You are on page 1of 211

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
 dE
c
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  Q
H
 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
 Q QH QL
 T  T  T  0
H L

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,
 Q 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  QH
 QL
 Wrev  QL   QL
 Wrev
Therefore for reversible HE    Q  0

  Q  QH
 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,
Q QH QL
 T   T  T  0
H L

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,
Q Q  Q 
      
T  T  boiler  T condenser

Thermodynamics – Chapter 8 28
The temperature remains constant in both the boiler
and condenser.
Q 1 2 1 4 Q Q4
 T  T 1  Q  T 3  Q  T  T
1 2 3

1 3 1 3

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


q
1 2
 h2
 h1
 2039.3 kJ/kg, T 1
 164.97 0
C
q
3 4
 h 4
 h 3
 463.4  2361.8   1898.4 kJ/kg, T3
 53.97 0
C
Q 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   2  T   (1)
A 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   2  T     (2)
C 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 
S 2  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  S f  Sg
ms  m f s f  mg sg
s  1  x  s f  x s g
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
S 2  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 Q4
S 4  S3       Q 
3

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.

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

Thermodynamics – Chapter 8 50
State 2: 20oC, 400 kPa (Table B3.2)
u2 = 180.57 KJ/kg, s2 = 0.7204 KJ/kg K
 Q 1Q 2
At const T , dS =  
T T
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 51
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 52
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 53
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 
 on unit mass basis
T ds  dh  v dP 

Thermodynamics – Chapter 8 54
The T dS relations are valid for both
reversible and irreversible processes and for
both closed and open systems.
Thermodynamics – Chapter 8 55
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 2Q  1
Q 
 T  1  T   2  T   0
A B
1
Q  2
Q 
      
2  T B 1  T A

Thermodynamics – Chapter 8 56
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 57
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 58
 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 59
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 60
Consider a reversible process, for which the entropy
generation is zero, and the heat transfer and work
terms therefore are
 Q  TdS and  W  P dV
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  Qirre  dU   Wirre
TdS  T  S gen  dU   Wirre

Thermodynamics – Chapter 8 61
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 62
The entropy balance equation for a control mass is
2 2
Q
S 2  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 63
Some Important Conclusions from the Entropy
Generation 2 2
Q
S 2  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 64
 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 65
 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 66
Principle of the Increase of Entropy (7 th
ed)
 We considered irreversible processes in which the
irreversibilities occurred inside the system of
control mas.
 It is found that 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.
 Now, we would emphasize the difference between
the energy and entropy equations and point out that
energyThermodynamics – Chapter
is conserved but entropy is not. 8 74
Principle of the Increase of Entropy (7 th
ed)

Consider two mutually


exclusive control volumes
A and B with a common
surface and their
surroundings C such that
they collectively include
the whole world.
Let some processes take
place so that these control
volumes exchange work
and heat transfer.

Thermodynamics – Chapter 8 75
Principle of the Increase of Entropy (7 th
ed)

 Energy equations:

 Entropy balance equations:

Thermodynamics – Chapter 8 76
Principle of the Increase of Entropy (7 th
ed)

 Energy change of the total world:

 Entropy balance equations:

Thermodynamics – Chapter 8 77
Principle of the Increase of Entropy (7 th
ed)

 Total energy is conserved. The energy is not


stored in the same form or place as it was
before the process, but the total amount is the
same.
 Total entropy increases due to positive
entropy generation terms and is thus not
conserved.
 This concept is referred as the Principle of
the increase of entropy.
Thermodynamics – Chapter 8 78
Principle of the Increase of Entropy (7 th
ed)

 We get the total entropy generation


(increase), but we didn’t talk about where in
the world the entropy was generated.
 Consider a heat transfer
process in which
energy flows from a
higher temperature
domain to a lower
temperature domain

Thermodynamics – Chapter 8 79
Principle of the Increase of Entropy (7 th
ed)

 The transfer goes


through walls, control
volume B, that separates
domains A and C.

Thermodynamics – Chapter 8 80
Principle of the Increase of Entropy (7 th
ed)

 From the point of view of control volume B,


the walls, there is no change of state in time,
but the state is nonuniform in space (it has T0
on the outer side and T on the inner side)
 Energy equation:

 Entropy equation:

Thermodynamics – Chapter 8 81
Principle of the Increase of Entropy (7 th
ed)

 Since T0 > T for the heat transfer to move in the


indicated direction, we see that the entropy
generation is positive.
 Suppose T0 < T, then to have positive entropy
generation, must be negative, that is, move in
the opposite direction.
 The direction of heat transfer from a higher to a
lower temperature is thus a logical consequence of
the second law.
Thermodynamics – Chapter 8 82
 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 83
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

The total entropy generation can be written as,


SCM  S Surr  1 S 2 gen  0

Thermodynamics – Chapter 8 84
(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
S Sur    2.013 kJ/K
Tsur (25  273) K

1
S2 gen  SCM  SSurr
  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
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 90
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 91
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 92
 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 93
(Ex 8.4) Consider 1 kg of liquid water is heated
from 20oC to 90oC. Calculate the entropy change
assuming constant specific 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 94
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 95
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 96
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 97
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,
0 Cp0 T

sT   dT
T T 0

The entropy change between any two states 1


and 2 is then given by,

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

Thermodynamics – Chapter 8 98
(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

 0 0
s2  s1  sT 2  sT 1   R ln
P2
P1
 150 
s2  s1  (8.0649  6.4168)  0.2598ln  
 200 
 1.7228 kJ/kg.K

Thermodynamics – Chapter 8 99
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 8 100


If constant specific heat is considered at temperature 300K

T2 P2
s2  s1  C p 0 ln  R ln
T1 P1
 1500   150 
s2  s1  0.922 ln    0.2598ln  
 300   200 
 1.5586 kJ/kg.K
This value is high by 4%

Thermodynamics – Chapter 8 101


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
C p0

C p0
Thermodynamics – Chapter 8 102
 1

T2  P2  
 
T1  P1 
From this expression and the ideal gas equation of state,
 1  
1
T2  T2 v1  
 T2  T2 v1
 1
 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 8 103
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.
 Then the relation between P and V is PVn=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 8 104


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 8 105


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 8 106


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


n
W
1 2  PdV and PV  constant
1
2 2
dV
1W2   PdV = constant  n
1 1 V
P2V2  PV
1 1
W
1 2 
1 n
mR(T2  T1 )

1 n

Thermodynamics – Chapter 8 107


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 8 108


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 8 109


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 8 110


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 8 111


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 8 112


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 8 113
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 8 114
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 8 115


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 8 116


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 8 117
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 8 118


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 8 119


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 8 120
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 8 121


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 8 122
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 8 123


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 8 124


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 8 125


(8.24/8.17) A heat engine receives 6 kW from a 2500C
source and rejects heat at 300C. Examine each of the
three cases with respect to inequality of Clausius.
. .
a. W  6 kW, b. W  0 kW, c. Carnot Cycle
.
a) W  6 kW, Q H  6 KW
 Q  0 L
.
δ Q Q H Q L
 T  TH  TL
6000 W 0
   11.47 W/K > 0
523 303
Impossible

Thermodynamics – Chapter 8 126


.
  6 KW
b) W  0 kW, QH
.
 QL  6 kW
.
δ Q 6000 6000
 T  523  303  8.33 W/K < 0
Possible Process
. .
δ Q 6000 QL
c)     0 Carnot cycle
T 523 303
. 303
QL   6000  3.476 kW
543
. . .
W  QH  QL  2.529 kW

Thermodynamics – Chapter 8 127


(8.31/8.32) Consider a Carnot-cycle heat engine with
water as the working fluid. The heat transfer to the
water occurs at 3000C, during which process water
changes from saturated liquid to saturated vapor. The
heat us rejected from the water at 40oC. Show the
process on a T-s diagram and find the quality of water at
the beginning and at the end of the heat rejection
process. Determine the net work output per kg water
and cycle thermal efficiency. qH
Solution: T 1 2

o
w
State 1: sat. liquid, T1  TH  300 C 4 3

s1  s f @ 300o C  s4  s f @ 40o C  x4  s fg @ 40o C qL s


3.2533 kJ/kgK  0.5724  x4  7.6845  x4  0.3489

Thermodynamics – Chapter 8 128


State 2: Saturated vapor, T2  TH  300o C
s2  sg @ 300o C  s3  s f @ 40o C  x3  s fg @ 40o C
5.7044 kJ/kgK  0.5724  x3  7.6845
 x3  0.6678
qH  TH (s2  s1 )
 573.15 K  (5.7044  3.2533) kJ/kgK
 1405 kJ/kg
wnet TH  TL 260
ηTH     0.4536
qH TH 573.15
wnet  ηTH  qH  637.3 kJ/kg
129
(8.33/8.35) Water is used as the working fluid in a
Carnot cycle heat engine, when it changes from
saturated liquid to saturated vapor at 2000C as heat
is added. Heat is rejected in a constant pressure
process (also constant T) at 20 kPa. The heat engine
powers a Carnot cycle refrigerator that operates
between -150C and +200C as shown in Fig. Find the
heat added to the water per kg water. How much
heat should be added to the water in the heat engine
so that refrigerator can remove 1 kJ heat from the
cold space.

Thermodynamics – Chapter 8 130


TH2
TH1=200 C
o

Carnot cycle referigerator


QL TL 2
COPR,rev  
W TH 2  TL 2
15  273

20  (15)
 7.37
QL
W 
COPR,rev
1 kJ
TL1= 60.06oC TL2   0.136 kJ
7.37

131
State 1 & 2: (2000 C) and sat liquid to sat vapour
qH  TH1 (s2  s1 )  TH1 s fg  h fg
 473.15  4.1014  1940 kJ/kg

State 3 & 4: two-phase, P  20 kPa, T  constant


Table B.2.1 
TL1  T3  T4  Tsat @ 20 kPa  60.060 C
TL1 333 W
ηHE  1   1  0.296 
TH1 473 QH
W 0.136
QH to H 2 O    0.46 kJ
ηHE 0.296
132
(8.39/Q) One kilogram of ammonia in a piston cylinder
at 500C and 1000 kPa ia expanded in a reversible
isothermal process to 100 kPa. Find the work and heat
transfer for this process.
C.V.: NH 3
Energy Eqn.: m(u2  u1 )  1 Q2  1W2
2
δQ 1 Q2
Entropy Eqn.: m(s2  s1 )    ( reversible)
1
T T

1W2   PdV , 1 Q2   TdS  mT (s2  s1 )

State 1: m = 1 kg , T1  500 C, P2  1000 kPa  Psat@500 C


Table B.2.2  u1  1391.3 kJ/kg, s1  5.265 kJ/kgK

Thermodynamics – Chapter 8 133


State 2: T2  500 C, P2  100 kPa
Table B.2.2  u2  1424.7 kJ/kg, s2  6.494 kJ/kgK

1 Q2  mT (s2  s1 )
 1 kg  (50+273)K  (6.494  5.265) kJ/kgK
= 396.97 kJ
W2  1 Q2  m(u2  u1 )  363.75 kJ
1

P 1
2
T 1 2

v s

Thermodynamics – Chapter 8 134


(8.45/8.47) One kilogram of water at 3000C expands
against a piston in a cylinder until it reaches ambient
pressure, 100 kPa, at which point the water has a
quality of 90.2 %. It may be assumed that the
expansion is reversible and adiabatic. What was the
initial pressure in the cylinder and how much is
done by water

C.V.: Water, reversible and Q  0


Energy Eqn.: m( u2  u1 )  Q  1W2   1W2 1
T
δQ
Entropy Eqn.: m(s2  s1 )  
T 2

Process: Reversible, Adiabatic s2  s1


s
Thermodynamics – Chapter 8 135
State 2: m  1 kg, P2  100 kPa, x2  0.902
Table B.1.2 
s2  1.3026  0.902  6.0568  6.7658 kJ/kgK
u2  417.36  0.902  2088.7  2301.4 kJ/kg

State 1: s2  s1  sg @ 3000 C
T1  3000 C, s1  s2  6.7658 kJ/kgK > s g @ 3000 C
From B.1.3  P1  2000 kPa, u1  2772.56 kJ/kg
W  m(u1  u2 )  1 kg  (2772.56  2301.4)  471.2 kJ
1 2

136
(8.51/8.51) A heavily-insulated cylinder fitted with
a frictionless piston, contains ammonia at 5°C,
92.9% quality, at which point the volume is 200 L.
The external force on the piston is now increased
slowly, compressing the ammonia until its
temperature reaches 50°C. How much work is done
by the ammonia during this process?

C.V.: Water, reversible and Q  0


Energy Eqn.: m( u2  u1 )  Q  1W2   1W2
δQ
Entropy Eqn.: m(s2  s1 )  
T
Process: Reversible, Adiabatic s2  s1

Thermodynamics – Chapter 8 139


State 1: m  1 kg, T1  50 C, x  0.929, V1  200 L  0.2 m 3
From table B 2.1 
v1  0.001583  0.929  0.2414  0.2258 m 3 / kg
u1  202.77  0.929 1119.2  1242.5 kJ/kg
s1  0.7951  0.929  4.4715  4.9491 kJ/kg
V1 0.2
m   0.886 kg
v1 0.2258
State 2 : T2  500 C, s2  s1  4.9491 kJ/kg.K  sg @500 C
 it is superheated  Table B.2.2 for s2  4.95 kJ/kg.K
 P2  1600 kPa, u2  1364.9 kJ/kg
1W2  m u1  u2   0.886 kg 1242.5  1364.9  kJ/kg
 108.4 kJ
Thermodynamics – Chapter 8 140
(8.54/8.48) Water at 1000 kPa, 250°C is brought to
saturated vapor in a rigid container. Find the final T
and the specific heat transfer in this isometric
process.

Continuity equation : m2  m1  m
Energy equation: m u2  u1   1 Q2  1W2
Q
Entropy equation: m  s2  s1   
T
Process : V  constant

Thermodynamics – Chapter 8 141


State 1: T1  2500 C, P1  1000 kPa
From table B 1.3  it is superheated
v1  0.23268 m3 / kg, u1  2709.91 kJ/kg, s1  6.9246 kJ/kg.K

State 2 : v2  v1  0.23268 m3 , saturated vapor


Table B 1.1  1700 C  T2  1750 C
T2  171.950 C 
 by interpolation
u2  2577.9 kJ/kg, s2  6.651kJ/kg.K 
Energy equation  u2  u1  1 q2  1 w2
Process: v2  v1  1 w2  0
q  u2  u1  2577.9  2709.91  132 kJ/kg
1 2

Engineering Thermodynamics 142


(8.59/8.54) A rigid, insulated vessel contains
superheated vapor steam at 3 MPa, 400°C. A valve
on the vessel is opened, allowing steam to escape.
The overall process is irreversible, but the steam
remaining inside the vessel goes through a
reversible adiabatic expansion. Determine the
fraction of steam that has escaped, when the final
state inside is saturated vapor.

Engineering Thermodynamics 143


State 1 : P1  3 MPa, T1  400o C
Table B 1.3  it is superheated
v1  0.09936 m 3 / kg
u1  2932.75 kJ/kg
s1  6.9211 kJ/kg.K
State 2 : saturated vapor
Process: reversible adiabatic expansion  s2  s1  6.9211 kJ/kg.K
Table B.1.1  140o C  T2  145o C
Table B 1.1 (page 675)  T2  1410 C,v2  0.4972 m 3 /kg
m1  m2 m2 v1 0.09936
1 1 1  0.8
m1 m1 v2 0.4972

Engineering Thermodynamics 144


(8.61/8.101) One kg water at 500oC and 1 kg
saturated water vapor both at 200 kPa are mixed in
a constant pressure and adiabatic process. Find the
final temperature and the entropy generation for the
process.
Continuity equation  m2  mA  mB
Energy equation  m2u2  mAu A  mB u B  1 Q2  1W2
 0  1W2
Q
Entropy equation  m2 s2  mA s A  mB sB    1 s2 gen
T
 0  1 s2 gen

Engineering Thermodynamics 145


Process : P  c
W2   pdV  p (V2  V1 )
1

where V1  mA v A  mB vB and V2  m2 v2

Energy equation  m2 u2  mA u A  mB uB  1 Q2  1W2


 0  1W2
m2 u2  mA u A  mB u B   p ( m2 v2  mA v A  mB vB )
m2 u2  m2 p2 v2  (mA u A  mA p A v A )  (mB u B  mB pB vB )
m2 h2  mA hA  mB hB
mA hA  mB hB
 h2 
m2
146
State A1:
mA =1kg PA =200kPa TA =500O C
From table B 1.3  it is superheated
hA =3487.03 kJ/kg, s A = 8.5132 kJ/kg K

State B1:
m B  1 kg, PA  200 kPa, TA  120.23O C
and saturated vapor
From the table B 1.2 
hB =2706.63 kJ/kg, sB =7.1271 kJ/kg K

147
mA hA mB hB
h2  
m2 m2
1 1
  3487.03   2706.63  3096.83 kJ/kg
2 2

State 2: P2  200 kPa, h2  3096.83 kJ/kg


From the tables B 1.2 (679)  it is a superheated vapour
T2 =312.2o C and s2  7.9328 kJ/kg K

s
1 2 gen  m2 s2  mA s A  mB sB
 2  7.9328  1  8.5132  1  7.1271  0.2253 kJ/K

148
(8.65/8.106) An insulated cylinder/piston contains
R-134a at 1 MPa, 50oC, with a volume of 100 L.
The R-134a expands moving the piston until the
pressure in the cylinder has dropped to 100 kPa. It is
claimed that R-134a does 190 kJ of work against the
piston during the process. Is that possible?

C.V: R-134a in cylinder, insulated  Q = 0


Energy equation  m2 ( u2  u1 )  1Q2  1W2   1W2
δQ
Entropy equation  m2 ( s2  s1 )    1 s2 gen
T
 0  1 s2 gen
State 1: P1  1000 kPa ,T1  50o C , V1  0.1m 3
 Superheated Table B.5.2
v1  0.02185 m 3 /kg, u1  409.39 kJ/kg,
s1  1.7494 kJ/kgK,
V 0.1
m=   4.577 kg
v1 0.02185

Energy Equation: m(u2 -u1 ) = 1 Q2 - 1W2 = 0-190 kJ


1W2 (-190 kJ)
u2 = u1 -  409.39 kJ/kg 
m 4.577 kg
 367.89 kJ/kg
State 2 : P2  100 kPa , u2  367.89 kJ/kg
u2  ug @100 kPa  superheated
0
Table B.5.2: T2  19.25 C , s2  1.7689 kJ/kgK

δQ
m(s2  s1 ) =   1 S2 ,gen  0  1 S2 ,gen
T
1 S 2 ,gen  m(s2  s1 ) = 0.0893 kJ/K

=> The process is possible since 1 S 2 ,gen  0


(8.71/Q) A piston/cylinder has ammonia at 2000
kPa, 80oC with a volume of 0.1 m3. The piston is
loaded with a linear spring and outside ambient is at
20oC. The ammonia now cools down to 20oC at
which point it has a quality of 10%. Find the work,
the heat transfer and the total entropy generation in
the process.

Engineering Thermodynamics 155


Energy equation: m(u2  u1 )  1 Q2  1W2
Q2
1
Entropy equation: m( s2  s1 )   1 s2 gen
Tamb
1
Proces: P  A  bV  1W2 = m( P1  P2 )(v2  v1 )
2
State 1:
P1  2000 kPa, T1  80o C, V1  0.1 m 3
From table B 2.2  it is superheated
v1  0.07595 m3 /kg, u1 =1421.6 kJ/kg, s1 =5.0707 kJ/kg.K
V1 0.1
m   1.3167 kg
v1 0.07595

156
State 2:
T2  20o C , x2  0.1
From the tables B 2.1  it is a saturated mixture
P2  857.5 kPa
v2  0.001638  0.1  0.14758  0.016396 m 3 /kg
u2  272.89  0.1  1059.3  378.82 kJ/kg
s2  1.0408  0.1  4.0452  1.44532 kJ/kg K

1
1W2   1.3167(2000  857.5)(0.016396  0.07595)
2
 112 kJ

157
1 Q2  m(u2  u1 )  1W2
 1.3167  (378.82  1421.6)  (112.04)
 1485 kJ

1 Q2
s
1 2 gen  m( s2  s1 ) 
Tambient
 1485 
 1.3167  (1.44532  5.0707)   
 293 
 0.295 kJ/K  Total entropy generation

158
(8.72/Q) A cylinder/piston assembly contains water
at 200 kPa and 200oC with a volume of 20 L. The
piston is moved slowly, compressing the water to a
pressure of 800 kPa. The loading on the piston is
such that the product PV is a constant. Assuming
that room temperature is 20oC, show that this
process does not violate second law.

C.V.: Water + Cylinder out to room at 20 o C


Energy Eq. : m(u2  u1 ) = 1 Q2  1W2
1 Q2
Entropy Eq. : m(s2  s1 ) =  1 S 2 ,gen
Troom
Process: PV = Constant = Pmv  v2  Pv
1 1 / P2

 v2 
1 w2   Pdv  P1v1 ln  
 v1 
State 1: P1  200 kPa, T1  200o C, V1  0.02 m 3
Table B.1.3  superheated
3
v1 = 1.0803 m /kg, u1  2654.4 kJ/kg, s1  7.5066 kJ/kgK
Pv 200  1.0803
State 2: P2  800 kPa, v2 = 1 1
  0.2701 m 3 /kg
P2 800
Table B.1.3  superheated  200o C  T2  250o C
By interpolation  u2  2655.0 kJ/kg, s2 = 6.8822 kJ/kgK
3  0.2701 
1 w2  200 kPa  1.0803 m /kg  ln    299.5 kJ/kg
 1.0803 
1 q2  u2  u1  1 w2  2655.0  2654.4  299.5

 298.9 kJ/kg
q
1 2
1 S 2,gen  s2  s1 
Troom
298.9
 6.8822  7.5066 
293.15
 0.395 kJ/kgK > 0
=> Satisfy Second Law
(8.77/8.60) A 4 L jug of milk at 25°C is placed in
your refrigerator where it is cooled down to the
refrigerators inside constant temperature of 5°C.
Assume the milk has the property of liquid water
and find the entropy generated in the cooling
process.
C.V. Jug of milk
Continuity equation : m2  m1  m
Energy equation: m u2  u1   1Q2  1W2
Entropy equation: dS net  dSc.m.  dSsurr    Sgen
1 Q2
m  s2  s1    1 S 2 gen
Tsurr
Process : P  constant  atmospheric pressure
Engineering Thermodynamics 169
State 1: T1  25o C, V  0.004 m 3
Table B 1.1  v1  v f @ 25o C  0.001003 m3 /kg,
h1  h f @ 25o C  104.87 kJ/kg,
s1  s f @ 25o C  0.3673 kJ/kg.K
V 0.004
m   3.988 kg
v1 0.001003
o
State 2 : T2  5 C
Table B 1.1  v2  v f @ 5o C  0.001 m3 /kg,
h2  h f @ 5o C  20.98 kJ/kg,
s2  s f @ 5o C  0.0761 kJ/kg.K

Engineering Thermodynamics 170


Process : P  constant  1W2  mP  v2  v1 
Energy equation: m u2  u1   1Q2  1W2
 1 Q2  m  h2  h1 
 3.988   20.98  104.87 
 334.55 kJ
The total entropy generation
1 Q2
1 S 2 gen  m  s2  s1  
Trefrig
 334.55 kJ 
 3.988 kg   0.0761  0.3673  kJ/kg   
 5  273 K 
 0.0421 kJ/K

Engineering Thermodynamics 171


(8.78/8.61) A foundry form box with 25 kg of 200oC
hot sand is dumped into a bucket with 50 L of water
at 15oC. Assuming no heat transfer with the
surroundings and no boiling away of liquid water,
calculate net entropy change for the process.
C.V.: Sand and water
Energy equation: m u2  u1   1 Q2  1W2
 0  1W2
Q
Entropy equation: m  s2  s1     1 S 2 gen
T
 0  1 S 2 gen
Process : P  constant  atmospheric pressure
msand (u2  u1 )sand  mH2O (u2  u1 )H 2O   P(V2  V1 )
msand Δhsand  mH2O ΔhH 2O  0
msand (C  ΔT )sand  mH2O (C  ΔT )H 2O  0
50  103
25  0.8  (T2  200) +  4.184  (T2  15) = 0
0.001001
 T2  31.20 C

 304.3   304.3 
ΔS = 25  0.8  ln    49.95  4.184  ln  
 473 .15   288.15 
 2.57 kJ/K  Entropy generated due to mixing
(8.81/8.67) Find the total work and the heat engine
can give out as it receives energy from the rock bed
as described in problem (7.61/7.63) : Write entropy
balance equation for the control volume that is the
combination of the rock bed and the heat engine.

2 m3 , 400 K  290 K To  290 K

C.V: Heat engine + Rock bed out to T0 , W and QL goes out


Energy Eq. : (U 2  U1 )rock = - QL  W
m  ρV  2750 kg/m3  2 m3  5500 kg
QL  T2 
Entropy Eq. : (S2  S1 )rock    mC ln  
T0  T1 
 290 
 5500  0.89  ln    1574.15 kJ/K
 400 
QL  T0 (S 2  S1 ) rock  290 K  (-1574.15) kJ/kg = 456504 kJ

The energy drop of the rock -(U 2  U1 ) rock equals QH of Heat engine
(U 2  U1 ) rock  mC (T2  T1 )
=5500  0.89  (290  400) = -538450 kJ
W  (U 2  U1 ) rock  QL  (  538450) - (456504)  81946 kJ
(8.83/8.65) A 12 kg steel container has 0.2 kg
superheated water vapor at 1000 kPa, both at 200oC.
The total mass is now cooled to ambient
temperature 30oC. How much heat transfer was
taken out and what is the total entropy generation?

C.V. Steel container and water


Energy equation: m u2  u1   1 Q2  1W2
Entropy equation: dS net  dSc.m.  dSsurr    Sgen
1 Q2
m  s2  s1    1 S 2 gen
Tsurr
Process : V  constant  1W2  0

Engineering Thermodynamics 176


State 1: m  0.2 kg, P1  1000 kPa, T1  25o C
Table B 1.3  it is superheated
v1  0.20596 m 3 /kg
u1  2621.9 kJ/kg
s1  6.6939 kJ/kg.K
State 2 : T2  30o C, saturated condition, constant volume process
v2  v1  0.20596 m 3 /kg
Table B 1.1  v f  v2  vg
v2  v f 2 0.20596  0.001004
x2    0.006231
v fg 2 32.8922
u2  125.77  0.006231  2290.81  140.04 kJ/kg
s2  0.4369  0.006231  8.0164  0.48685 kJ/kg.K

Engineering Thermodynamics 177


1 Q2  m u2  u1 
 msteel Csteel T2  T1   mH 2O u2  u1 H O
2

 12  0.42 30  200   0.2 140.04  2621.9 


 1353.2 kJ
The entropy generation
1 Q2
1 S 2 gen  m( s2  s1 ) 
Tsurr
 T2  1 Q2
 msteel Csteel ln    mH 2 O  s2  s1 H O 
 T1  Tsurr
2

 403   1353.2 
 12  0.42 ln    0.2  0.48685  6.6939    
 473   303 
 0.98 kJ/K

Engineering Thermodynamics 178


(8.88/8.88) A piston/cylinder setup contains air at
100 kPa, 400 K which is compressed to a final
pressure of 1000 kPa. Consider two different
processes (a) a reversible adiabatic process and (b) a
reversible isothermal process. Show both processes
in P-v and a T-s diagram. Find the final temperature
and the specific work for both processes.
C.V. Air
Energy equation: u2  u1   1 q2  1 w2
q
Entropy equation:  s2  s1     1 s2 gen
T
Process : reversible  1 s2 gen  0

Engineering Thermodynamics 184


P1  100 kPa, T1  400 K, P2  1000 kPa
(a) Reversible adiabatic process (Isentropic)
q  0  s1  s2
1 2
 1  1

T2  P2  
 P2  
   T2  T1   
T1  P1   P1 
0.4

 1000  1.4
 400     772.3 K
 100 
1
w2  u1@ 400 K  u2@772.3 K
 286.49  570.21 (from table A 7.1)
 283.72 kJ/kg

Engineering Thermodynamics 185


P1  100 kPa, T1  400 K, P2  1000 kPa
(b) Reversible Isothermal process  T1  T2
For this process, taking air as ideal gas 
 T
Cp0 
u2  u1 and sT 2  sT 1  sT  
0 0 0
dT 
 0 T T 
2 dT P2 
w  1 q2  T  s2  s1   T   C p 0
1 2
 R ln 
1 T P1 
 0 P2 
 T   sT 2  sT 1   R ln 
0

 P1 
P2
  R T ln  0.287  400  ln 10 
P1
 264.34 kJ/kg
Engineering Thermodynamics 186
Engineering Thermodynamics 187
(8.90/8.75) Consider a small air pistol with a cylinder
volume of 1 cm3 at 250 kPa and 270C. The bullet acts as
a piston initially held by a trigger, shown in Fig. The
bullet is released so that air expands in an adiabatic
process. If the pressure should be 100 kPa as the bullet
leaves the cylinder, find the final volume and work done
by the air. Hint: Assume reversible process

C.V. Air
Energy equation : m u2  u1   0  1W2
Q
Entropy equation: m  s2  s1     1 s2, gen  0
T
Process : Adiabatic reversible => 1 q2  0, 1 s2, gen  0 and s2  s1 Isentropic

Thermodynamics – Chapter 8 188


k 1 0.4
 P2  k
 100  1.4
T2  T1    300    300  0.40.28575
 P1   250 
 230.9 K

The ideal gas law: PV = mRT at both states leads to


V1 PT 1  250  230.9
V2  1 2
  1.92 cm3
P2T1 100  300

Polytropic work equation with polytropic exponent n  k:


6
P2V2  PV
1 1 (100  1.92  250  1)  10
W
1 2    0.145 J
1 k 1  1.4

Thermodynamics – Chapter 8 189


(8.95/8.83) A piston/cylinder contains air at 1380 K,
15 MPa, with V1 =10 cm3, Acyl = 5 cm2. The piston is
released, and just before the piston exits the end of the
cylinder the pressure inside is 200 kPa. If the cylinder is
insulated, what is its length? How much work is done by
the air inside? Hint: Assume reversible process
C.V. Air and cylinder is insulated
Energy equation: m u2  u1   1 Q2  1W2  0  1W2
q
Entropy equation:  s2  s1     1 s2 gen  0  1 s2 gen
T
P1  15000 kPa, T1  1380 K and P2  200 kPa, T2  ?
For getting T2 , assume this process is reversible.
1 s2 gen  0  s2  s1  0

Engineering Thermodynamics 190


State 1 : P1  15000 kPa, T1  1380 K
From table A.7 (by interpolation) 
u1  1095.2 kJ/kg, sT01  8.5115 kJ/kg.K
P1 V1 1500 10 106
m   0.00379 kg
RT1 0.287 1380
State 2 : P2  200 kPa
1 s2 gen  0  s2  s1  0
P2
s2  s1  0   sT 2  sT 1   R ln
0 0

P1
0 0 P2
sT 2  sT 1  R ln
P1
 200 
 8.5115  0.287 ln    7.2724 kJ/kg.K
 15000 
Engineering Thermodynamics 191
From table A.7, by interpolating 
T2  447.2 K and u2  320.92 kJ/kg

PV PV T2 P1
1 1
 2 2
 V2  V1  
T1 T2 T1 P2
100  447.2  15000
  243 cm 3
1380  200
V2 243
L2    48.6 cm
Acyl 5

1
W2  m u1  u2 
 0.00379 1095.2  320.92 
 0.2935 kJ

Engineering Thermodynamics 192


(8.96/8.87) Two rigid tanks shown in Figure each
contain 10 kg of N2 gas at 1000 K and 500 kPa. They
are now thermally connected to a reversible heat pump,
which heats one and cools the other with no heat
transfer to the surroundings. When one tank is heated to
1500 K the process stops. Find the final (P, T) in both
tanks and the work inputs to the heat pump, assuming
constant heat capacities.

1000 K  1500 K

1 3 1 2

Thermodynamics – Chapter 8 193


C.V.1: Hot tank B
Process : V  constant  1W2  0
Energy equation : m u2  u1   1 Q2  1W2 = 1 Q2  0
m u2  u1   mCv (T2  T1 )  10 kg  0.7448 kJ/kg.K  (1500  1000) K
1 Q2  3734 kJ
PT 500 kPa  1500 K
P2  1 2   750 kPa
T1 1000 K
C.V.2: Total (A+B)
For this C.V only WHP cross the control surface and no heat transfer
Entropy equation :  S2  S1 tot  0  mhot  s2  s1   mcold  s3  s1 
 T2   P2   T3   P3 
C p ,hot ln    R ln    C p ,cold ln    R ln    0
 T1   P1   T1   P1 
194
P3 T3 P2 T2
using  ;  and C p  Cv  R,
P1 T1 P1 T1

 T2   T2   T3   T3 
C p ,hot ln    R ln    C p ,cold ln    R ln    0
 T1   T1   T1   T1 

 T2   T3 
Cv ,hot ln    Cv ,cold ln    0, and C v is same
 T1   T1 
T1 1000 K
T3  T1   1000 K   667 K
T2 1500 K
P1  T3 500 kPa  667 K
P3    333.5 kPa
T1 1000 K
Qcold  1 Q3  mCv (T3  T1 ) = -2480 J
WHP  QHot  Qcold  1 Q2  1 Q3  1244 kJ

195
(8.99/8.128) A rigid tank contains 2 kg of air at 200
kPa and ambient temperature, 200C. An electric
current now passes through a resistor inside the
tank. After a total of 100 kJ of electrical work has
crossed the boundary, the air temperature inside is
800C. Is this possible?

C.V. Air in Tank


Process : Constant Volume and mass, v2  v1
Energy Eqn. : m(u2  u1 )  1 Q2  1W2 , 1W2   Welec
Q 1 Q2
Energy Eqn. : m(s2  s1 )    1 S2, gen   1 S2, gen
T Tamb

Thermodynamics – Chapter 8 196


State 1: T1  200 C, P1  200 kPa, m1  2 kg
State 2: T2  800 C, v2  v1
Ideal Gas Table A.5: R = 0.287 kJ/kgK, Cv  0.717 kJ/kgK
Assuming constant specific heat, energy equation becomes,
1 Q2  mCv (T2  T1 )  1W2  2  0.717(80  20)  100  14 kJ

s for the control mass:


 T2   v2  v2  v2 
s2  s1  Cv ln    R ln  ;  1 v2  v1  R ln    0
 T1   v1  v1  v1 
s2  s1  0.1336 kJ/kgK
1 Q2 14
1 S 2, gen  m( s2  s1 )   2  0.1336   0.315 kJ/K > 0
Tamb 293
Process is possible

Thermodynamics – Chapter 8 197


(8.100/8.84) Argon in a light bulb is at 90 kPa and
heated from 20oC to 60oC with electrical power. Do
not consider any radiation, nor the glass mass. Find
the total entropy generation per unit mass of argon.

C.V. Argon gas and neglect heat transfer Q  0 


Energy equation: m u2  u1   1 Q2  1W2 (electrical)
m u2  u1   0  (  1W2 )
q
Entropy equation:  s2  s1     1 s2 gen  0  1 s2 gen
T
P2 T2
Process: V  contant and ideal gas  
P1 T1

Engineering Thermodynamics 198


T2 P2
s
1 2 gen
 s2  s1  C p 0 ln  R ln
T1 P1
T2 T2
 C p 0 ln  R ln
T1 T1
T2
 C p 0  R  ln
T1
T2
 Cv 0 ln
T1
 60+273 
 0.312 ln    0.04 kJ/kg.K
 20+273 

Engineering Thermodynamics 199


(8.103/8.127) Nitrogen at 2000C and 300 kPa is in a
piston/cylinder device of volume 5 L, with the piston
locked with a pin. The forces on the piston require a
pressure inside of 200 kPa to balance it without the
pin. The pin is removed and the piston quickly
comes to its equilibrium position without any heat
transfer. Find the final P, T and V and the entropy
generation due to this partly unrestrained expansion
C.V. Nitrogen gas
Energy equation : m u2  u1   1Q2  1W2    Peq dV   P2 (V2  V1 )
Entropy equation : m  s2  s1   0  1 S 2, gen
Process : 1 Q2  0, P = Peq after pin is removed

Thermodynamics – Chapter 8 200


State 1: T1 = 2000 C, P1 = 300 kPa
PV 300  0.005
m 1 1   0.01068 kg
RT1 0.2968  473.15

State 2: P2  Peq  200 kPa


Energy eqn.: mu2  P2V2  mu1  P2V1  mh2
P2V1 P2V1 RT1  P2 
h2  u1   u1   u1    RT1
m PV
1 1  P1 
Solving using constants C p and Cv :
 P2 
C pT2  CvT1    RT1
 P1 
  P2    
Cv    R   200 
 P  0.745    0.2368
  1   
  300    428.13 K
T2  T1  473.15
Cp 1.042
201
 T2   P1   428.13   300 
V2  V1       0.005    
 T1   P2   473.15   200 
 0.00679 m3

1 S2, gen  m(s2  s1 )


 T2   P2 
 mC p ln    R ln  
 T1   P1 
PV   T2   P2 
1 1 
 C p ln    R ln   
RT1   T1   P1  
  428.13   200  
 0.01068 1.042ln    0.2968ln  
  473.15   300  
 0.000173 kJ/K
202
Alternate method-1
State 1: T1 = 2000 C, P1 = 300 kPa  superheated Table B.6.2
u1  351.4 kJ/kg 

v1  0.4686 m3 /kg  by interpolation
s1  6.995 kJ/kg.K 
V1 0.005 m3
m   0.01067 kg
v1 0.4686 m /kg
3

State 2: P2  Peq  200 kPa


Energy eqn.: mu2  P2V2  mu1  P2V1  mh2
P2V1
h2  u1 
m
200 kpa  0.005 m3
 351.4   445.12 kJ/kg
0.01067 kg

203
State 2: P2 = 200 kPa, h2 = 445.12 kJ/kg
Since h2  hg @ 200 kPa  81.05 kJ/kg (Table B.6.2)  it is superheated
T2  155.3o C  428.45 K 
 by interpolation
s2  7.012 kJ/kg.K 

1 S2, gen  m( s2  s1 )
 0.01067 kg  (7.012  6.995) kJ/kg.K
 0.00018 kJ/K

204
Alternate method-2
State 1: T1 = 2000 C, P1 = 300 kPa
PV
1 1 300 kPa  0.005 m3
m   0.01068 kg
RT1 0.2968 kJ/kg.K  473.15 K

State 2: P2  Peq  200 kPa


Energy eqn.: mu2  P2V2  mu1  P2V1  mh2
P2V1
h2  u1 
m
P2V1 RT1
 u1 
PV
1 1
 P2  200 kpa  0.005 m 3
 u1    RT1  343.26
    
kJ/kg   445.12 kJ/kg
 P1  by interpolation
0.01067 kg
Table A.8 (T1 =473 K)

205
From Table A.8 for h2  445.12 kJ/kg
 T2  428 K by interpolation

T1  473 K  sT01  7.295 kJ/kg.K


T2  420 K  sT0 2  7.195 kJ/kg.K

s
1 2 gen  m  s2  s1   0
 0 P2 
 m  sT 2  sT 1   R ln
0

 P1 

  200  
 0.01068 kg  7.195  7.295  kJ/kg.K  0.2968 kJ/kg.K ln  
  300 
 0.0002173 kJ/kg.K

206
(8.104/Q) A rigid container with 200 L is divided into
two equal volumes by a partition, shown in Fig. Both
sides contain nitrogen; one side is at 2 MPa and 2000C,
while the other is at 200 kPa and 1000C. The partition
ruptures and nitrogen comes to a uniform state at 700C.
Assume the temperature of the surroundings to be 200C.
Determine the work done and net entropy change for the
process. Take constant specific heats.

C.V. Nitrogen in A + B ; 1W2  0

Thermodynamics – Chapter 8 207


State A1: VA1  100 L, PA1  2000 kPa, TA1  200 o C
PA1VA1 2000 kPa  0.1 m 3
mA1    1.424 kg
RTA1 0.2968 kJ/kg.K  473.2 K
State B1: VB1  100 L, PB1  200 kPa, TB1  100o C
PB1VB1 200 kPa  0.1 m 3
mB1    0.1806 kg
RTB1 0.2968 kJ/kg.K  373.2 K

State 2: T2  70  273.15  343.2 K


mtot RT2  mA1  mB1  RT2
P2  
Vtot VA1  VA1
1.6046 kg  0.2968 kJ/kg.K  343.2 K
 3
 817 kPa
0.2 m

Thermodynamics – Chapter 8 208


 TA 2 PA2   TB 2 PB 2 
S A B  mA C p 0 ln  R ln   mB C p 0 ln  R ln 
 TA1 PA1   TB1 PB1 

  343.2   817  
S sys  1.424 1.042  ln    0.2968  ln  
  473.2   2000 
  343.2   817  
 0.1806 1.042  ln    0.2968  ln  
  473.2   200 
 0.1894 kJ/K

1 Q2  U 2  U1  mACv 0 (TA 2  TA1 )  mBCv 0 (TB 2  TB1 )


 1.424  0.745(70  200)  0.1806  0.745(70  100)  141.95 kJ
Q2 141.95
1
S surr    0.4841 kJ/K
T0 293.2
Snet  S sys  S surr  0.1894  0.4841  0.2947 kJ/K

Thermodynamics – Chapter 8 209


(8.120/8.138) A reversible heat pump uses 1 kW of
power input to heat a 250C room, drawing energy
from outside at 150C. Assuming every process is
reversible, what are the total rates of entropy into the
heat pump from the outside and from the heat pump
to the room?
TH  25o C

CV: Total setup


Q H

Reversible HP W  1 kW

Q L

TH  15o C

Thermodynamics – Chapter 8 214


C.V. Heat Pump
. . .
Energy Eqn.: Q + W  Q
L H

dSc.m 1
Entropy Eqn.:   Q  S gen
dt T
. . . .

Q L
Q H
Q L
Q H
0   0    flux of entropy
TL TH TL TH

Q H TH .
TH .
 Q 
TH  TL W
COPHP  
W TH  TL H

. .
.
Q H W 1 QL
   0.1 kW/K 
TH TH  TL 25  15 TL

215
(8.122/8.143) Room air at 23oC is heated by a
2000W space heater with a surface filament
temperature of 700 K. The room at steady state
loses heat to the outside, which is at 7 0C. Find
the rate of entropy generation and specify where
it is made

dScm Q 
Entropy Eqn. for C.V. at steady state:  0    S gen
dt T

Thermodynamics – Chapter 8 218


C.V.1: Heater Element
Q  (2000) W
dQ
S gen ,C .V .1        2.857 W/K
T T 700 K

C.V.2: Space between heater 700 0C and room 230C


Q dQ  2000 W 2000 W 
S gen , CV 2           3.9 W/K
T T  700 K  23  273 K 

C.V.3: Wall between 230 C and 70 C outside


dQ  2000 W 2000 W 
S genCV 3        0.389 W/K
T   23  273 K  7  273 K 
 S gen is largest for largest change in 1/T

Thermodynamics – Chapter 8 219


(8.123/8.161) A small halogen light bulb receives an
electrical power of 50 W. The small filament is at
1000 K and gives out 20% of the power as light and
the rest as heat transfer to the gas, which is at 500
K; the glass is at 400 K. All the power is absorbed
by the room walls at 250C. Find the rate of
generation of the entropy in the filament, in the
entire bulb including glass, and in the entire room
including the bulb.

Thermodynamics – Chapter 8 220


Troom  25o C

Gas 500 K

W elec
Radiation +
Conduction
Filament1000 K

glass  400 K

W elec  50 W, Q radiation  10 W, Q conduction  40 W

Thermodynamics – Chapter 8 221


C.V.1: Filament steady-state
Energy Eqn. :dECV / dt = 0 =W elec -Q rad - Q cond
Q rad Q cond 
Entropy Eqn. : dSCV / dt  0     S gen
TFil TFil

S   
Q rad  Q cond

W elec

50 W
 0.05 W/K
gen
TFil TFil 1000 K
C.V.2: Bulb including glass

S   dQ  (10 W)   40 W   0.11 W/K



gen  T 1000 K   400 K 
C.V.3: Total Room, all energy leaves at 250 C
Q Total 50 W
S gen    0.168 W/K
Twall 298 K

Thermodynamics – Chapter 8 222


(8.128/155) Two tanks contain steam, and they are both
connected to a piston/cylinder as shown in Fig. Initially
the piston is at the bottom, and the maximum of the
piston is such that a pressure of 1.4 Mpa below it will be
able to lift it. Steam in A has a mass of 4 kg at 7 MPa
and 7000C, and B has 2 kg at 3 MPa, 3500C. The two
valves are opened and water comes to a uniform state.
Find the final temperature and total entropy generation.
Control Mass: All water, mA  mB
Continuity Eqn. : m2  mA  mB
Energy Eqn. : m2u2  mA1u A1  mBuB1  1Q2  1W2   1W2
Entropy Eqn. :m2 s2  mA1s A1  mB sB1  1 S 2, gen
State A1: PA1  7000 kPa, TA1  700o C, mA  4 kg
 Superheated (B.1.3)  vA1  0.06283 m 3 /kg, u A1  3448.5 kJ/kg
s A1  7.3476 kJ/kgK, VA  0.2513 m3
State B1: PB1  3000 kPa, TB1  350o C, mA  2 kg
 Superheated (B.1.3)  vB1  0.09053 m3 /kg, uB1  2843.66 kJ/kg,
3
sB1  6.7427 kJ/kgK,VB  0.1811 m
Assume V2  VA  VB  P2  Plift , 1W2  P2 (V2  VA  VB )
Substituting in Energy eqn.:
m2u2  mA1u A1  mBuB1   P2 (V2  [VA  VB ])
m2 h2  mAu A1  mBuB1  P2 (VA  VB )
 4  3448.5  2  2843.7  1400  0.4324
 h2  3347.8 kJ/kg
State 2 : h2  3347.8 kJ/kg, P2  1400 kPa  Superheated
v2  0.2323 m3 /kg, s2  7.433 kJ/kgK, T2  441.90 C
V2  m2v2  1.394 m3  VA  VB  0.4324 m3
S
1 2, gen  6  7.433  4  7.3476  2  6.7427  1.722 kJ/K
(8.130/8.150) Water in a piston/cylinder is at 1
MPa, 500°C. There are two stops, a lower one at
which Vmin = 1 m3 and an upper one at Vmax = 3 m3.
The piston is loaded with a mass and outside
atmosphere such that it floats when the pressure is
500 kPa. This setup is now cooled to 100°C by
rejecting heat to the surroundings at 20°C. Find the
total entropy generated in the process.

226
State 1: P1  1000 MPa, T1  5000 C, V1  Vmax  3 m 3
Table B 1.3  it is superheated
v1  0.35411 m / kg, u1  3124.34 kJ/kg, s1  7.7621 kJ/K
3

V1 3
m 
v1 0.35411
 8.472 kg
State 1a : P1a  500 MPa, v1a  v1  0.35411 m 3 / kg
At P1a  500 MPa  v f  0.35411  vg  Saturated mixture
T1a  151.86o C  Tfinal  100o C
Process 1  1a : V  C

227
Vmin 1 m3
State 1b : P1b  500 MPa, v1b  
m 8.472 kg
 0.11803 m3 /kg
At P1b  500 MPa  v f  0.11803  vg  Saturated mixture
Process 1a  1b : P  C and T1b  151.86o C  Tfinal  100o C

State 2 : v2  0.11803 m , T2  100 C


3 0

Table B 1.1  T2  1000 C  v f @100 C  v2  vg @100 C


0 0

 It is a mixture  x2  0.0699
u2  418.91  0.0699  2087.58  564.98 kJ/kg
s2  1.3068  0.0699  6.048  1.73 kJ/K

228
W2  1aW1b   PdV  P V1b  V1a 
1

 500  1  3   1000 kJ
1
Q2  m u2  u1   1W2
 8.472 564.98  3124.34    1000   22682.8 kJ
Q2
1
S2 gen  m  s2  s1   1

Tsurr
 22682.8 
 8.472 1.73  7.7621   
 293.15 
 26.27 kJ/K

229
(8.131/8.153) A cylinder fitted with a frictionless
piston contains water. A constant hydraulic pressure
on the back face of the piston maintains a cylinder
pressure of 10 MPa. Initially, the water is at 700°C,
and the volume is 100 L. The water is now cooled
and condensed to saturated liquid. The heat released
during this process is the Q supply to a cyclic heat
engine that in turn rejects heat to the ambient at
30°C. If the overall process is reversible, what is the
net work output of the heat engine?

230
231
Energy equation: m u2  u1   1Q2  1W2
Process : P  C 
1W2   PdV  Pm  v2  v1 

State 1: P1  10 MPa, T1  700o C, V1  100 L  0.1 m 3


Table B 1.3  it is superheated
v1  0.04358 m3 / kg,
h1  3870.52 kJ/kg
s1  7.1687 kJ/kg.K
0.1
m  2.295kg
0.04358

232
State 2 : P2  10 MPa, Saturated liquid
Table B 1.2  v2  v f  0.001452 m3 / kg,
h2  h f  1407.53 kJ/kg
s2  s f  3.3595 kJ/kg.K
1 Q2  m u2  u1   1W2
 m u2  u1   Pm  v2  v1 
 m  h2  h1 
 2.295 1407.53  3870.52 
 5652.6 kJ

233
Heat transfer to the heat engine:
QH   1 Q2  5652.6 kJ
Take control volume as total water and H.E.
Qc.v.
Process: Reversible  m  s2  s1   0
TL
Qc.v.  TL m  s2  s1 
 303.15  2.295 3.3595  7.1687 
 2650.2 kJ
QL  Qc.v.  2650.2 kJ
Wnet  QH  QL
 5652.6  2650.2  3002.4 kJ
234
(8.134/159) A piston/cylinder assembly contains 2 kg
of liquid water at 200C, 100 kPa and it is now heated
to 3000C by a source at 5000C. A pressure of 1000
kPa will lift the piston off the lower stop. Find the
final volume, work, heat transfer, generation entropy.

C.V.: Water out to source at 5000 C


Energy equ : m u2  u1   1 Q2  1W2
Q2
Entropy equ : m  s2  s1   1
 1 S 2, gen
Tsource
Process : V  V1 if P  Plift or P  Plift if V  V1
State 1: m  2 kg, T1  20o C, P1  100 kPa
 compressed liquid
v1  v f @ 20o C = 0.001002 m3 /kg, u1  u f @ 20o C  83.94 kJ/kg
V1  mv1  2  0.001002 m3 /kg = 0.002 m3
Final state:T =3000 C and on line in P - V
3
State 1a: assume P1a =1000 kPa, v1a  v1  0.001002 m /kg
 Compressed liquid T1a  179.90 C

State 2 : P2 = 1000 kPa, T2 = 3000 C  Superheated vapor


v2 = 0.25794 m3 /kg, u2  2793.2 kJ/kg, s2  7.1228 kJ/kgK
V2  mv2  2  0.25794  0.51588 m3  V1  0.002 m3
 assumption is correct, the piston will lift


3
W
1 2  PdV  P2 (V2  V1 )  1000 kPa (0.51588  0.002) m
 513.9 kJ
1 Q2  m u2  u1   1W2  2 kg (2793.2  83.94) kJ/kg  513.9 kJ
 5932 kJ
1Q2
1 S2, gen  m( s2  s1 ) 
Tamb
5932.4 kJ
 2 kg (7.1228  0.2966) kJ/kg.K 
773.15 K
 5.98 kJ/K  0
(8.138/Q) A vertical piston/cylinder contains R-22 at
-200C, 70% quality, and the volume is 50 L, as
shown in Fig. 8.138. This cylinder is brought into a
200C room, and current of 10 A is passed through a
resistor inside the cylinder. The voltage drop across
the resistor is 12 V. It is claimed that after 30 min
temperature inside is 400C. Is this possible?

C.V.: R  22 Control mass


Energy Eqn.: m(u2  u1 )  1 Q2  1W2
Q 1 Q2
Entropy Eqn.: m(s 2  s1 )    1 S 2, gen   1 S 2, gen
T Tamb
Process: P = Constant => P1  Psat @  200 C  P2  245 kPa
State 1: T1  Tsat  20o C, x1  0.7, V1  0.05 m3
B.4.1: v1  0.06521 m 3 /kg, h1  176 kJ/kg, s1  0.6982 kJ/kgK
V1 0.05 m3
m   3
 0.767 kg
v1 0.06521 m /kg

State 2: P2  245 kPa, T2  40o C  superheated


Interpolate between 200 and 300 kPa
h2  282.5 kJ/kg, s2  1.1033 kJ/kgK

12  10  30  60
Elec. Work: Wele   Ei t   -216 kJ
1000
Total Work: 1W2  Wele  Pm(v2  v1 )
Substituting in energy eq. and solving for 1 Q2 :
1 Q2  m(u2  u1 )  Pm(v2  v1 )  Wele
 m(h2  h1 )  Wele
 0.767 kg (282  176) kJ/kg  216 kJ
 -134.5 kJ

1Q2
1 S 2, gen  m( s2  s1 ) 
Tamb
134.5 kJ
 0.767 kg (1.1033  0.6982) kJ/kg.K 
293.15 K
 0.768 kJ/K  0 Claim is valid
Radiation - Emission
 Radiation is the energy emitted by matter in the form of
electromagnetic waves (or photons) as a result of the
changes in the electronic configurations of the atoms or
molecules.
 Heat transfer by radiation does not require the presence
of an intervening medium.
 In heat transfer studies we are interested in thermal
radiation (radiation emitted by bodies because of their
temperature).
 Radiation is a volumetric phenomenon. However,
radiation is usually considered to be a surface
phenomenon for solids that are opaque to thermal
radiation.
Radiation - Emission

• Thermal radiation: 0.1 – 100 µm


• Light is visible portion: 0.4 – 0.76 µm
• The electromagnetic radiation emitted
by the sun is known as solar radiation
and nearly falls into 0.3 – 3 µm
• The infrared radiation : 0.76 -100 µm
• The ultraviolet radiation includes the
low-wavelength end of the thermal
spectrum and lies between the
wavelengths 0.01 - 0.4 µm. About
12% of solar radiation is in this range

You might also like