8 Entropy: ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

ME1100 Thermodynamics Lecture Notes Prof. T.

Sundararajan

Chapter 8 Entropy

8.1 Introduction

Heat and work were earlier defined as energy interactions between a


system (or CV) and the surroundings. Although both are energy interactions,
there is a fundamental difference between heat and work. Directionality and
order at microscopic level are associated with each form of work. For
example, during gas expansion molecules move in a direction perpendicular
to the system boundary. When a piece of iron is magnetized, tiny molecular
level magnets align in the direction of the applied field. When current flows in
a conductor, electrons move in the direction along the wire. On the other
hand, heat transfer occurs due to temperature difference and temperature
itself is related to the kinetic energy of random molecular motion. When heat
is conducted from a high temperature object to another low temperature
object in contact, energy transfer occurs due to random molecular collisions.
Even radiative heat transfer occurs due to random transitions of surface
molecules from a higher energy level to a lower energy level. Due to such
random energy transitions at molecular level, the radiation emitted by a hot
object contains radiation of all wavelengths, in general.

The work done during the reversible (fully resisted, frictionless)


expansion of a gas can be expressed as δW = p.dV, in terms of the properties
such as the instantaneous pressure and the change in volume. Similarly, the
heat transfer during a reversible process can be expressed as δQ = T.dS,
where T is the instantaneous temperature and dS is the change in a property
known as “Entropy”. Entropy gives a measure of the randomness in a
system. For example, let us consider the three states of the same substance,
say- ice, water and steam. The entropy of ice is less than that of water and
the entropy of water is less than that of steam. This is in view of the fact that
in any solid state, the mean positions of molecules are fixed although the
molecules may undergo random vibrations. In liquid state, molecules slowly
move; in the vapor state, molecules move at high velocities of the order of a
few hundred meters per second. Therefore, the degree of randomness is very

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

high for the gaseous state. In terms of entropy, it can be said that S gas > Sliquid
> Ssolid, in general.

When heat is transferred from an object A to another object B, the


level of randomness in A will decrease and that of B will increase. In other
words, change in entropy for B (∆SB) will be positive and the change in
entropy of A (∆SA) will be negative. The entropy change of a system
undergoing a reversible process can be defined as

δQrev
dS = (8.1)
T

for an infinitesimal heat addition, and for the whole process 1-2, the entropy
change is given by the integral

2
δQrev
∆S = ∫ (8.2)
1
T

While we are at liberty to define any quantity of interest to us such as


entropy, in order to declare that entropy is a thermodynamic property, we
need to establish that the quantity given in Eq. (8.2) is indeed dependent
only on the end states 1 and 2. In other words, we need to show that ∆S as
defined in Eq. (4.2) is path independent. This will be done in two steps: (i)
First we will establish an inequality known as the “Clausius inequality” for
cyclic processes (ii) In the second step, it can be shown that the integral of
δQrev/T is independent of path for given end states 1 and 2, by the use of the
Clausius inequality.

8.2 Clausius Inequalty for a cyclic process

The Clausius inequality states that

δQ
∫T ≤0 (8.3)

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

for any cyclic process. In the above expression, equality sign applies in the
case of reversible cycle and < sign applies for an irreversible process. In
other words,

δQ
∫T = 0 for reversible cycle and
(8.4)
δQ
∫T < 0 for irreversible cycle.

Let us provide a simple proof for the Clausius inequality, using the concept of
reversible heat engine that we learnt in Chapter 7. 1
QH at TH
Considering the cycle 1-2-3-4-1, the above
integral can be evaluated as follows: p 2

2
δQrev QH
For isothermal process 1-2, ∫
1
T
=
TH 4

For the adiabatic expansion process 2-3, QC at TC 3

3
δQrev V

2
T
=0
Fig. 8.1 Heat Engine cycle

4
δQrev QC
For the isothermal process 3-4, ∫ =− , because heat is rejected.
3
T TC

1
δQrev
For the adiabatic compression process 4-1, ∫
4
T
=0

δQ QH Q
Thus, for the reversible cycle 1-2-3-4-1, ∫T =
TH
+ 0 − C + 0 = 0 , since
TC
(QH/TH) = (QC/TC) for the reversible heat engine cycle. Now, if the engine is
irreversible, then for the same heat input Q H, the rejected heat Q’C of the
irreversible engine will be larger in magnitude that the rejected heat Q C of
reversible engine (since irreversible engines are less efficient and produce
less work than reversible engine for the same values of T H, TC and heat input

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

QH). Therefore, for the irreversible engine having two heat transfer and two

δQ QH Q'
adiabatic processes, ∫ T
=
TH
+0 − C +0 <0.
TC

The above proof, given in the context


of a heat engine shows that at least
for heat engine cycles, the
expressions given by Eq. (8.4) are
valid. Let us now consider a general p
cycle, which may not be simple heat
engine cycle, as shown in Fig. (8.2).
Consider a portion of the cycle which
is intercepted by two adiabatic lines
(dotted lines) as shown in the figure.
V
Now for the cycle bounded by the two
Fig. 8.2 General thermodynamic cycle
adiabatic processes (dotted lines) and
two infinitesimal heat transfer
processes (solid lines), one can say

δQ
that ∫T ≤ 0 , depending on whether

it is reversible or irreversible.

In a similar manner, dividing the above general cycle into several closely
spaced adiabatic processes (with infinitesimal heat addition/ rejection
processes in between), it can be shown that for any general cycle,

δQ
∫T ≤0 (8.5)

where, equality sign applies if the entire cycle is reversible and the inequality
sign applies if any portion of the cycle has irreversibilities.

8.3 Proof that S is a thermodynamic property

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

2
δQrev
Now let us consider the quantity ∫
1
T
. It is to be noted here that between

given states 1 and 2, an infinite number of reversible processes are possible.


We want to show that irrespective of the actual process, the value of the
integral is the same.

Consider two reversible cycles 1-A-2-C-1 and 1-B-2-C-1 (as shown in Fig. 8.3)
with different paths from state 1 to state 2, but the same return path 2-C-1.
For both cycles, the cyclic integral is zero.

2
B
C
1

Fig. 8.3 Proof for path independence of ∆S

For the reversible cycle 1-A-2-C-1, we have

δQrev δQrev δQrev



1−A−2 −B −1
T
=0 = ∫
1−A −2
T
+ ∫
2 −C −1
T

(8.6)

Similarly, for the reversible cycle 1-B-2-C-1,

δQrev δQrev δQrev



1−B −2 −C −1
T
=0 = ∫
1−B −2
T
+ ∫
2 −C −1
T

(8.7)

Subtracting Eq. (8.6) from Eq. (8.7), we get:

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

δQrev δQrev

1− A −2
T
− ∫
1−B −2
T
= 0 , since the return path is the same.

(8.8)

Since 1-A-2 and 1-B-2 are arbitrary paths, it is evident that the integral

2
δQrev

1
T
is independent of path. This quantity is defined as the change in the

2
δQrev
thermodynamic property “Entropy”. Thus, ∆S = ∫ .
1
T

8.4 Entropy changes for various processes

In section 8.1, it was discussed that entropy is a measure of randomness in a


substance. For example, consider a crystalline solid in which the molecules
are arranged in a regular lattice. When heat is added to the solid, molecular
vibrations increase due to increase in temperature and eventually the solid
starts melting and the regular crystalline arrangement of molecules breaks
down. Thus, it is evident that due to the addition of heat, the randomness in
the material has increased. Similarly, when heat is added to the liquid, the
vapor molecules break free of intermolecular forces and start moving in
different directions at high speeds. The directions of motion of these
molecules are modified only through collisions. It is clear that heat addition to
liquid and the subsequent formation of vapor, also increases molecular
chaos.

Equation (8.2) described earlier can be understood as a statement that


molecular chaos (i.e. entropy) of a substance increases upon the addition of
heat. On the other hand, if work is added to a system, because work involves
order, there is no increase in entropy. However, in irreversible processes,
useful work is destroyed and converted into random molecular motion. For
example, when brakes are applied in a car moving at high speed, the kinetic
energy of the car will be converted to thermal form by friction. In turn, the
heat will be eventually dissipated to the environment, thereby increasing the

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

molecular chaos in the environment. If the brake had not been applied, the
car would have moved further and useful work would have been done. In an
irreversible process like this, entropy is produced, by the conversion of
‘possible work’ into random kinetic energy of molecules.

In general, entropy change for system consists of two contributions, as shown


below:

∆S = ∆S transfer + ∆S production (8.8)

where entropy transfer takes place along with heat transfer, and entropy
production occurs in irreversible processes by the dissipation of useful work
into molecular chaos. In a reversible process, there is no production of
entropy. Therefore, in a heat interaction between the system and the
surroundings, entropy increase for the system will correspond to an entropy
decrease of equal magnitude for the surroundings, and vice versa.

Surroundings System

Qrev

Fig. 8.4

δQrev
∆S transfer = ∫ = ∆S system = −∆S surroundings (8.9)
T

For the reversible process, therefore, ∆S system + ∆S surroundings = 0

(8.10)

Since the system and surroundings put together constitute the whole
Universe, it can be stated that for a reversible process,

∆S universe = ∆S system + ∆S surroundings = 0 (8.11)

For an irreversible process, ∆S system + ∆S surroundings = ∆S production > 0 (8.12)

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

In terms of the entropy change of the Universe, it can be stated that:

∆SUniverse ≥ 0 for any process, with = sign for reversible process and > sign
for irrversible process.

The above statement is sometimes viewed as an alternative statement of the


II law of Thermodynamics. We can explain this as follows. All processes in
practice, tend to be less efficient than a corresponding ideal situation,
because some of the useful energy gets dissipated in the form of entropy due
to irreversible phenomena. This is the reason for the lower performance
parameters observed in the case of irreversible processes, as compared to
those of reversible processes.

Let us look at the performance of the heat engine and heat pump (or
refrigerator) in the context of entropy, in slightly greater detail. For a
reversible heat engine, since it operates on a reversible cycle, there is no
production of entropy. Thus, the entropy added to the engine during heat
input = (QH/TH) = entropy transferred from the engine to the sink during heat
rejection = (QC/TC). Net entropy change for the engine operating on cycle =
(QH/TH) - (QC/TC) = 0. The work produced by the reversible engine is equal to
(QH-QC).

For the irreversible heat engine, entropy added during heat input = (Q H/TH).
However, due to irreversibilities (such as friction), some entropy is produced
within the engine. For the engine to operate in a cycle, net entropy change
should be zero. Thus, entropy rejected should be (Q’ C/TC) = (QH/TH) +

δQ
∆Sproduction. Or, (Q’C/TC) > (QH/TH). For this reason, ∫ < 0 , in the case of an
T
irreversible heat engine cycle. Also, the work output = Q H - Q’C, which is less
than the work output of the corresponding reversible engine. In other words,
energy that could have been delivered as useful work, gets dissipated in the
engine as molecular chaos (entropy) due to irreversibilities such as friction.
During the cyclic process undergone by the working fluid, the entropy thus

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

produced is transferred to the environment as additional heat rejected. This


is why the thermal efficiency of an irreversible heat engine is always less
than that of a reversible engine. In a similar manner, it can be shown that the
COP of an irreversible heat pump (or refrigerator) is less than that of the
corresponding reversible system.

In reversible processes, heat addition results in the increase of entropy of the


system and heat rejection results in decrease of entropy of the system, as
can be seen from Eq. (8.2). In the case of a reversible adiabatic process, ∆S =
0. Referring to Eq. (8.8), it is clear that in the case of a reversible adiabatic
process, entropy production is zero (because of reversible process) and
entropy transfer is also zero (for adiabatic process). Thus slow expansion in a
frictionless adiabatic turbine will be a process with ∆S = 0. Such a process is
known as an “Isentropic Process” or process with constant entropy. In
general, adiabatic devices such as turbine, compressor, pump, nozzle and
diffuser can be assumed undergo isentropic processes in the ideal case.
Practical adiabatic devices will have some entropy increase during the
process due to irreversibilities such as friction.

An important factor that needs to be pointed out is that for irreversible


processes,

δQrev δQ
∆S = ∆S transfer + ∆S production = ∫ + ∆S production ⇒ ∆S > ∫ rev (8.13)
T T

The above inequality is to be understood as follows. The ratio of heat added


by temperature gives only the entropy transferred to the system from the
surroundings. In addition, during the irreversible process, if entropy is
generated, the total entropy change of the system will be greater than
∆Stransfer.

8.5 Entropy calculations for various processes1and substances

Consider the end states 1and 2, as shown


in Fig. 8.5. The entropy change between p

2
Dept. of Mechanical Engineering Indian Institute of Technology Madras

V
Fig. 8.5
ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

these end states can be calculated with the


help of any reversible process, connecting
the two states. The entropy change for an
infinitesimal part of the process is given by

δQrev
dS = ⇒ T .dS = δQrev = dU + p.dV (8.14)
T

for a simple compressible system. The above expression has to be viewed as


a property relationship between, T, U, p and V, since it can be applied along
any reversible path between 1 and 2. Therefore, entropy can be calculated
from

T .dS = dU + p.dV (8.15)

Using the definition of enthalpy (H = U + pV), it is also possible to rewrite the


above equation as

T .dS = dU + p.dV = dH −V .dp (8.16)

Equations (8.15) and (8.16) are called as T-dS relations or “Gibbs Relations”
and they can be employed for calculating the entropy change in a process.
Although Eq.(8.14) and hence Eqs. (8.15) and (8.16), have been derived on
the basis of a reversible process, the entropy change ∆S between given
states 1 & 2 is independent of what the actual process is. Even if an
irreversible process took place between the end states 1 and 2, still the
entropy change ∆S will be the same, because entropy is a thermodynamic
property and it is only a state function. In the case of a reversible process, ∆S
will be made up of ∆Stransfer (that associated with heat transfer) only. For an
irreversible process between states 1 and 2, total ∆S (= S2-S1) will be the
same as in the case of the reversible process, but there will be contributions
from both entropy transfer and entropy production.

Now let us calculate the entropy change for an ideal gas undergoing some
process.

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

mRT
T .dS = dU + p.dV = mCV dT + dV
V
 dT dV 
dS = m CV +R  (8.17)
 T V 
 T  V 
∆S = m CV ln 2  + R ln 2 
  T1   V1 

Using the second T-dS relationship, the entropy change of an ideal gas can be
obtained as:

mRT
T .dS = dH − V .dp = mC P dT − dp
p
 dT dp 
dS = m C P −R 
 T p  (8.18)

 T   p 
∆S = m C P ln 2  − R ln 2 
  T1   p1 

For an ideal gas undergoing a reversible adiabatic process for which ∆S = 0


(isentropic process), we get:

CP γ
 T   p  p T  R T  γ −1
∆S = m C P ln 2  − R ln 2  = 0 ⇒ 2 =  2  =  2  (8.19)
  T1   p1  p1  T1   T1 

For a solid or liquid undergoing change in temperature, the T-dS relationship


can be simplified, assuming negligible change in volume. In fact, for solids
and liquids, there is no difference between C P and CV. One can consider a
common specific heat C and the T-dS relationship Eq. (8.15) can be simplified
and the entropy change can be obtained as:

T .dS = dU + 0 = mCV dT
 dT  T  (8.20)
dS = m CV  ⇒ ∆S = mCV ln 2 
 T   T1 

For a phase change process such as boiling of a liquid occurring at constant


pressure (and temperature), Eq. (8.16) can be simplified as follows:

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

h fg
T .dS = H − V .dp ⇒ ∆S = (8.21)
Tsat

For a reversible isothermal heat transfer process,

δQrev Q
dS = ⇒ ∆S = rev (8.22)
T T

The above relationship can be rewritten in the form T.dS = δQrev


(8.23)

Thus, the area under the curve in T-S coordinates gives the heat transfer, for
a reversible process (similar to the fact that area under the curve in p-V
diagram represents the displacement work done in a quasi-static, frictionless
process). In figure 8.6 a, the shaded area represents the amount of heat
transfer during the associated reversible process. Also, as discussed earlier,
in the case of a reversible, adiabatic process

dS = 0; Or, S1 = S2 = constant. (8.24)

Hence, a reversible adiabatic process is isentropic (i.e. entropy of the


system remains constant during such a process). In T- S coordinates, the
reversible adiabatic process is shown as a vertical line.

1
1

T T
2
2

S S
Fig. 8.6 a Reversible heat transfer Fig. 8.6 b Reversible
adiabatic process

For devices such as turbine, compressor, pump, nozzle, diffuser etc, ideal
conditions would correspond to adiabatic, frictionless operation of the device.

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

For example, in an ideal steam turbine, one can assume that steam expands
reversibly and adiabatically between the boiler and condenser pressures. The
ideal expansion process can be shown as a vertical line between the inlet and
exit pressures (p1 and p2) for the turbine. However, in an actual turbine,
although the heat losses may still be neglected (i.e. adiabatic process),
irreversibilities such as friction may be present and hence the entropy may
increase due to entropy production. The performance of the turbine would be
affected (work output of the turbine would be less) and this can be expressed
with the help of a parameter known as isentropic efficiency of the turbine,
ηt. The isentropic efficiency of the turbine is defined (see Fig. 8.7 a) as

actual turbine work h − h2


ηt = = 1 (8.25)
ideal turbine work h1 − h2 s

Similarly, in the case of expansion in an adiabatic nozzle, the isentropic


efficiency of the nozzle can be defined as

actual kinetic energy change h1 − h2


ηn = = (8.26)
ideal kinetic energy change h1 − h2 s

For an adiabatic compressor operating between inlet and exit pressures p 1


and p2, the ideal compression process can be taken as isentropic (vertical line
in T-S diagram as shown in Fig. 8.7 b), while the actual process may have
some entropy generation. The isentropic efficiency of the compressor can
therefore be defined as

ideal compressor work input h − h2 s


ηc = = 1 (8.27)
actual compressor work input h1 − h2

2
1 2s

T T
2

Dept. of Mechanical Engineering2s Indian Institute


1 of Technology Madras

S S
ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

Fig. 8.7 a Adiabatic expansion Fig. 8.7 b Adiabatic


compression (in turbine / nozzle)
(in compressor/ diffuser)

It is clear from the above discussions that ideal adiabatic machines such as
turbines, compressors, nozzles, diffusers etc. may operate on isentropic
processes, while in the corresponding practical devices entropy will increase
due to irreversibilities. However, if there is some heat loss from the device,
entropy could also decrease from the inlet to the exit (due to entropy
transfer).

8.6 Representation of processes in T-S diagram

The representation of various processes in p-V diagram and the associated


displacement work expressions were discussed in Chapter 2. In a similar
manner, it is possible to represent different processes in a T-S diagram and
the heat transfer associated with each process can be evaluated as the area
under the curve. Let us for example, consider polytropic processes of the
form p.Vn = constant for an ideal gas substance. Here, n = 0 corresponds to a
constant pressure process, n= 1 represents isothermal process, n = γ
represents a reversible adiabatic process and n  ∞ represents a constant
volume process. In the T-S diagram, the isothermal process becomes a
horizontal line and the adiabatic process becomes a vertical line (isentropic).
Therefore, the constant pressure process will be a curve above the isothermal
p = const (n=0)
line and the constant volume process will be a curve to the right of the
adiabatic line (Fig. 8.8 b).
p = const (n=0)

p T T = const (n=1)
T = const (n=1)

V = const adiabatic
(nα) (n=γ )
V = const adiabatic (n=γ )
Dept. of Mechanical Engineering Indian Institute of Technology Madras
(nα)
V S
ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

Fig. 8.8 a Fig. 8.8b

The curves shown in the pictures above correspond to expansion processes


(volume increasing) or constant volume cooling. The curves corresponding to
compression processes or constant volume heating, can be similarly obtained
extending these curves in the opposite direction from the initial state.

The reversible heat engine cycle (Carnot cycle) becomes a simple rectangle
in the T-S diagram.

1 2 1-2 Reversible isothermal heat addition

T 2-3 Reversible adiabatic expansion


4 3
3-4 Reversible isothermal heat
rejection
S
Fig. 8.9 4-1 Reversible adiabatic compression

The area under the line 1-2 represents the heat input Q H and the area under
the line 3-4 represents the rejected heat Q C. The area enclosed by the
rectangle = QH – QC = Wnet for the cycle. Instead of traversing in a clockwise
sense, if the cycle (rectangle shown in Fig. 8.9) is traversed in the anti-
clockwise sense, the corresponding cycle will correspond to a Carnot heat
pump or a Carnot refrigerator.

8.7 Available and unavailable energy components

In any energy interaction, one can associate an available part and an


unavailable part of the energy interaction. The available part is the portion of
energy which can be converted into work by operating a reversible cycle and

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

returning the material to its initial state. The unavailable part is the portion of
energy that necessarily has to be rejected to the ambient as heat (since the
ambient is the final heat sink for most of the energy interaction processes
carried out by human beings). In Fig. 8.10, we consider a heat addition
process 1-2. The area below the curve 1-2 is the total amount of heat
transferred in the process. Now, if we carry out a reversible cycle (shown by
the dash lines which complete the cycle and bring back the system to its
initial state), the heat rejected to the ambient is the area under the line
corresponding to T = T amb. Now, the area enclosed between the curve 1-2 and
the isotherm T = T amb, is the available energy and the area below T=T amb
isotherm (the hatched portion) is the unavailable energy which has to be
rejected as heat to the sink. Thus, the unavailable energy = T amb. ∆S and the
available part = Q - Tamb. ∆S.

T
1
Tamb

Fig. 8.10 Available and unavailable energy components

S
The above discussions bring another important aspect into focus. In any
irreversible process, useful work gets destroyed by irreversible phenomena
and the corresponding energy is eventually rejected as heat to the ambient.
Thus, a work equivalent of Tamb. ∆Sproduction is wasted. This wasted work or “Lost
work” is also known as Irreversibility, I. In any process, one can evaluate I as
follows.

Lost Work = I = Tamb. ∆Sproduction = Tamb. { ∆ Ssystem + ∆Ssurroundings}


(8.28)

Dept. of Mechanical Engineering Indian Institute of Technology Madras


ME1100 Thermodynamics Lecture Notes Prof. T. Sundararajan

In a reversible process, ∆ Ssystem = - ∆Ssurroundings and ∆ Sproduction = 0. Therefore,


no useful work is lost. In an irreversible process or irreversible cycle, the
overall entropy of the universe increases (entropy of system + entropy of
surroundings) because of entropy production due to irreversibile mechanisms
such as friction. The entropy thus produced eventually increases the
molecular chaos of the universe.

Dept. of Mechanical Engineering Indian Institute of Technology Madras

You might also like