Sie sind auf Seite 1von 19

CHAPTER 3

HEAT ENGINES
AND
THE SECOND LAW OF THERMODYNAMICS
INTRODUCTION
In the last chapter we utilized an atomic kinetic model to develop an equation for the internal energy of
an ideal polyatomic gas and to develop the equation of state of such an ideal gas (the ideal gas law). We then
discussed the different modes of energy transfer across the boundary of a system, in particular heat and work.
Utilizing these concepts and the basic assumption of conservation of energy, we developed the first law of
thermodynamics. Using the first law, we were able to derive some useful relationships between the various heat
capacities of ideal gases.
In this chapter we wish to investigate the second law of thermodynamics: heat is always observed to
flow from a hotter system to a colder system, never the other way around. Likewise, if we use a paddle wheel to
stir a container filled with water, we find that the water will heat up, but heating up the water will never cause the
paddle wheel to rotate. This later example would seem to indicate that the kinetic energy associated with the
organized motion of the paddle wheel can be changed into a random form of internal energy characterized by
molecular motion, whereas the random form of internal energy can never be reorganized to produce the
organized motion of the paddle wheel. The connection of the second law to the microscopic motion of particles is
a fundamental one. However, the early study of thermodynamics was undertaken before the atomic theory was
well established. Early workers in this field found that a macroscopic property called entropy could be used to
predict the direction in which energy would flow in a given process. Although the second law in some way
determines the direction of energy transfer, the first law puts some limits on the amount of energy that can be
exchanged in a certain process. A real process cannot occur, however, unless both the first and second laws are
satisfied.
As mentioned earlier, much of the present theory of thermodynamics was developed during the age of
industrialization in an attempt to develop more efficient engines. In this chapter we will discuss heat engines,
refrigerators, and heat pumps, and the techniques for calculating the efficiency of these devices. We will find
that the first and second laws taken together place some limitations upon the efficiencies of these devices.
HEAT ENGINES
As mentioned in the introduction, a paddle wheel in a water bath can be used to convert organized
mechanical motion into heat. In fact all the mechanical motion of the paddle wheel can be converted into heat. It
is not such an easy task, however, to convert heat energy into mechanical energy (organized energy which can be
used to do work). A device which accomplishes this task is called a heat engine.
In the last chapter, we actually introduced a process which acts like a heat engine. Let's look again at
this process (Fig. 3.1). Process E p F is an isobaric expansion, where the work done by the gas is given by
[EF TE ZF  ZE 

(3.1)

During this process, the temperature of the gas also increases (as can be seen from the isothermal curves),
increasing the internal energy of the gas. We can find the heat added to the system during this process using the
first law
UEF ?IEF [EF

(3.2)

Since the work done by the gas and the change in internal energy of the gas are both positive, the heat added to
the system is positive, i.e., UEF !. Process F p G is isochoric, with no increase in the volume of the system.
In this process, there is no work done by (or on) the gas. However, the internal energy decreases due to the drop
in temperature of the gas. Thus, in process F p G heat is removed from the system, so that UFG  !. Process G
p E is an isothermal compression, where work is done on the system as the volume decreases. But, since this
process is isothermal, the internal energy remains constant. Thus, the heat added to the system in this process
must be equal to the work done by the system. Since the work done by the system is negative, heat must be
removed from the system during this process, so we find that UGE  !.

Chapter 3: The Second Law of Thermodynamics

P
TB > TA
A

PA

TB
C

PB

TA

VA

VB

Fig. 3.1 A T -Z diagram for an ideal gas as it is carried through the cyclic process EpFpG pE. Process
EpF is isobaric, process FpG is isochoric, and process G pE is isothermal. The work done in each
process is the area under the T -Z curve. The work is positive if the volume increases and negative if the
volume decreases during a process (it is zero if the volume does not change). When this cycle is
completed, the positive work done by the gas is greater than the negative work done on the gas, so that
the net work done in the complete cycle is positive. This process, then, acts like a heat engine.

The work done in the process G p E is negative, since the volume decreases, but is still just the area
under the T -Z curve for that process. Obviously, when this cycle is completed, the positive work done by the gas
in process E p F is greater than the negative work done on the gas in process G p E, so that the net work done
in the complete cycle is positive. This cyclic process, therefore, provides us a mechanism whereby we can supply
an amount of heat energy to a gas and accomplish a net positive mechanical work. Any cyclic process which
absorbs heat energy and accomplishes a net positive mechanical work is known as a heat engine.
But you will notice that we were not able to convert all the heat energy supplied to the system into useful
mechanical work - some heat was removed from the system. Heat is added to the system during process E p F,
but removed from the system during processes F p G and G p E. Since the total change in the internal energy
of the system must be zero in this cyclic process, the net positive work done by the system must be equal to the
net positive heat added to the system. This means that |Uin | |Uout |, or lUEF l lUFG l lUGE l, where |Uin | and
|Uout | represent the magnitude of the heat flow into and out of the system.
As you might imagine, there are a large number of different processes which could be used as heat
engines. It would obviously be desirable to devise a system whereby all the heat energy supplied to the system
could be used to do useful mechanical work. In fact, one of the major tasks of classical thermodynamics was to
determine which processes were the most efficient. In order to determine this, we need to examine some of the
more fundamental aspects of heat engines.
No matter what the actual process, we find that the following characteristics are true of all heat engines:
1.
2.
3.
4.

They receive energy from one or more high-temperature sources (such as solar heat, burning coal,
or a nuclear reaction).
They convert part of this energy into useful (organized) mechanical work (or energy).
The remaining energy is released to one or more low-temperature sinks (such as a river or cooling
tower).
They all work in a cycle.

A schematic of a simple heat engine with only one source and one sink is shown in Fig. 3.2. The best practical
example of a heat engine is a steam power plant (see Fig. 3.3).

Chapter 3: The Second Law of Thermodynamics

High-Temperature
Source
Qin
Worknet,out

Heat
Engine

Qout
Low-Temperature
Sink

Fig. 3.2 A schematic of a ideal heat engine. Energy in the form of heat is added to the heat engine from a
high-temperature source. Part of this added energy is then converted into useful mechanical work, while the
remaining amount is dumped into a low-temperature sink. The circle which designates the actual heat engine
represents the fact that the engine operates in a continual cycle.

Energy Source
(coal fire or nuclear reactor)
System Boundary

Qin
Boiler

Wout

Win
Turbine

Pump

Condenser

Qout
Energy Sink
(river or cooling tower)

Fig. 3.3 Schematic of a steam power plant. The net work out is the shaft work done by the turbine less the
work necessary to drive the pump. The net heat added to the system is the heat added at the high temperature
source less the heat removed at the low temperature sink. This might represent either a coal-fired steam plant
or a nuclear steam plant. The only real difference is the operating temperatures. The maximum allowed
temperatures for coal-fired steam plants is approximately 500 C, whereas the maximum allowed
temperatures for nuclear power plants is approximately 300 C.

Chapter 3: The Second Law of Thermodynamics

A steam power plant is an external combusion engine, as opposed to the internal combution gasoline
engine. In this external combustion engine, the fuel is burned outside the engine and the heat of this combustion
is transfered to the boiler in the form of heat. In the boiler, the working fluid (water and steam) is heated to
produce high temperature steam that is directed by a nozzle onto a turbine. This turbine produces shaft work
which turns a generator and produces electricity. The steam from the turbine is then cooled in a condensing unit
and then pumped back into the boiler, where it is again heated.
You might wonder why one would want a condenser unit since some of the energy is lost from the
system at this point. If there were no condenser unit to cool the steam, the temperature of the steam leaving the
turbine would rise until it becomes nearly equal to the temperature of the entering steam. This would produce a
large back-pressure and reduce the difference in pressure across the turbine. If this occurred the efficienty of the
nozzle-turbine subsystem would drop.
Although each component of the steam power plant is actually an open system, with mass flowing into
and out of it, the system as a whole is a closed system, with a constant amount of steam contained within the
subsystems and the connecting pipes. If we take the whole system as our thermodynamic system, we are dealing
with a closed system for which the total internal energy change in a complete cycle is zero! This means,
according to the first law,

* .I * $ U  * $ [ !

(3.3)

or

* $U * $[

|U38 |  |U9?> | |[9?> |  |[38 | |U8/>38 | |[8/> 9?> |

(3.4)

where, again, the symbols |Uin | and |Uout |, as well as |[38 | and |[9?> | represent magnitudes. This equation simply
states that the net work done by the system is equal to the net heat added to the system. That is, the heat added to
the system at the high-temperature source less the heat given up at the low-temperature sink is equal to the shaft
work done by the turbine less the shaft work done on the pump.
The amount of heat that is removed from the system at the low-temperature sink represents the amount of
heat wasted in order to complete the cycle. For those situations where there is wasted heat, the net work done
by the system is obviously less than the energy put into the system at the high-temperature source. Thus, only part
of the heat transfered to the system is converted to useful work. We define the fraction of the heat energy which
is converted into useful work in a heat engine as the thermal efficiency ( of the heat engine, or
(

R /> [ 9<5 S?>


|[8/>9?> |

L/+> M 8
|U38 |

(3.5)

Notice that this definition can also be expressed in the form


(

H/=3</. S?>:?>
V/;?3</. M 8:?>

(3.6)

In our discussion so far, we have implied that the temperature of the high-temperature source and the
low-temperature sink do not change during the operation of the system. At first thought, this might seem difficult
to accomplish in practice. However, a large enough body of water (say the ocean) can be treated as a constanttemperature source or sink, since the energy removed from, or added to, the ocean would not appreciably increase
or decrease the average temperature of the ocean. The same is true of the atmosphere, although local heating or
cooling would occur. In practice, we don't really need an ocean to maintain a reasonably constant temperature
source. All we need is a body which has a large enough thermal mass (the product of mass and specific heat
capacity) that the temperature of the body can be maintained at an approximately constant temperature as heat
energy is added to or removed from the body, or a thermistatically controlled heater. Such bodies are called heat
reservoirs. Thus, we will often assume that the sources and sinks of our heat engine are heat reservoirs which are
maintained at constant temperature.
We will designate the temperature of the high-temperature reservoir as XL and the temperature of the
low-temperature reservoir as XP . Likewise, we will use |UL | to designate the magnitude of the heat added to or
removed from the system at temperature XL and |UP | as the magnitude of the heat added to or removed from the

Chapter 3: The Second Law of Thermodynamics

system at temperature XP . Using this notation, we can express the net work done by a heat engine as it moves
through a complete cycle as
|[8/> 9?> | |UL |  |UP |

(3.7)

and the thermal efficiency as


(

|[8/> 9?> |
|UL |  |UP |
|UP |

"
|U L |
|UL |
|UL |

(3.8)

This equation for the thermal efficiency of a heat engine is very interesting. It tells us that the maximum
thermal efficiency will occur when |UP | p ! or when |UL | p . To obtain the latter condition might be
somewhat difficult, but perhaps we might be able to achieve the former, or at least approach it. In the next section
we will consider whether or not this is possible.
THE NECESSITY OF EXHAUST HEAT AND THE KELVIN-PLANCK STATEMENT OF THE
SECOND LAW
Consider an ideal gas confined in a cylinder with a movable lid (see Fig. 3.4). The system is
originally at an equilibrium temperature of 30 C. At this point, the lid of the system is at the lower stops, and we
place a load on the lid. If we now take the cylinder and put it in contact with a heat reservoir at 100 C, heat will
enter the system and the pressure will increase. As the volume of the gas increases, the system does work lifting
the lid and load. As long as the process proceeds slowly enough that it can be considered quasi-static, the
pressure of the gas inside will remain constant (equal to the outside pressure, and the pressure due to the lid and
load). Under these circumstances, the system is always in a state of thermodynamic equilibrium and the ideal gas
law requires that the temperature of the gas must rise. Thus, the energy added to the gas while it is in contact with
the heat reservoir is equal to the work done as the gas expands plus the increase in internal energy of the gas as
the temperature rises. We wait until the lid reaches the upper stops, and find that the gas temperature is 90 C. At
this point, we remove the cylinder from the heat reservoir. At this point we also remove the load, having
accomplished the work objective. Let's assume that 100 Joules of energy is added to the system from the hightemperature reservoir during this process, and that 25 Joules were required to lift the lid and load to the upper
stops, while the internal energy has been increased by 75 Joules.

Load
Load

Gas
o
30 C

Gas
o
30 C

Gas
o
90 C

Gas
o
30 C

Qin

Qout

20 C

100 C

Fig 3.4 A gas is contained in a cylinder with a movable lid. The lid rests on the lower stops when the gas is
at 30 C. A load is placed on the lid and the cylinder is placed in contact with a high-temperature reservoir.
Heat flows into the cylinder causing the pressure to rise and lift the lid. The internal energy also increases.
When the lid reaches the upper stops, the gas has reached 90 C and is removed from the high-temperature
reservoir. The load is now removed (useful work has been accomplished in lifting this load). We now wish
to return the system to its original state so that the lid rests on the lower stops and we can again place a load
on the lid to be lifted. To accomplish this, we must remove some of the internal energy from the gas. The
only way this can be done is to dump the heat into a low-temperature reservoir. This amounts to wasting
heat in this cycle.

For this system to act as a cyclic heat engine, we must restore the system to its original configuration in
order to repeat the cycle. To accomplish this, we must lower the temperature and pressure of the system (remove

Chapter 3: The Second Law of Thermodynamics

the 75 Joules of extra internal energy from this system now as 90 C) and restore the lid to its original position.
If we could remove the heat from our system and somehow put it back into the high-temperature reservoir, we
would be saving the heat energy we now have in our system. The second law, however, reminds us that it is
impossible to remove heat from a 90 C system and deposit that heat into a 100 C system. In order to remove the
necessary heat from our gas and return it to its original state (original temperature), we will need to bring it into
contact with a low-temperature reservoir which has a temperature less than or equal to 30 C. The heat removed
from our system in this process is exhaust heat and is essentially wasted.
Or is it? Can't we use this heat to do more work? After all, you have just added some extra energy to
the lower temperature reservoir. But that is just the point! In the process of moving through the cycle, we have
removed heat from a higher temperature reservoir and deposited heat to a lower temperature reservoir. To utilize
the excess heat deposited in the lower temperature (30C) reservoir, we would have to utilize this reservoir as a
high-temperature reservoir is some other process. This you might be able to do - you might even be able to
continue the process, transferring exhausted heat to a lower temperature reservoir further and further down the
temperature ladder. However, you will eventually reach a limit beyond which you cannot practically go. Thus,
we conclude that it is impossible to devise a system operating in a cycle that doesn't exhaust heat. This is the
foundation of the Kelvin-Planck statement of the second law. This statement of the second law can be expressed
as follows:
It is impossible for a heat engine, which operates in a cycle, to remove an amount of heat energy
from a single heat reservoir and to accomplish that same amount of mechanical work - there must
be some exhaust heat. Another way of saying this is that no heat engine can be 100% efficient.

REFRIGERATORS AND HEAT PUMPS


If you look back at the process described in Fig. 3.1, you will notice that this heat engine passes from
state E to F to G and back to E along a clockwise path enclosing a specific area on the T -Z graph. Since the
area under a T -Z curve is the work done by the system, it should be obvious that the net work done by the system
in the cylic process is just the area enclosed by this loop. Furthermore, it should be obvious that the enclosed
area corresponds to work done by the system if it is traversed in a clockwise direction, and corresponds to work
done on the system if it is traversed in a counter-clockwised direction.
If the system is operated in a counter-clockwise direction work is done on the system by the surroundings
and there is a net amount of heat removed from the system, i.e., more heat is removed from the system in the form
of heat than is added to the system in the form of heat. Schematically, we can represent a simple form of this
process by the drawing in Fig. 3.5

High-Temperature Reservoir

QH

Refrigerator

Worknet,in
QL
Low-Temperature Reservoir

Fig. 3.5 A schematic representation of a cyclic process working as a refrigerator.

Chapter 3: The Second Law of Thermodynamics

From Fig. 3.5 it should be clear that the net effect of this refrigeration process is to remove an amount
of heat from a low-temperature reservoir and deposit a somewhat larger amount of heat at a high-temperature
reservoir. This process is driven by the mechanical work done on the system. From the first law principle of
conservation of energy and the knowledge that the internal energy cannot change in a cyclic process we can see
from the schematic that |UP | |[8/>38 | |UL | (remember that these quantities all express magnitudes).
The amount of heat |UP | which can be removed from the low temperature reservoir in a cycle is
obviously related to the amount of mechanical work done on the system. Now if we are interested in operating a
refrigerator to keep something cool, the desired output of our device is the amount of heat that can be removed
from the low-temperature reservoir. The required input for this system is the mechanical work which must be
done on the system. We define the coefficient of performance of a refrigerator in terms of our previous definition
of thermal efficiency, i.e.,
GST</0

H/=3</. S?>:?>
V/;?3</. M 8:?>

(3.9)

or
GST</0

|UP |
|UP |

|[8/>38 |
|UL |  |UP |

"
"

(3.10)

|UL |
|UP |

Typically |UP | is larger that |[8/>38 |, so that the coefficient of performance of a refrigerator is greater than one.
(That is the reason this quantity is not called an efficiency, but a coefficient of performance - efficiencies are
typically less than one.) This means that the energy required to extract a certain amount of heat energy from the
low-temperature reservoir is less than the amount of heat extracted.
Energy Sink
(kitchen room air or heated space)
System Boundary

QH
Condenser
800 kPa
30 C
Expansion
Valve
120 kPa
-25 C

800 kPa
60 C

Wnet,in

Compressor

Evaporator

120 kPa
-20 C

QL
Cool Energy Source
(icebox or outdoors)

Fig. 3.6 Schematic of a typical refrigeration system indicating typcial operating conditions.

Notice also that the amount of heat added to the high-temperature reservoir must be larger than the
mechanical work done on the system. This means that we could use this refrigeration process to efficiently add
heat to the high-temperature reservoir. This is the principle behind the heat pump. A heat pump is simply a
refrigeration cycle in which the exhausted heat is what is utilized rather than the removal of heat from the lowtemperature reservior. As an example, if you take a conventional air-conditioning unit and mount it in the
window in reverse you will heat up your house and cool down the exterior. But the amount of heat supplied to
the house will typically be greater than the amount of mechanical energy required to run the air-conditioner! The

Chapter 3: The Second Law of Thermodynamics

coefficient of preformance for a heat pump is defined as


GSTL/+> T ?7:

H/=3</. S?>:?>
|UL |
|UL |
"

UP |
|[8/>38 |
|UL |  |UP |
V/;?3</ M 8:?>
"  ||U
L|

(3.11)

A typical refrigeration system is presented schematically in Fig. 3.6. Here the individual subsystems are
illustrated along with typical operating conditions. The refrigerant enters the compressor in the form of a vapor
and is compressed to a relatively high pressure and temperature. This heated refrigerant then enters a condenser
where it looses some of its heat energy to the surroundings. This cooled refrigerent then passes through a
throttling (or expansion) valve allowing the gas to expand adiabatically - thus cooling significantly. This cooled
gas then enters an evaporator where it absorbs heat from the space being cooled. The refrigerant is then drawn
back into the compressor and the cycle begins again.
As in the case of the steam power plant, each subsystem is actually an open system, but considered as a
whole, we can treat the entire system as a closed system and apply the first law conservation of energy principle.
It is this principle that Fig. 3.5 is based upon and upon which the definitions for the coefficient of performance are
based.
CLAUSIUS' STATEMENT OF THE SECOND LAW
The simplest statement of the second law of thermodynamics is to say that it is impossible for heat to
flow spontaneously from a cooler body to a warmer body. This is something we never observe. When an ice
cube is placed in a glass of water, the heat energy from the hotter water enters the ice and melts it - heat is never
extracted from the ice to heat up the hotter water. The refrigeration cycle that we mentioned above does not
violate this principle, even though heat is removed from a colder object and deposited into a hotter object. This is
because the cycle requires an input of mechanical work for the process to operate.
However, since more energy is added to the high-temperature reservoir than is added to the system in the
form of mechanical work, we might try to use some of this additional heat energy to actualy run the refrigerator.
One way to do this might be to operate a heat engine between the same two heat reservoirs and let the mechanical
output of the heat engine drive the refrigerator. If we could do something like that we might really have
something!
To test this idea, lets assume that the heat engine we operate between these two reservoirs has an
efficiency of 100%, i.e., no heat is exhausted to the low-temperature reservoir. This is clearly in violation of the
Kelvin-Planck statement of the second law. However, if this could be accomplished, then the refrigerator cycle
we have developed could be driven by this special heat engine and effectively remove heat from the lowtemperature reservoir without any additional input of mechanical energy (see Fig. 3.7). This leads us to another
/;?3@+6/8> statement of the second law, the Claussius statement: No device can be constructed which operates
in a cycle and produces no net effect other than the transfer of heat from a low-temperature reservoir to a hightemperature reservoir.
High-Temperature Reservoir

QH

Refrigerator

100% Efficient
Heat Engine

QL
Low-Temperature Reservoir

Fig. 3.7 A cyclic refrigerator driven by a 100% efficient heat engine effectively transfers heat from a lowtemperature reservoir to a high-temperature reservoir with no other net effect - a clear violation of the second
law.

Chapter 3: The Second Law of Thermodynamics

You can see that this system would effectively act as if you could extract heat from a cold object, and deposit that
same amount of heat at a higher temperature, violating the simple statement of the second law.
THE CARNOT CYCLE
To try and gain even better insight into the concepts of heat engines and refrigerators, as well as the
implications of the second law, we will now consider a very special cyclic process that was first proposed by Sadi
Carnot, a French engineer, in 1824. The Carnot cycle consists of an ideal, monatomic gas carried through four
reversible processes as shown in Fig. 3.8: 1) an isothermal expansion, 2) an adiabatic expansion, 3) an isothermal
contraction, and 4) an adiabatic contraction which returns the system to the original state. In this section we will
examine each of the four segments of this cycle and calculate, for each segment, the net change in the internal
energy of the gas, the work done, and the heat added to the system. We begin the cycle at state E where the
pressure, volume and temperature are designated TE , ZE , and XE , respectively.

Pressure

Qin
B

TH

D
Qout

TL

Volume
Fig. 3.8 T -Z diagram of a Carnot Cycle.
E F: First we carry out an isothermal expansion from state E to state F. During this process, the
temperature of the system remains the same (i.e., TF XE XL ) so that the change in internal energy of the gas
for this process is given by
?I

$
8VXF  XE !
#

(3.12)

Now from the first law, the work done by the system in expanding against its surroundings can only be
accomplished as a result of heat being added to the system during this process. Thus, we have
?IEF UEF  [EF 0
UEF [EF

(3.13)

and since positive work is done by the system in this process, the heat UEF added to the system must also be
positive. We will let the magnitude of the heat added to the system at the temperature XL be |UL | We can
calculate the work done by the ideal gas for this process from the equation
F

[EF ( T .Z 8VXL (
E

This equation can also be expressed as

.Z
ZF
8VXL ln
Z
ZE

(3.14)

Chapter 3: The Second Law of Thermodynamics

10

[EF ( T .Z 8VXL (
E

.Z
TE
8VXL ln
Z
TF

(3.15)

since
TE ZE TF ZF

ZF ZE TE TF

(3.16)

for an ideal gas during an isothermal process.


Since TE TF this equation demonstrates that the net work is positive, as we know it should be. Notice
that this process is a good example of a case where heat can be added to the system even though the temperature
may not change. Again, since the heat added to the system during this process must be equal to the work done
during this process, we have
UEF 8VXL ln

TE
ZF
8VXL ln !
TF
ZE

(3.17)

F G: The system is now allowed to expand adiabatically from state F to state G . In this process, no
heat is added to or removed from the system (i.e., UFG !). From the first law, we have
?IFG UFG  [FG

(3.18)

?IFG  [FG

(3.19)

But since UFG !, this reduces to

and we only need to find [FG or ?IFG . It is easier to find ?IFG in this case, since it is only a function of the
temperatures. Thus,
?IFG  [FG

$
$
8VXG  XF 8VXP  XL
#
#

(3.20)

Since XP  XL the internal energy of the system must decrease, but the work done during this process must be
positive, as the gas expands!
G H: Since this process is an isothermal process, we can find the changes in internal energy, work, and
heat added by using the equations developed in the first isothermal process, E F. We obtain:
?IGH !
[GH UGH 8VXP ln

TG
ZH
8VXP ln
TH
ZG

(3.21)
(3.22)

which is less than zero, showing that heat is removed from the system as work is done compressing the system
from ZG ZH . Thus, UGH (which is less than zero) is the heat removed from the system at temperature XP and
is given by
UGH 8VXP ln

TG
ZH
8VXP ln  !
TH
ZG

(3.23)

H E: Similarly, for the final adiabatic process, we can use the equations developed for the previous
adiabatic process:
UHE !
?IHE  [HE

$
$
8VXH  XE 8VXL  XP
#
#

(3.24)
(3.25)

Since XL XP , this last equation indicates that the internal energy of the system increases while the work done
by the gas is negative (i.e., work is done on the gas by the surroundings as the gas is further compressed)!
Net Changes in the Carnot Cycle. To find the net changes in internal energy, the net work done by the
system, and the net heat added to the system, we simply add the changes in each process together. This gives

Chapter 3: The Second Law of Thermodynamics

11

?I-C-6/ !

(3.26)

[-C-6/ U-C-6/ UEF UGH 8VXL ln

TE
TG
8VXP ln

TF
TH

(3.27)

This last equation can be simplified, using the relationships we derived earlier for adiabatic processes in ideal
gases. We showed that for an adiabatic process,
T "-# X # -98=>

(3.28)

which means that


"-#

"-#

(3.29)

"-#

"-#

(3.30)

TF XF TG XG
and

TE XE TH XH
Now, since
XE XF XL

and

XG XH XP

(3.31)

we can divide one of these last two equations by the other and obtain:
TF
TG

TE
TH

(3.32)

Thus, the net work done, which is equal to the net heat added to the system in one complete cycle is given by
[-C-6/ U-C-6/ UEF UGH 8VXL ln

TE
TG
8VXP ln

TF
TH

(3.33)

[-C-6/ U-C-6/ UEF UGH 8VXL ln

TE
TF
8VXP ln
TF
TE

(3.34)

TE
TE
 8VXP ln |UL |  |UP |
TF
TF

(3.35)

[-C-6/ U-C-6/ UEF UGH 8VXL ln

where we let |UL | and |UP | be the magnitude of the heat added to or removed from the heat reservoirs at
temperatures XL and XP , respectively.
Finally, we can express the work done in the Carnot cycle by the equation
[-C-6/ U-C-6/ UEF UGH |UL |  |UP | 8V ln

TE
XL  XP
TF

(3.36)

which demonstrates that the net work done by the system is positive, as it should be since this cyclic process acts
like a heat engine producing a net output of useful mechanical energy. If this process were operated in reverse,
the system would be a refrigerator or a heat pump and all the quantities we have calculated would be negative.
THE EFFICIENCY OF A CARNOT CYCLE
As we mentioned earlier, the thermal efficiency ( of a heat engine is defined to be the net work done by the
engine in a complete cycle divided by the amount of heat added to the system during that cycle. (Note that the net
heat added to the system is equal to the net amount of work done by the system.) For the Carnot cycle, we know
that the net work done in the cycle is equal to the net heat added during the cycle, or [-C-6/ |UL |  |UP | (see
Fig. 3.2). The efficiency of this cycle is given by
(

|[-C-6/ |
|UL |  |UP |
|UP |

"
|UL |
|UL |
|UL |

(3.37)

Chapter 3: The Second Law of Thermodynamics

12

Using the calculated values for |[-C-6/ |, and |UL |, we also find the efficiency to be
(

8V ln TTFE XL  XP )
|[-C-6/ |
XL  XP
XP

"
T
E
|UL |
XL
XL
8VXL ln TF

(3.38)

From these last two equations we discover an extremely important relationship between the magnitude of
the heat added to or taken from a system and the temperature of the system during the isothermal processes of a
Carnot cycle. These equations imply that for an ideal gas carried through a Carnot cycle
|UP |
XP

|UL |
XL

(3.39)

where |UP | and |UL | are the magnitudes of heat flow and XP and XL are the temperatures of the low- and hightemperature reservoirs, respectively. Since the ratio of the temperatures can be determined by simply measuring
the amount of heat transfered during the isothermal processes of a Carnot cycle, we can use the Carnot cycle to
define a temperature scale. This temperature scale, known as the thermodynamic temperature scale, has been
shown to be equivalent to the Kelvin scale. Thus, the ratio of the heat added to or removed from an ideal gas
system during the isothermal processes of a Carnot cycle is exactly equal to the ratio of the absolute
temperatures of the two heat reservoirs. This means that the equations developed earlier for the thermal
efficiency of a heat engine and the coefficient of performance of a refrigerator and of a heat pump can now be
evaluated for a Carnot cycle simply by knowing the temperatures of the heat reservoirs. One can show that now
heat engine, operating between two heat reservoirs, can be more efficient than a Carnot heat engine operating
between these same two heat reservoirs. Thus a Carnot heat engine is the most efficient heat engine operating
between any two heat reservoirs of fixed temperature.
As an example, let's apply what we have learned to the case of an ideal (Carnot) steam plant. The
efficiency of the steam plant is given by
(

|[-C-6/ |
|UL |  |UP |
|UP |
XP

"
"
|UL |
|UL |
|UL |
XL

(3.40)

We can achieve the largest possible efficiency by letting XP !, of by letting XL , or both. However, there
is usually a practical limit to these temperatures. If we consider the efficiency of a coal fired steam plant, for
example, the pipes which carry the heated steam to drive the turbine would melt if XL were too high (greater than
about 500G ). In addition, the heat exchangers which allow us to get rid of the excess heat do not function well
when XP gets too low (at temperatures lower than ! C, the water which passes over the heat exchanger freezes
and the ice acts as a thermal insulator decreasing the efficiency of the heat exchanger). For a typical coal fired
steam plant, then, the high temperature reservoir is at approximately 500 C (773 K) and the low temperature
reservoir is at approximately 27C (300K), so that the efficiency of this system can be no better than
( "

300
0.6119 or 61.19%
773

(3.41)

The total efficiency of a real coal-fired steam plant is much less than this, since we must take into account many
other factors: the efficiency of heating the water to produce steam, the loss of heat through the pipes, the friction
in the turbine, etc. In addition, a coal-fired steam plant is not exactly equivalent to a Carnot cycle. [Note:
Nuclear powered steam plants are operated at lower temperatures than steam plants - for safety reasons - with the
high temperature of the boiler restricted to about 300 C. This means that the efficiency of a nuclear powered
steam plant must necessarily be less than a comparable coal or gas fired steam plant.]
As a side note, notice that since no heat engine can ever by 100% efficient, we can never have |UP | !,
or XP !. Thus, absolute zero Kelvin is unattainable. Experiments have been performed in which temperatures
as low as several milliKelvin have been attained, but one can never reach absolute zero! This is often given as an
alternate way of stating the second law of thermodynamics.
We next look at the example of an icebox which is maintained at a temperature of &C in a room which is
maintained at a temperature of 23C. The coefficient of performance of a Carnot refrigerator operating between

Chapter 3: The Second Law of Thermodynamics

13

these two temperatures is given by


GST</0

|UP |
XP
268

9.57
|[ |
296  268
XL  XP

(3.42)

Notice that this number is greater than unity just as we said it would be! This means that the amount of heat
which can be extracted from the icebox is 9.57 times greater than the amount of energy input to the system in the
form of mechanical energy. This means that if energy were put into the compressor at a rate of 1 Watt, the rate at
which heat energy would be removed from the icebox is 9.57 Watts, provided that the compressor is 100%
efficient, which it isn't. However, you can see that even a 30% conversion of electrical energy to mechanical
energy in this case would yield a gain of .30 9.57 2.87 over the amount of energy supplied by the electrical
company.
Now consider a heat pump which maintains a home at 23C when the outside temperature is 5C (these
are the same temperatures we used for the refrigerator). The coefficient of performance of the heat pump is given
by
GST2:

XL
|UL |
296

10.57
|[ |
XL  XP
296  268

(3.43)

Notice that this value is different from the coefficient of performance of the refrigerator operating between the
same two reservoirs. In fact, notice that
GST2:  GST</0

|UL |
|UP |
|UL |  |UP |


"
|UL |  |UP | |UL |  |UP |
|UL |  |UP |

(3.44)

This equation is true for all systems, not just a Carnot system.
The actual efficiency of a heat pump is less than what we calculated for a Carnot system because of the
inefficiency of heat exchangers, and the fact that no mechanical motor is 100% efficient. This fact actually limits
the utility of heat pumps is some areas of the country. However, we can use the ideal system to determine under
what circumstances a heat pump might be most useful. For example, you should be able to see that the efficiency
of the heat pump goes up when the average outside temperature is close to the desired temperature inside the
house. The efficiency decreases when the average outside temperature drops. Thus, real heat pumps have been
more popular in the south, than in the north.

EFFICIENCIES OF REVERSIBLE AND IRREVERSIBLE PROCESSES


Carnot discovered two very important principles while studying the efficiencies of thermodynamic
systems:
1.

The efficiency of an irreversible heat engine is always less that the efficiency of a reversible
process when operating between the same two heat reservoirs.

2.

The efficiencies of all reversible heat engines must be the same when operating between the
same two heat reservoirs.

A proof of these two statements is left as an exercise for the student. However, the method of proof is
very similar to our discussion of the equivalence of the Claussius and Kelvin-Planck statements of the second law.
You can show that the assumption that there is an irreversible heat engine which is more efficient than a
reversible one leads to a contradiction of the second law. Once you show this, the proof of the second statement
follows easily.
The consequencies of these two principles are fairly straight forward. Since all reversible heat engines
operating between the same two heat reservoirs must have the same efficiency, the special nature of the Carnot
cycle is unimportant. Likewise, since no irreversible heat engine can be more efficient than a reversible one, the
Carnot cycle can be used as a theoretical limit on any real heat engine operating between two heat reservoirs.
Other types of heat engines, and their corresponding idealized models are discussed in the appendix.
There you will find descriptions of the Otto cycle (internal combustion engine with spark ignition [the gasoline

Chapter 3: The Second Law of Thermodynamics

14

engine]) as well as the Diesel cycle (internal combustion engines with compression ignition) and calculations of
their respective efficiencies.

THE DEFINITION OF ENTROPY


We discovered that when an ideal gas is carried through a Carnot cycle the ratio of the heat added to the
system at the high-temperature reservoir to the heat removed from the system at the low-temperature reservoir is
just equal to the ratio of the absolute temperatures of the two reservoirs, or
|UP |
XP

|UL |
XL

(3.45)

where |UP | and |UL | are the magnitudes of heat flow and XP and XL are the temperatures of the low- and hightemperature reservoirs, respectively. Rearranging this last equation, we find that
|UP |
|UL |

XP
XL

(3.46)

for the Carnot cycle. This equation serves to define a new quantity which we will call the entropy change of a
system during a reversible, isothermal process. More precisely, we define the change in entropy of a system
during a reversible, isothermal process by the equation
WEF

UEF

X
</@

(3.47)

where WEF will be positive if heat is added to the system, but negative if heat is removed from the system. Notice
that from the viewpoint of the thermal reservoirs, the heat removed from a reservoir is considered positive, while
the heat depositied in a reservoir is considered negative.
For the Carnot cycle, we have the total change in entropy of the ideal gas system for the complete cycle
given by
W-C-6/

UEF
UGH
|UL |
|UP |



0
XL </@
XP </@
XL </@
XP </@

(3.48)

where |UL | and |UP | are magnitudes, whereas UEF and UGH are either positive or negative depending upon the
direction of energy flow. This last equation indicates the special nature of the quantity which we defined as the
entropy. Since the change in entropy of the system is zero as the system is carried through a complete cycle, the
entropy of a system must be a state variable! Because the entropy is actually a state variable, it is more
appropriate to express the differential change in entropy by the following equation
.W

$U

X </@

(3.49)

where $U is the differential amount of heat added to the system during a reversible, isothermal process at
temperature X . If the system temperature in a process is changing, but the process is still a reversible one, we can
integrate to determine the total change in entropy of the system
?WEF (

$U

X </@

(3.50)

You might wonder how we can equate the change in a state variable with a change in a process variable.
Remember that we did the same thing for the differentical work done in a volume expansion. As long as the
process occurs very slowly (quasi-statically) this definition holds true. In a real, irreversible process, however,
the change in entropy is not equal to the amount of heat added to the system divided by the temperature. The
processes involved in the Carnot cycle from which we derived this expression are processes occurring for an ideal
gas as that gas moves through certain reversible, or quasi-static, processes on the T @ X surface of that gas. For
this reason, all the equations we have developed for the Carnot cycle are valid only for reversible processes
(including the equation relating the entropy to the heat transfer of an isothermal process)! However, if we change

Chapter 3: The Second Law of Thermodynamics

15

a system from state E to state F by some arbitrary process, the state variables must change by the same amount,
no matter if the process is reversible or not! This means that since the entropy of the system is a state variable,
changes in entropy will be the same between any two states whether or not the process which connects the two
states is reversible. However, we can actually calculate the change in entropy (or temperature, pressure, volume,
etc.) only if we choose an appropriate reversible process for which we know the equation of state of the system
(or the appropriate partials) and for which the equation
.W

$U

X </@

(3.51)

is valid!
THE SECOND LAW IN TERMS OF ENTROPY
We now want to consider how the change in entropy of a system is related to the second law of
thermodynamics. We want to demonstrate that the change in entropy can be used to indicate the direction of heat
(or energy) flow in a complex system. To see this, we will consider an arbitrary cyclic heat engine (not
necessarily reversible), in which a certain amount of heat, |UL |, flows from a high temperature reservoir at
temperature XL into the system, and an equal amount of heat, |UP |, flows out of the system and into a colder
reservoir at temperature XP . In this case the net work done by the system, [ , is zero, since |UL | |UP | |Uo |.
(This is not a very useful heat engine, since it will perform no useful work - but is helps in understanding the
relationship between entropy an heat flow.) This reversible engine is, therefore, equivalent to the case where heat
simply flows from a hotter reservoir to a cooler one. In this case, the ratio |UL |XL is not equal to the entropy
change, since the process is not necessarily reversible. However, we can still calculate this ratio for our arbitrary
system. Remembering that we take heat flowing out of a reservoir as positive and into a reservoir as negative, we
can write the sum of these ratios as

|UL | |UP |
"
"

|Uo |


XL
XP
XL
XP

(3.52)

This equation tells us that if XL XP , as we assumed, this sum is less than zero. If, however, we were to assume
that XP XL , (in direct violation of second law) this sum would be greater than zero! Thus, we conclude that the
sum of the amount of heat removed from the heat reservoirs divided by the absolute temperature of those
reservoirs must be less than zero, or

U
!
X

(3.53)

to be consistent with the second law. We, therefore, postulate that for any arbitrary process, carried through a
complete cycle, the sum of the heats added to the system divided by the corresponding absolute temperatures
must be less than or equal to zero (the equality occuring only in the case of a reversible system passing through a
cyclic process). This same principle can be extended to a system removing heat from any number of heat
reservoirs, and gives us the so-called Clausius inequality:

U
!
X

(3.54)

In the limit as we remove a small amount of heat at one temperature, and then a small amount of heat at another
adjacent temperature, the sum becomes an integral, and we write

* X !
$U

(3.55)

where the equal sign is true for a reversible process only. Note that this equation explicitly indicates that a cyclic
process is being considered by the inclusion of a circle over the integral sign. This integral is zero only for the
case of a reversible process and it is only for reversible processes that the ratio $UX is equal to the entropy, a
state variable.
To see how the Clausius inequality leads to a general statement of the law of entropy, consider a cyclic
process in which a system is carried from state E to state F and then back to E, as shown in Fig 3.9:

Chapter 3: The Second Law of Thermodynamics

16

I
B
A
II
Fig. 3.9 A closed-loop path for which we calculate the entropy change. Path I is an irreversible
process, while path II is a reversible one.

In this particular case, we consider a system which is allowed to go from state E to state F through some isolated,
adiabatic process, which is irreversible. It is then returned to its original state by means of some reversible
(quasi-static) process. The Clausius inequality for this cycle states that

* X !
$U

(3.56)

Since the cyclic process as a whole is irreversible, we must use the inequality, which gives
E
$U
$U
(
!
X </@
E X
F
M
MM

(3.57)

Since path II is a reversible process, the ratio $UX is equal to the entropy change, .W , so that the integral over
path II will give us the entropy change from state F to state E, ?WFE . But the process along path I was assumed
to be isolated (adiabatic), so that $U must be zero for all points along that path. This means that the integral
along path I (which is not equal to the entropy change since the process is irreversible) must be zero. Therefore,
we have
E
E
$U
(
( .W WE  WF  !
X </@
F
F
MM
MM

(3.58)

or
WE  WF

(3.59)

This equation asserts that the entropy of any isolated, irreversible process must increase! Notice that we can not
say that the entropy will always increase for every process. In fact, this is not generally true, as we will see in the
examples that follow. It is true only for isolated processes. Thus, if we can consider our galaxy as sufficiently
isolated from the rest of the universe, we can claim that the entropy of our isolated galaxy must be increasing if
there is anything, whatsoever, going on (any processes occuring)!
CALCULATION OF ENTROPY CHANGES IN VARIOUS PROCESSES
Since the entropy is defines by the equation
.W

$U

X </@

(3.60)

we can calculate the entropy change for any process E p F, so long as we can devise a reversible process which
will take us from E p F. In this section we will consider a process where the entropy of the system must increase

Chapter 3: The Second Law of Thermodynamics

17

because the process is inherently irreversible, and we will demonstrate how to calculate the entropy change for
this process. We will consider the irreversible heat flow that occurs when we mix water and ice.
As a specific example, consider the case where 1 kg of ice at ! C is placed in an adiabatic container
filled with 1 kg of water at 80 C. Since the temperature difference between the ice and water is fairly large, the
cooling of the water and the melting of the ice can not be considered a quasi-static (reversible) process. We can,
however, determine the amount of heat transferred from the water to the ice as the ice is melted, and we can
determine the final temperature of the system when it reaches equilibrium. (This can be done whether or not all
the ice is melted.) We know that melting 1 kg of ice at 0 C requires
U7/6> 7P " kg )! kcal/kg )! kcal

of energy. The amount of heat energy available in the water (i.e., the amount of energy that can be removed from
the water as it is cooled to the point of freezing) is equal to
U+@+36 7-?X " kg " kcal/kg C 8! C 8! kcal

so that for this particular example, all the ice is melted and the water's temperature is lowered to the freezing
point. We have transferred 80 kcal of thermal energy from the water to the ice, melting the entire 1 kg of ice, and
the water has been cooled to 0 C, the equilibrium temperature of the system.
Although the process we have described is not a reversible one, we can take 1 kg of ice and melt it in a
reversible process by bringing the ice in contact with a heat reservoir with a temperature infinitesimally above
0C so that heat is allowed to slowly enter the ice and melt it. Since this reversible process takes place at a
temperature of 273K, the heat added to the ice to melt it, divided by the absolute temperature of the heat reservoir
is just the change in entropy of the ice, given by
?W3-/

)! 5-+6
"#$ "!$ N O
#($ O

The entropy change of the heat reservoir in contact with the ice is the same magnitude as the entropy change of
the ice, but with an opposite sign, since heat is being removed from the reservoir. Thus to total change in entropy
of the universe for this reversible process is zero.
Likewise, we can take 1 kg of water at 80C and lower its temperature infinitesimally by bringing it into
contact with a heat reservoir infinitesimally cooler than 80C. We can continue this process, by moving the water
from one reservoir to another among an infinite number of reservoirs which are slightly cooler than the previous
one, until we remove all the heat in the water and the water reaches the equilibrium temperature of 0C. Since
this process is reversible, we can calculate the entropy change of the water from the equation
?WA+>/< (

X #($

X #($ )!

X #($
X #($
$U
7A -A .X
.X
7 A -A (
(
X </@
X
X #($ )!
X #($ )! X

assuming that the heat capacity of water is a constant over this temperature range. For our particular problem,
which gives
?WA+>/< 1 10$ g

%")'N
#($
$
68
 "!) "! N O
1O
$&$

This is negative since heat is being removed from the water. The change in entropy of all the heat reservoirs that
we had to use is just the negative of this quantity (the heat was added to the reservoirs), so that the total entropy
change in the universe for this reversible process is zero.
Now, the total entropy change of the water plus the ice is a state variable - independent of the process by
which you reach the end points. Thus, the total change in entropy of the isolated system (water plus ice) must be
given by
?W73B>?</ ?W3-/ ?WA+>/< "#$ "!$ N O  "!) "!$ N O
?W73B>?</ !"& "!$ N O
which is greater than unity as it must be for an isolated system!

Chapter 3: The Second Law of Thermodynamics

18

A Molecular View of Entropy


As we stated earlier, we can demonstrate several processes whereby mechanical energy can be converted
into heat. One such example is a block of wood sliding across the floor. We have also demonstrated that it is
impossible to convert a given amount of heat energy into an equivalent amount of mechanical energy. This oneway aspect of nature seems to be tied into the concept of entropy. We might expect, therefore, that the entropy
of a block of wood which is sliding across the floor is increasing as it comes to a stop. When the block of wood
come to a stop, we find that the block will be somewhat hotter than if there were no friction to slow the block
down. This increase in the thermal energy of the block with a subsequent loss of mechanical energy of motion
must be tied to the entropy change of the system. This would seem to indicate that the organized motion of all the
molecules that make up the block of wood has less entropy than the increased random motion of the molecules
within the block after it is heated by friction. Thus, we might conclude that the entropy of a system may be
related somehow to the random nature of the motion of the molecules within a material. We will attempt to
determine how the entropy of a system is related to the molecular structure of matter later in our study

Chapter 3: The Second Law of Thermodynamics

19

Appendix 3.1
EFFICIENCIES OF OTHER CYCLIC PROCESS COMPARED WITH THE CARNOT CYCLE
Diesel, Sterling, Joule, Otto, compared with Carnot

Das könnte Ihnen auch gefallen