You are on page 1of 9

Eciency of a Miller engine

A. Al-Sarkhi
a,
*
, J.O. Jaber
a
, S.D. Probert
b
a
Department of Mechanical Engineering, Hashemite University, Zarqa 13115, Jordan
b
School of Engineering, Craneld University, Bedford MK43 0AL, UK
Available online 9 June 2005
Abstract
Using nite-time thermodynamics, the relations between thermal eciency, compression
and expansion ratios for an ideal naturally-aspirated (air-standard) Miller cycle have been
derived. The eect of the temperature-dependent specic heat of the working uid on the
irreversible cycle performance is signicant. The conclusions of this investigation are of
importance when considering the designs of actual Miller-engines.
2005 Elsevier Ltd. All rights reserved.
Keywords: Finite-time thermodynamics; Miller cycle; Heat resistance; Friction; Temperature-dependent
specic-heat
Introduction
The Miller cycle, named after its inventor R.H. Miller, has an expansion ratio
exceeding its compression ratio. The Miller cycle, shown in Fig. 1, is a modern mod-
ication of the Atkinson cycle (i.e., a complete expansion cycle). The compression
ratios of spark-ignition, gasoline-fueled engines are limited by knock and fuel quality
to being in the range between 8 and 11, depending upon various factors, such as the
engines bore and stroke as well as engine speed.
Signicant achievements have ensued since nite-time thermodynamics was devel-
oped in order to analyze and optimize the performances of real heat-engines [13].
0306-2619/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apenergy.2005.04.003
*
Corresponding author. Tel.: +962 5 3826600x4574; fax: +962 5 3826348.
E-mail address: alsarkh@hu.edu.jo (A. Al-Sarkhi).
Applied Energy 83 (2006) 343351
www.elsevier.com/locate/apenergy
APPLIED
ENERGY
Homan et al. [4] and Mozurkewich and Berry [5] used mathematical techniques,
developed for optimal-control theory, to reveal the optimal motions of the pistons
in Diesel and Otto cycle engines, respectively. Aizenbud et al. [6] and Chen et al.
[7] evaluated the performances of internal-combustion engine cycles using the opti-
mal motion of a piston tted in a cylinder containing a gas pumped at a specied
heating-rate. Orlov and Berry [8] deduced the power and eciency upper limits
for internal-combustion engines. Angulo-Brown et al. [9], Chen et al. [10] and Wang
et al. [11] modeled the behaviors of Otto, Diesel and Dual cycles, with friction losses,
over a nite period. Klein [12] investigated the eects of heat transfer on the perfor-
mances of Otto and Diesel cycles. Chen et al. [13,14] and Lin et al. [15] derived the
relations between the net power and the eciency for Diesel, Otto and Dual cycles
with due consideration of heat-transfer losses. Chen et al. [16,17] determined the
characteristics of power and eciency for Otto and Dual cycles with heat transfer
and friction losses. Chen et al. [18], Al-Sarkhi et al. [19] and Sahin et al. [20] studied
the optimal power-density characteristics for Atkinson, Miller and Dual cycles with-
out any such losses. Qin et al. [21] deduced the universal power and eciency char-
acteristics for irreversible reciprocating heat-engine cycles with heat transfer and
friction losses. Parlak et al. [22] optimized the performance of an irreversible Dual
cycle: the predicted behavior was corroborated by experimental results. Fischer
and Homan [23] concluded that a quantitative simulation of an Otto-engines
behavior can be accurately achieved by a simple Novikov model with heat leaks.
Al-Sarkhi et al. [24] recently found that friction and the temperature-dependent spe-
cic heat of the working uid of a Diesel engine had signicant inuences on its
power output and eciency. This paper describes a corresponding analysis of the
behavior for an irreversible Miller-cycle with losses arising from heat resistance
and friction.
An air standard Miller-cycle model
As in Fig. 1, the compression 1 ! 2 process is isentropic; the heat addition 2 ! 3,
an isochoric process; the expansion 3 ! 4, an isentropic process; and the heat rejec-
3
1
2
4
5
v
P
S
T
2
1
3
4
5
q
in
q
o
u
t
q
o
u
t
6
Fig. 1. PV and TS diagram of a Miller cycle.
344 A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351
tion 4 ! 5, an isochoric process, while the rejection of heat 5 ! 1, an isobaric process.
Finally, the exhaust from 1 ! 6 is also an isobaric process. As is usual in nite-time
thermodynamic heat-engine cycle models, there are two instantaneous adiabatic-pro-
cesses namely 1 ! 2 and 3 ! 4 . For the heat addition (2 ! 3) and heat rejection
(4 ! 5 ! 1) stages, respectively, it is assumed that heating occurs fromstate 2 to state
3 and cooling ensues from state 4 to state 1: these processes proceed according to
dT
dt

1
C
1
for 2 ! 3;
dT
dt

1
C
2
for 4 ! 5; and
dT
dt

1
C
3
for 5 ! 1 1
where T is the absolute temperature and t is the time, C
1
, C
2
and C
3
are constants.
Integrating Eq. (1) yields
t
1
C
1
T
3
T
2
; t
2
C
2
T
4
T
5
; and t
3
C
2
T
5
T
1
2
where t
1
is the heating period and t
2
and t
3
the cooling periods. Then, the cycle per-
iod is
s t
1
t
2
t
3
C
1
T
3
T
2
C
2
T
4
T
5
C
2
T
5
T
1
3
In a real cycle, the specic heat of the working uid depends upon its temperature
and this will inuence the performance of the cycle. Over the temperature range gen-
erally encountered for gases in heat engines (i.e., 3002200 K), the specic-heat curve
is nearly linear, i.e., to close approximations
C
p
a k
1
T 4
C
v
b k
1
T 5
where a, b and k
1
are constants: C
p
and C
v
are the molar specic heats with respect to
constant pressure and volume, respectively. Accordingly, the constant, R, of the
working uid is
R C
p
C
v
a b 6
The heat added to the working uid, during the process 2 ! 3, is
Q
in
M
_
T
3
T
2
C
v
dT M
_
T
3
T
2
b k
1
T dT MbT
3
T
2
0.5k
1
T
2
3
T
2
2

7
where M is the molar number of the working uid.
The heat rejected by the working uid, during the process 4 ! 5, is
Q
out1
M
_
T
4
T
51
C
v
dT M
_
T
4
T
51
b k
1
T dT MbT
4
T
5
0.5k
1
T
2
4
T
2
5

8
The heat rejected by the working uid, during the process 5 ! 1, is
Q
out2
M
_
T
1
T
5
C
p
dT M
_
T
1
T
5
a k
1
T dT MaT
5
T
1
0.5k
1
T
2
5
T
2
1

9
A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351 345
Because C
p
and C
v
are dependent on temperature, the adiabatic exponent k = C
p
/
C
v
will also vary with temperature. Therefore, the equation often used for a revers-
ible adiabatic process, with constant k, cannot be used for a reversible adiabatic pro-
cess with a variable k. However, a suitable engineering approximation for the
reversible adiabatic process with a variable k can be assumed; i.e., the process can
be considered to occur in many innitesimally-small steps and for each of these,
the adiabatic exponent k can be regarded as a constant. For example, any revers-
ible-adiabatic process between states i and j can be regarded as consisting of a series
of numerous innitesimally-small processes, for each of which a slightly dierent va-
lue of k applies. For any of these processes, when innitesimally-small changes in
temperature dT, and volume dV of the working uid ensue, the equation for the
reversible adiabatic process with variable k can be written as follows:
TV
k1
T dTV dV
k1
10
From Eq. (10), we get
k
1
T
j
T
i
b lnT
j
=T
i
RlnV
j
=V
i
11
The compression, r
c
, and expansion, r
e
, ratios are dened as
r
c
V
1
=V
2
and r
e
V
4
=V
3
V
5
=V
2
12
Therefore, equations describing processes 1 ! 2 and 3 ! 4 are, respectively, as
follows:
k
1
T
2
T
1
b lnT
2
=T
1
Rln r
c
13
k
1
T
3
T
4
b lnT
3
=T
4
Rln r
e
14
For an ideal Miller-cycle model, there are no heat-transfer losses. However, for a
real Miller-cycle, heat-transfer irreversibility between the working uid and the cyl-
inder wall is not negligible. It is assumed that the heat loss through the cylinder wall
is proportional to the average temperature of the working uid and the cylinder wall,
and that, during the operation, the wall temperature remains approximately invari-
ant. The heat added to the working uid by combustion is given by the following
linear-relation [8,12,14,18,22]:
Q
in
MA BT
2
T
3
15
where A and B are constants related to the combustion and heat-transfer processes.
Taking into account the friction loss of the piston, as deduced by Angulo-Brown
et al. [9] and Chen et al. [16], and assuming a dissipation term resulting from the fric-
tion force as being a linear function of the velocity, then
f
l
lv l
dx
dt
16
where l is the coecient of friction and x is the piston displacement. Then, the lost
power is
P
l

dW
l
dt
l
dx
dt
dx
dt
lm
2
17
346 A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351
The pistons mean-velocity is
v
x
5
x
2
Dt
5!2

x
2
r
e
1
Dt
5!2
18
where x
2
is the pistons position corresponding to the minimum volume of the
trapped gases and Dt
5!2
is the time spent in the power stroke. Thus, the power out-
put P
output

W
s
P
l
_ _
can be written as
P
output

MbT
3
T
2
T
4
T
5
aT
5
T
1
0.5k
1
T
2
3
T
2
1
T
2
2
T
2
4

C
1
T
3
T
2
C
2
T
4
T
5
C
3
T
5
T
1

_
b
1
r
e
1
2
_
19
where b
1

lx
2
2
Dt
5!2

2
, and the eciency of the cycle g
th

P
output
Q
in
=s
. Thus
g
th

MbT
3
T
2
T
4
T
5
aT
5
T
1
0.5k
1
T
2
3
T
2
1
T
2
2
T
2
4

b
1
r
e
1
2
C
1
T
3
T
2
C
2
T
4
T
5
C
3
T
5
T
1

_ _
MbT
3
T
2
0.5k
1
T
2
3
T
2
2

20
When the values of r
c
, r
e
and T
1
are given, T
2
can be obtained from Eq. (13); then,
substituting from Eq. (7) into Eq. (15) yields T
3
, and T
4
can be found using Eq. (14).
The last unknown is T
5
, which can be deduced from the entropy change assuming an
ideal-gas: rst; the entropy change DS
3!2
between states 2 and 3, is equal to the en-
tropy change DS
4!1
between states 4 and 1. Thus
DS
3!2
DS
4!1
DS
4!5
DS
5!1
21
dS C
v
dT
T
R
dV
V
or dS C
p
dT
T
R
dP
P
22
Processes 2 ! 3 and 4 ! 5 occur at constant volume and 5 ! 1 is a constant-pres-
sure process. By substituting the specic heat from Eqs. (4) and (5) and integrating
from the initial to the nal state of the process, then:
b ln
T
3
T
2
_ _
ln
T
4
T
5
_ _ _ _
a ln
T
5
T
1
_ _ _ _
k
1
T
3
T
2
T
4
T
1
0 23
Substituting T
1
, T
2
, T
3
and T
4
into Eq. (23), we get T
5
and substituting T
1
, T
2
, T
3
, T
4
and T
5
into Eqs. (19) and (20) permits the eciency and power to be estimated.
Then, the relations between the power output and the compression ratio, as well
as between the thermal eciency and the expansion ratio, of the Miller cycle, can
be derived.
Numerical example and discussion
The following constants and parameter values have been used in this exercise:
A = 60,000 J/mol, B = 25 J/molK, b
1
= 33 kW, M = 1.57 10
5
kmol, T
1
= 300 K,
A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351 347
k
1
= 0.004 ! 0.008 J/mol K
2
, b = 20 ! 23 J/mol K, a = 27.5 ! 30 J/mol K, r
e
=
6 ! 13, C
1
= 8.128 10
6
s/K and C
2
= 18.67 10
6
s/K, and C
3
= 10 10
6
s/K
[4]. Cases were studied numerically for values of the expansion ratio (r
e
) from
6 to 13, for k
1
= 0.004, 0.006 or 0.008, for b = 20, 21 or 22, and for a = 27.5, 28,
29 or 30.
Figs. 27 show the eects of the temperature-dependent specic-heat of the work-
ing uid on the thermal eciency of the cycle with heat resistance and irreversible
friction-losses. The thermal eciency versus compression-ratio characteristics are
approximately exponential-like curves. The eciency versus expansion-ratio charac-
teristics approximate to parabolic-like curves. They reect the performance charac-
teristics of a real irreversible Miller-cycle engine.
Figs. 24 show the eects of k
1
, a and b on the performance of the cycle at an
expansion ratio of 10. The thermal eciency increases with increasing compression
ratio, and increasing value of a, but decreases with increases of k
1
and b. The eect of
changing k
1
is less than for b and even less than for a. This is due to the increase of (i)
the heat rejected by the working uid and (ii) the heat added by the working uid.
The magnitude of the thermal eciency becomes much smaller when the parameter
b increases (see Eqs. (7), (8) and (20)). Eq. (20) shows that the parameter b is multi-
plied by the highest temperature-dierence in the cycle: this indicates that the eect
of parameter b will be greater than the eects of the other parameters. Figs. 57 indi-
cate the eects of the parameters a, b and k
1
on the eciency of the cycle for dierent
values of the expansion ratio r
e
. The eciency increases with increasing expansion
ratio, reaches a maximum value and then decreases. The eect of the parameter b
on the eciency is the largest among all parameters for the same reason as men-
tioned earlier. The eciency decreases sharply even with only a slight increase of
b: when b increases by about 1%, the maximum eciency will decrease by about
33%. This is due to the increase in the heat rejected by the working uid as a result
of increasing b (see Eq. (8)).
According to the above analysis, it can be concluded that the eects of the tem-
perature-dependent specic heat of the working uid on the cycle performance are
signicant, and should be considered carefully in practical-cycle analysis and design.
a=28, b=20, r
e
=10
0
5
10
15
20
25
30
35
40
45
50
0 4 8
r
c


(
%
)
10
=0.004
=0.006
=0.008
k
1
2 6
Fig. 2. Eect of k
1
on the variation of the eciency with compression ratio.
348 A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351
r
e
=10, k
1
=0.008, b=20
0
5
10
15
20
25
30
35
40
45
50
55
0 4 8
r
c

(
%
)
10
a=28
a=29
a=30
2 6
Fig. 3. Eect of a on the variation of the eciency with compression ratio.
k
1
=0.008, r
e
=10, a=28
0
5
10
15
20
25
30
35
40
45
0 2 4 6 8
r
c

(
%
)
10
b=20
b=21
b=22
Fig. 4. Eect of b on the variation of the eciency with compression ratio.
r
c
=8, a=28, b=20
20
25
30
35
40
45
50
2 6 10 12 14
r
e

(
%
)
0.004
0.006
0.008
k
1
4 8
Fig. 5. Eect of k
1
on the variation of the eciency with expansion ratio.
A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351 349
Conclusion
An air-standard Miller-cycle model, assuming a temperature-dependent specic
heat of the working uid as well as heat resistance and irreversible friction losses,
has been investigated numerically. The performance characteristics of the cycle show
that there are signicant eects of the temperature-dependent specic heat of the
working uid. A slight increase in some parameters will have a signicant impact
on the thermal eciency of the studied cycle. The results obtained from this research
are compatible with those in the open literature, for other cycles, and may be used with
assurance to provide guidance for the analysis of the behavior and design of practical
Miller-engines. Future studies should discuss the possible eects of fuel additives in
order to achieve a less temperature-dependent specic heat of the working uid.
References
[1] Chen L, Wu C, Sun F. Finite-time thermodynamic optimization or entropy-generation minimization
of energy systems. J Non-Equil Thermodynamics 1999;24(4):32759.
r
c
=8, b=20, k
1
=0.004
20
25
30
35
40
45
50
2 10 12 1
r
e


(
%
)
4
a=27.5
a=28
a=28.5
4 6 8
Fig. 6. Eect of a on the variation of the eciency with expansion ratio.
r
c
=8, k
1
=0.004, a= 28
20
25
30
35
40
45
50
2 4 6 8 10 12 14
r
e


(
%
)
b=20
b=21
b=22
Fig. 7. Eect of b on the variation of the eciency with expansion ratio.
350 A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351
[2] Bejan A. Entropy-generation minimization: the new thermodynamics of nite-size devices and nite-
time processes. J Appl Phys 1996;79(3):1191218.
[3] Chen L, Sun F. Advances in nite-time thermodynamics: analysis and optimization. New
York: Nova Science Publishers; 2004.
[4] Homan KH, Watowich SJ, Berry RS. Optimal paths for thermodynamic systems: the ideal Diesel
cycle. J Appl Phys 1985;58(6):212534.
[5] Mozurkewich M, Berry R. Optimal paths for thermodynamic systems: the ideal Otto cycle. J Appl
Phys 1982;53(1):3442.
[6] Aizenbud BM, Band YB, Kafri O. Optimization of a model internal-combustion engine. J Appl Phys
1982;53(3):127782.
[7] Chen L, Sun F, Wu C. Optimal expansion of a heated working-uid with linear phenomenological
heat-transfer. Energ Conversion Management 1998;39(3/4):14956.
[8] Orlov VN, Berry RS. Power and eciency limits for internal-combustion engines via methods of
nite-time thermodynamics. J Appl Phys 1993;74(10):431722.
[9] Angulo-Brown F, Fernandez-Betanzos J, Diaz-Pico CA. Compression ratio of an optimized Otto-
cycle model. Eur J Phys 1994;15(1):3842.
[10] Chen L, Lin J, Luo J, Sun F, Wu C. Friction eects on the characteristic performance of Diesel
engines. Int J Energ Res 2002;26(10):96571.
[11] Wang W, Chen L, Sun F, Wu C. The eects of friction on the performance of an air standard Dual
cycle. Energy 2002;2(4):3404.
[12] Klein SA. An explanation for the observed compression-ratios in internal-combustion engines. Trans.
ASME J Eng Gas Turbine Power 1991;113(4):5113.
[13] Chen L, Wu C, Sun F, Cao S. Heat-transfer eects on the net work-output and eciency
characteristics for an air standard Otto cycle. Energ Conversion Management 1998;39(7):6438.
[14] Chen L, Zen F, Sun F, Wu C. Heat-transfer eects on the net work-output and power as functions of
eciency for an air-standard Diesel cycle. Energy 1996;21(12):12015.
[15] Lin J, Chen L, Wu C, Sun F. Finite-time thermodynamic performance of a Dual cycle. Int J Energ
Res 1999;23(9):76572.
[16] Chen L, Zheng T, Sun F, Wu C. The power and eciency characteristics for an irreversible Otto-
cycle. Int J Ambient Energy 2003;24(4):195200.
[17] Chen L, Sun F, Wu C. The optimal performance of an irreversible Dual-cycle. Appl Energy
2004;79(1):314.
[18] Chen L, Lin J, Wu C, Sun F. Eciency of an Atkinson engine at maximum power-density. Energy
Conversion Management 1998;39(3/4):33741.
[19] Al-Sarkhi A, Akash BA, Jaber JO, Mohsen MS, Abu-Nada E. Eciency of a Miller engine at
maximum power-density. Int Commun Heat Mass Transfer 2002;29(8):11579.
[20] Sahin B, Kesgin U, Kodal A, Vardar N. Performance optimization of a new combined power-cycle
based on a power-density analysis of the Dual cycle. Energy Conversion Management
2002;43(15):201931.
[21] Qin X, Chen L, Sun F, Wu C. The universal power and eciency characteristics for irreversible
reciprocating heat-engine cycles. Eur J Phys 2003;24(4):35966.
[22] Parlak A, Sahin B, Yasar H. Performance optimization of an irreversible Dual cycle with respect to
pressure ratio and temperature ratio: experimental results of a coated IDI Diesel-engine. Energy
Conversion Management 2004;45(7/8):121932.
[23] Fischer A, Homan KH. Can a quantitative simulation of an Otto engine be accurately rendered by a
simple Novikov model with a heat leak? J Non-Equil Thermody 2004;29(1):928.
[24] Al-Sarkhi A, Jaber JO, Probert SD. Eect of friction and temperature-dependent specic-heat of the
working uid on the performance of a diesel engine. Appl Energy [accepted].
A. Al-Sarkhi et al. / Applied Energy 83 (2006) 343351 351

You might also like