Heat Engines and the Second Law of Thermodynamics chapter 22.pdf

(966 KB) Pobierz
627022485 UNPDF
Chapter 22
Heat Engines, Entropy, and the
Second Law of Thermodynamics
CHAPTER OUTLINE
22.1 Heat Engines and the Second
Law of Thermodynamics
22.2 Heat Pumps and Refrigerators
22.3 Reversible and Irreversible
Processes
22.4 The Carnot Engine
22.5 Gasoline and Diesel Engines
22.6 Entropy
22.7 Entropy Changes in
Irreversible Processes
22.8 Entropy on a Microscopic
Scale
! This cutaway image of an automobile engine shows two pistons that have work done on
them by an explosive mixture of air and fuel, ultimately leading to the motion of the
automobile. This apparatus can be modeled as a heat engine, which we study in this chapter.
(Courtesy of Ford Motor Company)
667
627022485.014.png 627022485.015.png
T he first law of thermodynamics, which we studied in Chapter 20, is a statement of
conservation of energy. This law states that a change in internal energy in a system can
occur as a result of energy transfer by heat or by work, or by both. As was stated in
Chapter 20, the law makes no distinction between the results of heat and the results of
work—either heat or work can cause a change in internal energy. However, there is an
important distinction between heat and work that is not evident from the first law. One
manifestation of this distinction is that it is impossible to design a device that, operat-
ing in a cyclic fashion, takes in energy by heat and expels an equal amount of energy by
work. A cyclic device that takes in energy by heat and expels a fraction of this energy by
work is possible and is called a heat engine .
Although the first law of thermodynamics is very important, it makes no distinc-
tion between processes that occur spontaneously and those that do not. However,
only certain types of energy-conversion and energy-transfer processes actually take
place in nature. The second law of thermodynamics , the major topic in this chapter,
establishes which processes do and which do not occur. The following are examples
of processes that do not violate the principle of conservation of energy if they pro-
ceed in either direction, but are observed to proceed in only one direction, governed
by the second law:
Lord Kelvin
British physicist and
mathematician (1824–1907)
• When two objects at different temperatures are placed in thermal contact with
each other, the net transfer of energy by heat is always from the warmer object to
the cooler object, never from the cooler to the warmer.
• A rubber ball dropped to the ground bounces several times and eventually comes
to rest, but a ball lying on the ground never gathers internal energy from the
ground and begins bouncing on its own.
• An oscillating pendulum eventually comes to rest because of collisions with air mol-
ecules and friction at the point of suspension. The mechanical energy of the system
is converted to internal energy in the air, the pendulum, and the suspension; the
reverse conversion of energy never occurs.
Born William Thomson in Belfast,
Kelvin was the first to propose
the use of an absolute scale of
temperature. The Kelvin
temperature scale is named in
his honor. Kelvin’s work in
thermodynamics led to the idea
that energy cannot pass
spontaneously from a colder
object to a hotter object.
(J. L. Charmet/SPL/Photo
Researchers, Inc.)
All these processes are irreversible —that is, they are processes that occur naturally in
one direction only. No irreversible process has ever been observed to run backward—if
it were to do so, it would violate the second law of thermodynamics. 1
From an engineering standpoint, perhaps the most important implication of the
second law is the limited efficiency of heat engines. The second law states that a ma-
chine that operates in a cycle, taking in energy by heat and expelling an equal amount
of energy by work, cannot be constructed.
1 Although we have never observed a process occurring in the time-reversed sense, it is possible for it to
occur. As we shall see later in the chapter, however, the probability of such a process occurring is
infinitesimally small. From this viewpoint, we say that processes occur with a vastly greater probability in
one direction than in the opposite direction.
668
627022485.016.png 627022485.017.png
SECTION 22.1 • Heat Engines and the Second Law of Thermodynamics
669
22.1 Heat Engines and the Second Law
of Thermodynamics
A heat engine is a device that takes in energy by heat 2 and, operating in a cyclic
process, expels a fraction of that energy by means of work. For instance, in a typical
process by which a power plant produces electricity, coal or some other fuel is burned,
and the high-temperature gases produced are used to convert liquid water to steam.
This steam is directed at the blades of a turbine, setting it into rotation. The mechani-
cal energy associated with this rotation is used to drive an electric generator. Another
device that can be modeled as a heat engine—the internal combustion engine in an
automobile—uses energy from a burning fuel to perform work on pistons that results
in the motion of the automobile.
A heat engine carries some working substance through a cyclic process during
which (1) the working substance absorbs energy by heat from a high-temperature en-
ergy reservoir, (2) work is done by the engine, and (3) energy is expelled by heat to a
lower-temperature reservoir. As an example, consider the operation of a steam engine
(Fig. 22.1), which uses water as the working substance. The water in a boiler absorbs
energy from burning fuel and evaporates to steam, which then does work by expand-
ing against a piston. After the steam cools and condenses, the liquid water produced
returns to the boiler and the cycle repeats.
It is useful to represent a heat engine schematically as in Figure 22.2. The engine
absorbs a quantity of energy ! Q h ! from the hot reservoir. For this discussion of heat en-
gines, we will use absolute values to make all energy transfers positive and will indicate
the direction of transfer with an explicit positive or negative sign. The engine does
work W eng (so that negative work W !" W eng is done on the engine), and then gives up
a quantity of energy ! Q c ! to the cold reservoir. Because the working substance goes
Hot reservoir at T h
Q h
Figure 22.1 This steam-driven
locomotive runs from Durango to
Silverton, Colorado. It obtains its
energy by burning wood or coal.
The generated energy vaporizes
water into steam, which powers the
locomotive. (This locomotive must
take on water from tanks located
along the route to replace steam
lost through the funnel.) Modern
locomotives use diesel fuel instead
of wood or coal. Whether old-
fashioned or modern, such
locomotives can be modeled as
heat engines, which extract energy
from a burning fuel and convert a
fraction of it to mechanical energy.
W eng
Engine
Q c
Cold reservoir at T c
Active Figure 22.2 Schematic
representation of a heat engine.
The engine does work W eng . The
arrow at the top represents energy
Q h # 0 entering the engine. At the
bottom, Q c $ 0 represents energy
leaving the engine.
2 We will use heat as our model for energy transfer into a heat engine. Other methods of energy
transfer are also possible in the model of a heat engine, however. For example, the Earth’s atmosphere
can be modeled as a heat engine, in which the input energy transfer is by means of electromagnetic
radiation from the Sun. The output of the atmospheric heat engine causes the wind structure in the
atmosphere.
At the Active Figures link
at http://www.pse6.com, you
can select the efficiency of the
engine and observe the
transfer of energy.
627022485.001.png 627022485.002.png 627022485.003.png
670
CHAPTER 22 • Heat Engines, Entropy, and the Second Law of Thermodynamics
P
through a cycle, its initial and final internal energies are equal, and so % E int ! 0.
Hence, from the first law of thermodynamics, % E int ! Q & W ! Q " W eng , and with
no change in internal energy, the net work W eng done by a heat engine is equal
to the net energy Q net transferred to it. As we can see from Figure 22.2,
Q net ! | Q h | " | Q c |; therefore,
Area = W eng
W eng ! ! Q h ! " ! Q c !
(22.1)
V
If the working substance is a gas, the net work done in a cyclic process is the
area enclosed by the curve representing the process on a PV diagram. This is
shown for an arbitrary cyclic process in Figure 22.3.
The thermal efficiency e of a heat engine is defined as the ratio of the net work
done by the engine during one cycle to the energy input at the higher temperature
during the cycle:
Figure 22.3 PV diagram for an
arbitrary cyclic process taking place
in an engine. The value of the net
work done by the engine in one
cycle equals the area enclosed by
the curve.
e ! W eng
! Q h ! " ! Q c !
! Q h !
! Q c !
! Q h !
! Q h ! !
! 1 "
(22.2)
Thermal efficiency of a heat
engine
We can think of the efficiency as the ratio of what you gain (work) to what you give
(energy transfer at the higher temperature). In practice, all heat engines expel only a
fraction of the input energy Q h by mechanical work and consequently their efficiency
is always less than 100%. For example, a good automobile engine has an efficiency of
about 20%, and diesel engines have efficiencies ranging from 35% to 40%.
Equation 22.2 shows that a heat engine has 100% efficiency ( e ! 1) only if
! Q c ! ! 0—that is, if no energy is expelled to the cold reservoir. In other words, a heat
engine with perfect efficiency would have to expel all of the input energy by work. On
the basis of the fact that efficiencies of real engines are well below 100%, the
Kelvin–Planck form of the second law of thermodynamics states the following:
Hot reservoir at T h
Q h
It is impossible to construct a heat engine that, operating in a cycle, produces no
effect other than the input of energy by heat from a reservoir and the performance
of an equal amount of work.
W eng
Engine
This statement of the second law means that, during the operation of a heat engine,
W eng can never be equal to ! Q h !, or, alternatively, that some energy ! Q c ! must be
rejected to the environment. Figure 22.4 is a schematic diagram of the impossible
“perfect” heat engine.
Cold reservoir at T c
Quick Quiz 22.1 The energy input to an engine is 3.00 times greater than
the work it performs. What is its thermal efficiency? (a) 3.00 (b) 1.00 (c) 0.333
(d) impossible to determine
The impossible engine
Figure 22.4 Schematic diagram of
a heat engine that takes in energy
from a hot reservoir and does an
equivalent amount of work. It is
impossible to construct such a
perfect engine.
Quick Quiz 22.2 For the engine of Quick Quiz 22.1, what fraction of the en-
ergy input is expelled to the cold reservoir? (a) 0.333 (b) 0.667 (c) 1.00 (d) impossible
to determine
Example 22.1 The Efficiency of an Engine
An engine transfers 2.00 ' 10 3 J of energy from a hot reser-
voir during a cycle and transfers 1.50 ' 10 3 J as exhaust to a
cold reservoir.
Solution The efficiency of the engine is given by Equation
22.2 as
e ! 1 " ! Q c !
! Q h ! ! 1 " 1.50 ' 10 3 J
(A) Find the efficiency of the engine.
2.00 ' 10 3 J !
0.250, or 25.0%
627022485.004.png 627022485.005.png 627022485.006.png
 
SECTION 22.2 • Heat Pumps and Refrigerators
671
(B) How much work does this engine do in one cycle?
Answer No, you do not have enough information. The
power of an engine is the rate at which work is done by the
engine. You know how much work is done per cycle but you
have no information about the time interval associated with
one cycle. However, if you were told that the engine oper-
ates at 2 000 rpm (revolutions per minute), you could relate
this rate to the period of rotation T of the mechanism of the
engine. If we assume that there is one thermodynamic cycle
per revolution, then the power is
Solution The work done is the difference between the
input and output energies:
W eng ! ! Q h ! " ! Q c ! ! 2.00 ' 10 3 J " 1.50 ' 10 3 J
!
5.0 ' 10 2 J
What If? Suppose you were asked for the power output of
this engine? Do you have sufficient information to answer
this question?
! ! W eng
T
! 5.0 ' 10 2 J
"
" 1 min
# ! 1.7 ' 10 4 W
22.2 Heat Pumps and Refrigerators
! PITFALL PREVENTION
22.1 The First and Second
Laws
Notice the distinction between
the first and second laws of
thermodynamics. If a gas under-
goes a one-time isothermal process
% E int ! Q & W ! 0. Therefore,
the first law allows all energy in-
put by heat to be expelled by
work. In a heat engine, however,
in which a substance undergoes a
cyclic process, only a portion of
the energy input by heat can be
expelled by work according to
the second law.
In a heat engine, the direction of energy transfer is from the hot reservoir to the cold
reservoir, which is the natural direction. The role of the heat engine is to process the
energy from the hot reservoir so as to do useful work. What if we wanted to transfer en-
ergy from the cold reservoir to the hot reservoir? Because this is not the natural direc-
tion of energy transfer, we must put some energy into a device in order to accomplish
this. Devices that perform this task are called heat pumps or refrigerators. For exam-
ple, we cool homes in summer using heat pumps called air conditioners. The air condi-
tioner transfers energy from the cool room in the home to the warm air outside.
In a refrigerator or heat pump, the engine takes in energy ! Q c ! from a cold reser-
voir and expels energy ! Q h ! to a hot reservoir (Fig. 22.5). This can be accomplished
only if work is done on the engine. From the first law, we know that the energy given up
to the hot reservoir must equal the sum of the work done and the energy taken in from
the cold reservoir. Therefore, the refrigerator or heat pump transfers energy from a
colder body (for example, the contents of a kitchen refrigerator or the winter air out-
side a building) to a hotter body (the air in the kitchen or a room in the building). In
practice, it is desirable to carry out this process with a minimum of work. If it could be
accomplished without doing any work, then the refrigerator or heat pump would be
“perfect” (Fig. 22.6). Again, the existence of such a device would be in violation of the
second law of thermodynamics, which in the form of the Clausius statement 3
states:
Hot reservoir at T h
Q h
W
Heat pump
Q c
Cold reservoir at T c
Active Figure 22.5 Schematic diagram of a heat pump,
which takes in energy Q c # 0 from a cold reservoir and
expels energy Q h $ 0 to a hot reservoir. Work W is done
on the heat pump. A refrigerator works the same way.
At the Active Figures link
at http://www.pse6.com, you
can select the COP of the heat
pump and observe the transfer
of energy.
3
First expressed by Rudolf Clausius (1822–1888).
2 000 min #
1
60 s
627022485.007.png 627022485.008.png 627022485.009.png 627022485.010.png 627022485.011.png 627022485.012.png 627022485.013.png
Zgłoś jeśli naruszono regulamin