You are on page 1of 10

Proceedings of the ASME 2011 International Mechanical Engineering Congress & Exposition

IMECE2011
Proceedings of the ASMENovember
2011 International
Mechanical
11-17, 2011,
Denver, Engineering
Colorado, USA
Congress & Exposition
IMECE 2011
November 11-17, 2011, Denver, Colorado, USA

IMECE2011-62423
IMECE2011-62423

THERMODYNAMIC ANALYSIS OF A MULTI- FUELED SINGLE CYLINDER SI


ENGINE

M. Z. Haq
Department of Mechanical Engineering
Bangladesh University of Engineering and Technology
Dhaka-1000
Bangladesh
Email: zahurul@me.buet.ac.bd

NOMENCLATURE
ach Chemical exergy (kJ/kg)
ath
Thermomechanical exergy (kJ/kg)
atot Total exergy (kJ/kg)
B
Engine cylinder bore (m)
g
Specific Gibbs energy (kJ/kg)
h
Specific enthalpy (kJ/kg)
k
Specific heat capacity ratio (-)
N
Engine speed (rpm)
P
Gas pressure inside cylinder (kPa)
R
Specific gas constant (kJ/kg-K)
rc
Compression ratio (-)
s
Specific entropy (kJ/kg-K)
St
Engine stroke length (m)
T
Gas temperature inside cylinder (K)
u
Specific internal energy (kJ/kg)
V
Volume (m3 )
xi
Mole fraction of species i (-)
yi
Mass fraction of species i (-)

Crank angle (deg)

Specific chemical potential (kJ/kg)

ABSTRACT
The paper presents a thermodynamic analysis of a single
cylinder four-stroke spark-ignition (SI) engine fuelled by four
fuels namely iso-octane, methane, methanol and hydrogen. In SI
engines, due to phenomena like ignition delay and finite flame
speed manifested by the fuels, the heat addition process is not
instantaneous, and hence Weibe function is used to address
the realistic heat release scenario of the engine. Empirical
correlations are used to predict the heat loss from the engine
cylinder. Physical states and chemical properties of gaseous
species present inside the cylinder are determined using first
and second law of thermodynamics, chemical kinetics, JANAF
thermodynamic data-base and NASA polynomials. The model
is implemented in FORTRAN 95 using standard numerical
routines and some simulation results are validated against data
available in literature. The second law of thermodynamics is
applied to estimate the change of exergy i.e. the work potential
or quality of the in-cylinder mixture undergoing various phases
to complete the cycle. Results indicate that, around 4 to 24% of
exergy initially possessed by the in-cylinder mixture is reduced
during combustion and about 26 to 42% is left unused and
exhausted to the atmosphere.
Keywords: SI engine modelling, alternative fuels, thermodynamic analysis, exergy.

Address

M. R. Mohiuddin
Department of Mechanical Engineering
Bangladesh University of Engineering and Technology
Dhaka-1000
Bangladesh
Email: reaz.mohiuddin@gmail.com

1 INTRODUCTION
Improvement in energy utilization can be achieved in many
instances without elaborate analysis through common-sense,
good housekeeping and leak-plugging practices. But as thrifty

all correspondence to this author.

c 2011 by ASME
Copyright

available to produce useful work, while the remaining portion is


unavailable due to the developed irreversibility. Exergy represents the upper limit on the amount of work an engine can deliver
without violating any thermodynamic law, and irreversibility can
be viewed as wasted work potential or the lost opportunity to do
useful work [4]. This lost opportunities manifest themselves in
environmental degradation and avoidable emissions. The performance of an engine can be improved by minimising irreversibility associated with it [1].

engineering methods become more complex, the need for indepth analysis becomes apparent. Often this must go far beyond energy accounting [1]. The first law of thermodynamics
has widely been used as a primary tool to assess the performance
of energy conversion devices. It is an expression of energy conservation which asserts that energy is a thermodynamic property which can be transformed from one form to another but its
total amount is conserved. The second law of thermodynamics
asserts that energy has quality as well as quantity. It provides
an alternative and revealing mean of assessing and comparing
processes and systems rationally and meaningfully by introducing exergy. Exergy based analysis yields a true measure of how
nearly actual performance approaches the ideal, and clearly identifies the causes and the location of thermodynamic losses and
consequent impact on the environment. Exergy analysis can assist in improving and optimizing design [2].
Exergy is defined as the maximum theoretical work that can
be obtained from a combined system (combination of a system
and its reference environment) when the system comes into equilibrium (as thermally, mechanically and chemically) with the environment without violating any law of thermodynamics. The
maximum available work from a system emerges as the sum of
two contributions: thermomechanical exergy, ath and chemical
exergy, ach [2]. Thermomechanical exergy is defined as the maximum extractable work from the combined system as the system comes into thermal and mechanical equilibria with the environment, and its value is calculated with respect to restricted
dead-state condition. At the restricted dead-state conditions, no
work potential exists between the system and the environment
due to temperature and pressure differences [3]. At the restricted
dead-state, the control mass is in thermomechanical equilibrium
with the environment, but not necessarily in chemical equilibrium with it. In principle, the difference between the compositions of the system at the restricted dead-state and the environment can be used to obtain additional work to reach chemical
equilibrium. The maximum additional work obtained in this way
is called the chemical exergy [1].
In the present work, actual processes during spark-ignition
engine operation is developed by incorporating phenomenons
like phased burn combustion, spark advance, heat exchange of
the cylinder gases with the chamber walls and calculation of ten
chemical species (and fuel) concentration during combustion, at
a resolution of one-tenth of an engine crank angle degree. Unlike most previous studies on air-standard power cycles in which
air was assumed as the working fluid and an ideal gas with constant specific heat, the model developed in this study takes into
account the temperature and composition dependence of the specific heats. The results from the first law analysis of the real cycle
are compared favourably with the relevant experimental results.
The thermodynamic state points, determined from the first law
analysis are used to determine the exergy leading to the revelation of the fact that a portion of a given amount of energy is

2 MODELLING, SIMULATION and ANALYSIS


2.1 Governing Equations of the SI Engine Cycle
A simulation model of a single cylinder SI engine is developed which is fuelled by four different fuels. In SI engines, compression, combustion and expansion of premixed gaseous charge
occurs after the inlet valve close (IVC) and before exhaust valve
open (EVO) in a closed system. The closed-system differential
energy equation for a small crank-angle, d , is

Qnet W = dU

(1)

since, W = PdV and dU = mCv dT , and assuming ideal gas


behaviour (PV = mRT ), the cylinder pressure, P is expressed
by [5] as a function of the crank angle, , cylinder pressure, P,
cylinder volume, V and the net heat input, Qnet :
P dV
1 dQnet
dP
= k
+ (k 1)
d
V d
V d

(2)

The value of specific heat ratio, k = C p /Cv is important for an


accurate heat release and engine pressure development analysis.
It couples the systems energy to other thermodynamic quantities. It influences the size and shape of the heat release trace
and cylinder pressure development [6]. However, most analyses
assume constant values of k or use some simple correlation to
estimate k as a function of temperature, and neglects the effect of
constituents and their composition. In the present study, realistic
values of ks are estimated.
In order to integrate Eq. 2, the cylinder volume, V ( ) is required that is given by [5] as:

V ( ) =

VD
VD 2L
+
+ 1 cos
rc 1
2
St

s

2L
St

2

sin2

(3)
where, VD (= 4 B2 St ) is the displacement volume, B is the bore
and St is the stoke, rc is the compression ratio and L is the connecting rod length.
2

c 2011 by ASME
Copyright

Net heat transfer Qnet can be expressed as the heat release


from fuel, Q f uel minus the net heat loss

Qnet ( ) = Q f uel ( ) Qloss ( )

The heat transfer coefficient, hg ( ) is the instantaneous area


averaged heat transfer coefficient, and is related to engine parameters in Woschni correlation as:

(4)
hg ( ) = 3.26 P0.8U 0.8 b0.2 T 0.55

In ideal Otto cycle, the fuel is assumed to burn instantaneously which results in constant volume combustion at TDC.
Actual engine pressure and temperature data do not match the
simple model as finite time is required for an actual fuel-air mixture to burn completely. The rate at which actual fuel-air mixture
burns increases from a low value immediately following spark
discharge to a maximum value about halfway through the burning process and then decreases to close to zero as the combustion
process ends [7]. A functional form, widely known as Wiebe
function, is often used to represent the mass fraction burned,
yb ( ):

yb ( ) =

where the units of hg , P, U, b, and T are W/m2 K, kPa,


m/s, m and K, respectively. Here, U is the characteristic gas
velocity which is proportional to the mean piston speed during
intake, compression, and exhaust. During combustion and expansion, it is assumed that the gas velocities are increased by the
pressure rise resulting from the combustion, so the characteristic
gas velocity is affected by both mean piston speed, U p = 2NSt ,
and cylinder pressure rise, Pc = P Pmotor . Motoring pressure,
Pmotor is obtained by simulating the code without heat release.

h
 i

s
1 exp
if s s + d
d
(5)
0
if < s , > s + d

ToV Pc
U = 2.28U P + 0.00324
VD Po

where is the crank angle, the start of combustion is at s ,


d is the total combustion duration (yb = 0 to yb = 0.99), is
Wiebe form factor, and is Wiebe efficiency factor. These
Wiebe parameters are adjustable parameters and values of = 5
and = 3 have been reported to fit well with experimental data
[7]. Heat release from the fuel, Q f uel ( ) can be written as:

Q f uel ( ) = QLHV yb ( )

(10)

2.2 Simulation of SI Engine Model


In the present study, governing equations of the SI engine
cycle are numerically solved for each time step, the time step
used is for 0.1o crank-angle and the solution marches in time.
The computer code is written in FORTRAN 95. Before the start
of the computation, the design characteristics of the engine are
given (Table 1) as well as the initial conditions at the start of
the cycle when the inlet valve closes (IVC). All liquid fuels are
considered vapour as they enter the engine. By not considering the vaporization energy of the liquid fuels, the comparison
becomes more meaningful. Since the vaporization process is a
complex combination of fluid flow, heat transfer and fuel properties, it is difficult to establish a consistent and fair technique
for including this process. For example, stoichiometric methanol
and iso-octane premixtures have vaporization energies of 148
and 19.1 kJ/kg-mix, respectively, and these vaporization energies
would represent about a 128 and 19 K, respectively decrease in
air temperature [8]. The calculation stops at the end of power
stroke when the exhaust valve opens (EVO).

(6)

The heat loss, Qloss ( ) to the exposed cylinder wall at an


engine speed, N can be determined with a Newtonian convection
equation:


d
Qloss ( ) = hg ( )Aw ( ) T ( ) Tw
2 N

(9)

(7)

The cylinder wall temperature, Tw is the area-averaged mean


temperatures of the exposed cylinder wall, the head and the piston crown. The exposed cylinder area, Aw ( ) is the sum of the
cylinder bore and head area, and the piston crown area.

Calculation of the value of specific heat capacity ratio, k in


Eq. 2 requires the estimation of thermodynamic properties of the
species and their equilibrium chemical composition. Estimation
of product equilibrium composition is reported in 2.3. Thermodynamic properties of individual species and their mixture are
calculated by suitable expressions, as explained in 2.4. To validate the predicted values, cylinder pressure is selected as the
comparison parameter, and predictions are in good agreement
with the experimental data, as discussed in 3.0.

Aw ( ) = Awall + Ahead + A piston

s 
2
1 2  2L 

2L
= VD
+
+ 1 cos
sin2
St B

St
St

(8)
3

c 2011 by ASME
Copyright

2.3

Estimation of Equilibrium Composition


A general form of chemical reaction can be written as:

i Mi = 0

The mole and mass fraction of the fuel in the stoichiometric fuelair mixture is
(11)

xs =

where, the i are the stoichiometric coefficients and by convention are positive for the product species and negative for the
reactant species [7]. For the chemical reaction, equilibrium constant at constant pressure, K p is related to standard state specific
Gibbs function, go :


go
ln K p =
RT

Kp =
i

Pi
Po

i

P
Po

 i


= (x3 + x5 )Nt
(14)

= (2x4 + 2x6 + x7 + x10 )Nt
(15)
O : + 0.42 = (2x2 + x3 + 2x5 + x6 + x8 + x9 + x10 )Nt (16)
N : + 1.58 = (2x1 + x9 )Nt
(17)

where, Nt (= i ) is the total number of moles and xi are the


moles fractions satisfying the condition:
(13)

xi = 1

In the combustion products of fuels, the major species


present at low temperature are N2 , H2 O, CO2 along with O2 or
CO and H2 for lean and rich mixtures respectively. At higher
temperatures (greater than about 2200 K), some major species
dissociate and react to form additional species in significant
amounts. In the chemical equilibrium modelling done in the
present work, ten species (and fuel) are considered to be present
in the product.
Chemical reaction of combustion of a fuel, C H O N in
dry air containing 21% Oxygen and 79% Nitrogen, by volume
can be written as [5]:

f uel

(18)

The introduction of the following six equilibrium constants


for six non-redundant reactions provide the required equations
to solve the ten unknown mole fractions, xi , and total number of
moles, Nt :

1
*H
H2 )
2
1
O2 *
)O
2

air

}|
{
z }| { z
C H O N + (0.21O2 + 0.79N2) = 1 N2 + 2 O2 + 3CO
+ 4 H2 + 5CO2 + 6 H2 O + 7 H + 8 O + 9NO + 10 OH

is the molar fuel-air ratio and is fuel-air equivalence ratio


which is defined as:
(F/A)a
=
(F/A)s

K p1 =

K p2 =

1
1
H2 + O2 *
) OH
2
2

K p3 =

1
1
O2 + N2 *
) NO
2
2

K p4 =

1
H2 + O2 *
) H2 O
2

K p5 =

1
CO + O2 *
) CO2
2

K p6 =

x7 P1/2
x4
x8 P1/2
1/2

x2
x10
1/2

1/2

1/2

1/2

x2 x4
x9
x1 x2
x6
1/2

x2 x4 P
x5
1/2

x2 x3 P1/2

(19)
(20)
(21)
(22)
(23)
(24)

In the above reactions, expressions of the equilibrium constants K p s are obtained using Eq. 13, values of K p s (Eqs. 19-24)
are estimated using Eq. 12. Values of Gibbs fuctions, as required
in Eq. 12 are estimated using detailed thermodynamic analysis
and the required equations are briefly presented in 2.4.

Hence, (F/A)a is the actual fuel air ratio and (F/A)s is the stoichiometric fuel-air ratio and is given as:
(F/A)s =

(F/A)s
(F/A)s + 1

C:
H:

(12)

xi i

ys =

Atom balancing of the combustion chemical reaction yields


the following four equations:

The effect of pressure on the equilibrium composition is related to mole fractions of the involved as:


,
+1

(12.01 + 1.008 + 16.00 + 14.01 )


28.85
4

c 2011 by ASME
Copyright

2.4

Thermodynamic Properties and Estimation


The NASA polynomials are used to determine thermodynamic properties of mixture constituent species at different temperatures and at the standard pressure. These polynomials are
curve fits that express thermodynamic properties as functions of
temperature. The temperature range over which these polynomials are applicable is 300 K to 5000 K. Equations 25-27 give the
polynomials for determining constant pressure specific heat, enthalpy and entropy for each species i at standard pressure and at
temperature T :
Cp
= a1 + a2 T + a3 T 2 + a4 T 3 + a5 T 4
R
ho
a2
a3
a4
a5
a6
= a1 + T + T 2 + T 3 + T 4 +
RT
2
3
4
5
T
so
a3 2 a4 3 a5 4
= a1 ln T + a2T + T + T + T + a7
RT
2
3
4

Hence, the expressions for these exergies are given by [2] as:
n

ath = u + Pov Tos yi i,0

ach = yi (i,0 i,00 )

s = R ln(P/Po ) +

(33)

i=1

The quantity i,0 is the chemical potential of species i in


the restricted dead state, whereas i,00 represents the value in the
environmental dead state. The subscript 0 is used to identify
properties of the thermomechanical or restricted dead state, and
the subscript 00 is used to represent the properties at the environmental dead state. The numerical values of (To , Po ) for the
dead state are those of dead environmental state, assumed 300 K
and 101.325 kPa in the present study. For ideal gas:

(25)
(26)
(27)

If the gaseous mixture is assumed to be an ideal gas, then the effect of pressure on specific heats and enthalpy are negligible, and
the value of entropy at pressures other than 1 atm is determined
by using the following equation:
n

(32)

i=1

i,T = goi,T

Pi
+ RT ln
Po

(34)

and,
yi (soi Ri ln xi )

(28)

Pi,0
i,0 i,00 = RTo ln
Pi,00

i=1

where, yi is the mass fraction of species i.


From the estimated values of enthalpy and entropy at a given
temperature, values of internal energy, u and Gibbs free energy,
g can be determined using the following equations:
g = h Ts

(35)

Details of the modelling and simulation are reported in [9].


Modelled engine specification is reported in Table 1, and some
of the estimated key thermodynamic parameters are reported in
Table 2. Results obtained in the present study are presented in
3.

(29)

TABLE 1. Engine specification, same as in [10].

u = h RT

(30)

Properties of the mixture are estimated using: u =


y
u
i i , h = yi hi , s = yi si and mass fraction, yi is estimated
using yi = xi Mi / xi Mi . Hence, Mi is the molecular weight of
species i and xi is the mole fraction of species i.

Stroke

Connecting Rod

0.0763 m

0.1111 m

0.160 m

335 deg CA

60 deg CA

TABLE 2. Estimated thermodynamic properties of the fuels.

Estimation of Exergy
Total exergy of a mixture at a given state is the thermomechanical exergy plus the chemical exergy at that state [2]. Hence,

(h)Po ,To

(g)Po ,To

(g)Po ,To (g/h)Po ,To

LHV

(A/F)s

(MJ/kg)

(-)

(kJ/m3 -mix)

(-)

Iso-octane

44.6

15.03

2854

2785

3423

1.025

Methane

50.0

17.13

2754

2759

3097

0.998

Methanol

21.1

6.43

2894

2839

3372

1.019

Hydrogen

119.9

34.1

3231

3420

2905

0.945

Fuel

2.5

atot = ath + ach

Bore

(kJ/kg-mix) (kJ/kg-mix)

(31)
5

c 2011 by ASME
Copyright

1.5

3 RESULTS and DISCUSSIONS


Shown in Fig. 1 are the indicator diagrams for four different cases: (a) air-standard Otto cycle with rc = 12, (b) ideal Otto
cycle with actual/modelled k and rc = 7; actual cycle using heat
release profiles without heat loss in (c) and with heat loss in (d).
The deviations due to idealizations are clearly evident. In case
of (a), the higher values of k and rc and instantaneous heat release results in significantly higher peak pressures than the actual cases. The effect of k and rc on ideal cycle is evident in
(b) where significantly lower cylinder pressure is obtained. With
realistic heat release profiles, cylinder pressures are significantly
reduced in (c) and (d), and the effect of heat loss considerations
are also evident when (c) and (d) are compared. The resultant
variations of the gas temperature are shown in Fig. 2 where values of T , k and Cv of the actual cycle are plotted as a function of
crank angle, and lower values of k and Cv result in lower values
of cylinder temperature and pressure.

k = C /C

1.4
1.3

(kJ/kgK),

1.2
1.1
1.0
0.9

C
v

0.8
0.7
5000

rc = 12.0

Actual cycle w/o HT

Temperature (K)

4000

Air standard Otto cycle

Actual cycle w/ HT

3000

2000
Start of Combustion
End of Combustion
1000

0
180

210

240

270

300

330

360

390

420

450

480

510

540

Crank Angle (deg)

FIGURE 2. Variations of k, Cv and T .

300
Iso-octane

10000

(a) Air-standard Otto cycle, k = 1.4, r = 12


c

100

End of Combustion

8000
(kPa)

(b) Ideal Otto cycle, r = 7, actual k


c

Ideal Otto cycle

Pressure

10
Start of Combustion

= 1.0
= 7.0
= 2500

s = 330 deg CA
6000

d = 90 deg CA

Actual cycle

(c) Actual cycle, r = 12, actual k , w/ HT


c

P/P

rc
N

(d) Actual cycle, r = 12, actual k , w/o HT


c

(simulated)

Experimental

4000

Data [10]

2000

V = clearance volum e
c

0
180

8 9 10

210

240

270

300

12

330

360

390

420

450

480

510

540

Crank Angle (deg)

V/V

FIGURE 3. Pressure-Crank angle diagrams, predicted and reported


in [10].

FIGURE 1. Non-dimensional indicated diagrams on logarithmic P-V


co-ordinates.

The total exergy content within fuel-air premixture is consumed as the exergy transfer due to work and heat, and exergy
destruction due to combustion and in exhaust. Shown in Fig. 6
are the variation of non-dimensional exergy with crank-angle for
an air-standard otto cycle. Hence, the total exergy (atot ) and its
components are normalized by net heat release density of fuelair premixture (hoPo,To ). Results reported in [7] are also shown.
Predicted non-dimensional exergy without heat transfer consideration is in good agreement with the reported values, which neglected heat losses. Consideration of heat losses results in lower
values of total exergy and higher values of exergy transfer due to
heat transfer interaction, aq , which is defined as:

Figure 3 shows, the estimated and experimental pressure


traces as a function of crank-angle. As can be seen, the agreement between the predicted values and the experimental results
[10] are good. Hence, slight deviation is observed during combustion phase and similar deviations are also reported in [3] for
the simulation of the same engine configuration.
Shown in Figs. 4 and 5 are the variation of cylinder pressure
and temperature, respectively, for four different fuels considered
in the present study. Some of the key thermodynamic parameters estimated for these fuels are reported in Table 2. Iso-octane
and methane premixtures exhibit similar pressure and temperature profiles, and hydrogen as a fuel deviated the most by showing lower values of pressure and higher values of temperatures.
Methanol is found to remain in-between. These results are inline with the observations reported in [8].

aq =
6

Z 

To
1
T

qloss

(36)

c 2011 by ASME
Copyright

7000
6000

1.2
Iso-octane

7000

1.1

Methane
Methanol

Exergy destroyed due to combustion

1.0

Non-dimensional exergy

Pressure (kPa)

Hydrogen

5000
6000

4000
3000
2000

5000
360

370

380

390

1000

0.9

Exergy gain from compression

Exergy delivered as work

0.8
0.7
0.6
Exergy in intake charge

Total Exergy w/o HT

0.5
0.4
Unutilized Exergy

0.3
0.2

Total Exergy w/o HT [7]

TDC

0.1

0
180

Total Exergy w/ HT

0.0

210

240

270

300

330

360

390

420

450

480

510

540

180

210

240

270

300

Crank Angle (deg)

330

360

390

420

450

480

510

540

Crank Angle (deg)

FIGURE 4. Pressure-Crank angle diagrams for four fuels.

FIGURE 6. Variations of the non-dimensional exergy with crank angle for an Otto cycle.

4000
Iso-octane

3500

4000

1.2

Methane
Methanol

1.1

Hydrogen

1.0

2500

Non-dimensional exergy

Temperature (K)

3000
3500

2000
1500

3000
360

390

420

1000
500

0.9

0.8
0.7

tot

0.6
0.5

w/o HT

Rakopoulos [10]

tot

0.4

work interaction

0.3

w/o HT

0.2
0.1
0.0

0
180

210

240

270

300

330

360

390

420

450

480

510

-0.1

540

-0.2

Crank Angle (deg)

180

210

240

270

300

330

360

390

420

450

480

510

540

Crank Angle (deg)

FIGURE 5. Temperature-Crank angle diagrams for four fuels.

FIGURE 7. Variations of the non-dimensional exergy and its components with crank angle.

gle of 30o aTDC and expansion stroke continues until the piston
reaches the BDC. During this part of the cycle, decrease in total exergy continues due to exergy transfers (both work and heat)
from the system. The remaining exergy in the cylinder at the end
of expansion stroke emits with the exhaust gases, which is called
the exergy transfer with the exhaust gases.
Shown in Fig. 8 are the variation of non-dimensional exergy of stoichiometric iso-octane and methane air premixtures,
for various values of the compression ratio (rc = 7, 9, 11) at N =
2500 rpm. It is seen that, as the compression ratio increases, the
compression work increases, and the heat loss due to exhaust decreases. At higher compression ratios, the combustion pressure
and temperature are higher. The peak value of the total exergy
in the cylinder charge is also high due to greater compression
work. When the engine is operated at higher compression ratios,
the amount unutilized exhaust of exergy is found lower. Similar

Shown in Fig. 6 are also the variation of exergy transfer due


to work interactions, aw . For work interaction, the exergy is equal
to the net useful work which is equal to the net work minus the
work done against the surroundings [2].
Shown in Fig. 7 are the variation of non-dimensional exergy
and its components as a function of crank-angle. As the compression process takes place, the exergy of the mixture increases
because of the work done in compressing it. During this phase,
there are no irreversibilities while the heat transfer effects are
negligible and effectively this phase is overall adiabatic. With the
start of combustion at 30o bTDC, the chemical exergy of the fuel
decreases rapidly with its exhaustion, and pressure and temperature increase within cylinder which results in increase in thermomechanical exery and exergy transfer due to heat transfer to the
wall. Both heat transfer, and especially, combustion give a rapid
increase in exergy destruction. Combustion ends at the crank an7

c 2011 by ASME
Copyright

1.2

1.2
1.10

1.1

1.1

Non-dimensional exergy

Non-dimensional exergy

1.05

1.0

1.00

0.9

Iso-octane

r
r
r

0.8

0.7

0.6

= 7

0.95
330

= 9

= 11

rC

Methane

r
r
r

0.5

0.4

0.3

360

= 7

= 9

0.9
Methane @3000 rpm
0.8
Iso-octane

0.7

N = 1000
0.6

N = 2000

0.5

N = 3000

0.4
Methane
0.3
N = 1000
0.2

N = 2000

Iso-octane @1000 rpm

= 11

N = 3000

0.1

0.2
180

1.0

0.0
210

240

270

300

330

360

390

420

450

480

510

540

180

210

240

270

300

Crank Angle (deg)

330

360

390

420

450

480

510

540

Crank Angle (deg)

FIGURE 8. Variations of the non-dimensional exergy at various compression ratios.

FIGURE 9. Variations of the non-dimensional exergy at various engine speeds.


1.2

observations are reported in [10].


Shown in Fig. 9 are the variation of non-dimensional exergy
of stoichiometric iso-octane and methane air premixture with
crank-angle, for various engine speeds. At higher speeds, heat
transferred to the cylinder wall decreases as the engine running
at higher speeds effectively reduces the cycle duration. The reduction of convective heat loss results in the slight increment of
peak cylinder pressure and temperature with a slight reduction
in exergy destruction. At higher speeds, both the exhaust exergy
and energy increase because of the higher exhaust temperature.
Similar observations are also reported in [3].
Shown in Fig. 10 are the variation of non-dimensional exergy of iso-octane and methane air premixture with crank-angle,
for some lean and stoichiometric iso-octane and methane air premixtures. As the equivalence ratio approaches stoichiometric
value, the exergy destruction due to combustion is less. This is
largely due to the increase of combustion temperatures [8]. For
lean and stoichiometric mixtures, equivalence ratio approaching
the stoichiometric condition results in higher values of exhaust
exergy. Similar observations are also reported in [3].
Detailed distribution of percentage of energy and exergy usage in key processes of the SI cycle for two different speeds for
four fuels in the present study are presented in Figs. 11-14. All
of these results show that a significant fraction of energy and
exergy are wasted with exhaust gas. In energy based analysis,
the energy that is not converted to either work or convective heat
transfer is exhausted. However, the second law provides a deeper
insight by finding out the fact that chemical and thermal interactions between reactants and products during combustion cause
exergy destruction at a considerable level. As a result, a portion
of the exergy is lost due to combustion irreversibilities and the
rest is exhausted. Therefore, the exhaust exergy is lower than the
exhaust energy. Exergy destruction with combustion is neglected
in case of energy balance, however exergy analysis reports a sig-

Non-dimensional exergy

1.1

1.0

0.9
Iso-octane

0.8

= 0.6
0.7

= 0.8
= 1.0

0.6

0.5

1.0

Methane
= 0.6

0.4

= 0.8
= 1.0

0.3

0.2
180

210

240

270

300

330

360

390

420

450

480

510

540

Crank Angle (deg)

FIGURE 10. Variations of the non-dimensional exergy at various


equivalence ratios.

nificant fraction of work potential reduction with combustion and


their values are found to vary with fuels. Net work done and work
potential values are very close in both cases.
From the analysis of four different fuels considered in the
present study, some of the key results obtained are:
1. At 1000 rpm, 33 to 37% of energy contained with fuel is
converted to useful work, and the figure changes to 35 to
40% in case of 3000 rpm.
2. At 1000 rpm, energy loss due to heat transfer is 20 to
27% and 12 to 16% at 3000 rpm. However, associated exergy losses are 17 to 24% at 1000 rpm, and 10 to 14% at
3000 rpm.
3. At 1000 rpm, energy loss with exhaust is 36 to 47%, and 44
to 53% at 3000 rpm. Exergy loss with exhaust are 26 to 34%
8

c 2011 by ASME
Copyright

Combustion

Heat Loss

Exhaust

Workdone

Combustion

Heat Loss

Exhaust

23%

80

80

40

23%

24%

26%

47%

31%

53%

17%

20%

10%

12%

33%

33%

35%

35%

20

Energy or Exergy (%)

Energy or Exergy (%)

24%

60

Workdone

100

100

27%

46%

32%

53%

40

17%

21%

10%

12%

20

33%

33%

35%

35%

60

0
Exergy

Energy

Exergy

1000 rpm

Exergy

Energy

Energy

Energy
3000 rpm

FIGURE 13. Percentage of fuels energy and exergy for stoichiometric methanol fuelling.

FIGURE 11. Percentage of fuels energy and exergy for stoichiometric octane fuelling.

Combustion

Combustion

Exergy

1000 rpm

3000 rpm

Heat Loss

Exhaust

Workdone

Heat Loss

Exhaust

Workdone

100

100
18%

5%

Energy or Exergy (%)

18%

Energy or Exergy (%)

80

28%

43%

34%

50%

60

40

19%

22%

35%

35%

11%

13%

37%

34%

36%

24%

37%

42%

44%

27%

14%

16%

37%

40%

40%

60

40

20

20

4%

80

37%

0
Exergy

0
Exergy

Energy
1000 rpm

Exergy

Energy
1000 rpm

Energy

Exergy

Energy
3000 rpm

3000 rpm

FIGURE 14. Percentage of fuels energy and exergy for stoichiometric hydrogen fuelling.

FIGURE 12. Percentage of fuels energy and exergy for stoichiometric methane fuelling.

probe and identify the sources of work potential losses in different phases of the SI engine cycle.

at 1000 rpm, and 31 to 42% at 3000 rpm.


4. At 1000 rpm, 5 to 24% of energy contained with fuel is
destroyed in combustion, and 4 to 24% is destroyed at
3000 rpm. Hydrogen has the lowest destruction for the fuels
considered, and this is attributed to the fact that it is the simplest molecule and combustion products are also simple and
few in number.

ACKNOWLEDGEMENT
The authors like to acknowledge the enlightening discussions with Major Md. Mizanuzzaman of Military Institute of Science and Technology (MIST), Dhaka during the present work.

4 CONCLUSIONS
A comprehensive first law and second law thermodynamic
analysis is carried out for a single cylinder SI engine fuelled by
four fuels, namely iso-octane, methane, methanol and hydrogen.
The results highlight the importance of exergy based analyses to

REFERENCES
[1] Moran, M. J., 1982. Availability Analysis: A Guide to Efficient Energy Use. Prentice-Hall.
[2] Wark, Jr., K., 1995. Advanced Thermodynamics For Engineers. McGraw-Hill.
9

c 2011 by ASME
Copyright

[3] Sezer, s., and Bilgin, A., 2008. Mathematical analysis of


spark ignition engine operation via the combination of the
first and second laws of thermodynamics. Proceedings of
the Royal Society A: Mathematical, Physical and Engineering Science, 464, pp. 31073128.
[4] Li, K. W., 1996. Applied Thermodynamics: Availability
Methods and Energy Conversion. Taylor & Francis.
[5] Ferguson, C. R., and Kirkpatrick, A. T., 2001. Internal
Combustion Engines: Applied Thermosciences. John Wiley
& Sons.
[6] Shehata, M., 2010. Cylinder pressure, performance parameters, heat release, specific heats ratio and duration of
combustion for spark ignition engine. Energy, 35(12),
pp. 4710 4725.

[7] Heywood, J. B., 1998. Internal Combustion Engine Fundamentals. McGraw-Hill.


[8] Caton, J. A., 2010. Implications of fuel selection for an SI
engine: Results from the first and second laws of thermodynamics. Fuel, 89, pp. 3157 3166.
[9] Mohiuddin, M. R., 2010. Exergy analysis of a four stroke
spark-ignition engine using different fuels. Masters thesis, Department of Mechanical Engineering, Bangladesh
University of Engineering and Technology.
[10] Rakopoulos, C. D., 1993. Evaluation of a spark ignition engine cycle using first and second law analysis techniques. Energy Conversion and Management, 34(12),
pp. 1299 1314.

10

c 2011 by ASME
Copyright

You might also like