You are on page 1of 25

Chapter

3. Reversibility and Entropy


We have seen in the previous chapter that systems can move from an initial equi-
librium state to a different final equilibrium state by exchanging heat and work
with their surroundings. A new state can also be generated when a constraint, such
as a barrier to mass or heat flow, is released. The concept of the directionality of
change, however, has not yet been addressed. Everyday experience suggests that
some kinds of processes, even though in full compliance with the First Law, never
occur spontaneously. Examples of spontaneous processes are given in the table be-
low; the reverse processes have never been observed in isolated systems.

Initial state

Final state

Inflated helium balloon

Helium balloon deflated

Two metal blocks in contact


at different temperatures

Two metal blocks in contact


at the same temperature

Sugar crystals at the bottom


of a glass of pure water

Sugar dissolved in the water

Dispersed droplets of oil in water

Oil floating as separate layer on top of


the water


The last two of these examples (sugar dissolving in water, but oil separating
from water) illustrate that it is not always obvious which direction (mixing or de-
mixing) corresponds to spontaneous change. However, once the spontaneous
change direction has been established, it is impossible to convert a system in the
final state column back to the initial state without the action of an external agent.
For example, sugar crystals will not precipitate out of solution unless the water is
evaporated.


Material from Essential Thermodynamics, Drios Press (2011), A.Z. Panagiotopoulos

36

3.1 Reversibility

37

3.1 Reversibility
Clearly, all processes observed naturally are spontaneous and thus irreversi-
ble. Spontaneous changes define the direction of the arrow of time and are rigor-
ously analyzed through the Second Law of thermodynamics, which is formally in-
troduced in 3.2. By eliminating friction and all gradients of temperature or pres-
sure, and by performing changes at an infinitesimally slow rate, one can approxi-
mate an idealized reversible process. A process in a system going from an initial to
a final state is formally defined as reversible if the system can be brought back to
its initial state from the final state with no change to any part of the universe. Re-
versibility is an idealized abstraction that can be approximated in real processes, if
care is taken to avoid friction, internal or external gradients of temperature, pres-
sure, or composition, as well as any conversion of mechanical work into heat. In
reversible processes, forces
across any boundary exact-
ly balance at all times. Pro-
cesses take an infinitely
long time to complete, be-
cause rates of heat transfer,
macroscopic flow of mate-
rial, and component diffu-
sion are, respectively, pro-

Figure 3.1 Rupture of a partition in an insulated con-
portional to gradients of
tainer.
temperature, pressure, and

chemical potentials (the
latter defined in Chapter 5).
Consider the process schematically illustrated in Fig. 3.1. A thin wall separates
two sides of an insulated container in the initial state (A); the left-hand side con-
tains a gas and the right-hand side is evacuated. The wall is removed and the gas
expands to fill all available space. After some time, the gas reaches a new equilibri-
um state, (B). A First Law balance on the total contents of the container gives U =
Q
/ + W
/ = 0. If the gas is ideal, its energy depends only on the temperature, and T =
0, but this is not an essential characteristic of the process; for real gases, the tem-
perature of state B can be
lower or higher than that
of state A. We can return
the gas to its original state
A by using a piston to
compress it back to the
original volume, as shown
in Fig. 3.2. For this reverse

process, the change of en-
Figure 3.2 Reversing the effect of expansion.
ergy of the gas is the same

38

Chapter 3. Reversibility and Entropy

as previously, U = 0 = Q + W
Q = W ; since we clearly
need to provide work to
compress the gas, the re-
verse process is only pos-
sible if we remove heat
from the system. If some-

how we could convert all
Figure 3.3 Reversible expansion.
of the heat removed into
work, then the overall pro-
cess has no net effect on the environment and would have been reversible accord-
ing to the definition above. However, countless experiments have shown that
complete conversion of heat into work is impossible; we thus conclude that the
original process !A B was indeed irreversible.
We now consider a different process (Fig. 3.3) starting from the same initial
state A as previously, with final state C of the same volume as B. However, we now
perform an adiabatic process (!Q = 0 ), using a well-lubricated, frictionless piston to
obtain work from the expanding gas. The amount of work produced by the gas can
be calculated from:
V

WAC = C P dV
VA
!
Gases generally cool as they expand adiabatically, so that the final tempera-
ture will be TC < TA. For the reverse process of adiabatic compression, we need to
input work, which in the absence of friction will be equal in magnitude and oppo-
site in sign, WCA = WA C. Thus, there is no overall net effect of the process
ACA on either system or its environment. This process is reversible.

3.2 The Second Law of Thermodynamics


The Second Law of thermodynamics places the concept of spontaneous changes on
a rigorous, quantitative basis. It allows for precise predictions about which trans-
formations can or cannot occur naturally in the absence of external driving
forces. It represents, in a highly compact fashion, the accumulated scientific
knowledge gained from countless experiments in carefully constructed systems.
While it cannot be proven from first principles and has to be postulated, it repre-
sents one of the most basic and well-established tenets of modern science.


This, however, is not true for all substances.

3.1 Reversibility

39

The Second Law can be stated in a variety of equivalent ways; in the following,
we use a traditional approach that expresses it in terms of heat fluxes around heat
engines. Heat engines are devices that exchange heat and work with their envi-
ronment in processes that leave no net effects on the engines. An internal combus-
tion engine undergoing a full cycle of compression-ignition-expansion is a classic
example of such a device. However, living cells, photovoltaic arrays, and distilla-
tion columns can also be considered as heat engines, provided that no net changes
accumulate within the devices over the period of interest.
Analysis of heat engines becomes particularly simple if they exchange heat
with heat reservoirs large enough so that their temperature can be considered con-
stant, irrespective of the amount of heat exchanged with the engines. In essence,
this implies that the amount of energy exchanged is small relative to the total heat
capacity of the reservoirs. Practical approximations of heat reservoirs include the
atmosphere, rivers, or the sea, when considered over short enough periods of time
so that their temperature does not vary significantly. Electrical power plants are
usually built near large bodies of water precisely so that they can use them as heat
reservoirs. We will assume that heat reservoirs do not perform, or require, any
work in the process of interest.
The Kelvin-Planck postulate (statement of the Second Law) is that:

It is not possible for a heat engine that interacts with a single heat reservoir to
convert all the heat transferred from the reservoir into work.

In other words, heat cannot be completely converted into work. However, the re-
verse of this process, converting work entirely into heat, is perfectly possible.
Fig. 3.4 depicts a process that is impossible according
to the Kelvin-Planck statement of the Second Law, assum-
ing that the signs of the flows of heat Q and work W are as
indicated by the arrows. If |Q| = |W|, such a device does not

violate the First Law, as a trivial balance around the en-
gine demonstrates. No such device has ever been con- Figure 3.4 An impossible
process.
structed, despite efforts by many people since the dawn

of technological civilization. The practical value of being able to extract useful en-
ergy in the form of electricity or mechanical work by cooling the sea or the atmos-
phere would be enormous; fossil fuel use could then be reduced to zero and a prac-
tically inexhaustible supply of energy would become available to humanity. Na-
ture, however, does not permit such a device ever to function. Even in microscopic
systems, where spontaneous fluctuations occur that seem to result in mechanical
work being produced at the expense of the thermal energy of a reservoir (e.g.
Brownian motion of small particles), the Second Law is found to hold when the av-
erage behavior of a system is determined over long times. Devices that violate the

Kelvin-Planck
statement of
the Second
Law

40

Chapter 3. Reversibility and Entropy

First or Second Laws which thus can never be constructed are called perpetual
motion machines of the first or second kind, respectively.

3.3 Carnot Engines and Absolute Temperature


The seemingly simple statement of
the impossibility of certain process-
es made by the Second Law has
enormous consequences on the be-
havior of natural and engineered
systems. To start analyzing its con-
sequences, consider the behavior of
heat engines operating between two
heat reservoirs of different thermo-
metric temperatures, H and C , as

shown in Fig. 3.5. We will assume Figure 3.5 Thermal engines operating between
that the top reservoir is at a higher two heat reservoirs. Left side (engine A) corre-
sponds to power production and right side
thermometric temperature, H > C.
(engine B) corresponds to refrigeration.
Thermometric temperature in phys-
ical systems is observed to increase as energy flows into them.
On the left side of Fig. 3.5, engine A removes heat from the hot reservoir and
dumps some of it to the cold reservoir, while producing some work. Engine B on
the right panel does the reverse, removing heat from the cold reservoir and dump-
ing it to the hot reservoir through the consumption (input) of work. Nothing in the
First or Second Laws of thermodynamics prohibits any of these processes; practi-
cal examples of such systems, in the form of thermal power generation systems
and refrigeration units are ubiquitous. Nevertheless, there must be some con-
straints in the magnitudes of the work and heat streams from these engines, be-
cause otherwise a combination of them could give rise to a perpetual motion ma-
chine. To see why this is the case, consider using the work output of a machine of
type A to provide the energy input to a machine of type B. Without loss of generali-
ty, we can assume that:

Q = QCB
! CA
This is the case since we can always adjust the size of the heat engine, or how
many engine cycles are performed per unit of time. First Law balance around the
system composed of the engines A+B then gives the net work input as:

The net heat input is:

W = WB WA
!

41

3.3 Carnot Engines and Absolute Temperature

Q = QHA QHB = W
!


If W < 0, then

Q > 0 QHA > QHB


!

Net work is being produced by the two engines by removing heat from the hot res-
ervoir. This is equivalent to a perpetual motion machine of the second kind, violat-
ing the Kelvin-Plank postulate, so it is an impossibility. Therefore, we must have:

W 0 Q 0 QHA QHB
!

for any two heat engines operating between two specific temperatures H and C .
The arguments of the previous paragraph also apply to the special case of re-
versible heat engines, known as Carnot engines. For reversible engines we can
simply switch the role of A and B at will, which gives:

rev
rev
rev
rev

QHA
QHB
!!and!! QHB
QHA

rev
rev

QHA
= QHB

This implies that for any reversible heat engine that operates between H and C

the ratio of heat fluxes is a universal function of the two temperatures:

QHrev
QCrev

= f (H ,C )

!
The specific form of the function in Eq.
3.1 can now be determined, by consider-
ing a cascade of heat engines operating be-
tween a high temperature H, an interme-
diate temperature M and a low tempera-
ture C, as shown in Fig. 3.6. On the left
side, two Carnot engines, 1 and 2, are used
between H and M and between M and

C , respectively. On the right side, a single
engine 3 operates directly between H and
C. We can adjust the heat flows for en-
gines 1 and 2 so that the same amount of
heat is taken and removed from the in-
termediate-temperature reservoir at M,
and we can also adjust engine 3 so that the
same amount of heat is removed from the
hot reservoir as for engine 1. Since these
are Carnot (reversible) engines, we must
have that |W1| + |W2| = |W3|. Therefore, the

(3.1)


Figure 3.6 Thermal engines operating
between three heat reservoirs.

42

Chapter 3. Reversibility and Entropy

same amount of heat is dumped to the cold reservoir on the two sides of the cas-
cade:

QC = QC

QM / f (M ,C ) = QH / f (H ,C )

QH

QH

f (H ,M ) f (M ,C ) = f (H ,C )

f (M ,C ) f (H ,M ) f (H ,C )
!
Since this relationship is valid for any temperatures H, M, and C, the func-
tional form of f must be:

!

g(H )

f (H ,C ) =

g(C )

(3.2)

The function g() is universal for all Carnot engines. Its form can be obtained
from analysis of the operation of any reversible heat engine. When this analysis is
performed for engines that have ideal gases as working fluids (see 4.1), we ob-
tain a particularly simple relationship: g() = T, where T is the temperature that
was defined through the ideal-gas equation of state in 2.4. This remarkable prop-
erty provides a connection between the efficiency of reversible engines and the
ideal-gas temperature scale. The absolute temperature scale T is of fundamental
importance in thermodynamics.
Given that f (H,C) = TH/TC, we can now show that reversible heat flows occur-
ring to and from constant-temperature reservoirs are linked in a particularly sim-
ple way. The heat flows for the cold and hot reservoirs are of opposite signs, so Eq.
3.1 can be expressed as:

QHrev

QCrev

TH
TC

QHrev
TH

QCrev
TC

= 0

(3.3)

The efficiency of a Carnot engine that draws heat from a hot reservoir and
produces work (left side of Fig. 3.5) is defined as the fraction of heat removed from
the hot reservoir that is obtained from the device as work. From a First Law bal-
ance around a reversible engine, we have:

Carnot
Engine
efficiency

T
QHrev + QCrev +W = 0 QHrev C QHrev = W
TH
!

rev
!

W
QHrev

TH TC
TH

(3.4)

43

3.3 Carnot Engines and Absolute Temperature

The efficiency is between 0 and 1, the latter value being a theoretical limit
that can only be asymptotically approached if TC0 or TH+. Real engines, of
course, operate with efficiencies lower than those of reversible engines, since oth-
erwise an irreversible engine could be combined with a Carnot engine operating in
reverse, resulting in a violation of the Second Law. The performance of real en-
gines operating between reservoirs of given temperatures is not the same for all
irreversible engines it varies (sometimes greatly), depending on their design and
internal losses. The reversible engine efficiency provides a common upper limit for
all real engines.
Example 3.1 Efficiency of geothermal power production
In many parts of the world, especially near edges of tectonic plates, relatively high
temperatures can be reached by drilling to moderate depths. Taking advantage of
these high temperatures has been proposed as one possible technology for energy
generation without production of greenhouse gases. Assuming that heat can be
withdrawn from hot rock at H = 200 C and that cooling is available at C = 50 C,
what is the maximum possible fraction of heat removed that can be converted to
electricity?

The maximum possible fraction of heat conversion into work takes place using a
reversible Carnot engine, which has efficiency:
T T
200 50
rev = H C =
= 32%
TH
200 + 273
!
Note that only absolute temperatures [in K] can be used in the expression for ;
however, the difference between two temperatures in C is the same as the corre-
sponding absolute temperature difference in K, so the thermometric temperature
values can be used in the numerator.

For engines operating as refrigeration cycles (right side of Fig. 3.5, p. 40), with
heat removed from the cold reservoir through the net input of work, a measure of
performance different from is appropriate. In such cases, we are interested in
the amount of heat removed from the cold side (the interior of the refrigerator, or
the inside of a building in the case of air conditioning) per unit of work input. The
coefficient of performance, , is:

! = !
!

QC
W

QC

QC QH

rev !=

QCrev
QCrev

T
+ H QCrev
TC

TC

TH TC

(3.5)

Another possibility is the operation of a cycle as a heat pump, to maintain high-


er temperatures inside a building in the winter, by withdrawing heat from the cold

44

Chapter 3. Reversibility and Entropy

outside air. In this case, the coefficient of performance is defined as the ratio of
heat output (into the building) over the work input:

!=!

QH
W

QH

QC QH

rev =

QHrev
T
C QHrev + QHrev
TH

TH

TH TC

(3.6)

The two coefficients of performance defined in Eqs. 3.5 and 3.6 can be greater
than 1, if the difference between hot and cold temperatures is smaller than the ab-
solute value of the cold or hot temperature, respectively. Practical power and re-
frigeration cycles are discussed in Chapter 4.
Example 3.2 Air conditioner theoretical efficiency
Calculate the maximum possible coefficient of performance for an air conditioning
unit operating between an indoor temperature of 68 F and outdoor temperature
of 104 F.

We first need to convert thermometric to absolute temperatures: C = 68 F C =
293 K; C = 104 F TH = 313 K. The maximum possible coefficient of performance
is:

rev

!=!

QCrev
W

TC

TH TC

293
= 14.6
313 293

3.4 Entropy Changes in Closed Systems


We showed in the previous section that reversible heat flows occurring to and
from constant-temperature reservoirs satisfy the condition QHrev /TH + QCrev /TC = 0 ;
!
this in turn suggests that a heat-related property can be defined that is conserved
for a system undergoing a reversible process that starts and ends at the same state
point. This new property is called the entropy S. For a reversible process in closed
systems for which all exchange of heat is done at a single temperature T, the en-
tropy change S is defined as:

Q rev
S

(3.7)
T
!
For a differential change of state during a reversible process, the temperature
of the system is effectively constant, so we can write:

Q rev
dS =

T
!

(3.8)

3.4 Entropy Changes in Closed Systems

45

The overall entropy change for a general process that involves temperature
changes along its path is:
definition of
entropy

rev
B Q

S = SB S A =
A T
!

(3.9)

A reversible process can always be constructed between any two equilibrium


states of a system, provided that external reservoirs are available to exchange heat
and work as needed. Thus, Eq. 3.9, along with a choice of a reference state for
which S = 0, in principle provides a way to generate values for the entropy of any
equilibrium state of a thermodynamic system. As will soon become apparent, Eq.
3.9 is not usually a practical approach to obtain actual entropy changes. Thermo-
dynamic relationships incorporating volumetric and heat capacity data, to be dis-
cussed later in the present section and in 7.3, are more commonly used for en-
tropy change calculations.
A key property of the entropy is that it always increases for spontaneous pro-
cesses in closed, isolated systems. This property is often used as an alternative
statement of the Second Law
here, we simply derive it from the
Kelvin-Planck postulate as follows.
Consider an irreversible, adiabatic
process taking place between ini-
tial state A and final state D, along
the dashed-line path of Fig. 3.7.
Curves AB and CD represent re-
versible adiabatic paths from the
initial and final state respectively,
and curve BC is an isothermal path

at temperature T. For the reversible
Figure 3.7 Pressure-volume relationship for ir-
process DCBA that brings the sys-
reversible process (dotted line) and adiabatic-
tem from state D back to the initial
isothermal-adiabatic path.
state A, all heat transfer takes place
along the isothermal step BC, so that:
rev
rev
rev

U UD = QDA
+WDA
= T(S A SD )+WDA
! A
For the irreversible process AD

irr

U U A = WAD
! D
Adding these two expressions we obtain

rev
irr
T(S S )+WDA
+WAD
= 0
! A D

46

Chapter 3. Reversibility and Entropy

irr
rev
rev
+WDA
< 0 , then T(S A SD ) = QDA
> 0 ; the overall cycle has received
If WAD
!
!
!
heat input from a single reservoir at T and has produced net work. This is a viola-
tion of the Kelvin-Planck postulate and therefore impossible. The case
rev
W irr +WDA
= 0 implies that T(S A SD ) = Q rev = 0 and the final state could be re-
! AD
!


turned to the initial state with no net effect on the environment; this contradicts
the original statement that the process AD was irreversible. Therefore, we must
have, for any irreversible process in an isolated closed system from state AD,
T(S A SD ) < 0 , or equivalently:

S > S [for irreversible process A D]


!D A

(3.10)

irr
rev
irr
rev
+WDA
> 0 WAD
> WDA
In addition, we must have WAD
. By changing the
!
direction of the reversible process from DA to AD we then obtain:

rev
W irr > WAD
!!
! AD

(3.11)

Since the universe can be considered a closed isolated system, Eq. 3.10 is the
mathematical equivalent of the Clausius statement of the Second Law, the entropy
of the universe tends to a maximum. Eq. 3.11 states that the algebraic work
amount is always greater for an irreversible process. If the reversible process re-
rev
!>!0! ) then the irreversible process requires more work input;
quires work ( WAD
!
rev
!<!0! ), Eq. 3.11 suggests that the ir-
if the reversible process produces work ( WAD
!
reversible process will produce less work it may even require work input! Re-
versible processes are the best of all possible processes in achieving a given task
with the least expenditure of useful work and in extracting the maximum possible
amount of useful work out of a given change of state.
Entropy has been just derived from analysis of heat and work flows in reversi-
ble heat engines; however, once it has been established that entropy is a proper
thermodynamic function, it can be expressed for any equilibrium state in terms of
any convenient thermodynamic variables. This provides the starting point for de-
velopment of thermodynamic identities through the formal approach described in
Chapter 5. Let us take a look again at the differential form of the First Law of ther-
modynamics, written for a closed system undergoing a reversible process:

rev
rev
!dU = Q + W

(3.12)

A differential change of state does not substantially change the temperature of


the system, so we can set Q rev = TdS . Since the pressure also remains constant
!
(within a differential amount), the work performed by the environment on the sys-
tem can be expressed as !W rev = PdV . Combining the expressions for Q rev and
!
rev
!W we get the fundamental equation of thermodynamics, which provides a syn-
thesis of the First and Second Laws:

47

3.4 Entropy Changes in Closed Systems

!dU = TdS PdV

(3.13)

Eq. 3.13 has only proper thermodynamic state functions on both sides, contain-
ing no inexact differentials involving heat or work amounts. Changes in state func-
tions are independent of the path, so the equation is valid for all processes, re-
versible or irreversible. In irreversible processes, the first term (TdS ) is not equal
to the amount of heat, and the second term, (PdV ) does not equal the amount of
work; but their sum still gives the change in system energy.
Example 3.3 Entropy change for an ideal gas
10 mol of an ideal gas with CV = 20.8 J/mol at T0 = 300 K and P0 = 0.3 MPa occupy
the left half of an insulated vessel, as shown in Fig. 3.8. The other half is evacuated.
At time t = 0, a 1 kW electrical heating element is turned on. After 30 s, the parti-
tion dividing the vessel ruptures and the heating element is turned off. Calculate
(a) the final temperature Tf and pressure Pf of the gas in the vessel and (b) the en-
tropy change of the gas during this process.

The final temperature of the gas can be obtained from a First Law balance on the
contents of the vessel:

U = NCV T = Q + W Tf = T0 +

Q
NCV

30!s! !1000!J/s
Tf = 300!K! +
= 444!K
10!mol! 20.8!J/(mol!K)
!

The final pressure is obtained from the ideal-gas law,

Pf Vf
RTf

P0V0
RT0

Pf = P0

V0Tf
Vf T0

444!K
Pf = 0.3!MPa
= !0.222!MPa
2! !300!K
!
The entropy change cannot be directly obtained from
the definition (Eq. 3.9), as this is not a reversible pro-
cess. Instead, it can be obtained from Eq. 3.13:


Figure 3.8 Schematic of pro-
cess for Example 3.3.

1
P
dU = TdS PdV dS = dU + dV
T
T
!
NCV
C
T
V
NR
S
R
dS =
dT +
dV
= V dT + dV = CV ln f + Rln f
T
V
N
T
V
T0
V0
!

Valid for
any process
in a closed
system

48

Chapter 3. Reversibility and Entropy

S
444
J
J
= 20.8ln
+ 8.314ln2 !
= (8.16 +5.76)!

N
300
mol!K
mol!K

J
J
S = 10!mol! 13.9
S = 139!
mol!K
K
!
!
The entropy change is, of course, greater than zero during this irreversible pro-
cess.
Using the same approach as in Example 3.3, it is easy to prove that the general
relationship for the entropy change of an ideal gas with temperature-independent
heat capacities during a process that takes it from (T0 ,P0 ,V 0 ) (Tf ,Pf ,V f ) is:
Entropy
change for
an ideal gas

S
!

IG

= CV ln

Tf
T0

+ Rln

Vf
V0

= C P ln

Tf
T0

Rln

Pf
P0

(3.14)

Eq. 3.14 is of general applicability to changes of state for ideal gases the pro-
cess does not have to be reversible, at constant temperature or at constant pres-
sure.
Example 3.4 Forging
A common process in metallurgy is to immerse a red-hot item (e.g. a blade being
forged) in water to harden it. Let us consider a process in which a 1 kg steel blade
at 1 = 800 C is immersed in a large vessel filled with water at 2 = 25 C and imme-
diately quenched. Assuming that steel has heat capacity CV = 460 J/(kg K), inde-
pendent of temperature, calculate the entropy changes during this process of (a)
the steel blade, (b) the water, and (c) the universe. Assume that there is enough
water in the vessel so that its temperature does not increase appreciably during
the process of immersion.

The entropy change of the steel blade in this irreversible process can be calculated
from Eq. 3.13, ignoring the small volume change of the solid as it cools down:

dU = TdS PdV
!

NC dT
1
dS = dU = V

T
T

C
T
Sblade = N V dT = NCV ln f
T
T0
!

J
273+ 25
J
Sblade = 1!kg! !460!
!ln
= 589!
kg!K 273+ 800
K
!

49

3.4 Entropy Changes in Closed Systems

Even though this is an irreversible process, the blades entropy goes down the
blade is not an isolated system, so Eq. 3.10 does not apply.
For the water, we are not given enough information to calculate its temperature
change; since the process is clearly irreversible, it would seem that we cannot use
the definition of entropy change (Eq. 3.9) however, this is not true! Even though
the overall process is irreversible because of the large temperature difference be-
tween the blade and the water, we can devise a thought experiment in which the
heat transfer to the water is done in a reversible manner. The amount of heat
transferred is:

J
Q = U blade = NCV T = 1!kg 460!
(25 800)!K!=!356!kJ
kg!K
!
The entropy change for the water in a reversible process with the same Q is:
Q rev
356!kJ
J
Swater =
=
= 1196!
T
(273+ 25)!K
K
!
The entropy change of the universe is:

J
J
Suniverse = Swater + Sblade = (589+1196) = +607
K
K
!
Once again, the entropy change of the universe is positive, as it should be for an
overall irreversible process.

3.5 Entropy Changes in Open Systems


The entropy of an open system can change because of heat exchanged with the en-
vironment, because of flows into and out of the system and because of entropy
generation due to internal irreversibilities entropy is not a conserved quantity!
While entropy can be created, a process can never result in a net decrease in the
entropy of the universe; this would violate the Second Law of thermodynamics.
This condition translates into an inequality linking the entropy change of the sys-
tem Ssystem, the entropy change of the environment Senv, and the specific entro-
pies of entering and leaving streams:

Suniverse = Ssystem + Senv +


!

leaving
streams

Nout S out

entering
streams

Nin S in 0 (3.15)

Why are the terms for the entering and leaving streams in Eq. 3.15 of opposite
sign relative to Eqs. 2.16 or 2.19 (p. 18), the First Law balance for open systems?

50

Chapter 3. Reversibility and Entropy

Eqs. 2.16 and 2.19 are written from the point of view of the system, which gains
the streams that enter and loses the streams that leave. By contrast, Eq. 3.15 rep-
resents a balance for the entropy of the universe. For the universe, streams enter-
ing the system are lost and streams exiting the system are gained. A similar re-
lationship can also be written as a differential (rate of change):

Suniverse = Ssystem + Senv +

leaving
streams

N out S out

entering
streams

N in S in 0

(3.16)

For reversible processes, Eqs. 3.15-3.16 become equalities, with their right-
hand side equal to zero. For other special cases, specific terms can be set to zero,
significantly simplifying the equations. For example, the term Ssystem is zero at
steady state, as there is no net change in the system. Another example of simplifi-
cation is for an adiabatic process, for which we can set Senv = 0, since no heat flows

into the environment.
The fact that entropy generation is non-negative for all feasible processes im-
poses significant constraints on the amount of work that can be produced from (or
is required for) specific processes. When performing a thermodynamic analysis, a
typical approach is to apply in turn a First Law balance followed by a Second Law
balance, as illustrated in the examples that follow.
Example 3.5 Feasibility of a process
An inventor is proposing a black box device for producing electrical power that
operates on a stream of compressed air at P1 = 4 bar, T1 = 300 K. The input stream is
split into two equal flows of P2 = P3 = 1 bar at T2 = 280 K and T3 = 260 K, respectively.
The claim is made that power is produced at a steady-state rate of W = 3.4 kJ/mol
of air flowing through the device. Assuming that unlimited heat exchange with the
environment at Tenv = 300 K is allowed, you are asked to provide an analysis of the
thermodynamic feasibility of such a device. Assume that air can be approximated
as an ideal gas with CP = 29.1 J /(mol K), independent of temperature.

First, we apply an integral
First Law balance on this
open system over the pe-
riod of time it takes for 1
mol of air to flow through
the device, which is as-
sumed to be at steady
state. The flow of streams
2 and 3 is half the flow of
stream 1. We use a refer-
ence state for the enthalpy

Figure 3.9 Schematic of process for Example 3.5.


51

3.5 Entropy Changes in Open Systems

such that H = C PT .

U
= 0 = Q +W + NH 1

!steady!state

T +T
N
N
H 2 H 3 Q = W NC P T1 2 3
2
2
2

J
260+ 280
Q = (3400)!J! 1!mol!
!
!29.1!
300
K != 2527!J
mol!K
2
!


Since the heat calculated from the systems point of view is positive, heat flows
from the environment to the system during this process. We must now check
whether the entropy generation rate for the universe is non-negative. We use Eq.
3.14 (p. 48) to calculate the entropy change of the ideal gas streams and take into
account that the flow of heat into the environment is the opposite of the value cal-
culated from First Law balance on the system:

Suniverse =

Sdevice + Senv + Nout S out Nin S in =

leaving!
entering!

steady!state

!
= !

streams

streams

Q N
N
Q N
+ S 2 + S 3 N S1 =
+ (S 2 S 1 + S 3 S 1 )
Tenv 2
2
Tenv 2

TT
PP
Q N
+ C P ln 2 3 Rln 2 3 =
Tenv 2
T12
P12

2527!J 1!mol
J
260 280
J
1
!+!
29.1!
ln
8.314
ln 2 =

2
300!K
2
mol!K
mol!K 4
300

J
J
!!!!!!!!!!!!!!!!!!!!!!!!!!! 8.42 3.09!+11.53 ! = +0.02!
K
K
!

The entropy change of the universe is positive, so operation of the device is possi-
ble as described. However, since the entropy change is small relative to the terms
of which it consists, the device is operating near the thermodynamic limit for re-
versible processes.

Example 3.6 Solar power generation
Solar collectors are used to heat, continuously and at steady state, a molten salt
stream from 200 C to 600 C at a flow rate of 12 kg/s. The hot stream generates
electrical power by a complex process and then returns to the solar collectors.
What is the maximum amount of power that can be produced? Assume that the
environment is at 20 C. The heat capacity of the molten salt at constant pressure is
CP = 0.8 kJ/(kg K), independent of temperature.

52

Chapter 3. Reversibility and Entropy

Figure 3.10 Schematic of process for Example 3.6.


The system of interest is the power plant that
uses the salt stream. We label the

hot stream H and the cold one C. An open-system First Law balance for this sys-
tem gives:

dU
=0
dt

H HC )
= Q + W + N(H

steady!state

W = Q + NC P (TC TH )
!
The maximum amount of work is produced in a reversible process, for which

Suniverse = 0 =

Splant + Senv + Nout S out Nin S in =

leaving!
entering!

steady!state

streams

streams

T C
T
Q
ln C
+ N(S C SH ) Q = Tenv N C P dT = Tenv NC
P
T
Tenv
TH
H T

T
T ln H +T T
W = NC
P
env
C
H
TC

273+ 600
kg
kJ
W = 12 0.8
293!ln
+ 200 600 K!!= 2.12!MW

s
kg!K
273+ 200

!
The power is negative because it is being produced by the power plant.

Example 3.7 Analysis of a Steam Turbine


A power plant operates with supercritical steam at T1 = 900 K and P1 = 8 MPa as in-
put to a turbine. The steam exits at T2 = 430 K. Assuming that the turbine operates

53

3.5 Entropy Changes in Open Systems

reversibly and adiabatically, how much steam is required to generate power at a


rate of 1 MW?







Figure 3.11 Schematic of process for Example 3.7.


For reversible operation of the steam turbine, we must have:

Suniverse = 0 = Sturbine + Senv + N(S 2 S 1 ) S 2 = S 1

reversible!process

steady!state

adiabatic

The properties of steam at the entrance of the turbine (T1 = 900 K and P1 = 8 MPa)
can be obtained from the NIST WebBook as follows:

kJ
kJ
H 1 = 3707! !!!!!S 1 = 7.095

kg
kg!K
!
We now need to find the pressure for which S2 = S1 = 7.095 kJ/(kg K) at T2 = 430 K.
With a bit of trial-and-error with respect to the pressure range, we can obtain from
the NIST WebBook P2 = 0.31 MPa, H2 = 2775 kJ/kg. A First Law balance on the tur-
bine now gives:
dU

= 0 = Q + W + N(H
1 H2 )
dT

!steady!state
W
103 !kJ/s
N =
=

H 2 H 1 (2775 3707)!!kJ/kg
!
Example 3.8 Work for evacuating a tank

kg
N = 1.073
s
!

A tank of volume V0 initially contains air at a certain temperature T0 and pressure


P0. The environment is also air at T0 and P0. Obtain general expressions for the
minimum amount of work required to evacuate or compress the tank to a pressure
P1 = P0 (a) isothermally and (b) adiabatically. Perform illustrative calculations for
= 0.1 and 10, and obtain the work required for complete evacuation ( = 0). Air
can be considered an ideal gas with CV = 5R/2.

54

Chapter 3. Reversibility and Entropy

This problem can be handled most conveniently


by transformation from an open to an equivalent
closed system. For the case of evacuation, we
would like to remove material from the tank of
volume V0 , initially at T0 and P0, to bring the pres-
sure down to P1. This process reduces the number
of moles in the tank from N0 to N1. Instead of ac-
complishing this by action of a pump, in the
Figure 3.12 Schematic of pro-
closed system process, a frictionless piston is set
cess for Example 3.8.
up as shown in Fig. 3.12, at a position within the
tank that contains an amount of material N1, has
volume V1, and is originally at T0 and P0. At the right-hand side of the piston there
is air at T0 and P0. We will pick V1 so that after the piston has moved from V1 to V0,
the final pressure within the tank is P1 = P0. This reversible process accomplishes
the same task as pumping the original tank (of volume V0) from P0 to P1. The mini-
mum work for evacuation is obtained as the work to move the piston rightward
from V1 to V0. A similar approach can give the minimum work required for com-
pression to a certain pressure, or the maximum work that can be obtained from a
cylinder of compressed air.
(a) Isothermal operation
For this case, from the ideal-gas equation-of-state, PV = NRT, the volume V1 is ob-
tained as:

V1
V
! 0

P1
P0

N1
N0

The instantaneous pressure difference across the piston is P P0, where P is the
pressure within the expanding volume on the left side. The total work required to
move the piston to the right is:
V

V0 N RT
1

V1

V1

W = 0 (P P0 )dV =

dV + P0(V0 V1 ) = N1RT ln

V1
V0

+ P0(V0 V1 )

V PV PV
P V
W N1
=
RT ln 1 + 0 0 0 1 = RT ln + RT 0 1 = RT ln + RT RT
N0 N0
V0 N0
N0
N1

W
W
=
= ln +1
RTN0 P0V0

For = 0.1, this equation gives W/(RTN0) = 0.670.


Because ln0 as 0, the expression has a simple interpretation at the limit of
complete evacuation:

55

3.5 Entropy Changes in Open Systems

W
W
=
= 1 W = P0V0 for complete evacuation.
!
N0RT P0V0

!
In other words, the minimum work for complete evacuation is the work to push
back the atmosphere by a volume equal to the original tank volume.
n the case of pressurization, a more meaningful measure than the work per mole
initially in the tank is the work per mole of compressed air, obtained by dividing
the full expression above by = N1/N0,

W
W
1
=
= ln + 1
RTN1 P1V0

For = 10, this equation gives W/(RTN1) = 1.403. The work required is positive
(work input to the system) for both lowering and increasing the tank pressure to a
value different than atmospheric.
(b) Adiabatic operation
Here, we will use the expressions obtained in 2.4 for adiabatic compression and
expansion of an ideal gas. Even though we did not state so at the time, adiabatic
operation (Q = 0) at internal equilibrium implies also a reversible operation, since
the entropy change of the universe for such a process is zero.
As previously, we find V1 so that P1 = P0 after expansion. From Eq. 2.31, with ini-
tial volume and pressure V1 and P0 and final values V0 and P1 :

V
== 0
P
V1
! 0
P1

V1
V0

= 1/ =

CV /C P

The final temperature after expansion is obtained from Eq. 2.30:


1

P
R/C
= 1 = P
T
P
! 0 0
The work performed on the gas during the expansion is, from Eq. 2.32:
T1

T
R/C
Ugas = Q +Wgas Wgas = N1CV T = N1CV (T1 T0 ) = N1T1CV 1 0 = N1T1CV 1 P
T1
!

The net work required is the sum of the work performed on the gas and the work
to push back the atmosphere,

56

Chapter 3. Reversibility and Entropy

W = Wgas + P0 (V0 V1 ) = N1T1CV 1

R/C P

) + P (V V )
0

N1T1CV
PV
PV
PV C
W
R/C P
=
1
+ 0 0 0 1 = 1 0 V
N0RT0
N0RT0
N0RT0 N0RT0
P0V0R

W
W
1/
=
=
+1 1/
RT0N0 P0V0
1

V
1 +1 1

V0

For air, =7/5=1.4, so that for = 0.1 we obtain W/(RTN0) = 0.574. A little less
work is required for adiabatic evacuation than for isothermal one. At the limit
0, this expression also gives W = P0V0, as for the isothermal case.
For adiabatic pressurization, the work per mole of compressed air is:
1/

W
W V0 W
W
11/ 1
1/
=
=
=
+11/
=
+ 1/ 1
RT N
1
RT N P V V1 RT0N0
1

! 0 1
! 0 1 1 0

For = 10, this expression gives W / (RTN1) = 1.520, a little more than in the iso-
thermal case.

3.6 Microscopic Origin of Entropy


By this point, we are familiar with the definition of entropy from a macroscopic
thermodynamic viewpoint as S = Qrev/T. But what is the origin of entropy in the
microscopic, molecular world? Qualitatively, we expect that irreversible processes
(which result in an increase in total entropy) also increase the degree of disorder
in a system. It turns out that a quantitative measure of disorder, specifically the
number of microscopic states available to a system at a given internal energy U
and for specified number of molecules (or moles) N and volume V, can be quantita-
tively linked to the entropy. A distinct microscopic state (microstate) is defined by
all microscopic degrees of freedom e.g. positions and velocities of molecules in a
gas. A set of microstates with specified common properties (e.g. number of parti-
cles N, volume V and energy U) defines an ensemble. The expression of S in terms of
microstates is provided by the famous 1872 entropy formula of Ludwig Boltz-
mann,

S = kB ln (N ,V ,U )
!

(3.17)

Boltzmanns entropy formula is the foundation of statistical mechanics, con-


necting macroscopic and microscopic points of view. It allows calculations of mac-
roscopic thermodynamic properties by determining properties of microscopic

3.6 Microscopic Origin of Entropy

57

configurations. The constant kB is called Boltzmanns constant; it is the ratio of the


ideal-gas constant to Avogadros number, or equivalently the gas constant on a per
molecule basis: kB = R/NA =1.3806510-23 J/K.
To illustrate the concept of counting microstates, we will use the simple sys-
tem shown in Fig. 3.13. In the general case, it consists of N slots, each containing a
ball that can be at energy levels 0, +1, +2, +3, , measured in units of kBT0, where
T0 is a reference temperature. The specific case of N = 10 and 3 energy levels is
shown in the figure. The concept of discrete energy levels arises very naturally in
quantum mechanics. For this simple system, we can count states by using the
combinatorial formula giving the number of ways we can pick M specific distin-
guishable objects out of N total objects:

N
N!
N (N 1)(N M +1)

(3.18)

= M!(N M)! =
12M

M
!

There is only 1 state with internal energy U = 0. States with U = 1 have one ball
at level +1 and the others at level 0, so for N = 10, there are 10 such states. States
with energy U = 2 may have 1 ball at +2 and the others at 0, or 2 balls at +1, so there
are:
10

+10 = 55

2
!
such states. We can similarly obtain (3 )= 220; (4 )= 715, (5 )= 2002, and so
on. Note that the number of microstates increases rapidly with the total energy of
the system. This is generally the case for most systems.
Now we are in a position to show that S defined microscopically from Eq. 3.17
has the two key properties associated with the entropy of classical thermodynam-
ics:
1.

S is extensive: For two independent subsystems, A and B,

S = k ln( A+B ) = kB ln( A B ) = kB ln A + kB lnB


! A+B B
The reason is that each microstate of system A can be combined with a
microstate of system B to give a microstate of the combined system. This is
clearly true for the simple system illustrated on the previous page. However,
when mixing two
gases or liquids, we
only get the above
expression if we
assume that the
particles in the

systems
are Figure 3.13 A system of 10 spheres with 3 energy levels.
indistinguishable. If

58

Chapter 3. Reversibility and Entropy

particles are distinguishable, additional states are available to the combined


system resulting from the possibility of exchanging the labels of particles.
Although the indistinguishability of particles is really of quantum mechanical
origin, it was introduced ad hoc by Gibbs before the development of
quantum mechanics, in order to make entropy an extensive property.
2.

S is maximized at equilibrium: For a system with internal constraints (e.g.


internal rigid walls or barriers to energy transfer), the number of possible
microstates is always smaller than the number of microstates after the
constraints are removed.

S (N,V, U) > S (N,V, U; internal constraints)


To demonstrate this second property, consider the box with particles of the
example above, and think of any constraint to the system at a given total
energy (say U=+2). An example of a "constraint" would be to have that the
first five slots have exactly 1 unit of energy. The number of microstates in
this case is (5x5=25), less than the 55 states available to the unconstrained
system.

One clear distinction between macroscopic and microscopic definitions of en-


tropy is that the former is physically meaningful only as a difference of entropy be-
tween specified states, while the latter appears to provide a measure of absolute
entropy. This apparent discrepancy results from the inability of classical physics to
define uniquely when two nearby states (e.g. positions of a particle in free space
differing by a fraction of a nm) are sufficiently different to justify distinguishing
them from each other. Quantum mechanical methods, on the other hand, provide
precise ways to count states.
At low temperatures, the number of microstates available to any physical sys-
tem decreases rapidly. At the limit of absolute zero temperature, T 0 , most sys-
tems adopt a unique ground state for which = 1 S = kBln = 0. This is the basis
of the Third Law of thermodynamics postulated by Nerst in the early 1900s. The
NIST Chemistry WebBook lists absolute entropies for pure components and chem-
ical elements in the thermochemistry data section. However, using entropy values
calculated with respect to an arbitrary reference state gives the same results as
absolute entropies for heat and work amounts.
Example 3.9 Entropy of a lattice chain
A common model for polymers is the Flory lattice model, which represents chains
as self-avoiding random walks on a lattice (grid). Self-avoiding means that two
beads cannot occupy the same position on the lattice. When two non-bonded
beads occupy adjacent positions, they have energy of interaction equal to kBT0,
where T0 is a reference temperature. Obtain the number of microstates for a
two-dimensional square-lattice chain of 5 beads, as a function of the energy U of
the chain.

3.6 Microscopic Origin of Entropy

59


Fig. 3.14 shows the number of configurations for a square-lattice chain of 5 beads.
Without loss of generality, we have fixed the configuration of the first two beads of
the chain to be in the horizontal direction, with the second bead to the right of the
first. This reduces the total number of configurations by a factor of 4; such a multi-
plicative factor simply shifts the value of S obtained from Eq. 3.17 by a constant
factor, akin to the reference state for the entropy. The last bond is shown in multi-
ple configurations (arrows), along with their number and energy: !2(1) for the
top left image means there are 2 configurations, each of energy 1.
Overall, counting configurations of the same energy:
(U=0) = 3+2+2+3+2+2+3=17 ; (U=1) = 2+1+1+1+1+2=8
The number of microscopic configurations and energy levels increases rapidly
with chain length. Theoretical and Monte Carlo computer simulation techniques
are used for determining the properties of models of this type for longer chains.















Figure 3.14 Configurations for a two-dimensional chain of 5 beads

A key issue in statistical mechanics is the frequency of occurrence of different


microstates when the overall constraints of a thermodynamic system are specified.
A basic postulate, comparable in importance to the postulate about the existence
of equilibrium states in classical thermodynamics discussed in 1.4 (p. 6), is that
all microstates of a system at a given U, N and V are equally probable. The conse-
quences of this postulate will be fully explored in Chapter 5.

60

Chapter 3. Reversibility and Entropy

The basic postulate of statistical mechanics implies that the probability of any
microstate , P , is the same as that of any other microstate in the constant N ,V ,U
ensemble:
1

P = at constant N, V and U
(3.19)

!
From this postulate, we can now simply derive another famous expression, the
Gibbs entropy formula, by substituting Eq. 3.19 into Eq. 3.17:

Gibbs
entropy
formula

S = kB
!

P lnP

(3.20)

all!micro+
states!

The Gibbs entropy formula can be shown to be valid even for systems not at
constant energy U, volume V, and number of particles N. This is in contrast to Eq.
3.17, which is only valid at constant for microstates at constant U V and N. For ex-
ample, in 5.6 we prove Eq. 3.20 for systems at constant N, V, and T.

You might also like