You are on page 1of 117

Chapter 8: Entropy

Thermodynamics 1
Introduction
 The first law of thermodynamics lead to the
definition of a property called internal energy.
That enabled to use the first law quantitatively
for processes.
 The second law of thermodynamics leads to the
definition of a new property called entropy. That
enables to treat the second law quantitatively for
processes.
 To obtain the working definition of entropy, let's
derive the Clausius inequality.

Thermodynamics – Chapter 8 2
The Clausius Inequality
 The second law of thermodynamics often leads
to expressions that involve inequalities.
An irreversible heat engine, is less efficient
than a reversible one operating between the
same two thermal energy reservoirs.
An irreversible refrigerator or a heat pump
has a lower COP than a reversible one
operating between the same temperature limits.
 Another important inequality that has major
consequences in thermodynamics is the
Clausius inequality.

Thermodynamics – Chapter 8 3
The Clausius Inequality
It was first stated by the German Physicist R.J.E.
Clausius (1822-1888), one of the founders of
thermodynamics, and is expressed as

Q
 T
0

The cyclic integral of  Q / T can be viewed as the


sum of all differential amounts of heat transfer
divided by the temperature at the boundary. This
inequality is valid for all cycles, reversible or
irreversible.
Thermodynamics – Chapter 8
The Validity of the Clausius Inequality
Consider a system connected
to a thermal reservoir at a
constant thermodynamic
temperature of TR through a
reversible cyclic device. The
cyclic device receives heat
δQR from the reservoir and
supplies δQ to the system
whose temperature at that
part of the boundary is T.

Thermodynamics – Chapter 8 5
We apply the first law on an incremental basis to
the combined system composed of the heat engine
and the system.
Ein  Eout  Ec
 QR  ( Wrev   Wsys )  dEc

where δEc is the change in the total energy of the


combined system. Let δWc be the total work done
by the combined system. Then the first law
becomes
 QR   Wc  dEc
where  Wc   Wrev   Wsys

Thermodynamics – Chapter 8 6
Considering that the cyclic device is a reversible
one, then
 QR  Q Q
   QR  TR
TR T T
Eliminating δQR from the energy equation using the
above relation yields
Q
Wc  TR  dEc
T
Now let the system undergo a cycle while the cyclic
device undergoes an integral number of cycles,
then Q
Wc  TR    dEc
T
Thermodynamics – Chapter 8 7
If the system, as well as the heat engine, is required
to undergo a cycle, then
 dEc
0

and the total net work becomes


Q
Wc  TR 
T
 If Wc is positive, it appears that the combined
system is exchanging heat with a single heat
reservoir and producing an equivalent amount of
work; thus, the Kelvin-Planck statement of the
second law is violated.
Thermodynamics – Chapter 8 8
 But Wc can be zero (no work done) or negative
(work is done on the combined system) and not
violate the Kelvin-Planck statement of the
second law. Therefore, since TR> 0 (absolute
temperature), it must be

Q
 T
0

which is the Clausius inequality. This inequality is


valid for all thermodynamic cycles, reversible or
irreversible, including the refrigeration cycles.
Thermodynamics – Chapter 8 9
If no irreversibilities occur within the system as
well as the reversible cyclic device, then the
combined cycle undergone by the combined
system is internally reversible. As such, it can be
reversed. In the reversed cycle case, all the
quantities have the same magnitude but the
opposite sign. Therefore, Wc, which could not be a
+ive quantity in the regular case, can’t be a -ive
quantity in the reversed case. Then the work
WC,inte rev = 0, and therefore
Q 
  T   0
int rev

Thermodynamics – Chapter 8 10
for internally reversible cycles.
Thus, we conclude that the equality in the Clausius
inequality holds for totally or just internally
reversible cycles and the inequality for the
irreversible ones.

 Q 
  T   0 internally or totally reversible
 Q 
  T   0 irreversible

Thermodynamics – Chapter 8 11
Demonstration of the inequality of Clausius
Consider a reversible
(Carnot) heat engine
cycle operating between
reservoirs at temperatures
TH and TL. For this cycle,

 Q  QH
 QL  0

Since TH and TL are constant, from the definition of


the absolute temperature scale and from the fact it is
a reversible cycle,
Thermodynamics – Chapter 8 12
dQ QH QL
 T
 
TH TL
0

If the cyclic integral of δQ, approaches zero by


making TH approaches TL and the cycle remain
reversible, the cyclic integral of δQ/T remains zero.
Q
  Q  0 and 
T
0
Thus, we conclude that for all reversible heat
engine cycles,
Q
  Q  0 and  T
0

Thermodynamics – Chapter 8 13
Consider an irreversible heat engine cycle operating
between same TH and TL as the reversible engine
and receiving the same quantity of heat QH.
Comparing the irreversible cycle with reversible
one, we can conclude from the second law that,
Wirre  Wrev
Since QH  QL  W for both reversible and irreversible
cycles,
QH  QL irre  QH  QL rev
and therefore
QL irre  QL rev

Thermodynamics – Chapter 8 14
Consequently, for the irreversible cyclic engine,
dQ QH QL irre
 Q  QH
 QL irre  0 and 
T

TH

TL
0

Now causing the engine to become more and more


irreversible, but keep QH, TH, and TL fixed. The
cyclic integral of δQ then approaches zero, and that
for δQ/T becomes a progressively larger negative
values. Q
  Q  0 and   0 T
Thus, we conclude that for all irreversible heat
engine cycles,
Q
  Q  0 and  T  0
Thermodynamics – Chapter 8 15
Heat Engine
 Q  Q H
 QL
 Wrev  QL   QL
 Wrev
Therefore for reversible HE   Q  0

 Q  Q H
 QL irre
 Wirre  QL irre   QL irre
 Wirre
Therefore for irreversible HE   Q  0
Thermodynamics – Chapter 8 16
W TH  TL
 HE  and Carnot 
QH TH
W TH  TL

QH TH
QH  QL TH  TL

QH TH
QL TL
1  1
QH TH
QL TL QL QH
  
QH TH TL TH

Thermodynamics – Chapter 8 17
W TH  TL
Real HE  and Carnot 
QH TH
W TH  TL

QH TH
QH  QL irre TH  TL
<
QH TH
QL irre TL
1  1
QH TH
QL irre TL QL irre QH
  
QH TH TL TH

Thermodynamics – Chapter 8 18
Consider a reversible
refrigeration cycle operating
between reservoirs at
temperatures TH and TL. For
this cycle,

  Q  Q
H
 QL  0

From the definition of the absolute temperature


scale and from the fact it is a reversible cycle,
dQ QH QL
 T
 
TH TL
0

Thermodynamics – Chapter 8 19
As the cyclic integral of δQ, approaches zero by
making TH approaches TL and the cycle remain
reversible, the cyclic integral of δQ/T remains zero.

Q
  Q  0 and  T
0

Thus, for all reversible refrigeration cycles,

Q
  Q  0 and  T
0

Thermodynamics – Chapter 8 20
Let an irreversible cyclic refrigerator operating
between TH and TL and receive the same quantity
of heat QL as the reversible refrigerator. Comparing
the irreversible cycle with reversible one, we can
conclude from the second law that,
W irre  Wrev
Since QH  QL  W for both reversible and irreversible
cycles,
QH irre  QL  QH rev  QL
and therefore
QH irre  QH rev

Thermodynamics – Chapter 8 21
Consequently, for the irreversible refrigerator,
Q QH irre QL
  Q  QH irre  QL  0 and   
T TH

TL
0

Making the refrigerator to become more and more


irreversible, but keep QL, TH, and TL fixed. The
cyclic integral of δQ and δQ/T become larger in the
negative direction.
Q
  Q  0 and   0
T
Thus, we conclude that for all irreversible
refrigerators,
Q
  Q  0 and  T  0
Thermodynamics – Chapter 8 22
Refrigerator

  Q  QH
 QL
  Wrev  QL   QL
 Wrev
Therefore for reversible refrigerator   Q  0

  Q  QH irre
 QL
  Wirre  QL   QL
 Wirre
Therefore for irreversible refrigerator   Q  0
Thermodynamics – Chapter 8 23
QL TL
COPrefrigerator  and COPCarnot 
W TH  TL
QL TL

W TH  TL
QL TL

QH  QL TH  TL
1 1

QH TH
1 1
QL TL
QH TH QH TH QL QH
1  1    
QL TL QL TL TL TH

Thermodynamics – Chapter 8 24
QL TL
COPreal  and COPCarnot 
Wirre TH  TL
QL TL

Wirre TH  TL
QL TL
<
QH irre  QL TH  TL
1 1
<
QH irre TH
1 1
QL TL
QH irre TH QH irre TH QH irre QL
1  1    
QL TL QL TL TH TL

Thermodynamics – Chapter 8 25
Summary of the Demonstrations
We have considered all possible reversible cycles
 i.e.,   Q  0 &   Q  0  , and for each of these
reversible cycles
Q
 T 0
we have considered all possible irreversible cycles
 i.e.,   Q  0 &   Q  0  , and for each of these
irreversible cycles
Q
 T 0
Thermodynamics – Chapter 8 26
Thus, for all cycles we can write

Q
 T
0

where the equality holds for reversible cycles and


the inequality for irreversible cycles. This relation
in known as the inequality of Clausius.

Thermodynamics – Chapter 8 27
Does this Cycle Satisfy Clausius Inequality ?

Heat is transferred in two places, the boiler and the


condenser. Therefore,
dQ  dQ   dQ 
 T   T     T 
boiler condenser

Thermodynamics – Chapter 8 28
The temperature remains constant in both the boiler
and condenser.
dQ 1 2 14 Q 3 Q4
 T  T 1 dQ  T 3 dQ  T  T
1 2

1 3 1 3

Let us consider a 1 kg mass as the working fluid.


q
1 2
 h2
 h1
 2039.3 kJ/kg, T1
 164.97 0
C
q
3 4
 h4
 h3
 463.4  2361.8  1898.4 kJ/kg, T3
 53.97 0
C
dQ 2039.3 1898.4
 T 164.97  273.15 53.97  273.15  1.1487 kJ/kg.K
 

Thus, this cycle satisfies the inequality of Clausius,


which is equivalent to saying that it does not violate
the second law of thermodynamics.
Thermodynamics – Chapter 8 29
Entropy – A Property of a System
Let a system (control mass) undergo a reversible
process from state 1 to state 2 along a path A, and
let the cycle be completed along path B, which is
also reversible. Since it is reversible,

Q 2
 Q  1  Q 
 T
0     
1  T A
  (1)
2  T B

Now consider another reversible, which proceeds


first along path C and is then completed along path
B. For this cycle, we can write
Thermodynamics – Chapter 8 30
Q 2
 Q  1  Q 
 T
 0     
1  T C
    (2)
2  T B

Subtracting the second equation from the first, we


get 2
 Q 
2
 Q 
1  T   1  T 
A C

Since the  dQ / T is the same for all reversible paths


between states 1 and 2, this quantity is independent
of the path and it is a function of the end states only.
Therefore it is a property. This property is called
entropy and denoted by S.
 Q 
dS         (3)
 T rev
Thermodynamics – Chapter 8 31
The change in the entropy of a system as it
undergoes a change of state is found by integrating
Eq. 3. Thus,
2
 Q 
S2  S1     --------(4)
1  T  rev

 To perform the integration, one needs to know


the relation between Q and T during a process.
 Q is a path function, therefore δQ is an inexact
differential.
 1/T serves as the integrating factor in converting
the inexact differential to the exact differential
δQ/T for a reversible process.
Thermodynamics – Chapter 8 32
 Since entropy is a property, the change in
entropy of a substance in going from one state
to another is the same for all processes, both
reversible and irreversible.

Thermodynamics – Chapter 8 33
The Entropy of a Pure Substance
 Entropy is an extensive property of a system.
 The units of specific entropy in the steam tables,
refrigerant tables, and ammonia tables are
kJ/kg.K.
 The values are given relative to an arbitrary
reference state.
 In the steam tables the entropy of saturated
liquid at 0.01oC is given the value of zero and
for many refrigerants the entropy of saturated
liquid at -40oC is assigned the value of zero.
Thermodynamics – Chapter 8 34
 In the saturation region the specific entropy is
calculated using the quality.

s  1  x  s f  x sg
s  s f  x s fg

 Property diagrams serve as great visual aids in


the thermodynamic analysis of processes. We
have used P-v and T-v diagrams extensively
in previous chapters in conjunction with the first
law of thermodynamics.

Thermodynamics – Chapter 8 35
 In the second- law analysis, it is very helpful to
plot the processes on diagrams for which one of
the coordinates is entropy.
 The two diagrams commonly used in the second-
law analysis are the temperature-entropy and
the enthalpy-entropy diagrams.
 The thermodynamic properties of a substance
that are shown on an enthalpy-entropy diagram,
is also called a Mollier diagram, named after
Richard Mollier (1863-1935) of Germany.

Thermodynamics – Chapter 8 36
Temperature – entropy diagram for steam

Thermodynamics – Chapter 8 37
Enthalpy – entropy diagram for steam

Thermodynamics – Chapter 8 38
For most substances
the difference in the
entropy of a
compressed liquid
and a saturated
liquid at the same
temperature is so
small.

Thermodynamics – Chapter 8 39
Entropy Change in Reversible Processes
Now will consider the significance of entropy in
various process. Let the working fluid of a heat
engine operating on the Carnot cycle make up the
system.
 The first process is the isothermal transfer of
heat to the working fluid from the high-
temperature reservoir. For this reversible
process we can write,
2
 Q  1 2 Q
S2  S1       Q 
1 2

1  T  rev TH 1 TH

Thermodynamics – Chapter 8 40
 The second process is reversible adiabatic
expansion. From the definition of entropy,
 Q 
dS   
 T rev
 it is evident that the entropy remains constant
for this process. A constant-entropy process is
called an isentropic process.

 The third process is the reversible isothermal


process in which heat is transferred from
working fluid to the low-temperature reservoir.
For this reversible process we can write,

Thermodynamics – Chapter 8 41
4
 Q  1 4 Q
S 4  S3       Q 
3 4

3  T  rev TL 3 TL
 during this process the heat transfer is negative,
therefore the entropy of the working fluid
decreases.

 The final process is reversible adiabatic


compression, entropy remains constant for this
process (isentropic process).
 All these processes can be represented on a
temperature-entropy diagram.

Thermodynamics – Chapter 8 42
The Carnot cycle on the T-S diagram
 The area under line 1-2, area 1-2-b-a-1,
represents the heat transferred to the working
fluid.
 The area under line 3-4, area 3-4-a-b-3,
represents the heat transferred from the working
fluid.
Thermodynamics – Chapter 8 43
 Since the net work of the cycle is equal to the
net heat transfer. The efficiency of the cycle,
Wnet area 1-2-3-4-1
th  
QH area 1-2-b-a-1

 Increasing TH while TL remains constant


increases the efficiency.
 Decreasing TL as TH remains constant increases
the efficiency.
 The efficiency approaches 100% as the absolute
temperature at which heat is rejected approaches
zero.

Thermodynamics – Chapter 8 44
A Special Case: Internally Reversible Isothermal
Heat Transfer Processes

Consider the change of


state from saturated liquid
to saturated vapor at
constant temperature. Line
1-2 on the T-S diagram
represents this process.

1 2  Q  1 2 q h fg
s2  s1  s fg       Q  1 2

m 1  T  rev mT 1 T T

Thermodynamics – Chapter 8 45
If heat is transferred to the saturated vapor at
constant pressure, the steam is superheated along
line 2-3. For this process we can write,
13 3

q    Q   Tds
2 3
m2 2

Since T is not constant, this equation cannot be


integrated unless we know the relation between
temperature and entropy. However, the area under
3

line 2-3, area 2-3-c-b-2, represents  Tds and therefore


2
represents the heat transferred during this reversible
process.

Thermodynamics – Chapter 8 46
(Ex 8.1) Consider a Carnot-cycle heat pump with
R-134a as the working fluid. Heat is absorbed
into the R-134a at 0OC, during which process it
changes from a two-phase state to saturated
vapor. The heat is rejected from the R-134a at
60OC so that it ends up as saturated liquid. Find
the pressure after compression, before the heat
rejection process, and determine the coefficient
of performance for the cycle.

Thermodynamics – Chapter 8 47
State 4 (B.5.1)  s4  s3  s f @60o C  1.2857 kJ/kg K
State 1 (B.5.1)  s1  s2  sg @0o C  1.7262 kJ/kg K
State 2 (B.5.2)  60o C, s2  s1  1.7262 kJ/kg K
Interpolate between 1400 kPa and 1600 kPa in B.5.2:
1.7262  1.736
P2  1400  (1600  1400)  1487.1 kPa
1.7135  1.736
Thermodynamics – Chapter 8 48
TH 333.15
COP    5.55
TH  TL 60

Remark: Notice how much the pressure varies during


the heat rejection. Because this process is very
difficult to accomplish in a real device, no heat pump
or refrigerator is designed to attempt to approach a
Carnot cycle.
Thermodynamics – Chapter 8 49
(Ex 8.2) A cylinder/piston setup contains 1 L of
saturated liquid refrigerant R-12 at 20oC. The
piston now slowly expands, maintaining constant
temperature to a final pressure of 400 kPa in a
reversible process. Calculate the required work
and heat transfer required to accomplish this
process.

Thermodynamics – Chapter 8 50
State 1 : Saturated liquid, 20oC (Table B3.1)
u1 = 54.45 KJ/kg, s1 = 0.2078 KJ/kg K, P1 = 567.3 kPa
m = V/v1= 0.001 m3/ 0.000752 m3/kg = 1.33 kg
State 2: 20oC, 400 kPa (Table B3.2)
u2 = 180.57 KJ/kg, s2 = 0.7204 KJ/kg K
Q Q2
At const T , dS =  T

T
1

Thermodynamics – Chapter 8 51
1 Q2  mT ( s2  s2 )
 1.33  293.15  (0.7204  0.2078)
 200 kJ

1 W2  1 Q2  m(u2  u1 )
 1.33  (54.45  180.57)  200
 32.3kJ

Thermodynamics – Chapter 8 52
The Thermodynamic Property Relation
Two important thermodynamic relations for a
simple compressible system are,
T dS  dU  P dV
T dS  dH  V dP
Consider a simple compressible substance in the
absence of motion or gravitational effects. The first
law for a change of state under these conditions can
be written
dU   Q   W   Q  dU   W

Thermodynamics – Chapter 8 53
For a reversible process of a simple compressible
substance, we can write
 Q  T dS and W  P dV
Substituting these relations into the first-law
equation, we get
T dS  dU  P dV -------- (5)
This equation was developed for a reversible
process. However, the results obtained are valid for
both reversible and irreversible processes since
entropy is a property and the change in a
property between two states is independent of the
type of process the system undergoes.

Thermodynamics – Chapter 8 54
The property enthalpy is defined as
H  U  PV
it follows that
dH  dU  PdV  VdP 
dU  dH  PdV  VdP
Substituting this relation into Eq. 5, we get
T dS  dH  V dP -------- (6)
The two forms of the thermodynamic property
relation (Eq. 5&6) are frequently called Gibbs
equations.
T ds  du  P dv 

T ds  dh  v dP 
on unit mass basis

Thermodynamics – Chapter 8 55
The T dS relations are valid for both
reversible and irreversible processes and for
both closed and open systems.
Thermodynamics – Chapter 8 56
Entropy Change of a Control Mass During an
Irreversible Process

Consider a control mass that


undergoes the cycles. The
reversible cycle is made up of
the reversible processes A and B.

Q Q 
2 1
Q 
    
T 1  T  A 2  T B
 0
1
Q  2
Q 
      
2  T B 1  T A

Thermodynamics – Chapter 8 57
Entropy Change of a Control Mass During an
Irreversible Process

Another irreversible cycle is


made up of the irreversible
process C and reversible process
B. Q 2Q  1
Q 
      0
T 1  T C 2  T  B
2
Q  2
Q 
     0
1  T C 1  T A

Q 
2 2
Q 
     
1 T A 1  T C

Thermodynamics – Chapter 8 58
Since path A is reversible, and entropy is a property,
2
 Q  2 2

1  T   1 dS A  1 dSC
A

Therefore,
2 2
 Q 
1 dSC  1  T 
C

As path C was arbitrary, the general result is


Q 2
Q
dS  and S 2  S1  
T 1 T
This is one of the most important equations of
thermodynamics. It is used to develop a number of
concepts and definitions. It states the influence of
irreversibility on the entropy of a control mass.
Thermodynamics – Chapter 8 59
 If an amount of heat δQ is transferred to a
control mass at temperature T in a reversible
process, the change in entropy is given by,
 Q 
dS   
 T  rev
 If any irreversible effects occur while the
amount of heat δQ is transferred to a control
mass at temperature T, the change of entropy
will be greater than the reversible process.
 Q 
dS   
 T irrev
Thermodynamics – Chapter 8 60
Entropy Generation
The entropy change for an irreversible process is
larger than the change in a reversible process for the
same δQ and T. This can be written as
Q
dS    S gen
T
Provided the last term is positive   S gen  0

where δSgen, is the entropy generation in the


process due to irreversibilities occurring inside the
system. This internal generation can be caused by
friction, unrestrained expansions, and the internal
heat transfer over finite T .
Thermodynamics – Chapter 8 61
Consider a reversible process, for which the entropy
generation is zero, and the heat transfer and work
terms therefore are
 Q  TdS and W  PdV
For an irreversible process with a nonzero entropy
generation, the heat transfer is given by
 Qirre  TdS  T  S gen
Thus δQirre is smaller than for the reversible case
for the same change of state, dS. For an irreversible
process  Q  dU   W
irre irre
TdS  T  S gen  dU   Wirre

Thermodynamics – Chapter 8 62
and the property relation is valid,
T dS  dU  P dV
it is found that,
TdS  T  S gen  dU   Wirre
TdS  T  S gen  T dS  P dV   Wirre
 Wirre  PdV  T  S gen
showing that the work is reduced by an amount
proportional to the entropy generation. For this
reason the term T δSgen is often called “lost work”
although it is not a real work or energy quantity lost
but rather a lost opportunity to extract work.
Thermodynamics – Chapter 8 63
The entropy balance equation for a control mass is
2 2
Q
S2  S1   dS    1 S2 gen
1 1 T
2
Q
where  is know as entropy transfer
1 T

 Note that the entropy generation Sgen is always a


positive quantity or zero.
 Its value depends on the process, and thus it is
not a property of the system.
 Also, in the absence of any entropy transfer, the
entropy change of a system is equal to the
entropy generation.
Thermodynamics – Chapter 8 64
Some Important Conclusions from the Entropy
Generation 2 2
Q
S2  S1   dS    1 S 2 gen
1 1 T

 There are two ways in which entropy of a system


can be increased -
 By transferring heat to it.
 By having an irreversible process.
 For an adiabatic process, δQ = 0, and therefore
the increase in entropy is always associated with
the irreversibilities.

Thermodynamics – Chapter 8 65
 Since the entropy generation cannot be less than
zero, there is only one way in which entropy of a
system can be decreased, and is to transfer heat
from the system.

Change of entropy due to heat


transfer and entropy generation

 The presence of irreversibilities will cause the


work to be smaller then the reversible work. This
means less work out in an expansion process and
more work into the control mass in a
compression process.
Thermodynamics – Chapter 8 66
 The work for an irreversible process is not equal
to  P dV and the heat transfer is not equal to  T dS .
 Therefore, the area underneath the path does not
represent work and heat on the P-V and T-S
diagrams.

Reversible and irreversible processes on P-V and T-S diagrams

Thermodynamics – Chapter 8 67
Principle of the Increase of Entropy
 The entropy change of a control mass could be
either positive or negative, since entropy can be
increased by internal entropy generation and
either increased or decreased by heat transfer,
depending on the direction of the heat transfer.
 Let us examine the effect of heat transfer on the
change of state in the surroundings, as well as on
the control mass itself.

Thermodynamics – Chapter 8 68
Consider the process, in which a quantity of heat
δQ is transferred from the surroundings at
temperature To , to the control mass at temperature
T. Let the work done during this process be δW.

For the control mass,


Q
dSc.m. 
T
For the surroundings at T0 , δQ is negative, and
assume a reversible heat transfer
 Q
dSsurr 
T0
Thermodynamics – Chapter 8 69
The total net change of entropy is
Q Q
dSnet  dSc.m.  dSsurr  
T T0
1 1 
 Q  
 T T0 
Since T0 >T, [(1/ T) - [(1/ T0)] the quantity is
positive, and we conclude that
dSnet  dSc.m.  dSsurr  0

If T >T0 , both δQ and the quantity [(1/ T) - [(1/ T0)]


are negative, thus yielding the same result.

Thermodynamics – Chapter 8 70
1 1 
dSnet  dSc.m.  dSsurr  Q  
 T T0 
The right hand side of the equation represents an
external entropy generation due to heat transfer
through a finite temperature difference.

To amplify this point, take a control mass that


connects the surrounding at To , with the previous
control mass T, which typically is the wall. This
mass does not experience any change of state, yet
heat transfer causes fluxes of entropy to flow in and
out in an irreversible process. For this process,

Thermodynamics – Chapter 8 71
Q Q
dSc.m. 2  0     Sgen, 2
T0 T
The difference in the two δQ/T terms (fluxes of S)
is the entropy generated in this control mass.
1 1
 Sgen, 2  Q  
 T T0 
This is the location in space where the heat transfer
takes place over the finite temperature difference
To –T. This term is always positive (or zero for an
adiabatic process), but as the temperature difference
is made zero, this term also approaches zero.

Thermodynamics – Chapter 8 72
The total entropy generation can be written as,
dSnet  dSc.m.  dSsurr   Sgen  0
where the equality holds for reversible processes
and inequality for irreversible processes. This
equation is referred to as the principle of the
increase of entropy.

The increase of entropy can be stated in terms of an


isolated system. Since there is no change in the
surroundings, we can write
dSisolated system   Sgen, system  0

Thermodynamics – Chapter 8 73
Since no actual process is truly reversible,
some entropy is generated during a process, and
therefore the entropy of the universe, which can
be considered to be an isolated system, is
continuously increasing.

The entropy change of an isolated system is the


sum of the entropy change of its components.
Thermodynamics – Chapter 8 74
 The increase of entropy principle does not imply
that the entropy of a system cannot decrease.
The entropy change of a system can be negative
during a process, but entropy generation
cannot. The increase of entropy principle can
be summarized as follows:
 0 Irreversible process

 S gen  0 Reversible process
 0 Impossible process

This relation serves as a criterion in
determining whether a process is reversible,
irreversible, or impossible.
Thermodynamics – Chapter 8 75
Consider a control mass undergoing a process from
initial state 1 to final state 2, with an associated heat
transfer 1Q2, which may be known or calculated
from the first law. The heat transfer is to or from a
reservoir at temperature To. For this process,

Q2
Sc.m.  S2  S1 , Ssurr = 
1

T0
Snet  Sc.m.  Ssurr  0

Thermodynamics – Chapter 8 76
(Example) A frictionless piston–cylinder device
contains a saturated liquid–vapor mixture of
water at 100oC. During a constant-pressure
process, 600 kJ of heat is transferred to the
surrounding air at 25oC. As a result, part of the
water vapor contained in the cylinder condenses.
Determine (a) the entropy change of the water
and (b) the total entropy generation during this
heat transfer process.

Thermodynamics – Chapter 8
• There are no irreversibilities involved within
the system boundaries, and thus the process
is internally reversible.
• The water temperature remains constant at
100oC everywhere, including the
boundaries.

(a) The water undergoes an


internally reversible isothermal
process.
Q 600 kJ
SCM    1.61 kJ/K
Tsystem 373 K

Thermodynamics – Chapter 8
(b) Entropy change for the surrounding which is at
a temperature of 25oC.
Q 600 kJ
SSur    2.013 kJ/K
Tsur (25  273) K

1
S2 gen  SCM  S Surr
  1.61 kJ/K  2.013 kJ/K  0.403 kJ/K

The entropy generation in this case is entirely due


to irreversible heat transfer through a finite
temperature difference.
Thermodynamics – Chapter 8
Graphical representation of entropy generation
during a heat transfer process through a finite
temperature difference.
Thermodynamics – Chapter 8
The total entropy generated during a process can be
determined by applying the entropy balance to
an extended system that includes the system itself
and its immediate surroundings where external
irreversibilities might be occurring.

When evaluating the entropy


transfer between an extended
system and the surroundings,
the boundary temperature of
the extended system is simply
taken to be the environment Q2
m  s2  s1   1
 1 S2 gen
temperature. Tsurr

Thermodynamics – Chapter 8
Some Remarks about Entropy
1. Processes can occur in a certain direction
only, not in any direction. A process must
proceed in the direction that complies with the
increase of entropy principle, that is, δSgen ≥ 0.
A process that violates this principle is
impossible.
2. Entropy is a non-conserved property, and
there is no such thing as the conservation of
entropy principle. Entropy is conserved during
the idealized reversible processes only and
increases during all actual processes.

Thermodynamics – Chapter 8 82
3. The performance of engineering systems is
degraded by the presence of irreversibilities, and
entropy generation is a measure of the
magnitudes of the irreversibilities present during
that process.
4. The greater the extent of irreversibilities, the
greater the entropy generation. Therefore,
entropy generation can be used as a quantitative
measure of irreversibilities associated with a
process. It is also used to establish criteria for
the performance of engineering devices.

Thermodynamics – Chapter 8 83
Entropy Change of a Solid and Liquid
The liquids and solids can be approximated as
incompressible substances since their specific
volumes remain nearly constant during a process.
Thus, dv≈0 for liquids and solids, and property
relation equation for this case reduces to
du
T ds  du  P dv  ds
T
since Cp=Cv=C and du=CdT for incompressible
substances.
2
dT  T2 
Liquids, solids: s2  s1   C (T )  Cavg ln  
1 T  T1 
Thermodynamics – Chapter 8 84
 A relation for isentropic processes of liquids and
solids is obtained by setting the entropy change
relation above equal to zero. It gives
2
 T2 
Isentropic: s2  s1  0    Cavg ln    T2  T1
1  T1 
The temperature of a truly incompressible
substance remains constant during an
isentropic process. Therefore, the isentropic
process of an incompressible substance is
also isothermal. This explains why pumping
liquid does not change the temperature.

Thermodynamics – Chapter 8 85
(Ex 8.4) Consider 1 kg of liquid water is heated
from 20oC to 90oC. Calculate the entropy change
assuming constant heat and compare the result
with that found when using the tables.

 363.2 
s2  s1  4.184 kJ/kg.k  ln  
 293.2 
 0.8958 kJ/kg.K

s2  s1  s f @90o C  s f @ 20o C  1.1925  0.2966


 0.8958 kJ/kg.K

Thermodynamics – Chapter 8 86
Entropy Change of an Ideal Gas
An expression for the entropy change for an ideal
gas can be obtained as follows,
T ds  du  P dv
du P
ds   dv
T T
P R
For an ideal gas  du  Cv 0 dT and 
T v
dT R dv
Therefore, ds  Cv 0 
T v
2
dT v2
s2  s1   Cv 0  R ln ------- (7)
1 T v1
Thermodynamics – Chapter 8 87
Similarly, T ds  dh  v dP
v R
For an ideal gas  dh  C p 0 dT and 
T P
dT R dP
Therefore, ds  C p 0 
T P
2
dT P2
s2  s1   C p 0  R ln -------- (8)
1 T P1

To integrate equations 7 and 8, we must know the


temperature dependence of the specific heats. There
are three possibility to solve this.

Thermodynamics – Chapter 8 88
1. The simplest of which is the assumption of
constant specific heat.
T2 v2
s2  s1  Cv 0 ln  R ln
T1 v1
and
T2 P2
s2  s1  C p 0 ln  R ln
T1 P1

2. The second possibility is to use an analytical


equation for Cp0 as a function of temperature,
one of those listed in Table A.6.

Thermodynamics – Chapter 8 89
3. The third possibility is to integrate the results of
the calculations of statistical thermodynamics
from reference temperature T0 to any other
temperature T and define the standard entropy,
Cp0 T

sT  
0
dT
T T 0

The entropy change between any two states 1


and 2 is then given by,


s2  s1  sT 2  sT 1
0 0
  R ln
P2
P1

Thermodynamics – Chapter 8 90
(Ex 8.5) Oxygen in a piston cylinder assembly in
which is heated from 300K to 1500K. Assume
that during this process the pressure dropped
from 200 to 150kPa. Calculate the change in
entropy per kilogram.

From the ideal gas table A.8

 150 
s2  s1  (8.0649  6.4168)  0.2598ln  
 200 
 1.7228 kJ/kg.K

Thermodynamics – Chapter 8 91
Using the relation having specific heat in terms of
temperature
2
dT P2
s2  s1   C p 0  R ln
1
T P1

Put Cp0  C0  C1  C2 2  C3 3 and read values from A.6
we get after integrating
 2 1.5
 0.54 2 0.33 3 
s2  s1  0.88ln   0.0001     
 2 3 1 0.3
 150 
 0.2598ln    1.7058 kJ/kg.K
 200 
This value is within the 1.0% of previous value
Thermodynamics – Chapter 8 92
If constant specific heat is considered at temperature
300K

 1500   150 
s2  s1  0.922 ln    0.2598ln  
 300   200 
 1.5586 kJ/kg.K

This value is high by 4%

Thermodynamics – Chapter 8 93
The first of the three possibilities, constant specific
heat, is also worth analyzing as a special case.
T2 P2
s2  s1  0  C p 0 ln  R ln 
T1 P1
R

T2 R  P2  T2  P2  Cp 0
ln  ln     
T1 C p 0  P1  T1  P1 
R C p 0  Cv 0   1
However,  
Cp0 Cp0 
where  , the ratio of the specific heats, is defined as
Cp0

Cp0
Thermodynamics – Chapter 8 94
 1

T2  P2  
 
T1  P1 
From this expression and the ideal gas equation of state,
 1  
1
T2  T2 v1  
 T2  1
T2 v1  T2   T2 
 1
v1
           
T1  T1 v2   T1  T1 v2  T1   T1  v2
 1 
T2  v1  P2  v1 
  and  
T1  v2  P1  v2 
From the last expression, we note that for this process
Pv  constant
This is the special case of a polytropic process in
which polytropic exponent n = specific heat ratio γ.
Thermodynamics – Chapter 8 95
Reversible Polytropic Process for an
Ideal Gas
 When a gas undergoes a reversible process in
which there is heat transfer.
 The process take place such that the graph
between log P versus log V is a straight line.
n
 Then the relation between P and V is PV =C
 This type of process is called polytropic process.
 An example is the expansion of the combustion
gases in the cylinder of IC engine.

Thermodynamics – Chapter 8 96
If the pressure and volume are measured during the
expansion stroke of polytropic process, as might be
done with an engine indicator, and the logarithms of
the pressure and volume are plotted, the result
would be similar to the straight line as shown below.

Thermodynamics – Chapter 8 97
From the figure it follows that
d ln P
 n
d ln V

d ln P  nd ln V  0

If n is constant, this equation can be integrated to give the


following relation
PV n  constant  PV1 1
n
 P V
2 2
n

n 1
n 1 n
T2  P2  n  v1  P2  v1 
    and  
T1  P1   v2  P1  v2 

Thermodynamics – Chapter 8 98
For a control mass consisting of an ideal gas, the
work done at the moving boundary during a
reversible polytropic process can be derived from
the relations 2
   constant
n
W
1 2 PdV and PV
1
2 2
dV
1W2   PdV = constant  n
1 1 V
P2V2  PV
W  1 1

1 n
1 2

mR(T2  T1 )

1 n
Thermodynamics – Chapter 8 99
For a control mass consisting of an ideal gas, the
work done at the moving boundary during a
reversible isothermal process can be derived from
the relations
2
W2   PdV
1 and PV  constant
1
2 2
dV
1W2   PdV = constant 
1 1
V
V2 P1
1W2  PV1 1 ln  PV
1 1 ln
V1 P2
V2 P1
 mRT ln  mRT ln
V1 P2

Thermodynamics – Chapter 8 100


The polytropic processes for various values of n are
PV n  constant
Isobaric process n  0, Isothermal process n  1
Isentropic process n   , Isochoric process n 

Thermodynamics – Chapter 8 101


Entropy as a Rate Equation
 To track the process in time.
 Basis for development of entropy balance
equation in the general control volume analysis
for unsteady situation.
 Take the incremental change in S and divide by
δt, we get

dS 1  Q  S gen
 
t T t t

Thermodynamics – Chapter 8 102


If for the given control volume more than one
source of heat transfer, each at a certain surface
temperature, then the final form for the entropy
equation is

dSc.m 1
  Q  S gen
dt T

The rate of entropy change = flux of entropy into


the control mass from heat transfer + an increase
due to irreversible process inside the control mass

Thermodynamics – Chapter 8 103


What is Entropy ?
 It can be viewed as a measure of molecular
disorder, or molecular randomness.
 As a system becomes more disordered, the
positions of the molecules become less predictable
and the entropy increases.
 Therefore the entropy of a substance is lowest in
the solid phase and highest in the gas phase.
 In the solid phase, the molecules of the substance
continually oscillate about their equilibrium, but
they can’t move relative to each other.

Thermodynamics – Chapter 8 104


What is Entropy ?
 In the solid phase, the molecules of the substance
continually oscillate about their equilibrium, but
they can’t move relative to each other, and their
position at any instant can be predicted with good
certainty.
 In the gas phase, however, the molecules move
about at random, collide with each other, and
change direction, making it extremely difficult to
predict accurately the microscopic state of a
system at any time. Associated with this
molecular chaos is a high value of entropy.
Thermodynamics – Chapter 8 105
What is Entropy ?
 When viewed microscopically, an isolated system
that appears to be at a state of equilibrium may
exhibit a high level of activity because of the
continual motion of the molecules.
 To each state of macroscopic equilibrium there
corresponds a large number of possible
microscopic states or molecular configurations.
 The entropy of a system is related to the total
number of possible microstates of that system,
called thermodynamic probability p, by the
Boltzmann relation, expressed as
Thermodynamics – Chapter 8 106
What is Entropy ?

S  k ln p
Where k = 1.3806×10-23 J/K is the Boltzmann
constant. Therefore, from a microscopic point of
view, the entropy of a system increases whenever
the molecular randomness or uncertainty (i.e.,
molecular probability) of a system increases. Thus,
entropy is a measure of molecular disorder, and the
molecular disorder of an isolated system increases
anytime if it undergoes a process.
Thermodynamics – Chapter 8 107
What is Entropy ?
 The molecules of a substance in solid phase
continually oscillate, creating an uncertainty about
their position. These oscillations, however, fade
as the temperature is decreased, and the molecules
supposedly become motionless at absolute zero.
This represents a state of ultimate molecular
order. Therefore, the entropy of a pure crystalline
substance at absolute zero temperature is zero
since there is no uncertainty about the state of the
molecules at that instant. This statement is known
as the third law of thermodynamics.
Thermodynamics – Chapter 8 108
What is Entropy ?
 The third law of thermodynamics provides an
absolute reference point for the determination of
entropy.
 The entropy determined relative to this point is
called absolute entropy.
 Notice that the entropy of a substance that is not
pure crystalline is not zero at absolute zero
temperature. This is because more than one
molecular configuration exists for such
substances, which introduces some uncertainty
about the microscopic state of the substance.
Thermodynamics – Chapter 8 109
What is Entropy ?
 Molecules in the gas phase posses a considerable
amount of K.E. However, no matter how large
their kinetic energies are, the gas molecules don’t
rotate a paddle wheel inserted into the container
and produce work. This is because gas molecules,
and the energy they posses, are disorganized.
Probably the number of molecules trying to rotate
the wheel in one direction at any instant is equal
to the number of molecules that are trying to
rotate it in the opposite direction, causing the
wheel to remain motionless.

Thermodynamics – Chapter 8 110


What is Entropy ?

Therefore, we can’t extract


any useful work directly from
disorganized energy.

The level of molecular Disorganized energy does not


disorder (entropy) of a create much useful effect, no
substance matter how large it is.

Thermodynamics – Chapter 8 111


What is Entropy ?
Now consider a rotating shaft. This
time the energy of the molecules is
completely organized since the
molecules of the shaft are rotating in
the same direction together. This
organized energy can readily be used
to perform useful tasks such as
raising a weight or generating
electricity. Being an organized form
of energy, work is free of disorder
and thus free of entropy.
Thermodynamics – Chapter 8 112
What is Entropy ?
 There is no entropy transfer associated with
energy transfer as work.
 Therefore, in the absence of any friction, the
process of raising a weight by a rotating shaft (or
a flywheel) does not produce any entropy. Any
process that does not produce a net entropy is
reversible, and thus the process just described can
be reversed by lowering the weight.
 Therefore, energy is not degraded during this
process, and no potential to do work is lost.

Thermodynamics – Chapter 8 113


What is Entropy ?
Instead of raising a weight,
operate the paddle wheel in a
container filled with a gas. The
paddle-wheel work is converted
to the internal energy of the gas. This process is quite
different from raising a weight since the organized
paddle-wheel energy is now converted to a highly
disorganized form of energy, which can’t be
converted back to the paddle-wheel as the rotational
K.E. Only a portion of this energy can be converted
to work by the use of a heat engine.
Thermodynamics – Chapter 8 114
Mechanisms of Entropy Transfer
 Entropy can be transferred to or from a system by
two mechanisms: heat transfer and mass flow (in
contrast, energy is transferred by work also).
 Entropy transfer is recognized at the system
boundary as it crosses the boundary, and it
represents the entropy gained or lost by the
system during a process.
 The only form of entropy interaction associated
with a closed system (control mass) is heat
transfer.

Thermodynamics – Chapter 8 115


Mechanisms of Entropy Transfer
Heat is a form of
disorganized energy, and
some disorganization
(entropy) will flow with heat.
Heat transfer to a system
increases the entropy of that
system and thus the level of
molecular disorder or
randomness, and heat transfer
from a system decreases it.

Thermodynamics – Chapter 8 116


Mechanisms of Entropy Transfer
The ratio of the heat transfer Q at a location to the
absolute temperature T at that location is called the
entropy flow or entropy transfer and is expressed as
Q
Sheat  Where T  constant
T
When the temperature T is not constant, the entropy
transfer during a process 1-2 can be determined by
integration as
2
Q
S heat 
1
T

Thermodynamics – Chapter 8 117

You might also like