form to
another. It is known, from experience, that work may be continuously and
completely converted into heat. However, again from experience, it is found
that heat cannot be continuously and completely converted into work. The
first function of the 2nd Law is to express the limiting efficiency with which heat
can be converted into work.

(ii) The 1st Law expresses the principle of conservation of energy and places
restrictions on the changes of a system which are energetically possible.
However, not all such changes occur. For example, consider two bodies at
different temperatures, which are put into thermal contact in a rigid, thermally
insulated box. The 1st Law can be satisfied with the temperatures converging
or diverging, the only restriction being that the total energy remains constant.
However, it is known from experience that the temperatures do not diverge:
they converge until thermal equilibrium is reached. Thus there is a natural
direction in which processes proceed. The second function of the 2nd Law is
to provide a formal framework for determining this natural direction of change.

4.2 Thermal Reservoirs and Heat Engines

A thermal energy reservoir (RES) is a system of such large mass that a finite amount
of heat can be removed from it, or added to it, without causing any appreciable
change in its temperature. When heat is removed from a reservoir, it is termed a
source; when heat is added, it is termed a sink.

A heat engine (HE) is a device by whose agency a system (the working fluid)
receives heat and produces work while operating in a cycle. The purpose of a heat
engine is to deliver work continuously by performing the same cycle over and over
again.

For a cycle of a system (equation (2.8)):

 dW   dQ (4.1)

As given, this equation accords with the sign convention adopted in Section 1.6.
However, the arguments on the Second Law are normally formulated by giving
separate symbols to the magnitudes of the quantities of heat added to and rejected
from the system during the cycle. If Q1 is the heat added and Q 2 is the heat rejected
then, from the 1st Law,

Q1  Q2  W (4.2)

16
where W is the net work done by the system.

The thermal efficiency for the cycle is defined as

W Q1  Q 2 Q2
    1
Q1 Q1 Q1

If the direction of a heat engine cycle is reversed, the device is termed a heat pump
(HP). In this case, a net amount of work is done on the system during the cycle.

4.3 Statements of the Second Law

The 2nd Law has been stated in various forms. Two of the best known are the
Clausius and Kelvin-Planck statements.

Clausius: It is impossible to construct a device that, operating in a cycle,


produces no other effect than the transfer of heat from a cooler
body to a hotter body.

Kelvin-Planck: It is impossible to construct a device that, operating in a cycle,


produces no other effect that the extraction of heat from a single
reservoir and the performance of an equivalent amount of work.

The devices prohibited by these statements are shown in Figs 4.1 and 4.2.

RES TH RES TH

Q2 Q1

W = Q1
H.P. H.E.

Q2
FIG 4.2

RES TC

FIG 4.1

17
The Clausius and Kelvin-Planck statements, although they appear quite different, are
entirely equivalent to each other and either can be used as a statement of the 2 nd
Law. This equivalence can be demonstrated by showing that the violation of either
statement results in the violation of the other (Appendix 1).

The Clausius statement does not say it is impossible to transfer heat from a lower
temperature to a higher one. Indeed, this is exactly what a heat pump (or
refrigerator) does. A heat pump does not operate, though, unless it receives an
energy input, usually in the form of work (Fig 4.3).

The Kelvin-Planck statement implies that some heat must always be rejected from
the system. This requires that at some point in the cycle the system be in contact
with a reservoir of lower temperature than the working fluid.

RES TH RES TH

Q1 Q1

W = Q1-Q2 W = Q1-Q2
H.P. H.E.

Q2
Q2

RES TC RES TC

FIG 4.3 FIG 4.4

Thus if a system undergoes a cycle and produces work it must operate between at
least two reservoirs of different temperature (Fig 4.4), with a resulting thermal
efficiency   1.0 . The following questions naturally arise:

(a) what is the maximum possible efficiency, and


(b) what factors govern this value of η.

Before we can proceed to answer these questions, and derive other statements of
the 2nd Law (2nd Law Corollaries), the question of reversibility must first be
considered.

4.4 Reversibility

A reversible process is one that is performed in such a way that, at the conclusion of
the process, both the system and the surroundings can be restored to their original
states, such that there is no evidence of the process having taken place.

18
The requirements for a process to be reversible are:

(i) the process must be quasistatic,


(ii) heat must flow by virtue of an infinitesimal temperature difference,
(iii) there must be no frictional effects.

Such a process is obviously highly idealised and no real process can be completely
reversible. However, the concept of reversibility is essential to the development of
classical thermodynamics in that it leads to the limiting efficiency of work producing
devices, the absolute thermodynamic temperature scale and the property called
entropy.

A reversible process is an ‘optimum’ work process. It can be shown that the work
which can be extracted from a system in a given change has a maximum value when
the change is made reversibly. Conversely, if the change requires that work is done
on the system, the reversible process requires a smaller work input than an
irreversible one.

We can now think of a cycle made up of completely reversible processes i.e. a


reversible cycle for which the work and heat transfers effected when the cycle
operates in one direction will be equal in magnitude but opposite in sign when the
cycle operates in the reverse direction.

4.5 Corollaries of the Second Law

Two valuable statements pertaining to the efficiency of reversible engines can now
be deduced:

(1) It is impossible to construct an engine operating between two reservoirs which


will have a higher efficiency than a reversible engine operating between the
same two reservoirs.
(2) All reversible engines operating between the same two reservoirs have the
same efficiency.

Proof

(1) Consider a reversible engine RE and an irreversible engine X operating


between RES T1 and RES T2 (Fig 4.5). Let both engines receive heat Q1 and
assume that engine X has a higher efficiency than RE . Thus

 X   RE

W X  WRE

Q1  W RE  Q1  W X

19
RES TI RES TI

Q1 Q1 Q1 Q1

WX WRE W X - W RE WRE
X RE X RE

Q1-W X Q1-W RE Q1 - W X Q1 - W RE

RES T2 RES T2

FIG 4.5 FIG 4.6

Let RE be reversed to operate as a heat pump (Fig 4.6). It now receives heat
Q1  W RE , work W RE and rejects heat Q1 . As W X  W RE , engine X could drive
reversed RE and still deliver work W X  WRE .

For the combined plant  X  reversed RE  , RES T1 becomes superfluous as


reversed RE could reject its heat directly to X . Note also that the net heat
removed from RES T2 is W X  WRE . Thus together X and reversed RE
constitutes a device which operates in a cycle, extracts W X  WRE units of heat
from a single reservoir and delivers an equivalent amount of work. This
violates the Kelvin-Planck statement and the assumption  X   RE cannot be
true. Consequently statement (1) must be true.

(2) Consider that both engines operating between RES T1 and RES T2 are
reversible engines, RE1 and RE2. Assume that their efficiencies are different
so that their work outputs are different for the same heat input Q1 . Thus,

 RE 2   RE1
WRE 2  WRE1
Q1  WRE1  Q1  WRE 2

Let the less efficient RE1 be reversed. The more efficient RE2 could drive the
reversed engine and still deliver a work output W RE 2  WRE1 , while extracting
W RE 2  WRE1 units of heat from RES T2 . Again, RES T1 is superfluous. This is
the same violation of Kelvin-Planck as above and the assumption of different
efficiencies cannot be true. Consequently, statement (2) must be true.

20
Thus the efficiency of a reversible engine operating between two thermal
energy reservoirs depends only on the temperature of the reservoirs. The
efficiency does not depend on the working substance.

4.6 Thermodynamic Temperature

From the previous result regarding the efficiency of reversible engines operating
between constant temperature thermal energy reservoirs, a temperature scale can
be established which is related only to the efficiency of a reversible engine i.e. a
temperature scale independent of the properties of any substance.

Consider some arbitrary temperature scale  and that three reversible engines
operate between three reservoirs at temperatures  1 ,  2 and  3 (with  1   2   3 ),
as shown in Fig 4.7.

1
Q1
W1 Q1
RE1

Q2
W3
2 RE3 FIG 4.7
Q2
W2
RE2 Q3
Q3
3

Q2
For RE 1   1  f  1 ,  2 
Q1

where f is an unknown function. Rearranging this equation

Q1
 g  1 , 2 
1
 (4.4)
Q2 1  f  1 , 2 

where g is another unknown function.

Similarly for RE 2 and RE 3

21
Q2 Q
 g  2 , 3  and 1  g  1 , 3 
Q3 Q3
Q1 Q / Q3
Now  1
Q2 Q 2 / Q3

g  1 , 3 
and so g  1 , 2   (4.5)
g  2 , 3 

As the left-hand side of this equation is a function of  1 and  2 (and not of  3 ), so


the right-hand side must also be a function of  1 and  2 (and not of  3 ). The form
of function must therefore be such that

g 1 ,  3     1   3 
g 2 ,  3     2   3 

for in this way  3 will cancel from equation (4.5). Therefore

Q1   1 
 (4.6)
Q2   2 

where  is another unknown function. The form of the function    is completely


arbitrary. Lord Kelvin proposed that    simply be T , the ‘Thermodynamic’ or
‘Absolute’ temperature. Thus

Q1 T
 1 (4.7)
Q2 T2

On this scale (the Kelvin scale) the ratio of the temperatures of two reservoirs is
equal to the ratio of the quantities of heat exchanged at these reservoirs by a
reversible engine operating between them. To complete the definition of the scale, a
positive number is assigned to some easily-reproducible state. The chosen
reference is the ‘triple point’ of water, to which a value of 273.16 K is assigned. Thus
the numerical value in kelvins of any other temperature is given by

Q Q
T  TTP  273.16 (4.8)
QTP QTP

where Q and QTP are the quantities of heat transferred by a reversible engine
operating between a reservoir at T and a reservoir at TTP , the triple point of water.

The Celsius temperature scale is defined by

 C   T K   273.15

22
noting that 0C is assigned to the ‘ice-point’ temperature which is 0.01 K lower than
the triple-point.

4.7 The Carnot Cycle

As an example of a reversible engine operating between two thermal energy


reservoirs at temperatures T H and TC , consider the following cycle executed by a
gas in a closed system. The cycle is shown in Fig 4.8 and comprises the following
processes:

p
1
TH
Q=0 2

Q=0
4
TC 3

V
FIG 4.8

1→2 The system expands reversibly and isothermally at T H . Work is done by the
system and heat Q1 is transferred from RES TH .

2→3 The system expands reversibly and adiabatically doing work. The
temperature falls to TC .

3→4 The system is compressed reversibly and isothermally at TC rejecting heat Q 2


to RES TC .

4→1 The system is compressed reversibly and adiabatically to initial state 1.

This is the Carnot cycle.

For process 1→2 (equation (3.26))

23
V2
Q1  W12  mRTH ln
V1

For process 3→4


V3
Q2  W34  mRTC ln
V4

Therefore
Q2 T ln V3 / V4
  1  1  C
Q1 TH ln V2 / V1

For the adiabatic processes, from equation (3.35)

1
V3 T   1
  H 
V2  TC 

1
V4  T   1
and   H 
V1  TC 

V3 V 2
giving 
V 4 V1

Q2 T
Thus  C (4.9)
Q1 TH

in accordance with equation (4.7), and

TC T  TC
 1  H (4.10)
TH TH

This efficiency represents the maximum possible thermal efficiency between two
specified temperature limits and, as such, fulfils the first function of the 2 nd Law (page
33).

24
5. ENTROPY

5.1 Clausius’ Theorem

It is possible to approximate a reversible process between two equilibrium states by


a zigzag path consisting of an adiabatic process, followed by an isothermal process,
followed by another adiabatic process, such that the heat transferred during the
isothermal portion is the same as the heat transferred during the original process
(Fig 5.1). Obviously, the greater the number of adiabatic and isothermal steps, the
closer the zigzag path approaches that of the original process.

Consider now a reversible cycle of a system. A number of adiabatic lines can be


drawn dividing the cycle into a number of adjacent strips. A closed zigzag path may
now be drawn such that the sum of the heat transfers during all the isothermal
portions is equal to the heat transferred in the original cycle (Fig 5.2).

p
adiabatic
i isothermal
llll
FIG 5.1
a adiabatic
b
f
area under iabf
= area under if

Wiabf = Wif V
Qif = Qab

p
T3
T1 e
a
Q3 f
FIG 5.2
Q1 b

h
d g
Q4
c T4
Q2 T2

25
In the isothermal process ab at temperature T1 , heat Q1 is absorbed and in the
isothermal process at temperature T2 , heat Q 2 is rejected. Since a b c d constitutes
a Carnot cycle, then from equation (4.9)

Q1 Q
 2
T1 T2

If, instead of considering only the magnitudes of quantities of heat, as was done in
Section 4.2, the sign convention of Section 1.6 is re-adopted then

Q1 Q
 2  0 (5.1)
T1 T2

Similarly for e f g h

Q3 Q
 4  0 (5.2)
T3 T4

Adding the equations for all the Carnot cycles

Q1 Q2 Q3 Q4
     0 (5.3)
T1 T2 T3 T4

and since no heat is transferred during the adiabatic portions, we may write

Q
T  0 (5.4)

where the summation is taken over the complete zigzag cycle.

As the number of strips is increased, the zigzag path approaches ever closer to the
original cycle. When the isothermal portions become infinitesimal, the ratio dQ / T
for an infinitesimal isothermal between two adjacent adiabatics is the same as the
ratio of dQ / T for the infinitesimal part of the original cycle bounded by the same two
adiabatics. Therefore, in the limit, for any reversible cycle

dQ
 T
 0 (5.5)

This is Clausius’ theorem.

26
5.2 Entropy : The Property

Equation (5.5) is sufficient to identify A


dQ / T )R as the differential of a function of 2
state. This can also be demonstrated as
shown below. Consider that a system is
taken from an initial state 1 to a final state
2 along the reversible path A and then 1
back to 1 along the reversible path B . B

The two paths constitute a reversible cycle and by equation (5.5)

dQ
AB  T
 0

The cyclic integral may be expressed as the sum

2 dQ 1 dQ
A 1 T
 B  2 T
 0

2 dQ 1 dQ
or A 1 T
 B  2 T

As B is a reversible path

1 dQ 2 dQ
B  2 T
 B 
1 T

2 dQ 2 dQ
and so A 1 T
 B 
1 T
(5.6)

dQ  2
Since A and B are any reversible paths it follows that the integral
1
 is
T R 
independent of the reversible path connecting states 1 and 2 and depends only on 1
and 2. Therefore, there exists a thermodynamic property such that the change in its
2 dQ 
value between two equilibrium states 1 and 2 is equal to   .
1 T R

This property is called entropy, denoted S . Thus

2 dQ 
S  S 2  S 1  1

T R
(5.7)

27
or for two states infinitesimally close

dQ 
dS   (5.8)
T R

For a reversible adiabatic process dQ  0 and dS  0 i.e. a reversible adiabatic


process is an isentropic process.

In terms of specific entropy


2 dq 
s  s 2  s1    (5.9)
1 T R

dq 
and ds   (5.10)
T R

5.3 General Entropy Equation

The differential form of the 1st Law for a reversible process is

dU  dQ  pdV

Also for a reversible process

dQ  TdS

and so dU  TdS  pdV

or TdS  dU  pdV (5.11)

Since by definition
H  U  pV

dH  dU  pdV  Vdp
then
 TdS  Vdp

and TdS  dH  Vdp (5.12)

Equations (5.11) and (5.12) are two forms of the General Entropy Equation. In terms
of specific entropy

Tds  du  pdv (5.13)

Tds  dh  vdp (5.14)

28
5.4 Entropy Changes for Gases

For a perfect gas du  CV dT and equation (5.13) becomes

Tds  CV dT  pdv

dT p dv
or ds  CV 
T T

dT dv
 CV  R (5.15)
T v

Integrating between states 1 and 2, taking CV as constant

2 T2 v
s 2  s1  
1
ds  CV ln
T1
 R ln 2
v1
(5.16)

Similarly from equation (5.14) using dh  C P dT

dT dp
ds  C P  R
T p

T2 p
and s2  s1  CP ln  R ln 2 (5.17)
T1 p1

From the perfect gas equation of state

T2 p v
 2 2
T1 p 1 v1

and substituting into equation (5.16)

p v  v
s 2  s1  CV ln  2 2   R ln 2
 p1 v1  v1

Expanding this equation, noting that

R  C P  CV ,
p2 v (5.18)
s2  s1  CV ln  C P ln 2
p1 v1

29
5.5 Entropy as a Coordinate

For an infinitesimal portion of a reversible process

dQ  TdS

and it follows that the total amount of heat transferred in a reversible process from
state 1 to state 2 is given by

2
Q12  
1
TdS (5.19)

This integral can be interpreted graphically as the area under the process path
plotted on T  S (or T  s ) coordinates (Fig 5.3). The area enclosed by a cycle
therefore represents the net cycle heat transfer and from the 1st Law this is equal to
the net cycle work transfer (Fig 5.4).

T 2

Q12
T
1

S
dS
FIG 5.3

1
W net

S
FIG 5.4

30
As the Carnot cycle consists of two reversible adiabatic (isentropic) and two
reversible isothermal processes, it appears as a rectangle on T  S (or T  s )
coordinates (Fig 5.5).

T
1 2
TH

FIG 5.5

TC
4 3

a b S
S

Heat transfer to fluid 1→2  QH  area 12 ba 1  TH  S .

Heat rejected from fluid 3→4  QC  area 34 ab 3  TC  S .

net work Q  QC
   H
heat supplied QH

TH  TC  S T
  1 C
TH S TH

as must be the case.

Entropy can also be used as a coordinate in the construction of property diagrams


(see Section 3.1). The general form of a T  s diagram for a substance with a phase
change is given in Fig 5.6. It should be noted that for a constant pressure line on
such a diagram that since

Tds  dh  vdp

then as dp  0

so dh  Tds

31
Thus areas under constant pressure lines represent the enthalpy change between
some point on the line and some chosen datum condition.

CP Constant Pressure Line


T
Subcooled
Liquid
Region Constant Volume Line

FIG 5.6
Superheated
Vapour
Region
Wet
Vapour
Sat. liquid line Region
Sat. vapour line

5.6 Principle of Increasing Entropy

For a reversible cycle it was shown that

dQ 
   0
T R

From a consideration of irreversible cycles it can be deduced that

dQ 
   0
T I

Thus for any cycle we can write

dQ
 T
 0 (5.20)

where the equality holds for a reversible cycle. This is the Clausius Inequality.

Consider now that a system changes from state 1 to state 2 by either a reversible
path or an irreversible path and is returned to state 1 by a reversible path. From the
Clausius Inequality

32
R or I
2 dQ  1 dQ 
1
  2
T I
  0
T R 2
2 dQ  1 dQ 
1
  2
T R
  0
T R
1
R
2 dQ  2 dQ 
Subtracting gives 1
 
T I 
1
  0
T R

2 dQ  2 dQ 
or 1
 
T I 
1

T R

2 dQ 
But by definition 1
  S 2  S1
T R

2 dQ 
and so S 2  S1   
1 T I

dQ 
or dS  
T I

Thus it can be stated that

dQ
dS  (5.21)
T

with the equality holding for a reversible process.

For an isolated system dQ  0 and

dS  0 (5.22)

Thus entropy increases in an irreversible adiabatic process. This is the Principle of


Increasing Entropy which can be stated as

‘The entropy of an isolated system cannot decrease’.

This principle can be extended to a system which is not thermally isolated but may
exchange heat with other systems (the surroundings). These systems can also
exchange heat amongst themselves but not with any others. Thus together with the
original system they form a combined isolated system within which everything which
interacts thermally during the process is contained.

33
SURROUNDINGS
Q=0
SYSTEM
W=0

Thus dS SYS  dS SURR  0

or in integral form

STOT  S SYS  S SURR  0

where S implies S final state  S initial state

The 2nd Law can now be expressed in terms of entropy as:

If all bodies involved in a thermodynamic interaction are considered then S TOT  0


with the quality holding for a reversible interaction.

The fact that the total entropy can never decrease in a process provides a direction
for the sequence of natural events – the natural direction is that in which the total
entropy increases. This ‘arrow of time’ for the evolution of natural processes fulfils
the second function of the 2nd Law (page 33).

5.7 Entropy Generation

Consider a reversible process of a closed system between two end states. The work
transfer is

dWrev  pdV

Now from equation (5.11)

TdS  dU  pdV

and TdS  dU  dWrev

For any process of the system

dU  dQ  dWact
TdS  dQ  dWact  dWrev

dQ dWrev  dWact 
and dS   (5.23)
T T

34
For a reversible process

dQ
dS  and dWact  dWrev
T

For an irreversible process

dQ
dS  and dWrev  dWact  0
T

so that dWact  dWrev i.e. less work is done in the irreversible process. Designating
this lost work due to irreversibility as dWi then

dQ dWi
dS  
T T
 de S  di S
entropy entropy
transfer generation

Thus we see that the entropy change consists of a term related to heat transfer
across the system boundary and a term related to the loss of work due to presence
of irreversibilities.

d e S can be positive, negative or zero; d i S can only be positive (or zero if process is
reversible).

5.8 Isentropic Efficiency

T-s diagrams are useful for comparing actual irreversible processes with ideal
reversible processes. Many steady flow processes encountered in practice, such as
the expansion of a gas turbine, involve very little heat transfer (see example in
section 2.3) and are essentially adiabatic. The actual process, which will involve an
increase in entropy, is compared to the ideal isentropic process, by a parameter
called the isentropic efficiency ηs

Wact
s 
Wisen

Where Wact is the actual work of expansion and Wisen is the work for an isentropic
expansion between the same pressures ( Fig 5.7)

35
FIG 5.7

36
6. EXERGY

The exergy (availability) of a system, a stream of fluid or a heat transfer is the


maximum possible useful work which could be obtained from the system, stream of
fluid or heat transfer interacting with surroundings at po , To and zero velocity (the
‘dead’ state).

The maximum work would be obtained if the interaction with the surroundings was
completely reversible i.e. if the system or fluid stream is brought into equilibrium with
the surroundings in a series of reversible processes.

6.1 Heat Transfer

The simplest case to consider is a thermal energy reservoir at temperature T . We


know that the maximum efficiency with which heat drawn from the reservoir can be
converted into work is the Carnot efficiency. Thus the exergy of a heat transfer Q
from a reservoir at T in surroundings at To is

 T 
Q 1  o 
 T 

6.2 Closed System

Consider a closed system at state p1 , T1 being brought to equilibrium with the


surroundings. The system could go from state 1 to state 0 in a reversible adiabatic
process followed by a reversible isothermal process:

FIG 6.1

37
i.e. isentropic 1  To , then heat transfer with the surroundings at To . Two
possibilities are shown in Fig 6.1

From the 1st Law

dq  du  pdv

By the 2nd Law dq  Tds , and since heat transfer occurs only at T0 then

To ds  du  pdv

Integrating

o o o
To 
1
ds  
1
du  
1
pdv

Some of this pdv work is ‘useless’ work done in pushing back the surroundings at
po . Writing

p   p  po   po

To so  s1   u o  u1     p  p dv  po vo  v1 


o
Then o
1

  p  p  dv
o
o is the (specific) exergy of the system at state 1 and so
1

ex 1  u1  u o   To s1  so   po v1  vo  (6.1)

6.3 Open System

Fig 6.2 shows an imaginary and completely reversible device through which a fluid
stream is passing from p1 , T1 and velocity C1 to po , To and zero velocity. Again it is
assumed that any heat transfer occurs at To .

1 0
1 0

p1 , T1 p ,T
, o o
vel  C1 vel  0

Wmax possible = ex1 Imaginary completely reversible device

FIG 6.2

38
From the 1st Law

dq  dh  d KE   dw

where dw is the ‘shaft’ work and as such is the useful work available. Again
dq  To ds and

To s o  s1   ho  h1  
1 2
C1  ex 1
2

so that
 
ex 1   h1  ho  C12   To s1  s o 
1
(6.2)
 2 

6.4 Change of State

Consider a system going from state 1 to state 2.

ex 1 = maximum possible useful work 1→0


ex 2 = maximum possible useful work 2→0

Going reversibly from 1→0→2 gives the maximum possible work output for the state
change 1→2:

i.e. maximum possible work = ex 1  ex 2

For a closed system

ex 1  ex 2  u1  u o   To s1  so   po v1  vo 


 u 2  u o   To s 2  so   po v2  vo 

 u1  u 2   To s1  s 2   p o v1  v 2 


 u1  To s1  p o v1   u 2  To s 2  p o v 2  (6.3)

For an open system

   1 
 h1  ho   To s1  so   C12   h2  ho   To s 2  so   C 22 
1
ex 1  ex 2
 2   2 

 h1  h2  
2
 
C1  C 22  T0 s1  s 2 
1 2
(6.4)

39
6.5 Irreversibility

For any process the irreversibility i (per unit mass) can be defined

i maximum possible useful work actual useful work obtained


which could have been obtained (6.5)

Considering a closed system

i  ex 1  ex 2  wact
1

where w1act  wact  p0 v2  v1 

Therefore

i  u1  u 2   To s1  s 2   po v1  v2   wact  po v2  v1 

By the 1st Law

q  u 2  u1   wact

and wact  q  u 2  u1 

 q  u1  u 2 

This gives

i  u1  u 2   To s1  s 2   po v1  v 2   q  u1  u 2   po v 2  v1 

 u1  u 2   To s1  s 2   p o v1  v 2   q  u1  u 2   p o v1  v 2 

and i   To s1  s 2   q (6.6)

For a system of mass m , the total irreversibility is

I  mi   To S1  S 2   Q
 To S 2  S1   Q
 To S SYS  Q

Now Q   To S SURR

and I  To  SYS  To S SURR

giving I  To STOT (6.7)

40
It can be shown that equation (6.6) is also valid for an open system. In this case it
can be modified by a mass flow rate m to give

 To s2  s1   Q
I  m

 To SSYS  To SSURR

 To STOT (6.8)

APPENDIX 1

Mathematical Logic shows that two statements A and B are completely equivalent
 A  B  if, and only if, the negation of one  A implies the negation of the other
 B  and vice versa.
i.e. A B
if  A   B (  means implies)
and B   A

i) Consider that the Clausius statement were false. Fig A 1.1a shows a heat
pump which requires no work input to transfer Q 2 units of heat from RES TC
to RES TH . A heat engine, which does not violate the 2 nd Law also operates
between the reservoirs as shown.

For the combined plant HP  HE  the cold reservoir becomes superfluous.
Also the net heat removal from RES TH is Q1  Q2 . Thus together the HP and
HE constitute a system which operates in a cycle, extracts Q1  Q2 units of
heat from a reservoir and delivers an equivalent amount of work (Fig A 1.1b).
This violates the Kelvin-Planck statement.

ii) Consider that the Kelvin-Planck statement were false. Fig A 1.2a shows a
heat engine which extracts Q1 units of heat from RES TH and rejects no heat
to RES TC . A heat pump, which does not violate the 2 nd Law, also operates
between the reservoirs as shown.

For the combined plant HE  HP  , the HP could be driven by the HE . Also
the net addition of heat to RES TH is Q 2 . Thus together the HP and HE
constitute a system, which operates in a cycle, and whose sole effect is the
transfer of Q 2 units of heat from a cold reservoir to a hotter one (Fig A 1.2b).
This violates the Clausius statement.

41
TH
Equivalent System

Q2 Q1
TH

Q1 – Q2
W = Q1 – Q2
HP HE

W = Q1 – Q2
Q2
Q2
(b)

TC

FIG A1.1

Equivalent System
TH
TH

Q1 Q1+ Q2 Q2

W = Q1 W = Q1
HE HP
Q2

Q2 TC

TC
(b)
(a)
FIG A1.2

42
TUTORIAL

Unless otherwise stated, take atmospheric pressure  1 bar  100 kPa

Thermodynamic Revision

1. A cylinder fitted with a piston contains 1 kg of methane gas (molar


mass 16.04 kg / kmol ) at 700 kPa , 40C . The piston cross-sectional area is
0.5 m 2 , and the total external force restraining the piston is directly
proportional to the cylinder volume squared. Heat is transferred to the
methane until its temperature reaches 1100C . Determine the final pressure
inside the cylinder, the work done by the methane and the heat transfer during
the process. Cp  4.072 kJ / kgK

Answer: 1876 kPa; 183 kJ; 3950 kJ

2. The specific heat at constant pressure of a certain gas may be taken to vary
linearly with temperature being 1.0 kJ / kgK at 0C and 1.085 kJ / kgK at
170C . Determine the equation for Cp in terms of temperature.

The gas flows through a duct which it enters with a velocity of 650 m / s and
temperature 0C and leaves with a velocity of 65 m / s .

Neglecting heat losses, calculate the temperature of the gas at exit.

Answer: Cp = 1.0 + 0.0005t, where t is ºC; 199.2 ºC

3. A spherical balloon has a diameter of 0.25 m and contains air at a pressure of


1.3 bar . The diameter of the balloon increases to 0.3 m due to heating and
during the process the pressure inside the balloon is directly proportional to its
diameter. Calculate the work done by the air in the balloon during the
process.

Answer: 0.858 kJ

4. A mass of 0.05 kg of carbon dioxide (molar mass 44 kg / kmol ,   1.3 ),


occupying a volume of 0.03 m 3 at 1.025 bar , is compressed until the pressure
is 6.15 bar . Calculate the final temperature and the work and heat transfers:

(i) when the process is according to the law pv1.4  cons tan t
(ii) when the process is isothermal
(iii) when the process takes place in a perfectly insulated cylinder.

Answer: 543 K, 5.2 kJ, 1.72 kJ; 325.5 K, 5.5 kJ, 5.5 kJ; 492 K, 5.26 kJ

43
Gas Mixtures and Psychrometry

5. A vessel of volume 0.4 m 3 contains 0.45 kg of carbon monoxide and 1 kg of air


at 15C . Calculate the partial pressure of each constituent and the total
pressure in the vessel. The gravimetric analysis of air is to be taken as 23.3%
oxygen and 76.7 % nitrogen. Take the molar masses of carbon monoxide,
oxygen and nitrogen as 28 , 32 and 28 kg / kmol .

Answer: 96.2, 43.6, 164, 303.8 kPa

6. Air in an environmental chamber has a dry bulb temperature of 32C , a


relative humidity of 40 % and a total pressure of 98 kPa . Determine

(i) the moisture content


(ii) the enthalpy

Answer: 0.0123 kg/kgda, 63.48 kJ/kgda

7. For air with a dew point temperature of 6C , an enthalpy of 35.68 kJ / kgda
and a total pressure of 95 kPa , determine

(i) the dry bulb temperature


(ii) the specific volume
(iii) the relative humidity

Answer: 20°C, 0.894 m3/kg, 40%

8. Determine the dewpoint temperature of air with a dry bulb temperature of


38C , a thermodynamic wet bulb temperature of 27C and a total pressure of
95.56 kPa .

Answer: 23.4°C

44
Second Law – Reversibility – Carnot Cycle

9. A cyclic device operates between two energy reservoirs at different


temperatures. The work transfer to or from the device is 100 kJ , and the heat
transfer between one of the energy reservoirs and the device is 100 kJ (can
be either to or from the reservoir). Sketch the eight plant arrangements and
for each apply the First Law of thermodynamics to determine the magnitude
and direction of the heat transfer between the device and the second
reservoir. Use the Second Law of thermodynamics to determine which of the
arrangements are possible.

Answer: 5 possible arrangements

10. An inventor claims that a new engine cycle will develop 3.75 kW for a heat
addition rate of 315 kJ per minute. The maximum and minimum temperatures
of the cycle are 1700 K and 500 K respectively. It is possible for the claim to
be correct?

11. On the basis of the First Law of Thermodynamics complete the following table
of hypothetical heat engine cycles. On the basis of the Second Law discuss
the possibility of each.

Temperature °C Heat Flow kJ / s Power Efficiency


source sink supply rejection kW %
a 280 4 100 55
b 540 38 1000 65
c 560 80 26 60

12. Consider a cyclic device connected between two energy reservoirs at


temperatures T A and T B where T A  TB . Heat transfer from the reservoir at
T A to the device is Q A and heat transfer to the reservoir at T B from the device
is QB . The work output from the cyclic device is W .

If the temperatures of the reservoirs remain constant write down an


expression for the total change in entropy, S TOT , of the complete system and
hence show that the thermal efficiency  TH of the cyclic device can be
expressed as:

W T  TB T S
 TH   A  B TOT
QA TA QA

T A  TB
Explain why the thermal efficiency reaches a maximum value of  TH 
TA

45
when the cyclic device executes only reversible processes.

If T A  800 K , TB  290 , Q B  8000 kW and the total entropy generated is


12 kJ / Ks , calculate the work output and the thermal efficiency of the cyclic
device.

Answer: 4470 kW, 0.359

Entropy

13. Air in the cylinder of an internal combustion engine at the beginning of


compression occupies 0.1 m 3 at a pressure of 1 bar , and temperature 100C .
Compression takes place until the volume is 0.008m3 and the pressure is then
26.7 bar . Heat transfer to the air takes place at constant volume, until the
pressure reaches 53.3 bar . Calculate the change in entropy during each
operation stating whether there is an increase or decrease. Assume constant
specific heats Cp  1.002 and CV  0.715 kJ / kgK .

Answer: -0.017, + 0.046 kJ/K

14. A rigid vessel of total volume 0.5 m 3 contains 1 kg of air at 15C initially
constrained in a volume of 0.06 m 3 by a diaphragm. The remaining volume is
completely evacuated (pressure = 0). If the diaphragm is burst, allowing free
adiabatic expansion of the air, determine the change of entropy of the air and
of the surroundings. If instead of a free expansion, the process had been a
reversible expansion between the two states, what would then be the change
of entropy of the air, and of the surroundings? Compare the net change of
entropy for the system and surroundings in the two cases. Surrounding
temperature 15C .

Answer: +0.609, 0, +0.609 – 0.609 kJ/K

15. The specific heat at constant volume CV for a perfect gas is related to the
absolute temperature T by the expression

Cv  a  b / T  c / T 2

where a , b and c are constants.

Show that the change in specific internal energy u 2  u1 and the change in
specific entropy s 2  s1 are respectively given by

46
T2  1 1
u 2  u1  aT2  T1   b log e  c  
T1  T2 T1 

T2  1 1 c 1 1  v
s 2  s1  a loge  b      2  2   R log e 2
T1  T2 T1  2  T2 T1  v1

where v  volume and R  gas constant.

A closed vessel of 10 m 3 capacity contains such a gas at a pressure of 15 bar


and temperature 700 K . If a heat transfer occurs such that the gas
temperature falls to 300 K , determine

(a) the heat transfer to the atmosphere


(b) the entropy change of the gas
(c) the entropy change of the surroundings

Assume R  0.189 kJ / kgK


a  1.855 kJ / kgK
b  748 kJ / kg
c  11.658  10 4 kJK / kg
surrounding temperature 15C

Answer: 37449 kJ, -76.6 kJ/K, 130 kJ/K

16. An inventor claims to have a device, involving no moving parts, which is


capable of producing simultaneously a stream of hot air and a stream of cold
air, using only a supply of compressed air at atmospheric temperature. The
device is adiabatic.

It is claimed that when the device is supplied with 0.02 kg / s of air at 4.0 bar ,
temperature 15C , it produces a hot air stream of 0.01 kg / s at 30C and a
cold air stream of 0.01 kg / s at 0C . The kinetic energy of the supply air is
negligible and both hot and cold air streams leave the device with negligible
velocity at pressure 1.0 bar .

Treating the air as a perfect gas, use the First and Second Laws of
thermodynamics to decide whether the claim is justified, stating your
reasoning.

47
Exergy and irreversibility

17. An oil engine consumes 40 kg fuel per hour, with an air-fuel ratio of 20 kg air
per kg fuel. The exhaust gas temperature is 550 C and Cp  1.05 kJ / kgK
for the gas. The surroundings are at 20C . Determine the exergy of the
engine exhaust and the maximum power which may be recovered from the
exhaust gas.

Answer: 238.7 kJ/kg, 55.7 kW

18. A hydrogen-air heat exchanger is used to cool a stream of hydrogen gas,


entering at the rate of 100 m 3 per minute, from 350 C to 15C at a constant
pressure of 1 bar . The cooling is effected by air entering the exchanger at
10°C and leaving at 150°C, also at a constant pressure of 1 bar. No mixing of
the hydrogen and air takes place, there are no external energy losses and
kinetic energy and friction effects are negligible. Surrounding temperature,
10C . Determine

(a) the volume flow rate of air at inlet to the exchanger,


(b) the decrease in exergy across the heat exchanger and
(c) the irreversibility resulting from the heat transfer

For hydrogen take Cp  13.2 kJ / kgK and R  4.16 kJ / kgK .

Answer: 98.5 m3/min, 45.8 kW, 45.8 kW

19. In a certain ammonia refrigeration plant, where the surroundings conditions


are pressure 1 bar and temperature 10C , the compressor takes dry
saturated ammonia at 255 K and compresses it adiabatically to 8.57 bar ,
365 K . Determine

(i) the irreversibility per kg ammonia


(ii) the rational efficiency of the compressor defined as the ratio, the
minimum possible work input for the change of state to the actual work
input
(iii) the isentropic efficiency of the compressor defined as the ratio, the
work input required for isentropic compression to 8.57 bar to the actual
work input.

Answer: 20.94 kJ/kg, 0.908, 0.88

20. A gas stream enters a turbine at pressure 4.5 bar , temperature 635 C with a
velocity of 100 m / s and expands adiabatically to pressure 1.0 bar ,
temperature 350 C . The gas leaves the turbine with a velocity of 50 m / s .
The gas is then passed to a heat exchanger where it is used to heat an equal

48
mass flow rate of air at constant pressure from 200 C to 290 C . For the gas
Cp  1.06 kJ / kgK and R  0.28 kJ / kgK . Surrounding temperature 17C .

Calculate per kg of gas flow

(i) the turbine work output,


(ii) the decrease in exergy of the gas in passing through the turbine,
(iii) the irreversibility in the turbine,
(iv) the irreversibility in the heat exchanger assuming not heat losses and
neglecting kinetic energy changes

Answer: 305.9 kJ/kg, 312.2 kJ/kg, 6.38 kJ/kg, 5.8 kJ/kg

49

You might also like