You are on page 1of 8

INTERNATIONAL JOURNAL OF ENERGY RESEARCH

Int. J. Energy Res., 22, 249—256 (1998)

EFFECT OF COMBUSTION ON THE ECONOMIC OPERATION OF


ENDOREVERSIBLE OTTO AND JOULE—BRAYTON ENGINE

NIMAI CHANDRA BERA1 AND SANTANU BANDYOPADHYAY*,2


1Department of Physics, A.B.N. Seal College, Coochbehar, West Bengal, India
2Energy Systems Engineering, Department of Mechanical Engineering, Indian Institute of Technology, Powai, Mumbai 400076, India

SUMMARY
To simplify analysis of an internal combustion engine, air-standard cycles are conceived. Air is assumed to behave like
an ideal gas. In practice, air-standard analysis provides useful indication of the trends that the engine is likely to follow.
Air-standard Otto and Joule—Brayton cycles are bona fide assumption and cannot represent the complex combustion
process occurring in the internal combustion engines. In this paper, the complex combustion process is represented by
a parameter called fuel-flame temperature. The effect of combustion on the thermoeconomic performances of Otto and
Joule—Brayton engines are studied. It is observed that the efficiency at maximum power is less than the Curzon—Ahlborn
efficiency. The economic performance of the engine deteriorates due to combustion. The efficiency of the engine
corresponds to maximum specific-power output, depends not only on the fuel-flame temperature, but also on the specific
heats of the air and fuel. Ideal gas assumption of the working fluid is relaxed in this paper. Although somewhat idealized,
the effect of combustion on the performance and economics of the internal combustion engines gives a reasonable design
goal and better understanding of the real-heat engine. ( 1998 John Wiley & Sons, Ltd.

KEY WORDS flame temperature; heat engine; thermoeconomics; endoreversibility; combustion

INTRODUCTION
The maximum possible efficiency of a heat engine can only be achieved when the heat engine operates
reversibly. The efficiency of a reversible heat engine can be determined solely by the temperatures of the heat
source and heat sink (g "1!¹ /¹ ). This is known as the Carnot efficiency of a heat engine. Reversible
C .*/ .!9
heat transfer demands either infinite heat exchange surface area or infinite cycle time or both. Thus, the
design and operation of reversible heat engine becomes uneconomical and necessitous. Slow isothermal heat
exchange in Carnot heat engine, leads the power production to zero. Considering a simplified model for
Carnot heat engine where all of the irreversibility associated with heat transfer from thermal reservoir to the
power cycle equipment, the power produced by the Carnot heat engine can be maximized. These type of heat
engines are called ‘endoreversible’ (Rubin, 1979). The efficiency of the heat engine, operating at its maximum
power point, interestingly depends solely on the reservoir temperatures. The efficiency at maximum power
point is given as

S
¹
g "g "1! .*/ (1)
CA .1 ¹
.!9
This is known as the Curzon—Ahlborn (1975) efficiency. The same efficiency can also be derived by
minimizing the total entropy generation rate associated with the endoreversible Carnot engine (Bejan,

*Correspondence to: Santanu Bandyopadhyay, Energy Systems Engineering, Department of Mechanical Engineering, Indian Institute
of Technology, Powai, Mumbai 400076, India. Email: santanu@me.iitb.ernet.in

CCC 0363—907X/98/030249—08$17.50 Received 6 May 1997


( 1998 John Wiley & Sons, Ltd. Accepted 20 May 1997
250 NIMAI CHANDRA BERA AND SANTANU BANDYOPADHYAY

1996a, b). This existence of a maximum power for simple cycle was recognised and extended by several
researchers (El—Wakil, 1962; Rubin, 1979; Salamon and Nitzan, 1981; Andresen et al., 1984; Gordon, 1988;
Bandyopadhyay and Bera, 1996). The maximum power delivered by other heat engines and the efficiency of
the heat engine at the maximum power point were also studied (Leff, 1987; Ladas and Ibrahim, 1994). The
efficiency of most of the engines, at maximum power point, is equals to or closed to the Curzon—Ahlborn
efficiency (equation (1)).
The optimal piston trajectory for the efficient operation of Otto engine was reported by Mozurkewich and
Berry (1982). Leff (1987) showed that the efficiency of air-standard Otto and Joule—Brayton engine at
maximum work output per cycle is equal to Curzon—Ahlborn efficiency. Ibrahim et al. (1991) observed that
the optimum power produced by closed air-standard Joule—Brayton and Otto engines can be greater than
the optimum power produced by Carnot engine for specified boundary conditions. Sahin et al. (1995) showed
that the efficiency of a air-standard Joule—Brayton engine at maximum power output and at maximum
power density differs significantly. Efficiency at maximum power density is always higher than Curzon—
Ahlborn efficiency. This leads to compact and more efficient engine.
The working fluid undergoes permanent chemical change, due to combustion, in any internal combustion
engine. To simplify the analysis of internal combustion engines, air-standard cycles are conceived. In an
air-standard cycle, a certain mass of air operates in complete thermodynamic cycle and the heat is added to
or rejected with external heat reservoir. The working fluid is assumed to be ideal gas. The practice of using
air-standard analysis provides useful indication of the trends likely to follow from a change in engine design
and operating variables. This approach is highly optimistic estimation of the real Otto and Joule—Brayton
engine.
The goal of this paper is to study the effect of combustion on the performance, work production,
economics and efficiency of an endoreversible Otto and Joule—Brayton engines. Ideal gas assumption of the
working fluid is relaxed in this paper. Although somewhat idealized, the effect of combustion on the
performance and economics of the internal combustion engines gives a reasonable design goal and better
understanding of the real heat engine.

ENDOREVERSIBLE ENGINE
Unlike Carnot engine, Otto and Joule—Brayton engines do not operate under isothermal heat addition and
heat rejection. Therefore, these engines always operate under finite time and hence, the power optimization
for Carnot engine is some way equivalent to per cycle work optimization of these engines. All the engines are
assumed to be ‘endoreversible’, i.e., the only way entropy is generated through the finite temperature driving
force and the cyclic entropy generation of both the engines is zero. That is the irreversibility due to friction,
heat loss through the cylinder, blowdown, etc. are neglected. Assuming Newtonian (or linear) heat transfer,
i.e., the rate of heat transfer is proportional to the temperature difference between reservoir and engine
working fluid, the power generated by the Carnot engine can be expressed in terms of equivalent thermal
conductivity and the reservoir temperature (Bandyopadhyay and Bera, 1996).
(1!q)(q¹ !¹ )
Q "C
¼ .!9 .*/ (2)
%2 q
where q"¹ /¹ . The efficiency of the engine is given by g"1!q. For thermodynamic operation of the
1 3
heat engine, q has to lie in between (¹ /¹ ) and unity as the power output becomes zero at both the limits.
.*/ .!9
q"¹ /¹ signifies the classical reversible engine with infinite time requirement for isothermal heat
.*/ .!9
transfer, producing zero power. On the other hand, q"1 indicates the case of total thermal leakage, when all
the heat directly flows from the hot reservoir to the cold.
In an air-standard Otto engine, at ¹ temperature, air—fuel mixture is injected to the engine and this is
1
compressed to temperature ¹ . Then heat is added at constant volume to raise its temperature ¹ . The
2 3
Int. J. Energy Res., 22, 249—256 (1998) ( 1998 John Wiley & Sons, Ltd.
ENDOREVERSIBLE OTTO AND JOULE—BRAYTON ENGINE 251

expansion stroke reduce the temperature to ¹ after which the heat is being rejected to the atmosphere.
4
QQ "C (¹ !¹ ) (3)
23 v 3 2
QQ "C (¹ !¹ ) (4)
41 v 4 1
and hence from the first law we can write,
¼ Q "QQ !QQ "C (¹ #¹ !¹ !¹ ) (5)
23 41 v 3 1 2 4
From the endoreversibility criteria of the Otto engine, we can write that

A B A B
¹ ¹
C ln 3 "C ln 4 (6)
v ¹ v ¹
2 1
It follows that
¹ ¹ "¹ ¹ (7)
3 1 2 4
The endoreversibility criteria for Otto engine can also be derived assuming the working fluid is ideal gas.
By denoting the maximum temperature of the cycle ¹ as ¹ , the minimum temperature of the cycle ¹ as
3 .!9 1
¹ , and combining the above equations, we can express the work produced by the Otto engine as
.*/
(1!q) (q¹ !¹ )
¼ Q "C .!9 .*/ (8)
v q
where q"¹ /¹ and the efficiency of the engine is given by g"1!q. A similar analysis for Joule—Brayton
1 2
engine leads to the expression for work output as
(1!q) (q¹ !¹ )
Q "C
¼ .!9 .*/ (9)
p q
For thermodynamic operation of the Otto and Joule—Brayton engine, q has to be in between (¹ /¹ )
.*/ .!9
and unity as the work output becomes zero at both the limits. q"¹ /¹ signifies the case when there is
.*/ .!9
no heat exchange and the working fluid is compressed to the maximum possible extent (¹ "¹ )
2 .!9
producing no work. On the other hand, q"1 indicates the case of no compression or expansion (¹ "¹ ).
2 .*/
In this case, all the heat used for combustion is rejected to the environment, producing zero work. Therefore,
for the maximum power production, q has to be in these two extreme conditions.
For maximum power production of the Carnot engine, Otto engine or Joule—Brayton engine, the temper-
ature ratio, q is estimated to be the square root of allowable temperature ratio, i.e., J(¹ /¹ ) and hence
.1 .*/ .!9
the efficiency of these engines are equal to the Curzon—Ahlborn efficiency (1975). At maximum power, the
intermediate temperatures become equal to the geometric mean of the extreme temperatures.

¹ "¹ "J¹ ¹ (10)


2 4 .*/ .!9
Knowing the equation of state of the working fluid, the volumetric compression ratio can be expressed in
terms of q . For ideal gas the volumetric compression ratio, corresponds to maximum power output, is
.1
given by

A B
¹ 1@2(c~1)
r "q~1@(c~1)" .!9 "Jr (11)
.1 .1 ¹ .!9
.*/
The maximum work delivered by the heat engine is given as

¼Q "C(J¹ !J¹ )2 (12)


.!9 .!9 .*/
( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 249—256 (1998)
252 NIMAI CHANDRA BERA AND SANTANU BANDYOPADHYAY

Remarkably, the maximum power, ¼ Q , and the corresponding efficiency, g , are independent of the
.!9 .1
details equation of state of the working fluid, requiring only the assumption of a constant heat capacity
C along the non-adiabatic paths. The work delivered by the heat engine increases with the increasing amount
of working fluid. The concept of increasing the amount of working fluid in the real heat engine is
accomplished by adjusting chock and accelerator.

ENDOREVERSIBLE THERMOECONOMICS
The efficiency corresponds to the economic operation of the heat engine is always higher than the efficiency
corresponds to the maximum power condition (Bejan, 1993; De Vos, 1995; Bandyopadhyay and Bera, 1996).
As the efficiency of the heat engine decreases from the reversible Carnot limit, the power production increases
but the fuel-utilisation decreases. This counteractive activities bring the efficiency of the heat engine to be in
between the maximum power point and reversible operating point with maximum efficiency. The approaches
considered in this paper is presented by Bandyopadhyay and Bera (1996).
Considering the proper cost fraction for input energy (g , depends on the fuel cost), output energy
1
(g depends on the power selling price) and the cost associated with heat rejection (g , depends on the
3 2
cooling utility cost), the operating cost of the engine can be written as

[ g !g #q( g #g )](q¹ !¹ )
p"C 1 3 2 3 .!9 .*/ (13)
q

The economic design and operation of Carnot, Otto, or Joule—Brayton engine is not possible whenever

g !g ¹
3 1) .*/ (14)
g #g ¹
2 3 .!9
The choice q for cost-optimum operation of the heat engine depends on the minimum p. Optimizing p with
respect to q, we get

S
( g !g ) ¹
q " 3 1 .*/ (15)
015 ( g #g )¹
2 3 .!9
p "!C[J( g !g )¹ !J(g #g )¹ ]2 (16)
.*/ 3 1 .!9 3 2 .*/
Corresponding volumetric compression ratio can be calculated from the working fluid’s equation of state.
For ideal gas:

r "q~1@(c~1)
015 015

CA BA BD
¹ g #g 1@2(c~1)
" .!9 3 2
¹ g !g
.*/ 3 1

SC A B D
g #g 1@(c~1)
r " 3 2 (17)
.!9 g !g
3 1
( g #g ) is always greater than ( g !g ). Therefore, q )q and hence.
3 2 3 1 015 .1
g *g (18)
015 .1
Economically designed and operated heat engines are always more efficient than engines that delivers
maximum power output.

Int. J. Energy Res., 22, 249—256 (1998) ( 1998 John Wiley & Sons, Ltd.
ENDOREVERSIBLE OTTO AND JOULE—BRAYTON ENGINE 253

In case of Otto and open cycle Joule—Brayton engine, the heat is rejected to the atmosphere, and so g is
2
zero. Therefore, the optimality condition simplifies to

SA B
g ¹
q " 1! 1 .*/ (19)
015 g ¹
3 .!9

EFFECT OF COMBUSTION
Due to material safety point of view the maximum temperature is fixed in most of the operating heat
engine. The working air—fuel mixture is assumed to have constant specific heats and they do not change due
to the combustion (Heywood, 1988). The actual combustion process is very complex and depends on many
design and operating parameters (Heywood, 1988). For simplicity, in this paper, we assumed that a constant
combustion efficiency (g ) depicts the combustion process. Therefore, from energy balance, we can write
#0."
QQ "(mR c #mR c ) (¹ !¹ )"g mR (*H ) (20)
23 ! ! & & 3 2 #0." & &
Eliminating mass flow rate of the fuel (mR ), from the above equations we can write
&

C D
¹
QQ "mR c (¹ !¹ ) 0 (21)
23 ! ! 3 2 ¹ #¹ !¹
0 2 3
and
mR c (¹ !¹ )
&" ! 3 2 (22)
mR c (¹ #¹ !¹ )
! & 0 2 3
where

¹ "g (*H )/c (23)


0 #0." & &
¹ signifies the maximum possible temperature attainable by combustion; we called it fuel-flame temper-
0
ature. For any given combustion efficiency, ¹ depends on the fuel-characteristics alone. Similarly, the heat
0
rejection can be written as

C D
¹
QQ "mR c (¹ !¹ ) 0 (24)
41 ! ! 4 1 ¹ #¹ !¹
0 2 3
From the endoreversibility of the heat engine, we can write

¹ ¹
q" 1" 4 (25)
¹ ¹
2 3
Writing ¹ as ¹ and ¹ as ¹ , the efficiency and the power output of the internal combustion engine are
3 .!9 1 .*/
given by g"1!q and

(1!q)(q¹ !¹ )
Q "mR c ¹
¼ .!9 .*/ (26)
! ! 0 (¹ !¹ )q#¹
0 .!9 .*/
Maximizing ¼Q with respect to q, the condition for maximum power output is obtained.

S C D
¹ J¹ (¹ #¹ !¹ )!J¹ ¹
q " .*/ 0 0 .*/ .!9 .!9 .*/ (27)
.1 ¹ ¹ !¹
.!9 0 .!9
( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 249—256 (1998)
254 NIMAI CHANDRA BERA AND SANTANU BANDYOPADHYAY

Now, q can be simplified and approximated by the following way:


.1

S C D
¹ J1#(¹ !¹ )/¹ !J(¹ ¹ /¹2)
q " .*/ .*/ .!9 0 .!9 .*/ 0 (28)
.1 ¹ 1!¹ /¹
.!9 .!9 0

S C D
¹ 1#(¹ !¹ )/2¹ !J(¹ ¹ /¹2)
+ .*/ .*/ .!9 0 .!9 .*/ 0 (29)
¹ 1!¹ /¹
.!9 .!9 0

S C D
¹ 1!¹ /¹ #(1/2)(J¹ /¹ !J¹ /¹ )2
" .*/ .!9 0 .!9 0 .*/ 0 (30)
¹ 1!¹ /¹
.!9 .!9 0

S
¹
.*/
* (31)
¹
.!9
Therefore, the efficiency of the heat engine is always less than the Curzon—Ahlborn efficiency. Equality holds
only when ¹ PR. For any fuel, the fuel-flame temperature, ¹ , is finite. Hence, the performance of real
0 0
engine deteriorates as the fuel-flame temperature decreases. For low calorific fuel or decrease in combustion
efficiency due to partial burning, misfire, etc. (Heywood, 1988). Flame temperature is low. As ¹ P¹ , g
0 .!9 .1
becomes half of the Carnot efficiency.
Similarly, the operating cost of the engines can be written in terms of limiting temperatures, fuel-flame
temperature, cost parameters and q.
[g !g #q(g #g )](q¹ !¹ )
1 3 2
p"mR c ¹ 3 .!9 .*/ (32)
! ! 0
(¹ !¹ )q#¹
0 .!9 .*/
Optimum operating cost corresponds to

S C D
¹ J¹ [(¹ !¹ )((g !g )/(g #g ))#¹ ]!J¹ ¹
q " .*/ 0 0 .!9 3 1 3 2 .*/ .!9 .*/ (33)
015 ¹ ¹ !¹
.!9 0 .!9
Though the efficiency of the heat engine, operating at its minimum operating cost, deteriorates due to
combustion, it is always higher than the efficiency corresponds to the maximum power output condition. As
¹ approaches R or ¹ the corresponding efficiency becomes 1!J((g !g )¹ )/((g #g )¹ ) and
0 .!9 3 1 .*/ 2 3 .!9
0.5g #0.5(g !g )/(g #g ), respectively.
C 3 1 2 3
As the fuel-flame temperature decreases power output of the engine increases. This is because, the mass
flow rate of the working-fluid increases. To design compact heat engine and to have a control over the capital
cost of the heat engine, specific power (defined as the power output per unit of working fluid flow) is also
studied:
¼Q (1!q)(q¹ !¹ )
u" "c c ¹ .!9 .*/ (34)
mR #mR & ! 0 [¹ c !¹ (c !c )]q#¹ (c !c )
! & 0 & .!9 & ! .*/ & !
Minimum specific power output corresponds to:

S C D
¹ $J¹ (¹ #¹ !¹ )!J¹ ¹
q " .*/ 9 9 .*/ .!9 .!9 .*/ (35)
.41 ¹ ¹ !¹
.!9 9 .!9
where

A B
c g (*H )
¹ "¹ & " #0." & (36)
9 0 c !c c !c
& ! & !
Int. J. Energy Res., 22, 249—256 (1998) ( 1998 John Wiley & Sons, Ltd.
ENDOREVERSIBLE OTTO AND JOULE—BRAYTON ENGINE 255

$depends on whether c is greater than or less than c , respectively. Whenever, c *c , ¹ is always greater
& ! & ! 9
than ¹ , as can be observed from the last equation. Therefore, it is observed that
0
0.5g )g )g )g )g (37)
C .1 .41 CA C
As a special case, when c "c , ¹ PR, irrespective of whether fuel-flame temperature is finite or not and the
& ! 9
efficiency becomes equal to the Curzon—Ahlborn efficiency. Interestingly, it becomes independent of fuel-
flame temperature and specific heats of fuel and air. Whenever, c is less than c , ¹ becomes negative and it is
& ! 9
noted that
0.5g )g )g )g )g (38)
C .1 CA .41 C
The corresponding volume—compression ratio, mean-effective pressure, residual-mass fraction, etc. can be
found out from these optimum equations of q at different condition, knowing the equation of state of the
working fluid.

CONCLUSIONS
The air-standard Otto and Joule—Brayton engines are very simplistic and cannot possibly represent
the complex-combustion processes occurring in internal combustion engines. The effect of combustion on
the thermoeconomic operation of endoreversible Otto and Joule—Brayton engine is studied in this paper.
The complex combustion process is represented by simplified parameter called fuel-flame temperature.
It is observed that the efficiency at maximum power is always less than the Curzon—Ahlborn efficiency.
The economic performance of the engine deteriorates due to combustion. The efficiency of the engine
corresponds to maximum specific power output, depends not only on the fuel-flame temperature, but also
on the specific heats of the air and fuel. Unlike earlier papers, the working fluid is not assumed to be an
ideal gas. Knowing the detail equation of state of the working fluid, the design parameters such as
volumetric compression, mean-effective pressure, residual mass fraction, specific-fuel consumption, etc. can
be determined.

NOMENCLATURE
c "specific heat of air
!
c "specific heat of fuel
&
C "specific heat
C "equivalent thermal conductivity
%2
g "utility cost
i
*H "heat of combustion
&
mR "mass flow rate of air
!
mR "mass flow rate of fuel
&
QQ "heat
r "volumetric compression ratio
r "volumetric compression ratio at maximum power output
.1
r "maximum volumetric compression ratio
.!9
r "volumetric compression ratio at optimal operating cost
015
¹ "state temperature
i
¹ "flame temperature
0
¹ "corrected flame temperature
9
¹ "maximum temperature
.!9
¹ "minimum temperature
.*/
( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 249—256 (1998)
256 NIMAI CHANDRA BERA AND SANTANU BANDYOPADHYAY

¼ Q "power
¼ Q "maximum power
.!9
c "ratio of specific heat
g "Carnot efficiency
C
g "Curzon—Ahlborn efficiency
CA
g "combustion efficiency
#0."
g "efficiency at maximum power output
.1
g "efficiency at maximum specific power output
.41
g "efficiency at optimal operating cost
015
p "operating cost
p "minimum operating cost
.*/
q "temperature ratio
q "temperature ratio at maximum specific power output
.41
q "temperature ratio at optimal operating cost
015
u "specific power
1, 2, 3, 4 "state points

REFERENCES
Andresen, B., Salamon, P. and Berry, R. S. (1984). ‘Thermodynamics in finite time’, Phys. ¹oday, September 62—70.
Bandyopadhyay, S. and Bera, N. C. (1996). ‘Cost optimum design of an endoreversible Carnot-like heat engine’, Paper Presented at 49th
Indian Chemical Engineering Congress (CHEMCON’ 96), Ankleshwar, 18—21 December.
Bejan, A. (1993). ‘Power and refrigeration plants for minimum heat exchanger inventory’, ¹rans. ASME: J. Energy Res. ¹echnol., 115,
148—150.
Bejan, A. (1996a). ‘The equivalence of maximum power and minimum entropy generation rate in the optimization of power plants’,
¹rans ASME: J. Energy Res. ¹echnol., 118, 98—101.
Bejan, A. (1996b). ‘Models of power plant that generates minimum entropy while operating at maximum power’, Am. J. Phys., 64(8),
1054—1059.
Curzon, F. L. and Ahlborn, B. (1975). ‘Efficiency of a Carnot engine at maximum power output’, Am. J. Phys., 43, 22—24.
Gordon, J. M. (1988). ‘On optimized solar-driven heat engines’, Solar Energy, 40(5), 457—461.
Heywood, J. B. (1988). Internal Combustion Engine Fundamentals, McGraw—Hill, New York.
Ibrahim, O. M., Klein, S. A. and Mitchell, J. W. (1991). ‘Optimum heat power cycles for specified boundary conditions’, ¹rans. ASME:
J. Engng Gas ¹urb. Power, 113, 514—521.
Ladas, H. G. and Ibrahim, O. M. (1994). ‘Finite-time view of the Stirling engine’, Energy, 19(8), 837—843.
Leff, H. S. (1987). ‘Thermal efficiency at maximum work output: new results for old heat engines’, Am. J. Phys., 55(7), 602—610.
Mozurkewich, M. and Berry, R. S. (1982). ‘Optimal paths for thermodynamic systems: the ideal Otto cycle’, J. Appl. Phys., 53(1), 34—42.
Rubin, M. H. (1979). ‘Optimal configuration of a class of irreversible heat engines I’, Phys. Rev. A, 19(3), 1272—1276.
Sahin, B., Kodal, A. and Yavuz, H. (1995). ‘Efficiency of a Joule—Brayton engine at maximum power density’, J. Phys. D, 28, 1309—1313.
Salamon, P. and Nitzan, A. (1981). ‘Finite time optimization of a Newton’s law Carnot cycle’, J. Chem. Phys., 74(6), 3546—3560.
De Vos, A. (1995) ‘Endoreversible thermoeconomics’, Energy Convers. Mgmt., 36(1), 1—5.
El-Wakil, M. M. (1962). Nuclear Power Engineering, McGraw—Hill, New York.

Int. J. Energy Res., 22, 249—256 (1998) ( 1998 John Wiley & Sons, Ltd.

You might also like