Sei sulla pagina 1di 19

1

Chap. 4 Heat Engines, Power Cycles, and the Thermodynamics


of Open Systems
4.1 Heat Engines.1
4.2 Statements of the Second Law as applied to Heat-engine Cycles3
4.3 The Carnot Cycle.5
4.4 Application of the First Law to Open Systems7
4.5 Application of the Second Law to Open Systems.12
1
4.1 Heat Engines
In Sect. 1.9, it was noted that the First law regards heat and work as completely
interchangeable; if a certain number of Joules of heat added to a system increases the
internal energy of a body by, say, U, the same number of Joules of work performed on
the body would produce the same U. In addition, work can be completely converted to
heat, as everyday experience with friction attests. However, the reverse is not true; heat
cannot be completely transformed into work. This limitation, which is a consequence of
the Second law, is best demonstrated by studying the properties of heat engines.
A heat engine is a system operating in a cycle that receives heat from a high-
temperature source (called a reservoir) and produces useful work. However, since the
efficiency of conversion must be less than 100%, some of the input heat must be rejected
(as heat) to a cold reservoir. Figure 4.1 shows a schematic of a heat engine and its
associated thermal reservoirs.
Hot reservoir
T
H
cold reservoir
T
L

W
heat
engine
Q
H
Q
L
Fig. 4.1 A schematic of a heat engine driven by a hot reservoir
The reservoirs have the feature of supplying or receiving heat without alteration of their
temperatures. Heat flows in the reservoirs are reversible, even though processes in the
heat engine may be contain irreversibilities.
The circle with the arrows in Fig. 4.1 is a shorthand representation of the heat
engine. It is intended to signify that the working substance (a fluid such as an ideal gas or
water) moves through many thermodynamic states in a never-ending cyclic process. The
detailed structure of the heat engine can vary greatly, but the simplest version contains the
following four steps:
1. one in which heat absorbed isothermally from the high-temperature reservoir
2. the next in which work is produced adiabatically
3. followed by isothermal rejection of heat to the low-temperature reservoir
4. the last in which work is done on the working substance to return it to the state
at the start of step 1
2
The heat engine can operate in either of two ways: i) as a single device that
sequentially moves through the four processes described above, or ii) with a fluid flowing
through for distinct devices, each assigned to one of the four steps.
The sequential type is shown in Fig. 4.2. It consists of a single piston-cylinder that
performs each of the four steps in sequence. (Read Van Ness, pp. 36 40 for a fuller
explanation of this cycle)
Proceeding from left to right in the diagram follows the device in time. If the gas in the
cylinder were ideal and if the four steps were conducted reversibly, the cycle would appear
as shown in the bottom of Fig. 4.2. The cycle consists of two isotherms (like the one in
Fig. 3.2) and a pair of isentropes (see Fig. 3.4). Work is exchanged with the surroundings
in each step. The net work produced by the cycle is: W = W
1-2
+ W
2-3
- W
3-4
-W
4-1
. Since
each of the component work terms is the integral of pdV, W is the area inside the four-
sided figure in Fig. 4-2.
The flow-cycle heat engine shown in Fig. 4.3 contains the four processes in
separate units that act on the circulating working fluid. Van Ness devotes Chapter 4 of his
book to this type of cycle. This heat engine is an idealized steam power plant, whether
nuclear or fossil. The boiler (or steam generator in a nuclear plant) receives heat from the
primary source. The condenser rejects heat to a sink such as a river, lake, or cooling
tower. The turbine produces shaft (i.e., rotating) work, most of which is eventually
converted to electric power with high efficiency. However, some of the turbine power is
Fig. 4.2 A heat engine consisting of a single piston
performing four different operations. The cycle on a p-v
diagram is shown at the bottom
3
consumed internally by the pump needed to keep the fluid circulating in the loop. The net
work of the cycle is W = W
2-3
- W
4-1
.
Fig. 4.3 A heat engine based on a fluid circulating through four heat and work devices
In all three illustrations of heat engines, the First law is:
Q
H
= Q
L
+ W (4.1)
where Q
H
and Q
L
are the quantities heat delivered by the hot reservoir and received by the
cold reservoir, respectively. W is the net work of the cycle. Because of the cyclic nature of
thesystem, there is no change in internal energy in each cycle, or U = 0.
Note that the sign conventions for work and heat adopted in Sect. 1.9 for use in
the First law have been temporarily abandoned for the present discussion of heat engines.
The reason is simply to avoid negative values or the use of absolute-value signs on the Qs
and Ws. These quantities are considered to be positive in the direction shown by the
arrows representing them in the diagrams.
4.2 Statements of the Second Law as applied to Heat-engine Cycles
The following qualitative constraints on heat-engine cycles were discovered in the
nineteenth century and eventually led to the concept of entropy. These constraints restrict
the functioning of cycles more severely than does the First law. Both are equivalent
statements of the Second law, and are ultimately based on empirical evidence.
The Kelvin-Planck statement is: No cycle can produce net work with only a single
thermal reservoir.
This is a formal phrasing of a fact that has been noted earlier in these notes: that heat
cannot be completely converted to work. The Kelvin-Planck statement says that Q
L
in Fig.
4.1 cannot be zero.
4
The Clausius statement is: No cycle can produce only transfer of heat from a cold
reservoir to a hot reservoir
This statement essentially prohibits heat from flowing from cold to hot bodies,
something that was demonstrated (also using the Second law) in Sect. 1.10. With
reference to Fig. 4.1, the Clausius statement says that the direction of Q
H
and Q
L
cannot
be reversed and at the same time W set equal to zero. This version of the Second law does
not, however, entirely prohibit transfer of heat from a cold body to a hot body; it simply
requires that external work must be expended in order to do so. The directions of all of
the arrows in Fig. 4.1 can be reversed and the heat engine becomes a heat pump.
The two statements of the Second law appear to be quite different, but in fact they
are equivalent. This equivalence can be shown with the aid of Fig. 4.4, which contains heat
engine A and heat pump B operating between the same hot and cold reservoirs.
Hot Reservoir
Q
L
Q
H
Q
L Q
L
B A
Cold Reservoir
W = Q
H
- Q
L
Fig. 4.4 Demonstration of the equivalence the two statements of the Second Law
The devices are arranged so that the heat withdrawn from the cold reservoir by
pump B is the same magnitude as the heat rejected by engine A. Therefore, the two cancel
and the cold reservoir experiences no net heat transfer. Cycle B violates the Clausius
statement. It remains to show that the combination of cycle A and cycle B violates the
Kelvin-Planck statement.
Cycle A receives a quantity of heat Q
H
from the hot reservoir that is greater than
delivered by cycle B. By the First law applied to the system within the dashed box,
Q
H
- Q
L
units of heat are completely converted to work without rejecting heat to the cold
reservoir, in violation of the Kelvin-Planck statement. The device in Fig. 4.4, which is
known as a perpetual-motion machine of the second kind, thus fails both the Clausius and
Kelvin-Planck versions of the Second law.
5
4.3 The Carnot Cycle
If the Kelvin-Planck form of the Second law prohibits complete conversion of heat
to work, what then is the maximum value of the efficiency of the cycle shown in Fig. 4.1?
The cycle efficiency is defined by:
= =
W
Q
Q
Q
H
L
H
1 (4.2)
where the second form arises from elimination of W using Eq (4.1).
The maximum-efficiency cycle is the one containing the four steps listed on the
bottom of p. 1 with the additional restriction that the heat engine be reversible; heat
exchange with the reservoirs must take place over infinitesimally small temperature
differences, all motions must be frictionless, and rapid or compression of the working fluid
is prohibited. This idealized heat engine was first investigated by Carnot in 1824, and the
cycle and its efficiency bear his name. Carnot showed that the efficiency of this cycle is
simply related to the temperature of the hot and cold reservoirs, and moreover, that no
other engine operating between the same two temperatures can have a greater efficiency.
Determination of the efficiency of the Carnot cycle starts from the form of the
Second law given by Eq(1.20) for reversible operation:
S
engine
+ S
surroundings
= 0
The thermodynamic system in this usage is the Carnot engine and the surroundings are the
two reservoirs. The surroundings also contain a mechanism for receiving the work
performed by the Carnot engine but this mechanism is not involved in exchange of
entropy. Since the working fluid returns to its original state after each cycle, its entropy
change is zero. Because all aspects of the heat engine are reversible, no entropy is
generated from this device. Consequently, S
engine
= 0, and by the above equation,
S
surroundings
= 0 as well.
The null entropy change of the surroundings consists of two canceling terms, each
of the form given by Eq (1.16). The hot reservoir delivers entropy in the amount Q
H
/T
H
to
the engine and the reject heat transfers entropy equal to Q
L
/T
L
to the cold reservoir. If the
surroundings do not experience an entropy change as a result of operation of the Carnot
engine, these two entropy flows must be equal, or:
Q
T
Q
T
H
H
L
L
= (4.3)
Combining this result with Eq (4.2) gives the Carnot efficiency:
6

C
L
H
T
T
= 1 (4.4)
The above derivation did not utilize the details of the four-step cycle that comprises the
Carnot engine. However, analysis of the cycle in Fig. 4.2, for example, using Eq (3.5) for
the isothermal steps and Eq(3.16) for the isentropic steps, results in the efficiency given by
Eq (4.4). (see Potter and Somerton, p. 104 for additional information)
Demonstration that the Carnot engine has the maximum possible efficiency is
based on the pair of engines shown in Fig. 4.5. The method posits a higher efficiency
Hot Reservoir
Q
H
Q
H
Q
LA Q
LC
C A
W
C
Cold Reservoir
W
A
- W
C
Fig. 4.5 Diagram to demonstrate that the Carnot efficiency is the highest possible
for cycle A than for the Carnot Cycle C operating as a heat pump, then shows that this
supposition violates the Second law.
Cycles A and C are set to exchange equal quantities of heat with the hot reservoir,
which therefore experiences zero net heat transfer. Because cycle A is presumed to be
more efficient than Carnot cycle C, the work W
A
is greater than the work W
C
needed to
operate cycle C. Overall, cycles C and A taken as a single system within the dashed box in
the drawing receive an amount of heat net Q
LC
- Q
LA
from the cold reservoir and produce
an equal amount of net work, W
A
- W
C
. This is accomplished without rejecting heat to
another reservoir, thus constituting a violation of the Kelvin-Planck statement of the
Second law. This result demonstrates that the initial premise was false; heat engine A
cannot have an efficiency greater than the Carnot efficiency.
Problems involving heat engines operating on the Carnot cycle include Nos. 4.1 4.5,
4.11 and 4.12, and 4.15 and 4.16.
7
Other Power Cycles
Not every power-producing cycle consists of the four steps of the Carnot cycle. In
fact, nearly all practical cycles differ from the Carnot cycle. All cycles contain four distinct
steps, or rather two pairs of steps. Each pair of steps is conducted as an iso-something
process. The cycles differ in the thermodynamic properties that are held constant in the
pairs of steps. Table 4.1 lists the principle power cycles of practical significance, along
with the Carnot cycle for comparison.
Table 4.1 Power Cycles
Cycle Component Steps Efficiency Reference*
Carnot 2 isentropic; 2 isothermal 1 T
L
/T
H
P/S, p. 101; A/VN, p. 37
Rankine(water) 2 isentropic; 2 isobaric - P/S, p. 149; VN, Chap. 4
Brayton(gas) 2 isentropic; 2 isobaric

/ ) 1 (
H L
) p / p ( 1 P/S, p. 201
Otto (gas) 2 isentropic, 2 isochoric
1
H L
) V / V ( 1

A/VN, p. 38
Stirling 2 isochoric; 2 isothermal 1 T
L
/T
H
P/S, p. 199
Ericsson 2 isobaric, 2 isothermal 1 T
L
/T
H
P/S, p. 199
P/S = Potter and Somerton; A/VN = Abbott and Van Ness; VN = Van Ness
The Rankine and Brayton cycles differ only in the working medium. The Rankine cycle is
close to the Carnot cycle because water is used as the working medium. It closely
resembles Fig. 4.3. In this case, the isobaric steps involve condensation or vaporization of
water, which also occurs isothermally. Cycles that include a pair of isothermal steps have
the same theoretical efficiency (i.e., if they are perfectly reversible) as the Carnot cycle.
The efficiency is defined as the work produced divided by the heat added. The Otto cycle
utilizes a gas as a working fluid, and is the basis of the internal combustion engine. The
Stirling and Ericsson cycles also operate using gases, but Problem 4.3 analyzes an
Ericsson cycle using water/steam as the working fluid
4.4 Application of the First Law to Open Systems
The laws of thermodynamics can be applied to fluids flowing through devices that
change the properties of the fluid and that may in the process exchange heat and work
with the surroundings. Examples of such devices include pumps, boilers, turbines, valves
nozzles (used for increasing fluid velocity) and orifices in pipes. Some of these devices are
included in the simple steam cycle shown in Fig. 4.3. In this cycle, the components are
connected in series with the working fluid circulating continuously through them. Each
device in the cycle is subjected to a First-law analysis, and in some, application of the
Second law provides additional information on their performance. Usually, the
thermodynamic study of a device is restricted to the case of a steady-state flow of fluid
through it. However, flow components in which the fluid undergoes an unsteady-state
process can also be analyzed thermodynamically. An example of an unsteady-flow process
is the depressurization of a gas cylinder when the valve is opened.
8
All of the devices mentioned in the preceding paragraph can be represented by the
schematic open system shown in Fig. 4.6. This generalized device has a rigid casing
inlet
exit
m
m
area = A
i
state: p
i
, T
i
velocity = V
i
area = A
e
state: p
e
, T
e
velocity = V
e
Q
W
s
Open system
Fig. 4.6 Schematic of an open system
forming a boundary that does no expansion work. However, the unit may perform or
accept shaft work, as do the pump and turbine in Fig. 4.3. It may also exchange heat with
the surroundings, which is the function of the boiler and condenser in the simple steam
power cycle.
The principal difference between the system of Fig. 4.6 and the closed systems
which have been analyzed in Chap. 3 is the presence of inlet and outlet flow ports in the
open system. The boundary of the open system in Fig. 4.6 is composed of the inner
surface of the physical device plus the imaginary meshes on the inlet and outlet ports
through which fluid enters and leaves the unit. The mass flow rate through the device,
expressed in kg/s is assumed to be constant in time. The properties of the fluid at the
entrance and exit are fixed by any two state functions at these locations. Pressure and
temperature are indicated in Fig. 4.6, although any two properties suffice for a single-
phase substance and any one property fixes all others if the fluid is a vapor-liquid mixture.
In addition, the velocities of the fluid entering and leaving are important because they
enter into the First law applied to the open system. High entrance and exit velocities often
arise because the cross sectional areas of inlet and outlet ports may by small.
The volumetric flow rate is the product of the area of the port, A, and the velocity
of the fluid, V:
volumetric flow rate = AV
The volumetric flow rate can be converted to the mass flow rate & m by dividing by the
fluids specific volume, giving:
v
AV
m = & (4.5)
9
In the following development, only steady-state operation is treated, so & m is
constant in time and the same at the inlet and exit ports. These are designated by
subscripts i and e, respectively. However, the quantities on the right hand side of Eq (4.5)
at the inlet may be different from those at the exit port.
Application of the First law to the device illustrated in Fig. 4.6 introduces several
features not involved in the First-law treatment of closed systems. These are:
1. The internal energy (U or u) that appears in equations such as (1.11) and (3.2)
must be supplemented by the kinetic energy of the moving fluid which is & mV
2
. Changes
in gravitational potential energy between the inlet and exit lines should also be added to
the total energy, but for simplicity, this component of the energy is not treated here.
2. The internal energy convected through the boundaries of the open system needs
to be included in writing the First law.
3. In addition to shaft work, pressure-volume work is inherently involved in
pushing the fluid into the inlet and out of the exit line. This work component is called the
flow work.
The total work performed by the system is
& &
(
& &
) W W W W
s fi fe
=
where
&
W
s
is the rate at which the system performs shaft work, and the terms in parentheses
are the inlet and outlet flow work terms. The dots over the work symbols indicate that
rates of performing work (i.e., power) is involved. We have also reverted to the sign
convention of the First law for closed systems: work is positive if performed by the system
and heat is positive if added to the system.
The flow work terms are of the pressure-volume type because they involve force
acting over a distance, or, equivalently, pressure displacing volume. Thus the rate of
performing flow work is the pressure times the volume flow rate, or
&
W
fi
= p
i
A
i
V
i
= & mp
i
v
i
and
&
W
fe
= p
e
A
e
V
e
= & mp
e
v
e
where Eq (4.5) has been used to replace the volume flow rate in terms of the mass flow
rate.
The First law for the steady-state open system equates the net rate of energy
transport across system boundaries at the flow ports to the net rate of energy input in the
form of heat and work exchanged with the surroundings:
( ) [ ]
& / /
& &
( & & m u V u V Q W mp v mp v
e e i i s i i e e
+ =
2 2
2 2
Note that the properties of the fluid inside the system do not enter the statement of the
First law. The reason for this absence is the restriction to steady-state operation.
10
An important simplification of the above equation is obtained by combining the
internal energy terms with the flow-work terms. The combination u + pv is the enthalpy of
the fluid, h, so the First law becomes:
( )
& &
& / / Q W m h V h V
s e e i i
= +
2 2
2 2 (4.6)
In steady-state operation, Eq (4.6) can be divided by the mass flow rate, thereby
converting the heat and shaft work terms to a per-unit-mass basis rather than a per-unit-
time basis:
q w h V h V
s e e i i
= +
2 2
2 2 / / (4.7)
Equation (4.7) is the usual starting point for the subsequent analyses of open
systems, starting with the following example utilizing the steam power plant in Fig. 4.3.
Note that the flow work terms no longer appear because they are contained in the enthalpy
terms. The units of V
2
are m
2
/s
2
. Because a Joule is one kg-m
2
/s
2
, the units of all terms in
Eq (4.7) are J/kg.
Example: Steam power plant
The state of water as it moves around the cycle in Fig. 4.3 is shown in Fig. 4.7, which is
the T - v projection of the equation of state of water (see Fig. 2.6). The abcissa is log(v) because of
the many order of magnitude variation of the specific volume of water between low-pressure steam
and liquid. Some of the water conditions at the numbered positions in the cycle are listed in Fig.
4.7. Using this information and the First law, in conjunction with the steam tables, the analysis of
the power plant performance can be completed. We first use the steam tables to complete the
properties of the fluid between the components of the cycle.
From the pressure and temperature specified at location 2, the superheated steam entering
the turbine is obtained from Table A.3 as: h
2
= 3023.5 kJ/kg.
1
4
3
2
log(v)
T
1 w
4-1
= 4 kJ/kg
2
2 MPa, 300
o
C
3
15 kPa, 90% quality
4 15 kPa, 40
o
C
Fig. 4.7 The steam power plant of Fig. 4.3 superimposed on the T-v EOS of water
The fluid at the turbine exhaust (location 3) is a two-phase mixture. Specification of the pressure
here determines the condition of the saturated liquid and vapor phases from Table A.2, and the
steam quality given provides the mixture enthalpy via Eq (2.20a):
11
h
3
= 225.9 + 0.9 x 2373.1 =2361.8 kJ/kg
Finally, the enthalpy at the exit of the condenser (location 4) is obtained from extrapolation
of the data in Table A.4 to low pressure: h
4
= 167.6 kJ/kg.
With the thermodynamic property information in hand, the First law in the form of Eq
(4.7) can be applied to each component of the power plant. In all cases, the kinetic energy terms
are neglected because V
2
is small compared to the enthalpy terms.
Turbine (2 - 3 in Fig. 4.7)
The heat loss from a large turbine is negligible compared to the shaft work produced.
Therefore, this component can be assumed to operate adiabatically, or q
2-3
= 0. With these
eliminations, Eq (4.7) reduces to:
w
2-3
= h
2
- h
3
= 661.76 kJ/kg
w
2-3
is the shaft work produced by the turbine. It is not the net shaft work of the cycle because
some of the turbine output is used to operate the pump.
Condenser (3-4 in Fig. 4.7)
The condenser is not a work-producing device. Rather it is responsible for supplying the
reject heat of the cycle to the low-temperature reservoir. The condenser completes condensation of
the exhaust steam from the turbine and subcools the liquid. This is necessary because the pump
cannot handle a two-phase mixture. The first law for the condenser is:
q
2-3
= q
C
= h
3
- h
4
= 2194.2 kJ/kg
Pump (4-1 in Fig. 4.7)
The information given in Fig. 4.7 includes the work required to operate the pump. The
First law utilizes this number to determine the enthalpy at the pump outlet:
h
1
= h
4
+ w
4-1
= 167.6 + 4.0 = 171.6 kJ/kg
Boiler (1-2 in Fig. 4.7)
This component is solely responsible for receiving the heat input to the cycle from the hot
reservoir. Inside this unit, the inlet subcooled liquid is heated to saturation, completely vaporized,
and the steam is superheated. The First law for the boiler is:
q
1-2
= q
H
= h
2
- h
1
= 2851.9 kJ/kg
Cycle Efficiency
The net work produced by the cycle is:
w
s
= w
2-3
- w
4-1
= 657.7 kJ/kg
and the efficiency of the cycle, as defined by Eq (4.2), is:
%) 23 ( 23 . 0
9 . 2851
7 . 657
q
w
H
s
= = =
12
4.5 Application of the Second Law to Open Systems
Only in some cases does the Second law provide an equation that is as universally
useful as is the First law in the form of Eqs (4.6) or (4.7). The reason is that any
irreversibilities in the device create entropy that cannot be quantitatively determined from
thermodynamics. The Second law can be applied to the open system in two ways. For
simplicity, steady-state flow through the device is assumed so that the entropy and heat
can be expressed on a per-unit-mass basis.
The first use of the Second law applies Eqs (1.16) and (1.17) (combined for
compactness) to the flowing fluid. The difference in entropy between the outlet and inlet
conditions in Fig. 4.6 is expressed by:
T / q s s s
i e
= (4.8)
where q is the heat added to the system (a unit mass of the flowing fluid) and T is the
system temperature. In the event that the temperature varies with location in the device,
the right hand side of Eq (4.8) would have to be replaced by

j j
T / q , the summation of
increments of heat q
j
added at local fluid temperature T
j
. However, the single-
temperature version is used here for simplicity.
A second method of applying the Second law to the open system is via the total
entropy change version of Eq (1.20). This involves the entropy changes of the
surroundings, s
surr
= -q/T
surr
, where the minus sign appears because heat addition to the
system (q) is heat removal from the surroundings. Using the equality in Eq (4.8) for s of
the system (the moving unit mass of fluid), the total system + surroundings entropy change
is:
s
total
= s + s
surr
= s
e
s
i
q/T
surr
0 (4.9)
The distinction between Eqs (4.8) and (4.9) lies in the level of irreversibility
present. If the process is internally reversible (see Chap. 1, p. 12), the equality in Eq (4.8)
applies. However, internal reversibility does not guarantee total reversibility; if the heat q
is transferred over a nonzero T, an external irreversibility is present. This is revealed by
substituting (from Eq (4.8)) q/T for s
e
s
I
in Eq (4.9):
0
T
q
T
q
s
surr
total
=
Only if the process is both internally reversible and externally reversible (T = T
surr
) is total
reversibility achieved.
Essentially all common applications of the entropy balance are for devices such as
pumps, turbines, nozzles, valves and in-line flow components. A subset of these in-line
13
flow devices operate in a nearly adiabatic manner (q = 0). If they also function reversibly,
then Eq (4.8) becomes:
s
e
= s
i
(4.10)
Equations (4.9) and (4.10) are the Second law analogs of the First law expressed
by Eq (4.7). To the extent that the device can be assumed to operate reversibly, the
Second law provides another equation for thermodynamic analysis of the open system.
Efficiencies of work-producing and work-consuming devices
Even in cases where an adiabatic flow component such as a pump or a turbine is
not reversible, Eq(4.10) provides a means of calculating its efficiency. The efficiency in
this context is the ratio of the actual work produced or consumed to that for perfectly
reversible operation.
Turbine example: What is the efficiency of the turbine in the steam power plant for the inlet and
exhaust conditions given in Fig. 4.7 for this unit?
The first step is to determine the exit steam quality if the turbine were reversible. This is
accomplished using Eq(4.10), in which s
e
= (s
3
)
rev
and s
i
= s
2
. The entropy of the inlet superheated
steam (location 2) at 2 MPa and 300
o
C is determined from Table A.3 as s
2
= 6.766 kJ/kg-K. For
reversible turbine operation, this is also the entropy of the two-phase mixture leaving the unit, or
(s
3
)
rev
= 6.766 kJ/kg-K. Using the specified exhaust pressure of 15 kPa in Table A.2, the entropies
of the saturated liquid and vapor phases are determined. The steam quality for reversible operation
is:
83 . 0
755 . 0 009 . 8
755 . 0 766 . 6
s s
s ) s (
x
f g
f rev 3
rev
=

=
A reversible turbine with the same inlet steam conditions and the same exhaust pressure would
condense 7% more steam than does the actual turbine (see Fig. 4.7).
The exit enthalpy is evaluated using the reversible steam quality and the enthalpies of the
saturated phases in Eq (2.20a):
(h
3
)
rev
= 225.9 + 0.83x2373.1 = 2192.4 kJ/kg
and the reversible work per unit mass of flowing steam is obtained from the first law by:
(w
2-3
)
rev
= h
2
- (h
3
)
rev
= 3023.5 - 2192.4 = 831.1 kJ/kg
The actual work produced was calculated in the example in the previous section for the steam
power plant. Using this value in conjunction with the reversible work calculated above, the
efficiency of the turbine is:
turbine efficiency
w
w
rev
= = =

2 3
2 3
6617
8311
080
( )
.
.
. (80%)
14
Additional problems involving steam turbines in tandem with other components of
a power generation system are Nos. 4.6 and 4.13.
Although not as common as steam turbines, some turbine designs utilize a gas as a
working fluid. Problems 4.9, 4.10 and 4.14 are exercises in analyzing gas turbines.
Reversible Work of a Flow System
If the flow device operates reversibly, the minimum work needed for the task can be
determined by application of the First and Second laws. The First law gives:
h
e
h
i
= q - w
s,rev
where w
s,rev
is the reversible shaft work. The differential of Eq (3.8), dh = Tds +vdp, can
be integrated to yield:

+ = + =
e
i
e
i
e
i
p
p
p
p
s
s
i e
vdp q vdp Tds h h
Because the process is reversible, the Tds integral is the heat added during the process.
From the above two equations, the reversible work is:

=
e
i
p
p
rev , s
vdp w (4.11)
This formula is the open-system analog of the pV work integral in a closed system (see Eq
(3.1)). The work done by the 100% efficient turbine can in principle be calculated from
Eq (4.11) by expressing v as a function of p at constant s using the steam tables. However,
this is a cumbersome method, and the calculation based on the enthalpy change (p. 13) is
much more convenient.
However, if the fluid is a liquid, the specific volume v is approximately independent of
pressure and the reversible work simplifies to:
w
rev
=- v(p
e
- p
i
) (4.12)
Example: What is the efficiency of the pump in the steam power plant cycle of Fig. 4.7?
For compressed liquid water, v ~ 0.001 m
3
/kg. The pumps inlet pressure is 15 kPa and
the outlet pressure is 2 MPa, assuming negligible pressure drop through the boiler. Using these
values in Eq (4.12) yields:
(w
4-1
)
rev
= -0.001(2000 - 15) = -2 kJ/kg
the minus sign appears because of the convention that work done on the system is negative. From
the data given in Fig. 4.7, the actual work required for pump operation is 4 kJ/kg, or the pump
efficiency is only 50%.
15
Flow Devices Without Shaft Work
Numerous categories of common industrial fluid flow equipment operate in
essentially adiabatic fashion and do not exchange work with their surroundings. Throttling
devices such as orifice plates or their more complicated cousins, valves, provide a pressure
drop without any changes in velocity. Nozzles, on the other hand, are designed to speed
up the flow of a gas. These devices are represented in Fig. 4.8.
8 MPa
300
o
C
1.6 MPa
2-phase
(a) Orifice Plate
1 MPa
300
o
C
30 m/s
0.3 MPa
superheated
(b) Nozzle
Fig. 4.8 Typical flow devices that are both adiabatic and do not perform or accept work
The two classes of flow devices differ in two principal ways: fluid velocity changes
and reversibility. In both cases, the open system consists of the section of pipe containing
the device and imaginary surfaces perpendicular to the flow direction upstream and
downstream of the device. The inlet fluid condition is completely fixed (thermodynamic
state and velocity) and the outlet conditions are partially specified. The First law, and if
applicable, the form of the Second law for reversible processes, are used to calculate the
remaining downstream conditions. The mass flow rate is constant in time, so the analysis is
on the basis of a unit of flowing mass.
For an orifice plate, the First law reduces to:
h
e
= h
i
(4.13)
Example: What is the outlet steam quality in Fig. 4.8(a)?
The outlet pressure is specified in this example
*
. The example also implies that some
condensation has occurred. This assumption needs to be verified. Using the upstream pressure and
temperature in the superheated steam Table A.3, the enthalpy here and following the orifice is h
i
=
h
e
= 2785 kJ/kg. This enthalpy is slightly less than the enthalpy of the saturated vapor at the
downstream pressure of 1.6 MPa, which is obtained from Table A.2. Because the outlet water is in
the two-phase region, the temperature is T
e
= 201
o
C, corresponding to saturation at 1.6 MPa. The
quality of the downstream steam is:
x
h h
h
e
e f
fg
=

=

=
2785 859
1935
0 995 .

*
The downstream pressure could be computed from the flow rate, the orifice geometry, and the upstream
steam conditions. However, such a computation involves fluid mechanics, which is not within the scope of
thermodynamics.
16
Suppose, however, that the downstream pressure is specified as 1.0 MPa, In this case, h
e
is
greater than the enthalpy of the saturated vapor, and the downstream steam remains superheated.
To determine the final state of the gas, Table A.3 is entered with the combination p
e
=1.0 MPa, h
e
=2785 kJ/kg. Interpolating in the table, the downstream temperature is 183
o
C.
That flow through an orifice is irreversible seem intuitive because of fluid turbulence (or at
least laminar friction) created by the abrupt reduction in flow cross section. This qualitative
assessment can be verified by comparing the entropies of the gas before and after passage through
the orifice. At 8 MPa and 300
o
C, Table A.3 gives s
i
= 5.79 kJ/kg-K. For the downstream condition
(p
e
= 1.0 MPa, h
e
= 2785 kJ/kg), Table A.3 gives (with interpolation in the 1 MPa section)
s
e
= 6.60 kJ/kg-K. Thus, s
e
> s
i
, proving that gas flow through an orifice is indeed an irreversible
process.
The First law applied to a nozzle is:
h V h V
i i e e
+ = +
2 2
(4.14)
Because of the smooth shape of this device, the flow is approximately reversible. Since the
system is also adiabatic, the Second law provides the additional relation:
s
i
= s
e
(4.15)
In order to relate the constant-entropy condition of this particular of open system
to the changes in pressure and temperature of the fluid, Eq (3.10) is employed. Even
though this equation was derived for changes in a closed system, it is applicable to open
systems as well. The reason is that this equation involves only thermodynamic properties,
and so depends only on the initial and final states of the fluid, not on the process that
caused the change. Another way of viewing isentropic flow is to imagine that the flow
consists of small packets of fluid acting as closed systems undergoing reversible adiabatic
expansion or compression as they move from the inlet to the outlet.
With the subscripts 1 and 2 denoting the nozzle inlet and outlet, respectively, the
isentropic condition s
1
= s
2
converts Eq (3.10) to Eq (3.17). Once T
e
is determined, the
First law can be applied to complete solution of the problem.
Example: What are the exit temperature and velocity of steam in the nozzle of Fig. 4.8(b)?
Applying Eq (3.17) to the conditions shown in Fig. 4.8(b) yields:
T T
p
p
K C
e i
e
i
o
=

1
0 252
573
0 3
10
423 150
.
.
( )
.
Since the enthalpy change h
e
- h
i
= C
P
(T
e
- T
i
) for an ideal gas, Eq (4.14) can be solved for the exit
velocity. From Fig. 2.3, C
P
of steam is 33 J/mole-K, or 1830 J/kg-K. The exit velocity is:
[ ] [ ] s / m 525 ) 423 573 ( 1830 30 ) T T ( C V V
2 / 1
2
2 / 1
e i P
2
i
e
= + = + =
17
The area reduction in the converging nozzle needed to generate this increase in steam
velocity is calculated using Eq (4.5) for steady-state flow:
AV
v
AV
v
i i
i
e e
E
=
The specific volume ratio is obtained from the ideal gas law:
v
v
T
T
p
p
e
i
e
i
i
e
= = =
423
573
10
0 3
2 46
.
.
.
and
14 . 0
523
30
46 . 2
V
V
v
v
A
A
e
i
i
e
i
e
= = =
Other applications of the First and Second laws to flow devices are Nos. 4.7, 4.8,
and 4.13.
18

Potrebbero piacerti anche