Sunteți pe pagina 1din 9

Comparative exergy analyses of gasoline and hydrogen

fuelled ICEs
Jonathan Nieminen*, Ibrahim Dincer
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology, 2000 Simcoe Street North,
Oshawa, ON L1H 7K4, Canada
a r t i c l e i n f o
Article history:
Received 4 July 2009
Received in revised form
1 September 2009
Accepted 2 September 2009
Available online 12 October 2009
Keywords:
Exergy
Efciency
Irreversibility
Hydrogen
Gasoline
ICE
a b s t r a c t
Comparative exergy models for naturally aspirated gasoline and hydrogen fuelled spark
ignition internal combustion engines were developed according to the second law of
thermodynamics. A thorough analysis of heat transfer, work, thermo mechanical, and
chemical exergy functions was made. An irreversibility function was developed as
a function of entropy generation and graphed. A second law analysis yielded a fractional
exergy distribution as a percentage of chemical exergy of the intake. It was found that the
hydrogen fuelled engine had a greater proportion of its chemical exergy converted into
work exergy, indicating a second law efciency of 41.37% as opposed to 35.74% for
a gasoline fuelled engine due to signicantly lower irreversibilities and lower specic fuel
consumption associated with a hydrogen fuelled ICE. The greater exergy due to heat
transfer or thermal availability associated with the hydrogen fuelled engine occurs due to
a greater amount of convective heat transfer associated with hydrogen combustion.
However, this seemingly high available thermal energy or thermal exergy is misleading
due to the higher cooling load which decreases the power of a hydrogen fuelled ICE. Finally,
a second law analysis of both hydrogen and gasoline combustion reactions indicate
a greater combustion irreversibility associated with gasoline combustion. A percentage
breakdown of the combustion irreversibilities were also constructed according to infor-
mation found in literature searches.
2009 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.
1. Introduction
With an increasing population placing greater strain on
current energy resources and the detrimental effects of the
pollution due to the combustion of fossil fuels, the use of
hydrogen as an alternative fuel such as hydrogen has become
a clean and realistic option for the future development of
internal combustion engines. The changing of fuel from
gasoline to compressed hydrogen gas demands a thermody-
namic analysis according to the second law of thermody-
namics in order to determine relative changes in performance
and efciency. Through analysis, it should be possible to
troubleshoot potential problems and predict performance
characteristics. The second law of thermodynamics will be
employed to model the different exergy functions [1]. Entropy
generation is modeled as a function of convective heat
transfer, cylinder temperature, and dead state temperature
[2]. It is then to be employed in the modeling of thermody-
namic irreversibility as a function of a varying crank angle.
Furthermore, the proportional distribution of the intake
charges chemical exergy in terms of work exergy output, heat
transfer exergy, combustion irreversibilities and others will be
* Corresponding author.
E-mail addresses: jonathan.nieminen@hotmail.com (J. Nieminen), ibrahim.dincer@uoit.ca (I. Dincer).
Avai l abl e at www. sci encedi r ect . com
j our nal homepage: www. el sevi er . com/ l ocat e/ he
0360-3199/$ see front matter 2009 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2009.09.003
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2
determined, displayed, commented upon and reasoning
provided [3]. Second law efciencies for both gasoline and
hydrogen fuelling will be derived from this analysis and
explanations presented.
Plenty of work has been done in the area of the applica-
tion of the second law of thermodynamics to internal
combustion engines to diagnose losses and suggest solutions
for improving engine performance and efciency. Rakopou-
los and Giakoumis [3] explores the development of models
based on the rst and second laws of thermodynamics for
both spark ignition (Gasoline) and compression ignition
(Diesel) engines. The conceptual basis was formed by
dening the exergy term and by developing various exergy
equations and applying them to various engine systems and
control volumes. Second law efciencies were modeled as
a function of compression and equivalence ratios. Thermo-
dynamic modeling and commentary was made on the use of
alternative fuels, bi-fuel combustion, oxygen-enriched (oxy-
fuel) combustion, and different injection techniques such as
indirect and direct injection. Also, the employment of engine
subsystems such as turbochargers, manifolds, and inter-
coolers were reviewed and commented upon. Dunbar and
Lior [4] explores the Second Law of Thermodynamics when
applied to combustion reactions for methane and hydrogen.
Four paths were explored. These paths are analogous to
alternative chamber designs and arrangements to explore
their effect on combustion irreversibilities. They had found
that there were four primary sources of combustion irre-
versibility and that these vary depending on numerous
factors. They had found that with increasing excess air,
combustion irreversibility increased and product tempera-
ture decreased. Furthermore, a percentage breakdown of
each sources contribution to the overall combustion irre-
versibility was calculated for each of the four paths. Graves et
al. [5] discusses potential methodologies for increasing ICE
fuel conversion efciency to that of fuel cells by explaining
that traditional diffusion and mixed ames are the major
reasons for the high degrees of irreversibility. They com-
mented on various other combustion technologies, such as
Homogeneous Charge Compression Ignition (HCCI) and Low
Temperature Combustion (LTC). Also, they commented on
matching work extraction rate to combustion reaction rate in
order to minimize irreversibility by approaching isothermal
combustion. With respect to increasing engine efciency,
they proposed heat recuperation combined with compound
work extraction, low temperature bottoming cycles for
exhaust heat, and increasing compression ratios. The corol-
lary to higher compression ratios would be the employment
of an over expanded engine cycle, in which the expansion
ratio is greater than the compression ratio. Furthermore, it
was illustrated that the choice of fuel is also critical to the
degree of irreversibility present in combustion reactions.
Gasoline was found to have the highest irreversibility relative
to fuel HHV. Methane combustion was shown to be slightly
more reversible and hydrogen was shown to be the most
reversible relative to HHV. Ribiero et al. [2] goe over the
development of an entropy generation model for use in spark
ignition engines. With entropy dened as a function of
temperature, the entropy generation function was a function
of the dead state temperature, cylinder temperature and
convective heat transfer within the cylinder. It was also
found that the entropy generation per cycle was higher for an
Otto Cycle and slightly lower for a Miller Cycle. Therefore, it
is logical to state that Miller Cycles that must employ rela-
tively high compression ratios when compared to the Otto
Cycle [2,5] could reduce the cycle irreversibility by reducing
the entropy generation. Due to this, it can be stated that
Miller Cycle engines are favourable cycles to begin con-
structing ICEs around. Zhang and Sobiesiak [6] investigated
the various exergy distributions relative to fuel exergy and
makes commentary on their ndings. They had found that
CNG fuelling was more efcient from a second law perspec-
tive than gasoline fuelling. It can be stated that if CNG fuel-
ling is more efcient from a second law perspective than
gasoline fuelling and hydrogen less irreversible than CNG [5]
then it could be argued that hydrogen fuelling should have
a much higher proportion of fuel exergy converted to useful
work compared to gasoline fuelling.
In this paper, comparative exergy analyses of gasoline and
hydrogen fuelled internal combustion engines (ICE) are
carried out and the numerical results are graphed and com-
mented upon.
2. Exergy analysis
To construct a thermodynamic model for the working uid
inside an internal combustion engine, and following assump-
tions are made:
The system is a closed system.
The working uid is modeled as an ideal gas.
All the heat losses during the cycle are caused by the engine
cooling systemand only convective heat transfer effects are
considered.
The compression and expansion processes are polytropic.
Valve timing is assumed to be at the beginning/end of each
respective stroke.
Combustion is assumed to be stoichiometric.
Based on the second law of thermodynamics, the exergy
balance of a closed system is given as follows:
_
Ex
w

_
Ex
tm

_
Ex
chem

_
Ex
ht
T
o
_
S
gen
(1)
Here, each termcan be evaluated as long as the corresponding
properties of the working uid have been obtained.
2.1. Exergy due to heat transfer
Since an ICE operates on four distinct cycles, the model
developed within this paper is a piecewise model applied to
both gasoline and hydrogen fuelling scenarios. Since
combustion in an ICE is not isochoric as the Otto cycle
suggests, the Wiebeburn function will be employed to
generate continuous temperature and pressure models over
the entire cycle range as shown in refs [7,8]. Finally, the
exergy due to heat transfer (thermal exergy) is described as
follows:
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5125
_
Ex
ht;i
q
_

_
_
1
To
T
in
q
_
_
Q
i
q; q
TDC
q < q
BDC
_
1
To
Tcompq
_
_
Q
i
q; q
BDC
q < q
SOC
_
1
To
T
comb
q
_
_
Q
i
q; q
SOC
q < q
EOC
_
1
To
Texpq
_
_
Q
i
q; q
EOC
q < q
BDC
_
1
To
T
exh
q
_
_
Q
i
q; q
BDC
q < q
TDC
(2)
Since the dominant form of in-cylinder heat transfer in an ICE
is the convective heat transfer, the Newtons Lawof Cooling is
used to accurately describe the heat transfer within an ICE.
The Woschni correlation is widely used in determining the
convective heat transfer coefcient within an ICE. It is
a function of in-cylinder pressure, gas speed, cylinder bore,
and temperature [9,10]. According Shudo et al. in ref [11], the
combustion in a hydrogen fuelled ICE releases four times
more heat than indicated by the Woschni correlation. To
compensate for this, the convective heat transfer coefcient
as determined by the Woschni correlationwas multiplied by 4.
_
Q
i
q
_

_
h
in;i
qA
w
q
_
T
in;i
q T
wall

; q
TDC
q < q
BDC
h
comp;i
qA
w
q
_
T
comp;i
q T
wall

; q
BDC
q < q
SOC
h
comb;i
qA
w
q
_
T
comb;i
q T
wall

; q
SOC
q < q
EOC
h
exp;i
qA
w
q
_
T
exp;i
q T
wall

; q
EOC
q < q
BDC
h
exh;i
qA
w
q
_
T
exh;i
q T
wall

; q
BDC
q < q
TDC
(3)
h
i
q
_

_
CP
in;i
q
0:8
U
in;i
q
0:8
b
0:2
T
in;i
q
0:55
; q
TDC
q<q
BDC
CP
comp;i
q
0:8
U
comp;i
q
0:8
b
0:2
T
comp;i
q
0:55
; q
BDC
q<q
SOC
CP
comb;i
q
0:8
U
comb;i
q
0:8
b
0:2
T
comb;i
q
0:55
; q
SOC
q<q
EOC
CP
exp;i
q
0:8
U
exp;i
q
0:8
b
0:2
T
exp;i
q
0:55
; q
EOC
q<q
BDC
CP
exh;i
q
0:8
U
exh;i
q
0:8
b
0:2
T
exh;i
q
0:55
; q
BDC
q<q
TDC
(4)
In equation (4), a coefcient was used and depends on the
fuelling scenario. If the fuel is gasoline, then the coefcient is
3.26 [9]. If the fuel is hydrogen, then the coefcient is made to
be 13.04 in order to account for the greater heat release
associated with the hydrogen fuelled ICE [11]. Below, the
relation describing in-cylinder gas speed was developed as in
refs [9,10].
U
i
q
_

_
6:18Ns ; q
TDC
q <q
BDC
4:56Ns ; q
BDC
q <q
SOC
4:56Ns3:2410
3
T
o
_
V
d
Vo
__
DP
comb;i
q
po
_
; q
SOC
q <q
EOC
4:56Ns3:2410
3
T
o
_
V
d
Vo
__
DP
exp;i
q
Po
_
; q
EOC
q <q
BDC
6:18Ns ; q
TDC
q <q
BDC
(5)
2.2. Exergy due to work
For both gasoline and hydrogen fuelled engines, work exergy
is dened as the availability of the systemto do actual work on
a changing control volume against its surroundings. With
respect to a piston-cylinder device, boundary work is the work
required to move the piston against the boundary conditions
and change the cylinder volume. The compression and
expansion processes are assumed to be polytropic and as
a function of cylinder volume. Finally the formula for exergy
due to work can be given by [1]:
Ex
w;i
q
_

_
W
poly;i
q W
o
; q
TDC
q < q
BDC
W
poly;i
q W
o
; q
BDC
q < q
SOC
W
poly;i
q W
o
; q
SOC
q < q
EOC
W
poly;i
q W
o
; q
EOC
q < q
BDC
W
poly;i
q W
o
; q
BDC
q < q
TDC
(6)
W
poly;i
q
_

_
nR
i
T
in;i
q
n1
_
_
P
in;i
q
P
in
_n1
n
1
_
; q
BDC
q < q
SOC
nR
i
T
comp;i
q
n1
_
_
P
comp;i
q
P
BDC
_n1
n
1
_
; q
BDC
q < q
SOC
nR
i
T
comb;i
q
n1
_
_
P
comb;i
q
P
SOC
_n1
n
1
_
; q
SOC
q < q
EOC
nR
i
T
exp;i
q
n1
_
_
P
exp;i
q
P
EOC
_n1
n
1
_
; q
EOC
q < q
BDC
nR
i
T
exh;i
q
n1
_
_
P
exp;i
q
P
BDC
_n1
n
1
_
; q
BDC
q < q
TDC
(7)
where
W
o
q P
o
Vq (8)
Vq V
C

1
2
V
D
_
1 R cos q

R
2
sin
2
q
_ _
(9)
2.3. Thermo mechanical exergy
Thermo mechanical exergy is the available work of
a substance in a system when taken from its initial temper-
ature and pressure to its nal temperature and pressure
relative to the dead state [1]. At any temperature and pressure,
any given substance has thermo physical properties, such as
internal energy, enthalpy and entropy. Since the thermo
mechanical exergy of a substance is a function of pressure,
temperature, and specic volume, the thermo mechanical
exergy changes within an ICE due to the changes in the
aforementioned. Therefore, these properties must be calcu-
lated in order accurately to develop a model for thermo
mechanical exergy change in an ICE. The thermo mechanical
exergy can be determined using the following formula [1]:
Ex
tm;i
q
_

_
u
in;i
q P
o
v
in;i
q T
o
s
in;i
q; q
TDC
q <q
BDC
u
comp;i
q P
o
v
comp;i
q T
o
s
comp;i
q; q
BDC
q <q
SOC
u
comb;i
q P
o
v
comb;i
q T
o
s
comb;i
q; q
SOC
q <q
EOC
u
exp;i
q P
o
v
exp;i
q T
o
s
exp;i
q; q
EOC
q <q
BDC
u
exh;i
q P
o
v
exh;i
q T
o
s
exh;i
q; q
BDC
q <q
TDC
(10)
The internal energy of any mixture was determined to be
a function of temperature from ref [3]. The empirical constant
can be taken from refs [9,10].
u
i
q
_

_
R
i
__

5
n1
an
n
T
in;i
q
n
_
a
6
T
in;i
q
_
; q
TDC
q <q
BDC
R
i
__

5
n1
an
n
T
comp;i
q
n
_
a
6
T
comp;i
q
_
; q
BDC
q <q
SOC
R
i
__

5
n1
an
n
T
comb;i
q
n
_
a
6
T
comb;i
q
_
; q
SOC
q <q0
R
i
__

5
n1
bn
n
T
comb;i
q
n
_
b
6
T
comb;i
q
_
; q0 q <q
EOC
R
i
__

5
n1
bn
n
T
exp;i
q
n
_
b
6
T
exp;i
q
_
; q
EOC
q <q
BDC
R
i
__

5
n1
bn
n
T
exh;i
q
n
_
b
6
T
exh;i
q
_
; q
EOC
q <q00
R
i
__

5
n1
an
n
T
exh;i
q
n
_
a
6
T
exh;i
q
_
; q00 q <q
TDC
(11)
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5126
Since this model assumes that all gases are ideal, therefore
a rearrangement of the ideal gas law can yield a specic
volume term:
v
i
q
_

_
R
i
T
in;i
q
P
in;i
q
; q
TDC
q < q
BDC
R
i
T
comp;i
q
P
comp;i
q
; q
BDC
q < q
SOC
R
i
T
comb;i
q
P
comb;i
q
; q
SOC
q < q
EOC
R
i
T
exp;i
q
P
exp;i
q
; q
c
q < q
BDC
(12)
The relationship describing mixture entropy was found to be
a function of in-cylinder temperature, pressure, and species
mass fraction as given below:
Since the heat addition is not isochoric as the Otto cycle
suggests therefore the mass fractions during the combustion
process vary as reactants are consumed and products are
formed. Due to this, the Weibe burn function was incorpo-
rated to model the entropy of a mixture as its composition
changed. The spark timings are taken from ref [7].
For stoichiometric combustion (with 100% theoretical air),
the internal energy and entropy during combustion and
expansion must be broken into two sections. The reason for
this was because at crank angles q and q the temperature
inside the cylinder became greater than or less than 1000 K
and the set of constants had to change from set a to set b
[9]. The values for q were found to be 5
o
BTDC and 1
o
ATDC for
gasoline and hydrogen respectively. Furthermore, the values
for q were found to be 42
o
and 54
o
BTDC for gasoline and
hydrogen, respectively [7].
2.4. Chemical exergy associated with the intake charge
Thechangeinexergyassociatedwiththefuel-air intakecharge
was broken into three pieces as a piecewise function (intake/
compression: [q
BDC
<q <q
SOC
], combustion: [q
SOC
<q <q
EOC
] and
expansion/exhaust: [q
C
< q < q
BDC
]). It was assumed to be
constant during compression and expansion due to no chem-
ical reaction taking place and the change in chemical exergy
associated with combustion was assumed to be a linear
decrease [6]. The ratio of Gibbs free energy to enthalpy was
foundtobe0.9697for hydrogenand1.0286for gasoline[12]. The
exergy associated with the air was assumed to be ow exergy
[1,13]. The chemical exergy of the intake charge of gasoline and
hydrogen can be calculated with the following equations:
Ex
chem;i
q
_

_
_
DG
o
DH
o
_
LHV
o
fuel
ex
o
air
; q
TDC
q<q
SOC
_
mqmq
SOC

_
DG
o
DH
o
_
LHV
o
fuel
ex
o
air
_
; q
SOC
q<q
EOC
Ex
dest
_
_
DG
o
DH
o
_
LHV
o
fuel
ex
o
air
_
; q
EOC
q<q
TDC
(14)
where
m
i

DG
o
DH
o
_
LHV
o
fuel
ex
o
air
1 Ex
dest

q
EOC
q
SOC
(15)
ex
o
air
R
u
T
o
_
InX
O
2
InX
N
2

(16)
The standard molar chemical exergy of air was determined in
equation[16] [13]. Finally, Ex
dest
is the unavailability associated
with both gasoline and hydrogen combustion reactions; and
they are determined from a second law analysis applied to
combustion reactions [1].
2.5. Irreversibility
The fourth term of equation [1] represents the exergy
destruction caused by the irreversibility. Irreversibility is
a function of entropy generation, which increases with
temperature and describes the unavailability of a system [1].
The combustionreactionitself has several mechanisms which
require some of that energy in order to drive the reaction.
According to Dunbar and Lior in ref [4], a combustion reaction
has four major sources of internal irreversibility. They are:
A chemical diffusion process in which air and fuel mole-
cules are drawn together.
Combustion of the fuel-air mixture (thermo chemical
reaction).
Internal energy exchange through molecular collisions
amongst the products and radiation heat transfer amongst
product constituents due to unequal heat distribution.
Mixing process whereby reactants mix before combustion,
and products mix with reactants during combustion due to
proximity.
Cengel and Boles [1] and Moran and Shapiro [12] deter-
mined irreversibility to be a product of dead state of temper-
ature and entropy.
s
i
q
_

_
R
i
__
a
i;1
InT
in;i
q
_

5
n1
a
i;n
n1
T
in;i
q
n1
_
In
_
y
i
qP
in;i
q
Po
_
__
; q
TDC
q < q
BDC
R
i
__
a
i;1
InT
comp;i
q
_

5
n1
a
i;n
n1
T
comp;i
q
n1
_
In
_
y
i
qP
comp;i
q
Po
_
__
; q
BDC
q < q
SOC
R
i
__
a
i;1
InT
comb;i
q
_

5
n1
a
i;n
n1
T
comb;i
q
n1
_
In
_
y
i
qP
comb;i
q
Po
_
__
; q
SOC
q < q0
R
i
__
b
i;1
InT
comb;i
q
_

5
n1
b
i;n
n1
T
comb;i
q
n1
_
In
_
y
i
qP
comb;i
q
Po
_
__
; q0 q < q
EOC
R
i
__
b
i;1
InT
exp;i
q
_

5
n1
b
i;n
n1
T
exp;i
q
n1
_
In
_
y
i
qP
exp;i
q
Po
_
__
; q
EOC
q < q
BDC
R
i
__
a
i;1
InT
exh;i
q
_

5
n1
b
i;n
n1
T
exh;i
q
n1
_
In
_
y
i
qP
exh;i
q
Po
_
__
; q
EOC
q < q00
R
i
__
a
i;1
InT
exh;i
q
_

5
n1
a
i;n
n1
T
exh;i
q
n1
_
In
_
y
i
qP
exh;i
q
Po
_
__
; q00 q < q
TDC
(13)
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5127
I
i
q
_

_
T
o
s
i
q; q
TDC
q < q
BDC
T
o
s
i
q; q
BDC
q < q
SOC
T
o
s
i
q; q
SOC
q < q
EOC
T
o
s
i
q; q
EOC
q < q
BDC
T
o
s
i
q; q
BDC
q < q
TDC
(17)
Based on the above models, the exergy changes of the working
uids can be calculated and graphed for during the intake,
compression, combustion, expansion, and exhaust processes
for both hydrogen and a gasoline fuelled engines.
3. Results and discussion
3.1. Exergy due to heat transfer
Fig. 2 illustrates the variation of exergy due to heat transfer as
a function of crank angle. It can be seen that the combustion
process has the greatest impact on the exergy due to heat
transfer for both engines. The combustion durations from ref
[7] were used in this study. They were found to be [30
o
BTDC,
40
o
ATDC] and [8
o
BTDC, 18
o
ATDC] for gasoline and hydrogen
fuelling, respectively. Compared to the combustion process,
the rate of heat transfer exergy changes are smaller during
compression and expansion processes for both engines. The
greater rate of increase of thermal availability in the hydrogen
engine when compared to gasoline can be best explained by
the shorter combustion duration coupled with the higher
combustion temperatures associated with the hydrogen
fuelled engine [1,12]. Furthermore, the greater thermal avail-
ability of a hydrogen fuelled ICE is misleading. The relation-
ship for convective heat transfer coefcient was adapted from
Woschnis relation which, when adapted to a hydrogen fuel-
led engine undershoots the measured amount of heat transfer
by approximately a factor of four [11,14]. Therefore, a greater
amount of the thermal availability associated with a hydrogen
fuelled engine will leave the system via heat transfer to the
cooling systemthus reducing the work output of the hydrogen
engine [11,14]. It was found that the exergy change due to heat
transfer occupied approximately 19.3% and 27.3% of the total
exergy for gasoline and hydrogen fuelled engines respectively.
The results for gasoline fuelling are relatively consistent with
results reported by refs. [3,6]. Therefore, the same process
applied to a hydrogen fuelled ICE should yield results with
a similar degree of accuracy.
3.2. Exergy due to work
The relationship between work exergy and crank angle is
depicted in Fig. 3. The compression process occupies [BDC, 30
o
BTDC] and [BDC, 8
o
BTDC] for gasoline and hydrogen fuelling
respectively. The compression of hydrogen transfers a greater
amount exergy to the system compared with the gasoline due
to the greater degree of compressibility associated with
hydrogen, the larger gas constant, and extended duration of
the compression stroke [1]. Also, this transfer of exergy to the
system via compression work is the reason for the negative
value of the curve during compression. The combustion of
hydrogen increases the engines potential to do less work than
gasoline [7]. During the expansion stroke it can be seen that
a greater decrease in exergy takes place in the hydrogen
fuelled engine compared to the gasoline fuelled engine.
Therefore, it can be expected that the hydrogen fuelled engine
will have a greater proportion of its fuel charge exergy
Fig. 1 Piston-cylinder diagram [modied from Ref. 5].
Fig. 2 Specic heat transfer exergy vs crank angle.
Fig. 3 Specic work exergy vs. crank angle.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5128
converted to work exergy. By extension, this will lead to the
increase in second law efciency for the hydrogen fuelled
engine when compared to a gasoline fuelled engine. It was
found that greater proportion of the intake charge exergy was
converted into work exergy. In a hydrogen fuelled ICE,
approximately 41.37% of the chemical exergy was converted
to work exergy. As expected, the gasoline fuelled ICE had
a slightly lower second law efciency. Approximately 35.74%
of the chemical exergy was converted to work exergy Since
a gasoline fuelled ICE with f1.1 and r
c
9.3:1 had a second
law efciency of approximately 30% [6], therefore it is
reasonable to state that the process used to calculate the
second law efciency for a gasoline fuelled engine accurately
would yield accurate results for the second law efciency for
an ideal hydrogen fuelled engine when applied.
3.3. Thermo mechanical exergy
In Fig. 4, thermo mechanical exergy of both hydrogen and
gasoline fuelled engines are plotted against crank angle. This
plot illustrates the ability of a systems mass at various state
points to do actual work. The hydrogen fuelled engine had
a lesser thermo mechanical exergy during expansion and
exhaust strokes due to a much lower cycle pressures and
exhaust gas temperatures [7]. Furthermore, the behaviour of
the function was found to match the behaviour of the thermo
mechanical exergy function in other literature analyzing
gasoline fuelled ICEs [6]. Therefore, when applied it is
reasonable to state that the process for determining the
behaviour of thermo mechanical exergy varying with crank
angle for a hydrogen fuelled ICE has been done correctly.
3.4. Chemical exergy associated with the intake charge
Fig. 5 illustrates the behaviour of the exergy associated with
the intake charge for both gasoline and hydrogen fuelled
engines. Equation (17) describes the exergy of the intake
charge for gasoline and hydrogen respectively. It also
describes the compression processes as applied to the
compression stroke duration. The combustion processes are
assumed to be linear decreases in fuel exergy [6] and are
described by equation (14) applied over the combustion
duration. Finally, the expansion processes for both gasoline
and hydrogen are described by equation (14) applied to the
expansion stroke duration. Hydrogen has a LHV that is
approximately triple the LHV of gasoline and a ratio of Gibbs
free energy to enthalpy which is approximately 90% of that of
gasoline [1,12]. These two properties of hydrogen account for
the greater exergy of its intake charge despite the lower
specic fuel consumption associated with hydrogen fuelled
ICEs. Finally, since the combustion of hydrogen is a much less
irreversible process than gasoline combustion, the amount of
exergy within the exhaust products is much less than that of
gasoline [4,5].
3.5. Irreversibility
Fig. 6 illustrates irreversibility as a function of crank angle by
the graphing of equation (17) over the entire four stroke cycle.
Since irreversibility was found to be a function of entropy
[1,2] and entropy was greatly affected by the combustion
reaction, therefore the effect of combustion was more
notable in the combustion strokes of both fuelling scenarios.
Therefore, as in Fig. 1, irreversibility is primarily inuenced
by the combustion process which contributes the most to
cylinder convective heat transfer [2]. Similar to Fig. 1, the
compression stroke contributes little when compared to the
Fig. 4 Specic thermo mechanical exergy vs. crank angle.
Fig. 5 Specic chemical exergy vs. crank angle.
Fig. 6 Specic irreversibility vs. crank angle.
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5129
combustion process. However, the expansion and exhaust
strokes in the hydrogen fuelled engine have a relatively lesser
irreversibility associated with it due to the lower tempera-
tures and pressures of the expanding gas [1,7]. Furthermore,
the greater overall irreversibility associated with the use of
hydrogen as a fuel is primarily due to the higher tempera-
tures causing higher amounts of convective heat transfer
taking place within the cylinder. However, due to the greater
availability associated with the hydrogen intake charge the
expected decrease in second law efciency due to greater
irreversibility is offset.
3.6. Combustion irreversibilities
An irreversibility analysis is done for both gasoline-air and
hydrogen-air combustion reactions using the approach given
in the combustion section of ref. [1] under stoichiometric
conditions. As shown in Fig. 7, it was found that that the
combustion of hydrogen is less irreversible than the combus-
tion of gasoline. Approximately 11.72% (14.07 MJ/kg) of the
LHV (120.1 MJ/kg) is unavailable for combustion in a hydrogen
fuelled engine. This is less than the 29.09% (12.8 MJ/kg)
unavailable for LHV (44.0 MJ/kg) associated with the gasoline
fuelled engine. Finally, as shown in Fig. 8, the intake manifold
on an internal combustion engine was approximated to be
a mixer and preheater and is therefore is treated as path4 as
outlined by Dunbar and Lior [4]. Within path 4, a breakdown
of irreversibilities was provided to be:
a) 77% of the total irreversibility is caused by preheating in
the intake manifold, reactants mixing with leftover
combustion products, and heating during the compression
stroke.
b) 14% of the total irreversibility is caused by the exothermic
nature of combustion reaction which requires some of its
own energy to act as a thermal driver for the reaction to
occur.
c) 9% of the total irreversibility is caused by the mixing of
reactants in the intake manifold, and the mixing of left-
over combustion products and reactants before
combustion.
In summary, it is really critical to investigate the exergetic
performance aspects of hydrogen fuelled ICEs and compare
with the corresponding parameters as obtained for a gasoline
ICE. The need for such works is emphasized in various refer-
ences [1517]. The present work has aimed to contribute to the
area by fullling such a need.
4. Conclusions
This comparative exergy analysis between gasoline and
hydrogen fuelled internal combustion engines has indicated
that a hydrogen fuelled engine is more efcient than a gaso-
line fuelled engine from a second law perspective, converting
41.37% of the intake charge exergy into useful work as
opposed to 35.74% for gasoline fuelling. Some reasons to
explain the greater second lawefciency of a hydrogenfuelled
engine would be that the compression of hydrogen requires
less work due to greater compressibility, the more efcient
combustion of hydrogen is able to convert more chemical
potential into work, and lower specic fuel consumption.
However, due to lower cycle pressures the hydrogen engine
will have a lower mean effective pressure (MEP) and therefore
produce less power for the same displacement of engine. The
Fig. 7 Proportional combustion irreversibilities for gasoline and hydrogen fuelled ICEs.
Fig. 8 Comparative exergy distributions.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5130
hydrogen fuelled ICE converts 27.3% of its intake charge into
thermal availability as opposed to 19.3% for a gasoline fuelled
ICE. This is because of a greater amount of heat transfer due to
higher combustion temperatures and a greater amount of
heat transfer fromthe engine to the coolant due to the shorter
quenching distance. This higher cooling load further
decreases the work output of a hydrogen fuelled engine rela-
tive to a gasoline fuelled engine. Furthermore, the combustion
of hydrogen was shown to be a less irreversible than the
combustion of gasoline. Only 11.72% of the hydrogen engines
intake charge is unavailable as opposed to the 29.09%
unavailability associated with the gasoline fuelled engine.
Finally, the analysis conducted in this study shows that
a hydrogen fuelled internal combustion engine has a greater
second law efciency, decreased work output, and less irre-
versibilities associated with the combustion reaction itself.
Acknowledgement
The authors acknowledge the support provided by the Natural
Sciences and Engineering Research Council.
Nomenclature
Abbreviations and symbols
A Exposed area (m
2
)
a Polynomial constants
ABDC After bottom dead center
ATDC After top dead center
b Bore (m)
BBDC Before bottom dead center
BDC Bottom dead center
BTDC Before top dead center
C HTC constant (C3.26 or 13.04)
d Displacement
EOC End of Combustion
Ex Exergy (J/g)
h Convective heat transfer coefcient
I Irreversibility (W)
k Ratio of specic heats(k 1.40)
LHV Lower heating value (MJ/kg)
m Rate of exergy decrease (J/g*

CA)
n Polytropic Constant(n1.30)
P Pressure (kPa)
Q Convective heat transfer (W)
R Ratio of connecting rod length to stroke
r
c
Compression Ratio (-)
R
u
Universal Gas Constant (J/mol*K)
s Stroke (m)
s Specic entropy (J/kg*K)
S
gen
Entropy generation (J/kg*K)
SOC Start of Combustion
T Temperature (

K)
TDC Top dead center
U Gas speed (m/s)
U
p
Mean piston speed (m/s)
V Volume (m
3
)
v Specic volume (m
3
/kg)
X Mole fraction
Y Mass fraction
Subscripts and superscripts
AIR Air
c Clearance
chem Chemical
comb Combustion
comp Compression
d Displacement
dest Destroyed
fuel Fuel
ht Heat Transfer
i Gasoline or Hydrogen
mix mixture
o Dead State
poly Polytropic
tm Thermomechanical
w Work
wall Wall
o Molar
Greek letters
q Degrees of crank angle (

CA)
Dq Combustion duration (

CA)
4 Equivalence ratio ()
DG

Gibbs free energy


DH

Enthalpy
r e f e r e n c e s
[1] Cengel Y, Boles M. Thermodynamics: an engineering
approach. McGraw-Hill Education; 2005.
[2] Ribiero B, Martins J, Nunes A. Entropy generation in spark
ignition engines (ISSN 1301-9724). International Journal of
Thermodynamics 2007;10(2):5360.
[3] Rakopoulos CD, Giakoumis EG. Second law analyses applied
to internal combustion engine operations. Progress in Energy
and Combustion Science 2006;32(1):247.
[4] Dunbar WR, Lior N. Sources of combustion irreversibility.
Combustion Science and Technology 1994;103(16):4161.
[5] Graves R, Daw S, Conklin J, Chakravarthy VK. Advanced
combustion and emission research for high efciency
engines. 2004. Advanced Combustion Engine R&D 2006;(21
28):2532.
[6] Zhang S, Sobiesiak A. The First and second law analyses of
a port injected, spark ignition engine fuelled with
compressed natural gas (CNG); 2002.
[7] Nieminen J. Comparative combustion characteristics of
gasoline and hydrogen fuelled ICEs (ICH2P-09 Edition),
Elsevier Press. International Journal of Hydrogen Energy 2009.
[8] ErikssonL, AnderssonI. Ananalytic model for cylinder pressure
in a four stroke S.I engine. 2002-01-0371. SAE; 2002. 19.
[9] Heywood J. Internal combustion engine fundamentals.
McGraw-Hill Education; 1988.
[10] Ferguson CR, Kirkpatrick AT. Internal combustion engines:
applied thermosciences. Wiley; 2004.
[11] Shudo T, Suzuki H. Applicability of heat transfer equations to
hydrogen combustion. JSAE Review 2002;23(3):3038.
[12] Moran MJ, Shapiro HN. Fundamentals of engineering
thermodynamics. 6th ed. New York: Wiley and Sons; 2007.
[13] Sato N. Chemical energy and exergy: an introduction to
chemical thermodynamics for engineers. Elsevier Press; 2004.
i nt e r na t i o na l j o ur na l o f hy d r og e n e ne r gy 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5131
[14] DSouza N. Comparative heat transfer analysis of
gasoline and hydrogen fuelled internal combustion
engines, Undergraduate UOIT Thesis Report.
Apr. 2009.
[15] Rakopolous CD, Michos CN. Generation of combustion
irreversibilities in a spark ignition engine under biogas-
hydrogen mixtures fueling. International Journal of
Hydrogen Energy 2009;34(10):442237.
[16] Verhelst S, Sierens R. A quasi-dimensional model
for the power cycle of a hydrogen fuelled ICE.
International Journal of Hydrogen Energy 2007;32(15):
354554.
[17] Szwaja S, Bhandary KR, Naber JD. Comparisons of hydrogen
and gasoline combustion knock in a spark ignition engine.
International Journal of Hydrogen Energy 2007;32(18):
507687.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 5 ( 2 0 1 0 ) 5 1 2 4 5 1 3 2 5132

S-ar putea să vă placă și