You are on page 1of 39

Chapter 3 Power Cycles

Despite recent progress in fuel cell technology which converts the chemical energy of
fuels into electrical energy directly, large scale production of power from fuels involves
the intermediate step of heat generation by combustion. This heat is then converted into
work by a cyclic process in which a working fluid such as water, air or some suitable
chemical absorbs heat at a high temperature and rejects part of this heat at a lower
temperature as shown below. Since the efficiency of converting heat into work is limited
by the second law of thermodynamics, different cycles with different working fluids have
been proposed which strive to approach this theoretical limit as much as possible.

3.1. Fundamentals
System Boundary: Although the heat generated by combustion of a fuel is the most
common heat source for power cycles, other heat sources are also available, such as the
solar radiation collected by concentrating thermal collectors or the heat recovered from
gas turbine exhaust gases. Hence it is important to analyze these cycles independent of
the origin of the heat source in order to evaluate their efficiency in converting heat into
work. For this purpose, the system boundary will be defined such that it is in contact with
the heat source and the heat sink as shown in the figure above. Thus irreversibilities
associated with heat transfer between the working fluid and the heat source/sink are
included but irreversibilities associated with the generation of the heat source itself are
excluded.

Qin

Qout

FIGURE 3.1: System boundary.

45
Carnot, Thermal and Second Law Efficiencies: Thermal efficiency is the work
produced by the cycle divided by the heat input:

η = W/Qin
I

Second law efficiency of the cycle can be calculated as:

η = W/ Wideal = W/(W + LW)


II

where lost work or irreversibility can be calculated according to Gouy-Stodola theorem


as the product of the environmental temperature To and the cyclic integral of dQ/T where
T is the temperature of the source(s) or sink(s) exchanging heat with the working fluid.

For each step in the cycle, the entropy balance for a closed or an open system can be
used to calculate entropy production. Then multiplication of this entropy production with
the environmental temperature gives irreversibility for this step.

For a closed system,

I = To[ ( Sfinal – Sinitial )sys – ∑ Qin/Tsource + ∑ Qout/Tsink ]

For a steady flow, open system,

I = To[∑(ms)out – ∑(ms)in – ∑ Qin/Tsource + ∑ Qout/Tsink ]

Now let’s consider a reversible heat engine whose low temperature sink is the
environment. Then the work produced by this heat engine is the same as the ideal work
and the Carnot and the second law efficiencies are:

η = Wrev/Qin
C
η = Wrev/ Wideal= 1
II
For an actual heat engine operating between the same source and sink as the reversible
one, thermal efficiency is the product of second law and Carnot efficiencies as shown
below:

η = Wactual/Qin
I
η = Wactual/ Wideal = Wactual/ Wrev
II
η = (Wactual/Qin )(Wrev/Wrev ) = (Wactual/Wrev)(Wrev/Qin)
I
η =η η
I II C

46
This relationship is applicable to all heat engines whose low temperature sink is the
environment. Note that if the temperature of the heat source is variable, Carnot efficiency
is not equal to 1 – To/T and should be properly modified as shown in the following
example.
.
Example 3.1: A hot exhaust stream from a process plant is at T1 and Po. Since T1 is
considerably greater than To, it is considered to construct a heat engine which will be
driven by the heat extracted from this stream as it is cooled down to T2. The heat engine
will use the environment as its sink. If this engine were a reversible one, what would be
its thermal efficiency?

Solution:

H1 H2
S1 S2
Q

Qo

To

FIGURE 3.2: Heat engine with a variable temperature heat source.

The engine is shown above. The condition for reversibility can be expressed as:

ΔS + Qo/ To = 0 → Qo = – To ΔS

W = Q – Qo

Q = ΔH = H1 – H2= ∫cPdT

ΔS = S2 – S1 = ∫cPdT/T

Assuming heat capacity is constant, integration yields:

47
Wrev = cP(T1 – T2) – TocPln(T1/T2)

η = Wrev/Q = cP(T1–T2) –TocPln(T1/T2) / cP(T1–T2)


C

Rearranging and simplifying,

η = 1 –Toln(T1/T2) /(T1–T2)
C

Ideal and Non-ideal Cycles: All ideal cycles are internally reversible. Hence all the
irreversibilities are external and associated with heat flow. Since frictional dissipation is
ignored, pipes connecting different components are assumed to have no heat loss and
processes involving work are assumed to be adiabatic, heat addition and heat rejection
processes are the only sources of irreversibility in ideal cycles. Provided that source and
sink temperatures can be properly defined, second law efficiency of the entire cycle can
be easily calculated and it is not necessary to know the entropy of the working fluid.
A suitable sink temperature is usually the ambient temperature which is constant whereas
the source temperature is usually not constant. Furthermore, assigning a numerical value
to the source temperature is difficult in the case of internal combustion engines where
actual combustion is replaced by external heat addition to simplify the analysis. One
approach is to assign the maximum temperature of the cycle as the source temperature.
Obviously, second law efficiencies calculated in this manner are only approximate and
are useful for a quick comparison of different cycles.
Calculating irreversibilities of heat addition and heat rejection steps separately requires
application of the entropy balance for a closed or steady flow, open system depending on
the characteristics of the cycle. It is customary to employ the steady flow entropy balance
for cycles such as the Rankine and the Brayton cycles whereas cycles such as the Otto
and the Diesel cycles are analyzed as closed systems. In either case, entropy of the
working fluid must be known.
Ideal cycles which are internally reversible do not exist in real life since fluid flow
always involves frictional dissipation. For incompressible fluids flowing in pipes
connecting different pieces of equipment, frictional dissipation is usually ignored by
assuming pressure drops are negligible. However, for compressible flow through
equipment such as compressors and turbines where the ratio between inlet and outlet
pressures is usually greater than 5, frictional dissipation can no longer be ignored, i.e.,
compression and expansion processes can not be assumed isentropic.
Now let us consider non-ideal cycles where compression and expansion processes are not
isentropic. The second law efficiency of the cycle can still be calculated in the same
manner, i.e., by the cyclic integral of dQ/T since the effect of internal irreversibilities are
reflected in the numerical values of heat flows. For example, if the turbine has an
isentropic efficiency of less than 100%, less work than the ideal cycle is produced with
the consequence of rejecting more heat than the ideal cycle.

48
3.2. Gas Power Cycles
In heat engines, working fluids undergo cyclic processes, extracting heat from a hot
source and rejecting heat to a low temperature sink (usually the environment), producing
some mechanical work in the way (usually shaft work). Gases are the best working
substances because they can easily convert thermal energy into mechanical energy by
compression or expansion, whereas liquids have little compressibility. Gas power cycles
are those where the working substance stays all the time in the gas phase, and vapor
power cycles are those where the gas condenses to liquid in some part of the cycle.
Many gas cycles have been proposed, and several are currently used, to model real heat
engines: the Otto cycle (approximates the actual gasoline engine), the Diesel cycle
(approximates the actual diesel engine), the mixed cycle (a hybrid of the Otto and the
Diesel that is better than both), the Brayton cycle (approximates very well the actual gas
turbine engine), and the Stirling cycle. We will restrict the analysis to the so called ‘air
standard’ model, which assumes air as an ideal gas with temperature-independent
properties is the working fluid, neglecting fuel addition effects (the air to fuel mass ratio
is around 15:1 in Otto engines, 30:1 in Diesel engines, and 60:1 in Brayton engines), and
considers an equivalent heat addition.
The Otto Cycle: In the ideal air standard Otto cycle, air with temperature independent
properties undergoes four main processes as shown in Figure 3.3: (i) isentropic
compression(1→2), (ii) constant volume heat addition(2→3), (iii) isentropic
expansion(3→4) and (iv) constant volume heat rejection(4→1).

3 3
P T
Qin
Wout Wout
Qin
4
2 4 2
Win
Win Qout
Qout
1 1

V S

FIGURE 3.3: The ideal Otto cycle.

Assuming a thermal reservoir is available at the maximum temperature of the cycle (T3)
and the heat sink is at T1, it can be shown that the thermal and second law efficiencies are
given by:

49
η = 1 – T1/T 2 η = (1 – T1/T 2)/( 1 – T1/T 3)
I II

Hence from a thermodynamics point of view, the main parameters are T3 and the
compression ratio V1/V2 which determines T2. In real Otto engines, the typical range of
this ratio varies between 8 and 11 and may go up to 14 in direct injection engines.
The Diesel Cycle: In the ideal air standard Diesel cycle there are four main processes: (i)
isentropic compression(1→2), (ii) constant pressure heat addition(2→3), (iii) isentropic
expansion(3→4) and (iv) constant volume heat rejection(4→1) as shown in Figure 3.4.

P Qin T
Qin 3
2 3

Wout Wout
2

Win Win 2 4
4
2
21 Qout
Qout
1

V S

FIGURE 3.4: The ideal Diesel cycle.


The main parameters of the Diesel cycle are the compression ratio which varies from 16
to 22 and the cutoff ratio, V3/V2.

Example 3.2: An air standard Otto cycle has a compression ratio of 8. Intake pressure
and temperature are 1 atm and 300 K, respectively. Air to fuel ratio is adjusted such that
the heat input in the combustion stroke is 2,000 kJ/kg of air. Find the temperature and
pressure at the end of each stroke, net work, heat rejection, thermal efficiency and second
law efficiency of this cycle. Then resolve this problem for an air standard Diesel cycle
which has the same compression ratio, the same intake pressure and temperature and the
same heat input. Which cycle has a higher thermal efficiency, higher maximum
temperature, higher exhaust temperature and higher second law efficiency?
Air is assumed as an ideal gas with constant specific heats of cP = 1.15 kJ/(kg-K) and
cV = 0.862 kJ/(kg-K)

50
Solution:
k = cP/cV = 1.150/0.8620 = 1.3345
For the Otto cycle: Compression ratio is V1/V2 and for an isentropic compression:

T2 = T1(V1/V2)k-1 and P2 = P1(V1/V2)k

T2 = 300( 8 )0.3345 = 601.5 K P2 = 1(8)1.3345 = 16.04 atm.

T3 = T2 + qin/cV = 601.5 + 2000/0.8620 = 2921.7 K

P3=P2T3/T2= 16.04(2921.7/601.5) = 77.92 atm

T4 =T3(V3/V4)k–1 = 2921.7(1/8)0.3345 = 1457.3 K

P4 = P1(T4/T1) = 1(1457.3/300) = 4.86 atm

qin = 0.862(2921.7 – 601.5) = 2000 kJ/kg

qout= 0.862(1457.3 – 300) = 997.6 kJ/kg


Thermal Efficiency = (2000 – 997.6 )/2000 = 0.5012

In order to calculate the second law efficiency, we will assume that the heat source is
available at the maximum temperature of the cycle and the heat sink is at 300K.

I = (–2000/2921.7 + 997.6/300 )300 = 791.6 kJ/kg

Second Law Efficiency = ( 2000 – 997.6 ) / ( 1002.4 + 791.6 ) = 0.559

η = η η = 0.5012*(1–300/2921.7) = 0.5016
I II C
Irreversibility can also be calculated for each step one by one using:

I = To[ ( Sfinal – Sinitial )sys – ∑ Qin/Tsource + ∑ Qout/Tsink ]

For the heat addition step:

I = 300[ 1.15ln(2922/601) – 0.2857ln(78/16) – 2000/2922] = 204.4 kJ/kg

For the heat rejection step:

I = 300[1.15(300/1454) – 0.2857ln(1/4.86) + 997.6/300] = 587.8 kJ/kg

For the entire cycle:

51
I = 204.4 + 587.8 = 792.2. kJ/kg

which is almost identical to the result found by the cyclic integral of ∫ dQ/T.

Note that more than 70% of the irreversibility is due to heat rejection. In fact, if it were
possible to utilize this heat to run a Carnot engine with a heat sink at To, irreversibility
will dramatically decrease since work equal to:

w = cV(T3–T4) – TocVln(T3/T4)

w = 0.862(1457.3 – 300) – 300x0.862ln(1457.3/300)= 587.8kJ/kg


will be produced by this engine. Then irreversibility will reduce to:

I = 791.6 – 587.8 = 203.8 kJ/kg

or

I = (–2000/2921.7 + 0.862 ln(1457.3/300) ) 300= 203.8 kJ/kg

Second Law Efficiency = 1002.4 /(1002.4 + 202.7) = 0.8318

Note that the installation of a Carnot engine is equivalent to rejecting heat to a variable
temperature heat sink. The variable temperature of this sink is equal to the temperature
of the working fluid, i.e., heat rejection takes place reversibly.

For the Diesel cycle:

T2 and P2 are the same → P2= P3= 16.04 atm.

T3 = T2 + qin /cP = 601.5 +2000/1.150 = 2341 K

V3/V2=T3/T2= 2341/601.5= 3.89

V3/V4= (V3/V2)(V2 /V1) = 3.89(1/8) = 0.4864

T4= T3(V3/V4)k-1 = 2341(0.4864)0.3345 = 1839.8 K

qout = cV (T4–T1) = 0.862 (1839.8 – 300) = 1327 kJ/kg


Thermal Efficiency = ( 2000 –1327 )/ 2000 = 0.3365

I = (–2000/2341 + 1327/300 )300 = 1069.8 kJ/kg

Second Law Efficiency = ( 2000 –1327 )/( 2000 –1327 + 1069.8 ) = 0.3862

52
η = η η = 0.3862(1–300/2341) = 0.3367
I II C
Irreversibility can also be calculated for each step one by one using:
I = To[ ( Sfinal – Sinitial )sys – ∑ Qin/Tsource + ∑ Qout/Tsink ]
For the heat addition step:
I = 300[ 1.15ln(2341/601) – 2000/2341] = 212.8 kJ/kg
For the heat rejection step:
I = 300[ 1.15ln(300/1840) – 0.2857 ln(1/6.13)+ 1327/300] = 856.7 kJ/kg
For the entire cycle:
I = 212.8 + 856.7 = 1069.5 kJ/kg
which is almost identical to the result found by the cyclic integral of ∫ dQ/T.
If rejected heat were utilized to drive a Carnot engine,
w = 0.862(1839.8 – 300) – 300x0.862ln(1839.8/300) = 858 kJ/kg
will be produced and the second law efficiency will improve to
Second Law Efficiency = ( 2000 –1327 + 858 )/ ( 2000 –1327 + 1069.8 ) = 0.8784
Note that since the heat source is at a lower temperature and heat rejection is at higher
temperatures, improvement in the second law efficiency for the Diesel engine is even
greater.

Diesel and Otto cycles with the same heat input, same compression ratio and same intake
conditions are compared in the figure below. Otto cycle has a higher maximum
temperature, a higher thermal efficiency and a higher second law efficiency while the
Diesel cycle rejects heat at a higher temperature.
At the end of the compression stroke (Point 2 in the figure) the working fluid is at the
same conditions in both cycles. However, the Otto cycle reaches a higher temperature
(Point 3) at the end of heat addition step simply because cP is greater than cV. Note that
the areas (a1234ba) and (a1234ca) represent the heat input and are therefore equal. On
the other hand, the area (a14ba) is the heat rejected by the Otto cycle whereas the area
(a14ca) is the heat rejected by the Diesel cycle. Since the latter is greater than the former
by an amount represented by the blue dotted area (44cb4) in the figure, net work of the
Diesel cycle is also less by the same amount. Consequently, the thermal efficiency of the
Otto cycle is higher.
For both of these cycles the major source of irreversibility is the heat rejection step. Since
heat rejection takes place at higher temperatures in the Diesel cycle, its total
irreversibility is also higher. Consequently, the second law efficiency of the Diesel cycle
is less than that of the Otto cycle.

53
T 3
Otto
3
Diesel
Common

4
4

a b c
S

FIGURE 3.5: Comparison of Otto and Diesel cycles.

3
T

1 P2
Qin Wout
PP
2
P2
4
P2

2
Win
1 P1 Qout

S
FIGURE 3.6: The ideal Brayton cycle.

54
The Brayton Cycle: In the ideal air standard Brayton cycle, there are four main
processes as shown in Figure 3.6: (i) isentropic compression(1→2), (ii) constant pressure
heat addition(2→3), (iii) isentropic expansion(3→4) and (iv) constant pressure heat
rejection(4→1). The main parameters are T3 and the compression ratio P2/P1. In a real
cycle, two additional parameters, namely the isentropic efficiencies of compression and
expansion are also involved. It can be shown that for the ideal cycle the thermal and
second law efficiencies are given by:

η = 1 – T1/T 2 η = (1 – T1/T 2)/( 1 – T1/T 3)


I II

Example 3.3: A gas turbine power plant operating on a real Brayton cycle has a pressure
ratio of 6 and a turbine inlet temperature of 1400oF. If the compressor and the turbine
have isentropic efficiencies of 89 and 86%, respectively, what will be the thermal and
second law efficiencies? The environmental temperature is 520oR. Assume that a heat
reservoir at the maximum temperature of the cycle is available.

k =1.40 cP =0.24Btu/(lb-oF)
Solution: The power plant and the corresponding Brayton cycle are shown in Figure 3.7.

3
T
CC 1 P2
Qin Wout

2 FUEL
3 4
C T G 2
G
Win
1 4 P1 Qout
1
C Compressor
CC Combustion Chamber
G Generator
T Turbine
S

FIGURE 3.7: Example 3.3.

The compressor work will increase and the turbine work will decrease due to internal
irrversibilities. As shown in the figure below, these internal irreversibilities also lead to a
decrease in the heat input (dotted area on the left) and an increase in the heat rejected
(dotted area on the right). Since the source and the sink temperatures do not change, the
irreversibility calculated by the cyclic integral of dQ/T will also increase.

If the compression were isentropic:

55
T2 = T1(P2/P1)(k-1)/k → T2 = 520(6)0.2857 = 867.6 oR

The actual compressor outlet temperature can be easily calculated using the compressor
isentropic efficiency:

T2 = 520 + (867.6 – 520)/0.89 = 910.6oR

w in = cP(T2 – T1) → w in = 0.24(910.6 – 520) = 93.7 Btu/lb

qin = cP(T3 – T2) → qin = 0.24(1860 – 910.6) = 227.8 Btu/lb


If the expansion were isentropic:

T4 = T3(P1/P2)(k-1)/k → T4 = 1860(1/6)0.2857= 1115 oR

The actual turbine outlet temperature can also be calculated using the turbine efficiency:

T4 = 1860 – (1860 – 1115)0.86 = 1219oR

qout = cP(T4–T1) → qout = 0.24(1219–520) = 167.8 Btu/lb

w out = cP(T3–T4) → wout = 0.24(1860–1219) = 153.8 Btu/lb


Thermal Efficiency = (153.8 – 93.7)/227.8 = 0.2640

I= ( –227.8/1860 + 167.8/520)520 = 104.11 Btu/lb

Second Law Efficiency = (153.8 –93.7)/(153.8 –93.7 + 104.11) = 0.3660

η = η η = 0.3660*(1–520/1860) = 0.2637
I II C
Note that compression work is more than 50 % of the expansion work, significantly
reducing the net work produced by the cycle.

Irreversibility can also be calculated for each step one by one using:

I = To[∑(ms)out – ∑(ms)in – ∑ Qin/Tsource + ∑ Qout/Tsink ]

For the heat addition step:

I = 520[ 0.24ln(1860/911) – 227.8/1860] = 25.45 Btu/lb

For the heat rejection step:

I = 520[ 0.24ln(520/1219) + 167.8/520] = 61.47 Btu/lb

56
For the compression step:

I = 520[ 0.24ln(911/520) – 0.06899ln6 ] = 5.64 Btu/lb

For the expansion step:

I = 520[ 0.24ln(1219/1860) + 0.06899ln6 ] = 11.54 Btu/lb

For the entire cycle:

I = 5.64 + 11.54 +25.45 + 61.47 = 104.1 Btu/lb

Again, this result is almost identical to the one found by the cyclic integral of dQ/T.

The thermal and second law efficencies of the Brayton cycle can be improved
significantly by the addition of a regenerative heat exchanger in which the compressor
outlet is heated with the hot gases leaving the turbine. Thus heat rejection and external
heat addition are reduced by the same amount which corresponds to the heat duty of the
regenerative heat exchanger.

Example 3.4: Consider the same Brayton cycle described in Example 3.3. A
regenerative heat exchanger is installed in which the compressor outlet is heated to
1140oR using the turbine outlet. Calculate the new thermal and second law efficiencies
and the external and internal irreversibilities.

Solution: The modified cycle is shown in the figure below. Note that the colored areas
in this figure represent the heat gained by the compressor outlet (the reduction in heat
addition) and the heat lost by the turbine outlet (the reduction in heat rejection).and they
are equal.

Note that the inlet and outlet temperature for the compressor and the turbine are not
affected. The installation of the regenerative heat exchanger changes only the combustor
inlet temperature and the temperature of the exhaust gases.

The exchanger duty can be easily calculated as :

q hex = cP(T3 – T2)= 0.24(1140 – 910.6) = 55.06 Btu/lb


The hot gases from the turbine leave the exchanger at T6 :

T6 = T5 – q hex/cP

T6= 1219–55.06/0.24 = 989.6o R

qin = cP(T4 – T3) = 0.24(1860 – 1140) = 172.8 Btu/lb

qout = cP(T6 – T1) = 0.24( 989.6–520)= 112.6 Btu/lb

57
2 3 4

5
1 6

Heat addition
Heat rejection
Regeneration
Work
T 4

3 5
6
2

FIGURE 3.8: Brayton cycle with regeneration.

I = 520(112.6/520 – 172.8/1860) = 64.32 Btu.lb

w in and w out will not change, therefore


Thermal Efficiency = 60.1/ 172.8 = 0.3478

Second Law Efficency = 60.1/( 60.1 + 64.32) = 0.4830

Irreversibility can also be calculated for each step one by one:

I = To[∑(ms)out – ∑(ms)in – ∑ Qin/Tsource + ∑ Qout/Tsink ]

58
For the heat addition step:

I = 520[0.24ln(1860/1140) – 172.8/1860] = 12.79 Btu/lb

For the heat rejection step:

I = 520[ 0.24ln(520/989.6) + 112.6/520] = 32.29 Btu/lb

Irreversibilities of the compression and expansion steps will not change. Therefore the
irreversibility due to heat transfer in the heat exchanger is:

I= 64.32 – ( 12.79 + 32.29 + 5.64 + 11.54 ) = 2.06 Btu/lb

Internal irreversibility = 5.64 + 11.54 + 2.06 = 19.24 Btu/lb

External irreversibility = 12.79 + 32.29 = 45.08 Btu/lb

Note that although it is reduced by almost 50%, heat rejection is still the major
irreversibility component and despite operating with a considerable temperature
difference, the contribution of the heat exchanger to the total irreversibility of the cycle is
minimal.
The Stirling Cycle: Paths of (i) isobaric, (ii) isothermal, (iii) adiabatic and (iv) isochoric
processes on a P-V diagram are shown in Figure 3.9. It is observed that isotherms are less
steep than adiabats. Therefore if adiabatic compression and expansion steps of a cycle are
replaced by isothermal ones, such a cycle will produce more work than the original cycle
for the same compression ratio, the same intake conditions and the same maximum
temperature, as shown ın Figure 3.10.

Isobaric process
Isothermal process
Adiabatic process
Isochoric process

FIGURE 3.9: Paths of different processes.

59
P 3
Adiabatic
Isothermal
Common

2
4

2
4

11

FIGURE 3.10: Isothermal vs. adiabatic compression and expansion.

Red dotted areas in this figure show extra work produced by switching to isothermal
expansion and compression. Unfortunately, additional heat must also be added during
isothermal expansion and the thermal efficiency of this cycle is still going to be lower
than the original cycle. However, for a working fluid with constant heat capacities, the
amounts of heat added and removed at constant volume are now equal. Therefore if heat
removed after the expansion step can somehow be recycled back to the cycle after the
compression step, heat input from an external source will be needed only during the
isothermal expansion and thermal efficiency will improve significantly. Such a cycle has
been proposed by Stirling who also invented the heat engine by the same name. The
unique feature of this engine which distiguishes it from other heat engines is the existence
of an internal heat exchanger to achieve the regenerative heat transfer..

The ideal Stirling cycle with constant volume regenerative heat transfer processes is
shown in the figure below. The cycle consists of four main processes: (i) isothermal
compression (1→2) (ii) regenerative heat addition (2→3) (iii) isothermal expansion
(3→4) and (iv) regenerative heat removal (4→1). It can be shown that with perfect
regeneration and an ideal gas as the working fluid, the thermal efficiency of the ideal
cycle is the same as that of a Carnot engine operating between Tmax and Tmin.

60
P 3
T Tmax 3 4

4
2 Tmin

2 1
1

V S

FIGURE 3.11: The ideal Stirling cycle.

The ideal Stirling cycle is compared with the ideal Otto cycle in the figure below. The
dotted areas in this figure represent the extra work produced by the Stirling cycle. Heat
addition and removal at constant volume are represented by areas (a2234fcba) and
(b144edcb), respectively. Since these areas are equal, with perfect regeneration the area
(cf434edc) represents the only heat input to the Stirling cycle from an external source.
The heat rejection is represented by the area (a21ba) which is equal to the area (cfedc).
Therefore the area (f434ef) must be equal to the net work produced by the cycle. It is
seen from the figure that this area is also equal to heat input (cf434edc) multiplied by a
factor of (Tmax – Tmin)/ Tmax, confirming the validity of Carnot efficiency.

T
Tmax 3 4

Stirling
Otto
Common
2 4

Tmin 2 1 f e

a b c d
S

FIGURE 3.12: Comparison of Stirling and Otto cycles.

61
Example 3.5: Calculate the thermal and second law effiencies of an ideal Stirling cycle
with a compression ratio of 8 and a maximum temperature of 2922 K.The temperature of
the environment is 300K. Working fluid is an ideal gas with a constant heat capacity of
cV = 0.862 kJ/(kg-K).

Solution:
q12 = RTminln(V2/V1) = 0.2887(300)ln(1/8) = – 180 kJ/kg
q23 = cV(T3 – T2) = 0.863(2922 –300) = 2260 kJ/kg = – q41
q34 = RTmaxln(V4/V3) = 0.2887(2922)ln(8) = 1754 kJ/kg

w net = 1754 + (–180 ) = 1574 kJ/kg

Thermal Efficiency = Wnet/ q34 = 1574/1754 = 0.8974

Carnot Efficiency = 1 – 300/2922 = 0.8974

Second Law Efficiency = 1574/[1754(1 – 300/2922)] = 1.0000

Note that without regeneration, these efficiencies would be 0.3921 and 0.4370,
respectively.

The preceding example illustrates the significance of regeneration in improving the


efficiency of the Stirling cycle. Hence the regenerator is the key component of any heat
engine utilizing this cycle. Obviously a perfect regeneration is not possible since it will
require an infinitely large heat transfer area and the absence of friction. The design
challenge for the Stirling engine regenerator is to provide sufficient area which will allow
transfer of heat across as small a temperature difference as possible without introducing
too much internal flow resistance. These inherent design conflicts are just one of many
factors which limit the efficiency of practical Stirling engines. A typical design is a stack
of fine metal wire meshes with low porosity, and with the wire axes perpendicular to the
gas flow to reduce conduction in that direction and to maximize convective heat transfer.

The working fluid in the Stirling cycle operates in a closed loop, therefore a gas such as
helium or hydrogen with better heat transfer characteristics may also be used instead of
air. The external heat source may be provided by the combustion of a fuel and since
combustion products to not mix with the working fluid, fuels that would damage internal
combustion engines may also be employed. Other external heat sources include
concentrated solar energy, nuclear energy, geothermal energy and waste heat.
The Ericsson Cycle: This is another regenerative cycle where the working fluid operates
in a closed loop. The ideal Ericsson cycle has the same efficiency as the Carnot cycle and
consists of four main processes: (i) isothermal compression (ii) regenerative heat addition
at constant pressure (iii) isothermal expansion and (iv) regenerative heat removal at

62
constant pressure. Hence the Ericsson cycle is very similar to the Stirling cycle, the only
difference is isobaric regeneration instead of an isochoric one. As depicted in Figure 3.13,
the Brayton cycle employing multistage compression with intercooling and multistage
expansion with regeneration approaches the performance of the Ericsson cycle.

4 9 5

2 6

6 7
2

8
10
1

T
Tmax 5 7

9 6 8

4 2 10
Tmin
3 1
External heating
External cooling
Regeneration

S
FIGURE 3.13: The Brayton cycle with multistage compression and expansion.

63
3.3. Vapor Power Cycles
The gas power cycles which employ internal combustion (Otto, Diesel and Brayton) have
two important features in common: The working fluid is air and the major source of
irreversibility is the heat rejection step. This is due to the fact that the working fluid must
be exhausted during the heat rejection step since it contains combustion products and
these exhaust temperatures are quite high. The irreversibility of the heat rejection step can
be significantly reduced if heat is rejected at a constant temperature which is much closer
to the temperature of the surroundings. Condensation occurs at a constant temperature
with much better heat transfer characteristics and therefore requires a smaller temperature
difference as the driving force. Furthermore, pressurizing a condensed phase consumes
much less work than compressing a gas. These factors favor the utilization of a working
substance which undergoes phase change during the cycle.
The Rankine Cycle: The ideal Rankine cycle, whose several variants are implemented
in modern steam power plants Rankine cycle consists of two isentropic and two isobaric
steps as depicted in Figure 3.14:

(i) Superheated vapor expands isentropically from P1 to P2 through a turbine (State


1→ State 2)
(ii) Heat is rejected in a condenser at constant pressure P2. (State 2→ State 3)
(iii) Saturated liquid is pumped isentropically into a boiler. (State 3→ State 4)
(iv) Liquid is heated in the boiler at constant pressure P1. (State 4→ State 1)

T
P1
1

4 Qin W

P2

3 2
Qout

FIGURE 3.14: The ideal Rankine cycle.

64
Note that the work consumed to pump the liquid from the condenser into the boiler is
usually negligible compared to the work produced in the turbine.
In the Rankine cycle, the working fluid is usually water which circulates in a closed loop
since it is separated from the heat source by a physical boundary. For example, in a fossil
fuel–fired plant, the combustion gases are separated from steam by boiler tube walls.
Example 3.6: Steam enters the turbine of a power plant operating on an ideal Rankine
cycle at 247 psia and 700oF. The condenser pressure is 0.949 psia. Calculate the thermal
and second law efficiencies. Temperature of the environment is 60oF.

Solution: The operating conditions and corresponding working fluid properties are given
below:

Point # T( oF ) P(psia ) h( Btu/lb ) s( Btu/lb–oR )

1 700 247 1372 1.700


2 100 0.95 949 1.700
3 100 0.95 68 0.123
4 100† 247 68† 0.123

Pump work is assumed to be negligible

qin = h1 – h4 = 1372 – 68 = 1304 Btu/lb

qout = h2 – h3 = 949 – 68 = 881 Btu/lb

w = h1 – h2 = 1372 – 949 = 423 Btu/lb


Thermal Efficiency = 423/ 1304 = 0.324

We assume a thermal reservoir at 700oF is available:

I = ( 881/520 – 1304/1160 )520 = 296.4 Btu/lb

Second Law Efficiency = 423/(423 + 296.4 ) = 0.5880


η = η η = 0.5880*(1–520/1160) = 0.3244
I II C

Irreversibility can also be calculated for each step one by one using:

I = To[∑(ms)out – ∑(ms)in – ∑ Qin/Tsource + ∑ Qout/Tsink ]

For the heat addition step:

65
I = 520[ (1.700 – 0.123) – 1304/1160] = 235.5 Btu/lb
For the heat rejection step:
I = 520[(0.123 – 1.700) + 881/520 ] = 60.96 Btu/lb
For the entire cycle:
I = 235.5 + 60.96 = 296.4 Btu/lb
which is identical to the result found by the cyclic integral of dQ/T.
External irreversibility due to heat rejection can also be calculated as:
I = 881[(560 – 520)/(560x520)]520 = 62.93 Btu/lb
Note that in contrast with the previous cycles, more than 75% of the external
irreversibility is due to heat addition.

A Rankine cycle where the turbine and the pump are not isentropic is shown below. Since
actual turbine work is less than the isentropic work, more heat will be rejected in the
condenser, as shown in the figure below where the blue dotted area represents the
additional heat rejection due to the internal irreversibilities in the turbine. Similarly,
irreversibilities in the pump will lower the heat addition (red dotted area in the figure),
but this is usually neglected. This means that points 4 and 4’ are identical and,
accordingly, red dotted area vanishes.

T 1

4
4’
4 2

3 2’
2
2

FIGURE 3.15: A non-ideal Rankine cycle.

66
Example 3.7: Consider the same Rankine cycle described in Example 3.6. If the turbine
isentropic efficiency is 88%, what will be the new thermal and second law efficencies ?

Solution: The operating conditions and corresponding working fluid properties are
shown in the table below where changes are highlighted:

Point # T( oF ) P(psia ) h(Btu/lb) s(Btu/lb–oR)

1 700 247 1372 1.700


2 >100 0.95 1000 1.790
3 100 0.95 68 0.123
4 100 247 68 0.123

qin = h1 – h4 = 1372 – 68 = 1304 Btu/lb

qout = h2 – h3 = 1000 – 68 = 932 Btu/lb

w = h1 – h2 = 1372 – 1000 = 372 Btu/lb

Thermal Efficiency = 372/ 1304 = 0.2852

I = ( 932/520 – 1304/1160 )520 = 347.4 Btu/lb

Second Law Efficiency = 372/(372 + 347.4) = 0.5171


η = η η = 0.5171*(1–520/1160) = 0.2853
I II C
Internal irreversibility = To(s1 – s2) = 520(1.790 – 1.700) = 46.8 Btu/lb
External irreversibility due to heat addition will not change, hence external irreversibility
due to heat rejection will be:
I = 347.4 – 235.5 – 46.8 = 65.1 Btu/lb
This value is in close agreement with
I = [932(560 –520)/(560x520)] 520 = 66.6 Btu/lb
or
I = 520[(0.123 – 1.790) + 932/520 ] = 65.2 Btu/lb
Note that the internal irreversibility in the turbine also affects external irreversibility due
to heat rejection, but irreversibility due to heat addition is still the major source of
irreversibility. This is due to the fact that the average temperature driving force for heat

67
addition (working fluid is heated from 100oF to 700oF by a constant temperature source at
700oF) is much greater than the driving force for heat rejection which takes place
between the working fluid at 100oF and a heat sink at 60oF.

In order to improve its performance, several modifications are introduced into the basic
Rankine cycle. One such modification involves passing the high pressure turbine outlet
through the boiler once again prior to the second stage of expansion, a process known as
reheat. In this configuration, more work is produced per unit amount of the working fluid.
Hence for a given power output, flow rate of steam is lower than that of the basic cycle
which implies that equipment sizes will also be smaller. Another advantage the reheat
configuration offers is the smaller liquid fraction in the low pressure turbine outlet.
The thermal efficiency may be higher or lower than that of the basic cycle. There are no
additional components introducing additional internal irreversibilities and external heat
addition is at a higher average temperature. Then the thermal efficiency will be higher
than that of the basic cycle if the heat rejection occurs at the same temperature. However,
if the second turbine outlet is superheated, average temperature of heat rejection increases.
Depending on the extent of superheat, this phenomenon may sometimes lead to a
decrease in thermal efficiency.

HP LP
TURBINE TURBINE

1 2 3 4

T
1 3
B
0
I
L
E
R
2
6
5 4 6
7

S
CONDENSER
5
PUMP

FIGURE 3.16: Rankine cycle with reheat.

68
Another highly common modification of the basic Rankine cycle involves preheating the
boiler feedwater with steam extracted from the turbine at an intermediate pressure. In this
configuration, less work is produced per unit amount of the working fluid since flow rate
through the second turbine is reduced by the amount of extracted steam. On the other
hand, the thermal efficiency of this regenerative system is expected to increase since heat
is now added at a higher average temperature, increasing the Carnot efficiency. However,
additional internal irreversibilities are introduced due to the heat transfer between
extracted steam and boiler feedwater. Therefore poor design of the regeneration system
may actually lead to a decrease in the second law efficiency to the extent that thermal
efficiency also decreases.
There are several variants of regenerative Rankine cycles. The one depicted in the next
figure involves a direct contact heat exchanger known as an open feedwater heater where
extracted steam is injected into boiler feedwater at constant pressure. Note that a second
pump is necessary to raise the pressure of the feedwater leaving the heater from the
intermediate level to the boiler level.

HP LP
TURBINE TURBINE
1
2 4

T
1

B
0
I
L
E
7 R
3 2
6
5 7
4 PUMP 3
H
E
X

S 6
5 CONDENSER

PUMP

FIGURE 3.17: Regenerative Rankine cycle with open feedwater heater.

69
Heat exchangers known as closed feedwater heaters where there is no direct contact
between extraction steam and boiler feedwater are also frequently employed in
regenerative Rankine cycles. In this configuration the hot and cold sides of the exchanger
can be operated at different pressures enabling the elimination of the second circulation
pump if desired. When this is the case, extraction steam leaving the heater is passed
through an expansion valve to reduce its pressure down to the condenser level as shown
in Figure 3.18.

HP LP
TURBINE TURBINE
1
2 4

T
1

B
0
I
L
E
R
8 3 2
7
8
6
5 4 H
E
X

7 3 5
S
CONDENSER

PUMP 6

FIGURE 3.18: Regenerative Rankine cycle with closed feedwater heater.

Example 3.8: Steam enters the turbine of a power plant operating on a simple
Rankine cycle at 250 psia and 700oF. The condenser pressure is 1.0 psia. This cycle is
modified by introducing reheat at 110 psia. The operating conditions and corresponding
properties of the working fluid for the modified cycle are given in the following table.
Calculate the thermal and second law efficiencies of the modified cycle and compare
with those of the basic cycle.

Note that turbines are isentropic, pump work and pressure drops through heat exchange
equipment are neglected. Enthalpies of compressed liquids are assumed to be equal to
that of the saturated liquid at the same temperature.

70
Point # T( oF ) P(psia ) h( Btu/lb ) s(Btu/lb–R) State

1 700 250 1371.6 1.6975 Vapor


2 500 110 1278.3 1.6975 Vapor
3 700 110 1378.9 1.7928 Vapor
4 101.74 1 1001.8 1.7928 Liq + Vapor
5 101.74 1 69.73 0.1326 Satd. liquid
6 101.74 250 69.73 0.1326 Liquid
7 101.74 1 948.3 1.6975 Liq + Vapor

Solution: For the original cycle(17561 in Figure 3.14):

W = 1371.6 – 948.3 = 423.3 Btu

Qin = 1371.6 – 69.73 = 1301.9

Qout = 948.3 – 69.73 = 878.6

Thermal Efficiency = 423.3/1301.9 = 0.3251

I = (878.6/520 – 1301.9/1160 )520 = 294.95

Second Law Efficiency = 423.3/(423.3 + 294.95) = 0.5893

For the reheat cycle(1234561in Figure 3.14):

W = 1371.6 – 1278.3 + (1378.9 – 1001.8) = 470.4 Btu

Qin = 1371.6 – 69.73 + (1378.9 – 1278.3) = 1402.5 Btu

Qout = 1001.8 – 69.73 = 932.1 Btu

I = 520(932.1/520 – 1402.5/1160) = 303.4 Btu

Thermal Efficiency = 470.4/1402.5 = 0.3354

Second Law Efficiency = 470.4/(470.4 + 303.4) = 0.6079


The second turbine outlet is not superheated and both efficiencies are higher than those of
the basic cycle.

71
Example 3.9: The basic Rankine cycle described in Example 3.8 is modified by
installing a closed feedwater heater where boiler feedwater is heated with steam extracted
from the turbine at an intermediate pressure of 110 psia. The extracted steam leaves the
heater as a saturated liquid and the boiler feedwater is heated to the temperature of the
condensing steam. The operating conditions and corresponding properties of the working
fluid for the modified cycle are given in the following table. Calculate the thermal and
second law efficiencies of the modified cycle and compare with those of the basic cycle.

Point # T( oF ) P(psia ) h( Btu/lb ) s(Btu/lb–R) State

1 700 250 1371.6 1.6975 Vapor


2 500 110 1278.3 1.6975 Vapor
3 334.79 110 305.80 0.4834 Satd. liquid
4 101.74 1 948.3 1.6975 Liq + Vapor
5 101.74 1 305.80 0.5531 Liq + Vapor
6 101.74 1 69.73 0.1326 Satd.liquid
7 101.74 250 69.73 0.1326 Liquid
8 334.79 250 305.80 0.4834 Liquid

Note that turbines are isentropic, pump work and pressure drops through heat exchange
equipment are neglected. Enthalpies of compressed liquids are assumed to be equal to
that of the saturated liquid at the same temperature.

Solution: The amount of extracted steam can be determined by an energy balance around
the heat exchanger:

m(1278.3) + 1.00*(69.73) = m*(305.80) + 1.00*305.80 → m = 0.2427 lb

The heat extracted from steam is:


Qhex = 0.2427*(1278.3 – 305.80) = 236.03 Btu
The irreversibility of the heater is given by:
I =  1.0*(0.4834 – 0.1326) + 0.2427*(0.4834 – 1.6975) 520 = 29.19 Btu

W = 1371.6 – 1278.3 + 0.7573*(1278.3 – 948.3) = 343.21 Btu

Qin = 1371.6 – 305.8 = 1065.8 Btu

Qout = 0.2427*(305.80 – 69.73) + 0.7573*(948.3 – 69.73) = 722.64 Btu

72
I = 520(722.64/520 – 1065.8/1160) = 244.87 Btu

Thermal Efficiency = 343.21/1065.80 = 0.3220

Second Law Efficiency = 343.21/(343.21 + 244.87) = 0.5836

It is observed that no improvement in the efficiencies has been achieved by modifying the
original cycle. The additional internal irreversibilities due to heat transfer in the heat
exchanger and due to frictional dissipation in the expansion valve offset any gain
obtained by regeneration. In order to eliminate the internal irreversibility introduced by
the heater, we will again consider a reversible, adiabatic and steady flow process where
inlet streams and extraction stream outlet are the same as those of the heater. For this
process there is no entropy production. An entropy balance yields the specific entropy of
the feedwater outlet:

s8 = 0.2427*(1.6975) + 1.00*(0.1326) – 0.2427*(0.4834) = 0.4273 Btu/(lb-R)


The corresponding outlet enthalpy and temperature are found from steam tables as 262.2
Btu/lb and 292.7oF, respectively. The energy balance then yields the theoretical work that
could be produced:

0.2427*(1278.3) + 1.00*(69.73) – 0.2427*(305.80) – 262.2 = W → W= 43.56 Btu

Similarly, if the extraction steam at point 3 expanded isentropically through an expansion


device instead of the expansion valve, its specific outlet enthalpy would be 266.67 Btu/lb
and work that could be produced is:

W = 0.2427*(305.80 – 266.67) = 9.50 Btu.

Therefore W, Qin and Qout for this hypothetical cycle are:

W = 343.21 + 43.56 + 9.50 = 396.27 Btu

Qin = 1371.6 – 262.2 = 1109.4 Btu

Qout = (266.67*0.2427 +0.7573*948.3) – 69.73 = 713.1 Btu

The corresponding irreversibility, thermal efficiency and second law efficiency are :

I = (713.1/520 – 1109.4/1160 )520 = 215.8 Btu

Thermal Efficiency = 396.27/1109.4 = 0.3572

Second Law Efficiency = 396.27/(396.27 + 215.8 ) = 0.6474

As expected, both efficiencies increase considerably when internal irreversibilities are


completely eliminated which is not possible in actual practice. However, it is possible to

73
reduce the irreversibility of the heat exchanger by reducing the average temperature
difference between the two fluids. This can be achieved by changing the the intermediate
pressure at which steam is extracted. Let’s assume that steam is extracted at 45 psia. The
new operating conditions and specific properties of the working fluid are shown in the
next table where changes are highlighted.

Point # T( oF ) P(psia ) h( Btu/lb ) s(Btu/lb–R) State

1 700 250 1371.6 1.6975 Vapor


2 320 45 1195.3 1.6975 Vapor
3 274.44 45 243.49 0.4021 Satd. liquid
4 101.74 1 948.3 1.6975 Liq + Vapor
5 101.74 1 243.49 0.4421 Liq + Vapor
6 101.74 1 69.73 0.1326 Satd.liquid
7 101.74 250 69.73 0.1326 Liquid
8 274.44 250 243.49 0.4021 Liquid

The energy balance around the heat exchanger yields:

m*(1195.3) + 1.00*(69.73) = m*(243.49) + 1.00*243.49 → m = 0.1826 lb

W = 1371.6 – 1195.3 + 0.8174*(1195.3 – 948.3) = 378.2 Btu

Qin = 1371.6 – 243.49 = 1128.1 Btu

Qout = 0.1826*(243.49 – 69.73) + 0.8174*(948.3 – 69.73) = 749.87 Btu

I = 520(749.87/520 – 1128.1/1160) = 244.12 Btu

Thermal Efficiency = 378.2/1128.1 = 0.3353

Second Law Efficiency = 378.2/(378.2 + 244.12) = 0.6077

Note that both efficiencies are now improved over the basic cycle.

HEX Condenser Boiler HEX Valve


Duty Irreversibility Irreversibility Irreversibility) Irreversibility
(Btu) (Btu) (Btu) (Btu) (Btu)
236.03 53.32 153.55 29,19 8.80
173.80 55.35 167.90 17.14 3.80

74
The irreversibilites of different components are compared in the above table for the two
alternative designs of the regeneration system. Note that in the first alternative, high
internal irreversibilities due to improper choice of extraction pressure prevent any
improvement in thermal efficiency despite a larger heat exchanger duty. In other words,
the increase in Carnot efficiency due to heat addition at a higher average temperature is
completely offset by the decrease in efficiency due to higher internal irreversibility. It can
be shown that for a fixed intermediate pressure, regeneration will improve the thermal
efficiency only when the following condition is satisfied:
(h2 – h4)/(h2 – h3) < η
where η is the thermal efficiency of the original cycle without regeneration.

Since reheat and regeneration configurations offer different advantages over the basic
cycle, many steam power plants feature combinations of both in order to benefit from
these advantages simultaneously. Such a combined configuration is depicted in Figure
3.19. In actual practice, the flow diagram of a power plant is much more complicated,
involving multi-stage turbines and applications of reheat and generation at more than one
intermediate pressure. Determination of these pressure levels is a crucial step in the
design of modern plants.

HP LP
TURBINE TURBINE

1 2 3 4

T
1 3
B
0
I
L
E
R
9 7 2
6
5 8 4 9
H
E
X

7 8
S 6
CONDENSER

PUMP
5

FIGURE 3.19: Rankine cycle with reheat and regeneration.

75
The Organic Rankine Cycle: Renewed interest in waste heat power recovery has
resulted in the development of Organic Rankine Cycle (ORC). The Organic Rankine
cycle allows power recovery of waste heat streams that are too low in temperature to be
cost effectively converted into power by conventional steam cycles. The use of organic
fluids, such as refrigerants, enables efficient and cost-effective power recovery of waste
heat streams.
The most common waste heat streams that lend themselves for power recovery by use of
organic Rankine cycle equipment are:
1. Exhaust gases from engines or incinerators. Engine exhaust temperatures, either
from a gas turbine or reciprocating engine, vary between 600 and 700 K.
Incinerator exhaust gas temperatures are even higher. A refrigerant evaporator
extracts the thermal energy from these waste heat streams by cooling them down.
The actual gas composition of the exhaust waste heat dictates how much heat can
be extracted from these waste heat streams before condensation of acidic
byproducts causes heat exchanger corrosion concerns. Practical limits for
minimum cooled-down gas temperature vary, depending on the chemical
composition of the waste heat gas stream, between 350 and 400 K.
2. Low temperature wet steam from industrial processes at just-above atmospheric
pressures. Under these conditions all heat is supplied at a constant temperature but
this temperature is at a much lower level, around 380 K.
3. Hot liquid waste heat. In these cases a secondary fluid, a brine or oil, is circuited
through a process or engine as a cooling medium and is traditionally reduced in
temperature in some type of radiator. Replacing the radiator with an organic
Rankine cycle enables power recovery from this waste heat. Typical temperatures
for these types of liquid waste heat streams are from around 380 K incoming to
about 340 K returning.
The maximum thermal efficiency of an Organic Rankine Cycle system with a sensible
waste heat source, which is cooled down from an inlet absolute temperature Tin to an
outlet absolute temperature Tout, and a heat rejection absolute temperature TL is:

η = 1 –TLln(Tin/Tout) /(Tin–Tout)
max

This equation shows that the maximum efficiency is higher when Tout is closer to Tin, in
other words, when less heat is extracted from the waste heat stream. In practice
maximizing waste heat power recovery is more important than thermal efficiency.
Assuming that there is no other potential use for the waste heat stream than power
recovery, Tout should be selected to maximize waste heat power recovery instead of
maximizing thermal efficiency of the system, i.e. as low as economically possible.
The condenser of an organic Rankine cycle can be either air-cooled or water-cooled.
Water-cooled condensation results in lower condensing temperatures and therefore higher
cycle efficiency than can be achieved with air-cooled condensers. The trade-off to the
higher efficiency advantage of water-cooled condenser systems is the need for a
secondary water loop with pump and cooling tower and its corresponding water

76
consumption. Also, the power consumption of the water pump and the cooling tower fans
is typically larger than the power required by the air cooled condenser fans.
For a given heat source and heat sink, the performance of an organic cycle is highly
dependent on the working fluid as shown in Figure 3.20 where two different substances
are compared. Irreversibility due to heat addition is much higher for the fluid with a
lower critical temperature while irreversibilities due to heat rejection are almost identical.
Therefore the cycle employing the fluid with the higher critical temperature will have a
higher efficiency. This result can not be generalized since there are other factors
influencing the relative performances of different fluids. For instance, the slope of the
saturated vapor line is negative for both fluids in Figure 3.20. When this slope is positive,
the working fluid is superheated at the end of expansion. The extent of this superheat may
be such that the additional irreversibility introduced may offset any advantage gained in
the heat addition step.

T SOURCE

SINK

FIGURE 3.20: Effect of working fluid on heat transfer irreversibility.

77
The Supercritical Rankine Cycle: When turbine inlet temperature and condenser
pressure are fixed due to several environmental, economical and metallurgical constraints,
the efficiency of the Rankine cycle can be improved by increasing the boiler pressure.
The effects of increasing the boiler pressure are depicted in Figure 3.21. The net work
output of the cycle may increase or decrease depending on the actual operating conditions
(relative magnitudes of tan and turquoise colored areas in the figure) but thermal
efficiency increases since the average temperature at which heat is added increases. Most
modern steam power plants are now operating at pressures around 240 bar which is above
the critical pressure of water (220.55 bar) to take full advantage of high pressures and
temperatures(up to 900 K) made possible by the advances in ferritic materials technology.

T 1 1 Tmax

4
4

3 2
2

Increase in Wnet Decrease in Wnet

Decrease in Wnet Decrease in Wnet


FIGURE 3.21: The effects of increasing boiler pressure.

As can be seen from figure 3.22, the effect on the average temperature of heat addition is
more pronounced at supercritical pressures since there are no latent heat effects. One
disadvantage of operating at supercritical temperatures is the low quality of steam at the
turbine exit. However, with the use of reheat techniques (most supercritical steam power
plants use single or double reheat) the turbine exit condition can be pushed to the right on
the temperature-entropy diagram so that the exit steam is completely dry. This also
simplifies the system by eliminating the need for steam separators, dryers and turbines
specially designed for low quality steam.

78
T 1 1
Tmax

3 2
2

S
Increase in Wnet Decrease in Wnet

Decrease in Decrease in Wnet


Wnet
FIGURE 3.22: The supercritical Rankine cycle.
The Kalina Cycle: One way to increase the efficiency of the Rankine steam cycle is to
reduce the heat transfer irreversibility in the boiler by better temperature matching
between the heat source and the working fluid. This can be achieved by replacing the
one-component working fluid with a binary mixture. Binary fluids boil and condense at
increasing and decreasing temperatures respectively, while one-component fluids boil
and condense at constant temperatures.
The change in temperature during boiling and condensation of a mixture makes it
possible to keep the temperature profile of a binary working fluid closer to a heat source
of decreasing temperature or a heat sink of increasing temperature. This is illustrated in
Figure 3.23 where the temperature profiles in a vapor generator are compared for two
different working fluids: Water and ammonia-water mixture It can be seen in the figure
that the ammonia-water mixture extracts more energy from the heat source and the heat
transfer irreversibility is reduced due to the smaller temperature differences between the
source and the working fluid. Note that the outlet temperature of the heat source is greater
when working fluid is water. Therefore with water, the Carnot efficiency is higher while
the second law efficiency is lower.

79
T Ammonia-Water

Heat Source

Water

FIGURE 3.23: The temperature profiles and extracted heat in a vapor generator.
In 1983, Dr. Alexander Kalina presented a power cycle with an ammonia-water mixture
as its working fluid. The use of a binary mixture to reduce heat transfer irreversibility in
the boiler was well known by that time, however the difficulty associated with the
condensation of the binary working fluid had not been solved: Condensation also occurs
over a temperature range and is not complete with normal cooling water temperatures
unless turbine exit pressure is increased. Then all gains achieved in the boiler is lost in
the turbine. Dr. Kalina solved this problem using absorption refrigeration principles: The
pressure of the working fluid leaving the turbine is elevated to a level where complete
condensation is achieved using the heat in the turbine exhaust. The first Kalina cycles
were designed for using the hot exhaust gas from diesel engines and gas turbines as heat
sources. Cycle configurations for other applications, such as geothermal and direct-fired
power plants, have also been developed. The flow diagram of the simplest possible
Kalina cycle configuration is shown in Figure 3.24.
The operation of the condensing system can be explained as follows: A basic solution at
high pressure which has an X (ammonia mass fraction) of 0.41 is heated by the turbine
exhaust which is a vapor at low pressure with X= 0.70. The basic solution is then sent to
a distillation column where it is separated into an enriched vapor with X= 0.96 and a lean
liquid solution with X= 0.34. After its pressure is reduced, the lean solution is mixed with
the turbine exhaust and forms the basic solution which is condensed in the absorber with

80
Strong solution Working fluid
Lean solution Basic solution

TURBİNE

SEPARATOR
VAPOR
GENERATOR

HEAT
EXCHANGER

VALVE

CONDENSER ABSORBER

PUMP PUMP

FIGURE 3.24: The simplest possible Kalina cycle configuration.


cooling water. By lowering the mass fraction of ammonia before the absorber the turbine
can be expanded to a lower pressure. This results from the fact that a mixture with low
mass fraction of ammonia has a lower condensing pressure than a mixture with high mass
fraction of ammonia at the same temperature. After the absorber the basic mixture is
pressurized and then split into two streams. One of the streams is the basic solution which
is heated by the turbine exhaust. The other stream is mixed with the enriched vapor from
the distillation column to reconstitute the working fluid at a higher pressure than the
turbine exit. This working fluid is then condensed with cooling water, pressurized and
sent to the vapor generator. Note that the amount of basic solution is almost five times
greater than that of the working fluid.
The actual configuration of a Kalina cycle may be much more complicated in order to
achieve better internal heat exchange. One such configuration is shown in the next figure.
Compositions of ammonia solutions may also be changed to achieve a better temperature
match with the available source and sink so that the performance of the cycle is
maximized.

81
Lean solution
Strong solution
C
Basic solution
Working fluid
O
Cooling Water L
U
M
N H
E
X
H
E
X
PUMP

H H
E E
X X

H
E
X

PUMP PUMP

FIGURE 3.25: Another Kalina cycle condensing system configuration.


Before closing this chapter, it will be appropriate to scrutinize the approach we have used
so far to analyze the thermodynamic cycles. The system boundary has been defined in
such a way that the cycles can be analyzed independent of the origin of the heat source.
In other words, any external or internal irreversibility associated with the generation of
the heat source are not accounted for when total irreversibility is calculated by the cyclic
integral of ∫dQ/T. Hence the second law efficiencies calculated with this approach reflect
the efficiency of different cycles in converting heat into work, but actual efficiencies may
be quite different depending on the primary source of heat.

82
EXERCISES

3.1. Steam enters the turbine of a power plant operating on an ideal Rankine cycle at
2,000 psia and 1,000oF. The condenser pressure is 1 psia. Calculate the thermal and
efficiencies and the irreversibility for each step. Temperature of the environment is 60oF.

3.2. Consider the same Brayton cycle described in Example 3.3. A 4% fractional
pressure drop occurs in the combustor. Calculate the new thermal and second law
efficiencies and the external and internal irreversibilities.

3.3. An Otto engine takes in an air-fuel mixture at 80°F and standard atmosphere
presssure. It has a compression ratio of 8. Using Air Standard cycle analysis, a heating
value of 20,425 Btu/lbm, and air to fuel mass ratio of 15, determine the thermal and
second law efficiencies. What is the major source of irreversibility?

3.4. A Diesel engine has a compression ratio of 20 and a peak temperature of 3000K.
Using an Air Standard cycle analysis, estimate the work per unit mass of air, the thermal
efficiency and the second law efficiency. What is the major source of irreversibility?

3.5. Steam enters the turbine of a power plant operating on a simple


Rankine cycle at 250 psia and 700oF. The condenser pressure is 1.0 psia. This cycle is
modified by introducing regeneration with an open feedwater heater at 45 psia. Assume
that the working fluid leaves the heater as a saturated liquid .Calculate thermal and
second law efficiencies of the modified cycle and compare with those of the original
cycle.

3.6. Reheat and regeneration at 45 psia are introduced to the basic Rankine cycle
described in Exercise 3.6. Calculate and compare thermal and second law efficiencies for
regeneration with an open feedwater heater and regeneration with a closed feedwater
heater.

3.7. Interpret the results of Exercise 3.5 and 3.6 in terms of the irreversibilities of system
components.

83

You might also like