You are on page 1of 14

4.

0 The Second Law Of Thermodynamics


According to the first law of thermodynamics, when a closed system undergoes a
complete cycle, then the net heat supplied to the system is equal to the net work done
by the system, Q − W = 0. This is based on the conservation of energy and follows
from the observation of natural events. The second law of thermodynamics, which is
also a natural law, indicates that although the net heat supplied in a cycle is equal to
the net work done, the gross heat supplied must be greater than the net work done,
meaning that some heat must always be rejected by the system. We will consider a
heat engine [2, 3].

4.1 The heat engine


The heat engine is a system operating in a complete cycle and which converts heat
input into work output. The piston-cylinder heat engine is the most common heat
engine.
The system for the piston-cylinder engine will be the air/gas contained therein.
We will assume the gas is ideal (or perfect) such that it obeys the equation of state,

pv = RT (4.1)
Ru
R is the specific gas constant and is calculated, R = where Ru is the universal gas
M
3
constant (8.314 x 10 J/kmol K) and M is the molecular weight of the gas (which is 29
8.314 x 103
kg/kmol for air). For air, R = = 2.87 x 102 J/kgK.
29

Figure 4.1: Heat engine. The gas inside the cylinder is the system

We will consider the system to be closed (no mass transfer across the boundary).
Let a hot plate supply heat to the gas in such a way that all the heat goes into the gas

26
and none is absorbed by the cylinder wall - refer to Fig. 4.1 (a). The system undergoes
energy changes; the potential energy and kinetic energy changes are however assumed
small, compared to the internal energy change [2].
When the gas is heated, it will expand and push against the piston. Let the piston
initially be restrained so that it does not rise. This causes the pressure of the gas to
increase between 1 and 2, Fig. 4.2. As the pressure reaches point 2, the restraint
is removed and the piston will rise thus lifting the weight mg and doing work. As
the piston does work, the process could proceed in an infinite number of ways e.g.
at constant pressure, 2-3a, or it could be polytropic, pv n = constant, where n is a
polytropic index (process 2-3b). The particular path taken affects the efficiency of the
engine. Let us assume the path taken is 2-3b. At point 3b, the weight is taken off;

Figure 4.2: pv diagram for the heat engine

work has already been done in lifting the weight.


The piston will now be returned to its original state in two processes because the
pressure acting on it may be high: First, some heat is taken away from the gas through
a cold plate. This is at constant volume, process 3b-4 (corresponding to Fig. 4.1 c).
When the pressure has come sufficiently down, at point 4, the piston can now move
down against the gas in the process 4-1 thus taking the system back to its initial state

27
[2].
There is also an infinite number of possible paths for process 4-1, just like in the
process 2-3.

• The work done by the system during process 2-3b is given as,
 3b
W2−3b = m pdv. (4.2)
2

Note that because the system is closed, m is constant.

• Work done on the system in the process 4-1 is,


 1
W4−1 = m pdv (4.3)
4

In both cases, the relationship between p and v must be known for the respective
equations to be integrated. The essential point to note is that the work done is
proportional to the area under the curve on the p-v plot.

• The heat input into the system, is represented by process 1-2 and is calculated,

Q1−2 = mcv (T2 − T1 ) (4.4)

cv is the specific heat capacity of the gas at constant volume (note that the
heating occurs at constant volume).

• The thermal efficiency is defined as the ratio of the net work output to the heat
input into the system. In this case,
Wnet W2−3b − W4−1
ηth = = (4.5)
Qin Q1−2
A basic issue in heat engine design has always been: What is the best path the
processes should take, so as to produce the most efficient engine? This is a central
concern of the second law of thermodynamics.

It is not possible to have an efficiency of 100% since some of the heat input is always
dissipated in such processes as overcoming friction. Nevertheless, engineers have always
tried to modify designs e.g. changing pressure ratios, varying the load on the piston
etc, to improve efficiency. Two of the most common cycles are:

1. The Otto cycle: this describes the typical automobile internal combustion (spark-
ignition) engine

2. The Diesel cycle: typical for large vehicles such as trucks and buses that use the
compression-ignition engines

28
4.2 Reversibility
The heat engine already described does not have a thermal efficiency of 100%. For
instance, some of the heat input is used to overcome friction. Such factors lead to
process irreversibilities. One way of simulating reversibility is by pulling the piston a
small amount to the right so slowly that we do not cause any large-scale motion of the
gas, Fig. 4.3 (a). The process can then be reversed, i.e., the piston may be pushed
back to its original position so that everything returns to the same state it started
from. Such an ideal work transfer is regarded a reversible work interaction, where all
properties are restored to their initial values after the pull-push motion is complete. In
this process, friction is assumed absent between the cylinder and the piston (friction
causes a process to be irreversible).

Figure 4.3: (a)Reversible and (b)Irreversible work transfer

In reality, all processes are irreversible, but the concept of reversibility is central to
thermodynamic analysis. Reversibility implies that everything is infinitesimally close
to equilibrium at all times throughout the process.
As far as heat interaction is concerned, if the heat interaction occurs at virtually
constant temperature, with only an infinitesimal difference to provide the heat transfer
in the required direction, the heat interaction is reversible. In Fig. 4.4 (a), if we heat
the object a little and cool it a little, the heat will flow first in one direction and then
in the other direction in a reversible way. In other words, to attain reversible heat
transfer, we decrease the temperature difference until in the limit, it is almost zero
(infinitesimally small difference). Heat transfer across a finite temperature difference
as in Fig. 4.4 (b) is irreversible.
In summary, the concept of reversibility means that the system and its surroundings
are always infinitesimally close to equilibrium [2]. A reversible process then is any
process performed so that the system and all its surroundings can be restored to their
initial states by performing the process in reverse [6].

4.3 Deriving the second law of thermodynamics


The essence of a heat engine is that it has a heat intake from a high temperature
source, and a heat rejection to a low-temperature sink. We will call these the high and

29
Figure 4.4: (a)Reversible and (b)Irreversible heat transfer

low temperature reservoirs respectively [2].

Figure 4.5: (a)Schematic diagram of a heat engine. (b)A reversible heat engine (c)A
heat pump - a heat engine working as a heat pump or refrigerator

Consider Fig. 4.5 (a) showing heat input from a high temperature reservoir, heat
rejection to a low temperature reservoir and net work done on surroundings. If the

30
cycle were reversible, all the arrows of Fig. 4.5 (a) can be reversed, as shown in Fig.
4.5 (b). A heat engine working in reverse is called a heat pump, requiring work input
to transfer heat from a low temperature to a high temperature reservoir. If a heat
pump is irreversible, it would be represented by Fig. 4.5 (c).

Figure 4.6: (a)A reversible heat engine (ER ) and an irreversible heat engine (E) both
operating at the same temperature difference, and both extracting the same amount of
heat from TH (b)The reversible engine is now reversed (c)The high temperature source
TH is unnecessary, and net work is produced by extracting heat from a low-temperature
reservoir only.

Now consider the following steps:

• Let us place a reversible engine ER and an irreversible one E side by side, to


operate between the same hot and cold plates (TH and TC respectively) as shown
in Fig. 4.6 (a). The engines do work WR and W respectively, while extracting
the same amount of heat, QH .

• Let us reverse ER (Fig. 4.6 (b)) and use symbol ∃R for it. The engine now
transfers heat QH to TH while an amount of work WR is done on it. Since the
heat QH rejected by ∃R to TH is the same amount extracted by E from TH , then
the reservoir TH becomes redundant. Again, if we assume W is greater than WR ,

31
then E can run ∃R and still have some work W − WR left over to do external
work. This situation is represented by Fig. 4.6 (c).

Such a system, though not violating the first law of thermodynamics, still violates
natural expectations: It does not require a temperature difference and so requires no
fuel to provide a high-temperature source. If it could exist, all energy and environmen-
tal problems would vanish! The first law is obeyed thus,

Q − W = 0, i.e., (QR − Q) − (W − WR ) = 0 (4.6)

For each of the engines, we would have:

• For the reversible engine ∃R ,

(QR − QH ) + WR = 0 (4.7)

• For the irreversible engine E,

(QH − Q) − W = 0 (4.8)

Adding Eqs. (4.7) and (4.8) gives Eq. (4.6). From Eq. (4.6), QR − Q = W − WR ,
and because we have assumed W > WR , then QR > Q.
The flaw is in assuming W > WR , i.e., assuming that an irreversible engine can
produce more work than the reversible engine operating between the same hot and
cold reservoirs (Fig. 4.6 (a)). If we assume W < WR , then it is fine, but it would
similarly mean that QR < Q. This would be a useless friction machine, that simply
converts work into heat.
The second law must then be stated: It is impossible for any system to
operate in a thermodynamic cycle and do net work on its surroundings,
while exchanging heat with a single reservoir [2].
There are many statements and corollaries of the second law that can be found in
thermodynamics literature. For example, two other statements are:

• Clausius statement: It is impossible to construct a device which operates in


a cycle and whose sole effect is to transfer heat from a cooler body to a hotter
body [6]

• Kelvin-Planck statement: It is impossible to construct a device which operates


in a cycle and produces no other effect than the production of work and exchange
of heat with a single reservoir [6]

32
4.4 Clausius inequality
The second law of thermodynamics often leads to expressions that involve inequalities.
An actual heat engine, for example, is less efficient than a reversible one operating
between the same two thermal reservoirs. An important inequality with major con-
sequences in thermodynamics is the Clausius inequality. This was first stated by a
German physicist called Clausius. It is expressed

δQ
≤ 0, (4.9)
T
δQ
that is, the cyclic integral of is always less than or equal to zero. This inequality is
T 
valid for all cycles whether reversible or irreversible. The symbol is used to indicate
that the integration is to be performed over the entire cycle.
Note that the equality in Eq. (4.9) holds for reversible cycles and the inequality for
the irreversible ones, i.e.,
  
δQ
= 0, (4.10)
T rev
  
δQ
< 0. (4.11)
T irrev

4.5 Entropy
4.5.1 Definition
It is worthwhile to note that the cyclic integral of a thermodynamic property is zero.

δQ
Looking again at Eq. (4.10), it can then be concluded that the quantity
T rev
represents a property. Based on this, Clausius realized in 1865 that he had discovered
a new thermodynamic property, which he named entropy. We can therefore infer
that the second law of thermodynamics gives rise to a property called entropy. Being
a thermodynamic property, entropy does not depend on the path taken. It is denoted
by S when we are referring to the total entropy, which is an extensive property (units:
kJ/K). Entropy per unit mass, also called specific entropy and denoted by s, is an
intensive property (units: kJ/kg K). Nevertheless, the term entropy is normally used
to refer to either of the two, and the context is what usually clarifies which of the two
is being referred to.
In simple terms, entropy is defined as a measure of irreversibility. In other appli-
cations, entropy is regarded as a measure of randomness. The defining equation for
entropy associated with a reversible process is
 
δQ
dS = . (4.12)
T rev

33
This applies to all working substances [3]. The change of entropy of a system during
a process can be determined by integrating Eq. (4.12) between the initial and final
states, thus
 2 
δQ
S = S2 − S1 = . (4.13)
1 T rev

For that reason, it is important to note that we have actually defined change of entropy
instead of the entropy itself. Absolute values of entropy are determined on the basis
of the third law of thermodynamics, which is discussed in advanced topics in thermo-
dynamics. In any case, Engineers are usually concerned with changes in entropy.

4.5.2 Principle of increase of entropy


Consider a cycle made up of two processes: Process 1-2 is arbitrary (i.e., can be
reversible or irreversible), while process 2-1 is reversible. From the Clausius inequality

Figure 4.7: Cycle composed of reversible and irreversible processes

(Eq. (4.9)), we can write the inequality for the cycle 1-2-1 thus,
 2  1 
δQ δQ
+ ≤ 0. (4.14)
T T rev
 
1
 2

A B

The second integral (B in Eq. (4.14)) when compared with Eq. (4.13), is recognized
as the entropy change in the process 2-1, i.e., S1 − S2 . We may therefore rewrite Eq.
(4.14) thus,
 2
δQ
+ S1 − S2 ≤ 0, (4.15)
1 T
or
 2
δQ
S2 − S 1 ≥ . (4.16)
1 T

34
 2 S2 − S1 = S represents the entropy change of the system while the
The quantity
δQ
quantity represents the entropy transfer. For a reversible process, the two
1 T
quantities are equal. For an irreversible process, the entropy change is larger than the
entropy transfer, i.e.
 2
δQ
S2 − S1 > . (4.17)
1 T
This simply means that some entropy is generated (or created) during an irreversible
process, and this generation is entirely due to the presence of irreversibilities. If we
denote the entropy generated as Sgen , we may convert the inequality of Eq. (4.17) to
an equation thus
 2
δQ
S = S2 − S1 = + Sgen . (4.18)
1 T
Sgen is always a positive quantity or zero: It is zero in the absence of irreversibilities
and positive for actual processes, which are always accompanied by irreversibilities.
We may then conclude that the entropy of a system during a process always increases,
or in the limiting case of a reversible process, remains constant. It never decreases.
This is known as the increase of entropy principle.
More often than not, this principle is deemed to apply to an isolated system, i.e.,
a system across whose boundary there is no heat, work or mass transfer. Ordinarily,
such an isolated system is considered to be constituted by the system itself and its
surroundings since both can be enclosed by a sufficiently large arbitrary boundary.
The entropy change of the isolated system during a process is then the sum of the
entropy changes of the system and its surroundings (which is equal to the entropy
generation since an isolated system involves no entropy transfer), thus

Sisolated = Ssys + Ssurr ≥ 0. (4.19)

Equation (4.19) is often used as the mathematical statement of the second law.
The performance of engineering systems is degraded by the presence of irreversibil-
ities, and entropy generation is then used to give a measure of the magnitude of the
irreversibilities present during the process. The greater the extent of irreversibilities,
the greater the entropy generation. Entropy generation is normally used to establish
criteria for the performance of engineering devices.

4.6 The Carnot cycle


It can be shown from the second law that no heat engine can be more efficient than
a reversible heat engine working between the same temperature limits. Sadi Carnot,

35
a French engineer, showed that the most efficient possible cycle is one in which all
the heat supplied is supplied at a fixed upper temperature, and all the heat that is
rejected is rejected at a lower fixed temperature. The complete cycle consists of two
isothermal processes and two adiabatic processes. Since all the processes are reversible,
the adiabatic processes in the cycle are termed isentropic (reversible and adiabatic).
The processes are,
1-2: isothermal heat supply
2-3: isentropic expansion
3-4: isothermal heat rejection
4-1: isentropic compression
The Carnot engine and p-v cycle are shown in Fig. 4.8, while the T-s diagram is
shown on Fig. 4.9.

Figure 4.8: The Carnot engine

The thermal efficiency of the cycle is,


Wnet Wnet
ηth = = (4.20)
Qin QH
From the first law of thermodynamics for a cycle, Q - W = 0, i.e.

(QH − QC ) − (Wout − Win ) = 0


QH − QC − Wnet = 0
QH − QC = Wnet (= Qnet ) (4.21)

Therefore,
QH − QC Qnet QC
ηth = = =1− , (4.22)
QH QH QH
but
QC TC (s3 − s4 )
= but (s2 − s1 ) = (s3 − s4 ), (4.23)
QH TH (s2 − s1 )

36
Figure 4.9: Carnot T-s cycle

therefore,
TC
ηth = 1 − (4.24)
TH
The Carnot cycle efficiency is the highest that can be achieved between any two fixed
temperatures. All others must be less than it.
EXAMPLE 6
What is the highest possible theoretical efficiency of a heat engine operating with a
hot reservoir of furnace gases at 20000 C when the cooling water available is at 100 C?
[87.54%]
Tutorial Problem 4
A Carnot engine operates between a source temperature of 7000 C and a sink temper-
ature of 200 C. Assuming that the engine will have a net output of 65 hp, determine
the thermal efficiency of the engine, the heat supplied and the heat rejected. [69.9%;
69.37 kJ/s; 20.88 kJ/s]
To increase the Carnot efficiency, the temperature difference TH − TC can be in-
creased but there are limitations e.g., for a fixed TC , TH can only be raised to as high
a value as the metallurgical capabilities of the material components of the engine.
Again, in practice, it is difficult to devise a system which can receive heat at a
constant temperature, and reject heat at a constant temperature. A wet vapour is
the only working substance which can do this conveniently, since for a wet vapour the
pressure and temperature remain constant as the latent heat is supplied or rejected.
Such a cycle is shown in Fig. 4.10 [4].

37
Figure 4.10: A T-s diagram for a Carnot cycle using steam

4.7 Absolute temperature scale


It is possible to establish a temperature scale that is independent of the working fluid.
Recall from the efficiency of a Carnot cycle (Eq. (4.22)) that
QH − QC Qnet
ηth = =
QH QH
The efficiency of an engine operating on the Carnot cycle depends only on the temper-
atures of the hot and cold reservoirs. Let us denote temperature on an arbitrary scale
by X such that

ηth = φ(XH , XC ) (4.25)

where φ is a function while XH and XC are the temperatures of the hot and cold
reservoirs respectively.
Combining Eqs. (4.22) and (4.25),
Qnet
= φ(XH , XC ) (4.26)
QH
The function φ cannot be determined analytically, for it is entirely arbitrary and many
temperature functions can satisfy it [6]. Kelvin proposed that the temperature function
may be chosen thus,
Qnet XC
1− = = 1 − ηth
QH XH
XC
∴ ηth = 1 − (4.27)
XH

38
From Eq. (4.24), we have

TC
ηth = 1 −
TH
or,
Qnet TC
1− = (4.28)
QH TH

Comparing Eq. (4.27) and (4.28), it can be seen that the temperature X is equivalent
to the temperature T . Thus, by suitably choosing the function φ, the ideal temperature
scale is made equivalent to the temperature scale based on the perfect gas. Such a scale
is the absolute thermodynamic temperature scale and does not depend on the working
substance [3, 6].

39

You might also like