You are on page 1of 10

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

Available online at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/he

A fractal-based quasi-dimensional combustion model for SI


engines fuelled by hydrogen enriched compressed natural gas
Fanhua Ma*, Shun Li, Jianbiao Zhao, Zhengliang Qi, Jiao Deng, Nashay Naeve, Yituan He,
Shuli Zhao
State Key Laboratory of Automotive Safety and Energy, Tsinghua University, Beijing 100084, PR China

article info

abstract

Article history:

A quasi-dimensional model based on the concepts of fractal geometry has been developed

Received 24 February 2012

for an SI engine fuelled with natural gas/hydrogen blends. The fundamentals of the

Received in revised form

thermodynamic model, the fractal combustion model and related equations are intro-

8 March 2012

duced. This paper investigates the influence of manifold absolute pressure, equivalence

Accepted 10 March 2012

ratio and hydrogen fraction on fractal dimension and improves the fractal dimension

Available online 11 April 2012

expression. Comparisons are conducted between the improved and original models by the
prediction outcomes. After the determination of model constants by calibration, the model

Keywords:

predictions of cylinder pressure histories and mass fraction burned of an HCNG engine are

Quasi-dimensional

then compared with experimental data over a wide range of loads, equivalence ratios,

Fractal dimension

engine speeds and hydrogen blending ratios. The pressure profiles show that predictions of

HCNG

the improved model match quite well with the experimental results except for the early
combustion stage. The improved model is proved to be more suitable for predicting HCNG
engine performance.
Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.

1.

Introduction

HCNG (hydrogen enriched natural gas) is a promising technology for reducing harmful exhaust emissions, increasing
engine thermal efficiency and promoting the development of
a hydrogen economy [1e9]. Quasi-dimensional model has
been proven as a useful way to predict the performance of the
HCNG engine. It is of great value for simulations to investigate
the characteristics of natural gasehydrogeneair mixtures
such as laminar burning velocities [10]. Verhelst and Sierens
[11] investigated the power cycle of a hydrogen-fuelled ICE
with a quasi-dimensional two-zone combustion model containing six turbulent burning velocity models and showed the
correspondence between simulation and measurement for
varying equivalence ratio, ignition timing and compression

ratio. Perini et al. [12] developed a predictive two-zone, quasidimensional model for the simulation of the combustion
process in spark ignition engines fuelled with hydrogen,
methane or hydrogenemethane blends. A two-zone quasidimensional model for simulating the performance of an
HCNG engine has been built in the authors previous work
[13,14], which was validated by different operating conditions.
The calibration coefficients have also been studied [15], thus
the prediction capability of the model can be greatly
improved.
In the quasi-dimensional models mentioned above, the
combustion process model is the eddy burning model and
the mass burning rate is assumed to be dominated by the
rate of entrainment of unburned mixture. Matthews and
Chin [16] developed a fractal-based quasi-dimensional SI

* Corresponding author. Tel./fax: 86 10 62785946.


E-mail addresses: mafh@mail.tsinghua.edu.cn, mafh@tsinghua.edu.cn (F. Ma).
0360-3199/$ e see front matter Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2012.03.045

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

engine model which assumes that the burning rate is


dominated by flame wrinkling (rather than entrainment) and
uses the concepts of fractal geometry to account for the
flame wrinkling process. The fractal burning model can yield
reasonable predictions of the pressure profiles over a range
of engine operating conditions. The model was further
developed for simulating effects of engine speed [17] and the
flame peninsulas [18].
In this work, a fractal-based combustion model for HCNG is
developed on the basis of the authors previous work. The
fractal dimension model is established in consideration of
hydrogen ratio influence and the effects of operating
conditions.

2.

Mathematical model

Quasi-dimensional model considers the combustion chamber


as several separate parts. In this work, we employed a very
popular and simple two-zone model which considers the
chamber as two parts divided by the flame front. This model
contains two main zones: one zone is the thermodynamic
model which can reproduce the cylinder pressure data by use
of an already known mass fraction burned profile; the other
zone is the turbulent combustion model used to predict the
above-mentioned mass fraction burned profile.

2.1.

Two-zone thermodynamic model

The two-zone thermodynamic model has been introduced in


previous papers [13,14], and will be briefly summarized here.
The two-zone thermodynamic model is built around certain
assumptions, as outlined as follows:
1. The cylinder charge during combustion is assumed to be
divided into two-zones: the burned products and unburned
reactants, as can be seen in Fig. 1.
2. The two zones are recognized as homogeneous ideal gas
and have uniform properties.

9893

Unburned gas is considered of air, H2 and CH4, ignoring the


residual gas.
3. The pressure at any time is uniform throughout the
cylinder and is just the function of crank angle.
4. The in-cylinder temperature is divided into two parts, as
burnt temperature and unburned temperature, which are
uniform in their own zone. There is no heat transfer
between the two-zones.
5. The analysis is restricted to the closed valve period. All
crevice effects are ignored.
These assumptions constitute the basis of the model. The
following three first order differential equations can be
derived from some manipulation of the Equation of State, the
First Law of Thermodynamics, and the Conservation of Mass.
The subscript u represents unburned zone and b represents
burnt zone.


dTu
1
dP dQu
Vu

dq
dq
mu cpu
dq

dTb
1
dV
dmb
P

 Rb Tb  Ru Tu
dq mu cpu
dq
dq



Ru
dP dQu
dP
Vu

V
cpu
dq
dq
dq



dP
1
cvb dV dQ
1
P


ub  uu
cvb Ru
cvb
Rb
dq cvu
dq dq
Vu 
Vu V
cpu
Rb cpu
Rb





Ru
dmb
cvu cvb Ru dQu


 cvb Tb  Tu
Rb
dq
cpu Rb cpu dq

(1)

(2)

(3)

In the above equations, V, cp and cv are calculated by using


empirical formulas. dQ/dq represents the heat transfer, which
can also be estimated by the Woschni equations [20]. Therefore, after dmb/dq is determined, the equations can be solved
by 4th-order Runge-Kutta method. The turbulent combustion
model which is detailed in the next section devotes to the
calculation of dmb/dq.
During the compression and expansion stroke, the above
equations are reduced to the following equations (single zone):


dP 1
dT dQ

mcp 
dq V
dq dq

(4)



dT
1 dQ
dP

V
dq mcp dq
dq

(5)

2.2.

Fractal combustion model

The mass burning rate is modeled on the basis of the flame


propagation [21] as:


mb ru AL Sb

Fig. 1 e Two-zone combustion chamber [19].

(6)

The fractal combustion model assumes that the burning


rate is dominated by flame wrinkling (rather than entrainment) and the wrinkled surface area of the flame can be
characterized by fractal geometry. The effect of the turbulence
is to wrinkle the flame, thereby increasing the surface area
and consequently increasing the flame speed. The general
expression of the correlation linking turbulent surface areas

9894

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

Eqs. (9) and (10) are based on the physical arguments that
turbulent convective motion produces flame wrinkles, which
is characterized by u0 ; and that this wrinkling is opposed by
the local flame propagation process, which is characterized
by SL. This model represents a weighted average of two
limiting fractal dimension. For a quiescent flow field, the
flame surface is smooth and the fractal dimension is equal to
the topological dimension of a smooth surface, D3 2.0.
Various theories have been proposed for the upper limit of
the fractal dimension for a premixed flame propagating in
a high Reynolds number flow field, these values ranges from
2.33 to 2.41 [16].
In this study, Eq. (9) is tried to be used to predict fractal
dimension, but the value of D3max should be chosen properly
according to the operating conditions. In the Ref. [28], several
operating conditions and the measured fractal dimensions in
a stainless steel cylindrical combustion chamber are listed.
Three conditions are selected and shown in Table 1, where
the equivalence ratio(F) and the pressure were kept
constant, the hydrogen mole fraction (x) was varied from 0 to
0.2 while the turbulence intensity was fixed. For case A, the
value of D3max was calculated, and by using Eq. (9) according
to given parameters in Table 1, this value approached to 2.16.
With the calculated D3max of 2.16, the values of D3 are predicted by using Eq. (9). These values are shown in Table 1,
represented by D3_pre(ori). The differences between the
predicted and experimental results (D3_exp) of the fractal
dimension are quite large for case B and case C, where the
methane is enriched with hydrogen and x is 0.1 and 0.2
respectively. From Table 1, it was found that the predicted
fractal dimension decreased with an increasing hydrogen
fraction, but the experimental results instead. It is recognized that if the turbulence intensity is kept unchanged, the
instantaneous turbulent flame front thickness decreases
with the increase of hydrogen fraction in HCNG. Therefore,
a larger fraction of the small-scale end of the turbulence
spectrum contributes to wrinkling of the instantaneous
flame fronts, thus increasing their fractalization. The original
D3 expression shown by Eq. (9) is not capable of predicting
the turbulent flame fractal dimension with various hydrogen
fractions. According to the fractal analysis results in [28], the
effect of hydrogen fraction on D3 is added to Eq. (9) and the
following equation is formed:

and speeds to the laminar surface areas and speeds is shown


in the following equation:
AT Sb

AL SL

D 2

Lmax 3
Lmin

(7)

where AT and AL represent the turbulent and laminar surface


areas, Sb and SL represent the turbulent burning speed and the
laminar flame velocity, Lmax and Lmin represent the maximum
and minimum flame wrinkling scales and D3 signifies the
fractal dimension, which characterizes the roughness of the
flame wrinkles.
Many investigations [22e24] have shown that Eq. (7) alone
can not accurately predict the turbulent burning velocity for
every value D3. It has also been recognized that small-scale
turbulence not only increases the flame surface area but also
significantly enhances the transfer of species and heat [25].
Baratta et al. [25] improved the modeling of turbulenceeflame interaction considering that the effect of species and
heat transfer at high turbulence levels, which are not taken
into account by Eq. (7), should be modeled as a function of
charge density, which increases racial species concentration
and heat transfer across the flame front. In the present model,
a multiplying factor connecting the burned mass density and
the unburned mass density has been added:
AT Sb

AL SL

D 2
 0:25 
rb
Lmax 3
ru
Lmin

(8)

where rb and ru are the burned and unburned mass density


and the exponent is set to 0.25.
Matthews and Chin [16] compared three assumptions
about the ratio of flame wrinkling scales and suggested the
best assumption for the relationship between the inner and
outer cut-off of the wrinkling scales was the ratio of the
integral length scale to the Kolmogorov scale: Lmax =Lmin li =h.
They also showed that the following model for D3 predicted
the measured fractal dimensions of flames in engines [26]
with an error of less than 3.6%:
D3 D3max

u0
SL
2:0 0
; D3max 2:35
u0 SL
u SL

(9)

where u0 is the turbulence intensity and SL is the laminar


burning speed. Eq. (9) is named the original expression of D3 in
this paper.
Perini [12] used the similar expression for the prediction of
D3 from Santavicca et al. [24] with different upper limits of D3,
u0
SL
2:0 0
; D3max CD3  2:35
D3 D3max 0
u SL
u SL

D3 D3max

(10)

1
1
2:0
1
SL
1
u0
1
1
0
1 0:1x u
1 0:1x SL

(11)

where x is the hydrogen mole fraction in HCNG. The predicted


results of the improved D3 expression are also listed in Table 1.
The differences between predictions D3_pre(imp) and experimental results D3_exp are quite small, hence the improved

where the calibration constant CD3 is set to 1.013. Eq. (10) is


proven acceptable for combustion simulation of HCNG and
hydrogen engines [11,12].

Table 1 e Differences between experimental and predicted D3.


Case
A
B
C

P
(MPa)

u0
(m/s)

u0 /Sl

D3_exp

D3_pre(ori)

D3_pre(imp)

0.1
0.1
0.1

0.6
0.6
0.6

0
0.1
0.2

0.17
0.17
0.18

1.53
1.44
1.33

2.097
2.101
2.112

2.097
2.0944
2.0913

2.0968
2.1044
2.1115

9895

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

Table 2 e Test engine specifications.

Table 4 e Operating conditions.

Item

Value

Displacement volume (L)


Compression ratio
Bore (mm)
Stroke (mm)
Rating power (kW)
Rating torque (N m)

6.234
10.5
105
120
154
620

expression can reflect the influence of hydrogen mole fraction


on turbulent flame fraction dimension.
In previous research, the upper limit of fractal dimension is
set constant [12,16,27]. Cecile et al. [28] show that the turbulent flames have a smaller and more complex structure under
conditions of higher pressure. Moreover, the scale of wrinkle
in the premixed turbulent flames becomes smaller as the
equivalence ratio increases. For fixed composition and
turbulence intensity, the higher pressure not only induces
a raise in u0 /Sl, it also decreases the small-scale turbulence
characteristic dimensions, generating smaller flame elements
and therefore a more fractalized flame contour with increased
D3. So, in Eq. (11) the influence of pressure on turbulence
scales should be reflected on the change of D3max. For conditions of fixed turbulence intensity and increasing equivalence
ratio, Sl increases, and u0 /Sl decreases, so it can be predicted
that D3 decreases by Eq. (11); but in fact D3 increases from the
experimental results in reference [28]. The reason for the
rising D3 is also the decreasing of the instantaneous flame
front thickness. Thus, the influence of equivalence ratio must
be added to D3max. In summary, the upper limit of fractal
dimension is set 2.35 in the present work for the reference
condition ( p0 70 kPa, l0 1.246). As the condition changes, it
is proposed that D3 is calculated by the following equation:
 0:008  0:035
p
l
(12)
D3max 2:35
p0
l0
where p represents MAP and l represents the excess air ratio.
The exponents 0.008 and 0.035 are determined by calibration
of the operating condition A, condition B respectively and
condition C as listed in Table 4. The Eq. (11) using the D3max in
Eq. (12) is named the improved model in this paper.
In the Comparison results section, the improved D3 equation is compared with the original Eq. (9) to find out whether
the improved method is more capable for HCNG engine
simulation.

Table 3 e Components of the natural gas used in the


experiments.
Components
CH4
C3H8
n-C4H10
O2
N2
C2H6
i-C4H10
i-C5H12
CO2

Volume fraction (%)


96.51
0.18
0.02
0.01
0.22
1.2
0.02
0.01
1.81

Operating
conditions

n (r/min)

MAP (kPa)

qig ( BTDC)

1600
1600
1600
2000
1600
1600
1600
1600
1600
1600
800
1200
1600
1600
1600
1600
1600
1600
1600
1200
1200

70
125
125
90
70
70
125
125
70
125
90
110
125
70
70
125
125
125
125
80
105

0
0
0
0
0.15
0.15
0.15
0.15
0.3
0.3
0.3
0.3
0.3
0.45
0.45
0.45
0.45
0.55
0.55
0.55
0.55

1.3
1.5
1.3
1.4
1.3
1.5
1.3
1.5
1.3
1.3
1.5
1.5
1.1
1.3
1.5
1.3
1.5
1.5
1.3
1.3
1.4

28
28
28
32
24
28
20
24
28
24
20
22
20
26
24
24
22
26
24
16
16

A
B
C
D
E
F
G
H
I
J
K
L
M
N
O
P
Q
R
S
T
U

The method of calculating turbulence parameters is briefly


described here:
Integral length scale L [29]:
L0 C L  H

(13)

L L0 ru0 =ru 1=3

(14)

where ru is the density of the gas in the unburned zone, H is the


chamber height and the subscript 0 represents ignition timing.
Taylor micro scale LT [30]:
LT CT Liv rin =ru0 3=4

(15)

where rin is the density of gas in the intake stroke, Liv is the lift
of the intake valve.
Kolmogorov scale h [17]:
1=4
h n3 =

(16)

where n is the laminar kinematic viscosity of the unburned


mixture and is the rate of dissipation of turbulence kinetic
energy.
Turbulence intensity u0 [29]:
u00 Cu  Cm

(17)

u0 u00 ru0 =ru 1=3

(18)

where Cm is the mean velocity of piston movement, Cu is


a constant and 0 represents ignition timing.
Ignition lag DTig [11,12]:


1=3 2=3 2=3
DTig Cig H1=3 Cm
LT SL

2.3.

(19)

Adiabatic flame temperature of HCNG

Adiabatic flame temperature is an important parameter of the


two-zone model, because the temperature of the flame kernel is

9896

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

evaluated as the adiabatic flame temperature. An accurate


value of the adiabatic flame temperature can be obtained by
iterating until the enthalpy of the products equals that of the
reactants, with the consideration of chemical equilibrium. Here,
for simplicity in the work the adiabatic flame temperature is
calculated by the following simplified empirical equations:

Hreac Ti ; p Hprod Tad ; p

(20)

Rearranging with the lower heat value;


1
Hu cp Tad  Ti =M
1 ll0

(21)

unburned zone and burnt zone respectively, constants B, E, F,


G, m and n are determined by fuel type [32].

2.5.

Flame geometry

In the quasi-dimensional model, the correlation between


flame radius and flame surface area as well as the burned
volume is needed. This correlation is closely related to the
geometry of the combustion chamber. Since the combustion
chamber in our engines is quite simple in shape, an integral
approach to deal with this problem is chosen. The calculation
methodology can be found in Ref. [19] and will not be detailed
here.

Rearrange the above equation:


Tad Ti

M
Hu
1 ll0 cp

(22)

where l is the excess air ratio, l0 is the stoichiometric air/fuel


ratio and M is the mean molar mass.
The stoichiometric ratio and lower heating value of the
HCNG fuel can be written as the following:
l0 34:48

Hu

4  3x
8  7x

(23)

120x 500:48  8x
MJ=kg
8  7x

(24)

where x represents the hydrogen blending ratio.


Until now, the adiabatic flame temperature for HCNG with
various hydrogen blending ratios has been determined by
Eq. (21).

2.4.

Laminar burning velocity of HCNG

Laminar burning velocity is a fundamental characteristic for


a fuel, and also an essential parameter, directly influencing
the speed of flame spread. In order to obtain a correlation of
the laminar burning velocity of HCNG that is valid at both
intermediate and high hydrogen contents and at different
values of equivalence ratio, Sarli and Benedetto [31] have
tested the feasibility of a Le Chateliers Rule-like formula
expressed as:
Sl

HCNG f; x

x=Sl

H2 f

1
1  x=Sl

CH4 f

(25)

where Sl H2 f is the laminar burning velocity of hydrogen and


Sl CH4 f is that of methane. They can both be calculated by the
empirical formula described in Ref. [32]:


 m Tu Tb  T0 n
SL A T0 YF;u
T0 Tb  Tu

(26)

 p
T0 E=ln 
B

(27)




G
A T0 Fexp  0
T

(28)

m
where YF;u
is the fuel fraction in unburned gas, p is the pressure of reaction, Tu and Tb represent the temperature of

3.
Comparison of the experimental and
prediction data (model validation)
3.1.

Experimental system

In order to validate the above predictive model against


experimental data, a lot of experiments were conducted with
engine running on HCNG with various hydrogen blending
ratios. The engine operating conditions have also been varied
extensively.
The experiments were carried out on an in-line 6 cylinders
turbocharged engine (specifications are shown in Table 2). The
components of the natural gas used in the experiments are
listed in Table 3. According to the data listed in Table 3, the
relative molecular mass of the natural gas which is 15.911 is
approximate to the relative molecular mass of methane which
is 16.04. Moreover, the lower calorific value of the natural gas
used in the experiments which is 35.63 MJ/Nm3 is nearly equal
to the lower calorific value of methane which is 35.88 MJ/Nm3.
Thus, the natural gas is represented by methane in the
simulation.
The test engine was coupled to an eddy-current dynamometer for engine speed and load measurement and
control. An electronic control system provided access to all
calibration parameters, allowing the user to set a desired
equivalence ratio (by adjusting fuel injection duration) and
ignition time. The air/fuel ratio was monitored by an HORIBA
wide range lambda analyzer. In-cylinder pressure data was
taken by a Kistler 6117B piezoelectric high-pressure transducer which was connected via a Kistler 5011B charge
amplifier to the YOKOGAWA ScopeCorder for signal A/D
conversion and data acquisition. Crankshaft position was
measured by a Kistler 2613B crank angle encoder with a resolution of 1  CA.

3.2.

Comparison results

As seen from the above equations, there are several constants


that need to be determined including the following: Cu for the
calculation of turbulence intensity u0 , CT for the Taylor length
scale LT, CL for the Integral length scale and Cig for ignition lag.
These constants need to be calculated with reference to
a certain engine operating condition by matching the simulated pressure trace under that condition to the measured one.
In this work, we chose the operating condition of engine

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

Fig. 2 e Comparisons between experimental pressure, MFB and predicted ones.

9897

9898

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

Fig. 2 e (continued).

speed 1600 r/min, MAP 70 kPa, hydrogen fraction 0%,


excess air ratio 1.25 and spark advance 28 as the reference
condition for calibration. The calibration result is that Cu 2.35,
CT 3, CL 0.3 and Cig 66.5, among them Cig is the only
parameter that is varied when the operating condition changes
for an equivalent ignition delay time, with the experimental
result and the other parameters remaining constant. Therefore, the models prediction ability can be tested.
In order to test the applicability of the quasi-dimensional
model, Table 4 shows 21 operating conditions which are
selected for both experiments and simulation. The range of the
21 operation conditions are quite wide, in which speed (n),
manifold absolute pressure (MAP), hydrogen mole ratio (x),
excess air ratio (l) and ignition timing (qig) are all changed
widely, in order to validate the applicable range of the proposed
model.
Fig. 2 shows the predicted and experimental pressure and
mass fraction burned curves. For the sake of saving space, only
a few typical conditions are listed. In Table 5, detailed
comparison data are listed, in which all of the operating
conditions are included. From the figures, it can be seen that
the predicted data of the improved model are approximately
consistent with the experiment data. While simulation using

the original model doesnt correlate well with the experimental results. However, at the early combustion stage, the
predicted pressure rises faster than experimental pressure for
both models. From the mass fraction burned curves, it can be
seen that the predicted mass fraction is larger than experimental one at the early combustion stage, which is why the
pressure curves are not coincident at this stage. Also after the
mass fraction burned being larger 90%, the predicted curves
rise much faster. This is because the fractal characteristics of
the flame front in the flame developing stage is different from
the fully combustion stage. And at the late burning stage, flame
front surface has reached the cylinder, so the fractal dimension may changes irregularly. Therefore, the fractal dimension
equation may not be used for the entire combustion process.
Fractal dimension should be further investigated for the early
and late combustion stages. From the comparison results in
Table 5, it can be seen that the improved model can provide
much closer predictions to the experimental results than the
original model. For the main comparison parameters, such as
the maximum pressure, the crank angle at the maximum
pressure, fast burning angle and the degree for 50% mass
fraction burned, the improved simulation results are much
better, while the indicated mean effective pressures are close

9899

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

Table 5 e Differences between predicted and experimental results.


Operating
conditions
A(exp)
A(imp)
A(ori)
B(exp)
B(imp)
B(ori)
C(exp)
C(imp)
C(ori)
D(exp)
D(imp)
D(ori)
E(exp)
E(imp)
E(ori)
F(exp)
F(imp)
F(ori)
G(exp)
G(imp)
G(ori)
H(exp)
H(imp)
H(ori)
I(exp)
I(imp)
I(ori)
J(exp)
J(imp)
J(ori)
K(exp)
K(imp)
K(ori)
L(exp)
L(imp)
L(ori)
M(exp)
M(imp)
M(ori)
N(exp)
N(imp)
N(ori)
O(exp)
O(imp)
O(ori)
P(exp)
P(imp)
P(ori)
Q(exp)
Q(imp)
Q(ori)
R(exp)
R(imp)
R(ori)
S(exp)
S(imp)
S(ori)
T(exp)
T(imp)
T(ori)
U(exp)
U(imp)
U(ori)

Pmax (MPa)

q Pmax ( CA)

2.96
2.95
2.95
2.12
2.12
2.00
6.01
6.01
5.58
3.23
3.24
2.99
2.88
2.84
2.84
2.36
2.28
2.13
5.43
5.36
4.98
4.45
4.00
3.47
3.41
3.39
3.37
6.35
6.20
5.72
3.61
3.55
4.90
4.13
4.00
3.48
6.54
6.53
6.53
3.51
3.47
3.55
2.85
2.80
2.62
6.96
6.84
6.31
5.45
5.40
4.63
6.09
5.64
4.78
6.82
6.85
6.36
4.11
4.11
3.98
4.53
4.50
4.18

377
375
375
377
375
372
375
374
374
376
375
375
378
378
378
377
376
373
378
380
381
379
380
370
373
374
374
375
375
377
378
379
373
379
380
379
376
376
376
372
374
374
376
374
374
372
374
376
376
374
375
372
370
373
369
373
375
372
374
375
376
380
382

Pmax (%)

0.16
0.16
0.10
5.57
0.02
7.17
0.42
7.45
1.37
1.64
1.37
1.64
1.18
8.16
10.10
22.12
0.42
1.17
2.34
9.99
1.69
35.51
3.06
15.56
0.24
0.24
1.19
0.92
1.98
8.32
1.70
9.21
0.79
14.92
7.42
21.47
0.40
6.79
0.14
3.21
0.59
7.75

qcd ( CA)

q 50% ( CA)

Pi (MPa)

28.25
28.61
28.61
35.61
38.55
41.35
24.66
23.78
26.52
33.69
35.68
39.45
27.03
28.13
28.15
32.18
37.57
40.11
25.17
26.12
27.51
30.97
31.22
35.21
26.16
26.13
26.15
22.85
24.57
26.98
23.08
23.06
18.45
26.27
27.08
30.33
21.31
23.08
23.08
23.60
23.74
23.41
26.92
31.63
33.13
21.30
22.12
24.27
24.24
29.97
33.99
20.00
31.01
35.68
19.73
20.72
22.77
15.18
18.32
19.57
18.59
18.11
19.32

371
369
369
378
379
381
367
365
367
374
374
378
372
370
370
375
376
379
372
371
373
375
378
384
366
364
364
367
366
368
371
371
362
373
374
378
368
366
366
365
363
363
370
369
371
364
363
366
370
368
373
364
366
371
360
363
365
364
364
365
368
372
373

0.57
0.60
0.60
0.48
0.51
0.50
1.08
1.08
1.07
0.68
0.74
0.72
0.57
0.61
0.60
0.48
0.51
0.50
1.12
1.14
1.13
0.96
0.99
0.94
0.57
0.60
0.59
1.13
1.12
1.11
0.65
0.66
0.75
0.83
0.86
0.84
1.26
1.21
1.21
0.56
0.57
0.57
0.50
0.52
0.52
1.14
1.12
1.12
0.99
1.00
0.98
0.93
0.98
0.96
1.01
1.08
1.08
0.63
0.64
0.64
0.77
0.84
0.83

Pa (%)

5.31
5.31
6.62
3.70
0.15
0.41
8.38
5.68
6.13
4.86
5.31
4.86
2.12
0.65
2.95
1.98
4.49
4.05
0.36
1.03
2.34
15.44
4.34
1.40
3.23
3.83
1.12
1.37
4.06
2.80
2.15
2.21
1.01
1.22
5.50
3.53
7.62
7.51
2.38
2.18
9.60
7.77

9900

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

to the experimental results for both models. Overall, the newly


developed model can be used to predict performance of HCNG
engines, and the improved model is proven to have more
advantages over the original model. It should be noted that
some assumptions, such as a complete chemical reaction, and
no heat transfer between burned and unburned zone also
influence the computing precision.

[10]

4.

[11]

Conclusion

A quasi-dimensional model for a turbocharged SI HCNG


engine based on fractal geometry has been developed. The
equations of a two-zone model are recognized as the basic
functions, which are then coupled to the fractal combustion
model and used to simulate the combustion process. An
improved fractal dimension model is developed considering
the influence of operating condition and hydrogen volume
fraction. Comparisons are conducted between the improved
and original models through prediction outcomes.
The models predictions are compared with experimental
data over a range of loads, equivalence ratios, engines speeds
and hydrogen blend ratios. From the results it can be
concluded that the improved model can more accurately
predict the HCNG engine performance, but further investigation of fractal dimension in the early flame stage and the late
combustion stage should be carried on in the future. After all,
the model established in this paper is able to conduct the
performance simulation of an SI HCNG engine under various
hydrogen blending ratios.

[9]

[12]

[13]

[14]

[15]

[16]

[17]

[18]

references
[19]
[1] Hoekstra RL, Collier K, Mulligan N, Chew L. Experimental
study of a clean burning vehicle fuel. Int J Hydrogen Energy
1995;20(9):737e45.
[2] Sierens R, Rossel E. Variable composition hydrogen/natural
gas mixtures for increased engine efficiency and decreased
emissions. J Eng Gas Turbines Power 2000;122:135e40.
[3] Huang ZH, Wang JH, Liu B, Zeng K, Yu JR, Jiang DM.
Combustion characteristics of a direct-injection engine
fueled with natural gas-hydrogen blends under different
ignition timings. Fuel 2007;86:381e7.
[4] Zuohua Huang, Bing Liu, Ke Zeng, Yinyu Huang,
Deming Jiang, Xibin Wang, et al. Experimental study on
engine performance and emissions for an engine fueled with
natural gas-hydrogen mixtures. Energy & Fuels 2006;20(5):
2131e6.
[5] Akansu SO, Dulger Z, Kahraman N, Veziroglu TN. Internal
combustion engines fueled by natural gas-hydrogen
mixtures. Int J Hydrogen Energy 2004;29:1527e39.
[6] Ma FH, Wang Y, Liu HQ, Li Y, Wang JJ, Zhao SL. Experimental
study on thermal efficiency and emission characteristics of
a lean burn hydrogen enriched natural gas engine. Int J
Hydrogen Energy 2007;32:5067e75.
[7] Ortenzi F, Chiesa M, Scarcelli R, Pede G. Experimental tests of
blends of hydrogen and natural gas in light-duty vehicles. Int
J Hydrogen Energy 2008;33:3225e9.
[8] Hu Erjiang, Huang Zuohua, Liu Bing, Zheng Jianjun,
Gu Xiaolei. Experimental study on combustion
characteristics of spark-ignition engine fuelled with natural

[20]

[21]

[22]

[23]

[24]

[25]

[26]

[27]
[28]

gas-hydrogen blends combining with EGR. Int J Hydrogen


Energy 2009;34(2):1035e44.
Hu Erjiang, Huang Zuohua, Liu Bing, Zheng Jianjun,
Gu Xiaolei, Huang Bin. Experimental investigation on
performance and emissions of a spark-ignition engine
fuelled with natural gas-hydrogen blends combined with
EGR. Int J Hydrogen Energy 2009;34(1):528e39.
Huang Zuohua, Zhang Yong, Zeng Ke, Liu Bing, Wang Qian,
Jiang Deming. Measurements of laminar burning velocities
for natural gas-hydrogen-air mixtures. Combust Flame 2006;
146(1e2):302e11.
Verhelst S, Sierens R. A quasi-dimensional model for the
power cycle of a hydrogen-fuelled ICE. Int J Hydrogen Energy
2007;32:3545e54.
Federico Perini, Paltrinieri Fabrizio, Mattarelli Enrico.
A quasi-dimensional combustion model for performance
and emissions of SI engines running on hydrogen-methane
blends. Int J Hydrogen Energy 2010;35:4687e701.
Ma FH, Liu HQ, Wang Y, Wang JJ, Ding SF, Zhao SL. A quasidimensional combustion model for SI engine fuelled by
hydrogen enriched compressed natural gas; 2008. SAE paper
no 2008-01-1633.
Ma FH, Wang Y, Wang MY, Liu HQ, Wang JJ, Ding SF, et al.
Development and validation of a quasi-dimensional
combustion model for SI engines fuelled by HCNG with
variable hydrogen fractions. 2008;33:4863e4875.
Ma FH, Deng J, Qi ZL, Li S, Chen RZ, Herman Y, et al. Study on
the calibration coefficients of a quasi-dimensional model for
HCNG engine. 2011;36:9278e9285.
Matthews Ronald Douglas, Chin Young-Wook. A fractalbased SI engine model: comparisons of predictions with
experimental data; 1991. SAE paper no 910079.
Wu C-M, Roberts CE, Matthews RD, Hall MJ. Effects of engine
speed on combustion in SI engines: comparisons of
predictions of a fractal burning model with experimental
data; 1993. SAE paper no 932714.
Matthews RD, Hall Matthew J, Dai Wen, Davis George C.
Combustion modeling in SI engines with a peninsula-fractal
combustion model; 1996. SAE paper no 960072.
Hamori F. Exploring the limits of hydrogen assisted jet
ignition. PhD thesis, The University of Melbourne, 2006.
Woschni G. A universally applicable equation for the
instantaneous heat transfer coefficient in the internal
combustion engine; 1967. SAE paper no 670931.
Gerhard R, Ho T, Peter VW, Jae I, Soo Y, Dae J. Performance
analysis and valve event optimization for SI engines using
fractal combustion model; 2006. SAE paper no 2006-01-3238;2006.
Sadami Y, Eiji T, Zhong Z, Yoshisuke H. Measurement and
simulation of turbulent flame propagation in a spark ignition
engine by using fractal burning model; 2001. SAE paper no
2001-01-3603.
L, Smallwood GJ, Wong R, Snelling DR, Smith R,
Gulder O
Deschamps BM, et al. Flame front surface characteristics in
turbulent premixed propane/air combustion. Combust
Flame 2000;101:461e70.
Lee TW, Lee SJ. Direct comparison of turbulent burning
velocity and flame surface properties in turbulent premixed
flames. Combust Flame 2003;132:492e502.
Baratta M, Catania AE, Spessa E, Vassallo A. Development
of an improved fractal model for the simulation of
turbulent flame propagation in SI engines; 2005. SAE paper
no 2005-24-082.
Chin Y-W, Matthews RD, Nichols SP, Kiehne TM. Use of
fractal geometry to model turbulent combustion in SI
engines. Combust Sci Technol; 1989.
Liou D, North GL, Santavicca DA. A fractal model of turbulent
flame kernel growth; 1990. SAE paper no 900024.
L. Fractal
Cecile C, Fabien H, Christian C, Iskender G, Gulder O
characterization of high-pressure and hydrogen-enriched

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 7 ( 2 0 1 2 ) 9 8 9 2 e9 9 0 1

CH4-air turbulent premixed flames. Proc Combust Inst 2007;


31:1345e52.
[29] Tabaczynski RJ, Ferguson CR, Radhakrishnan K. A turbulent
entrainment model for spark-ignition engine combustion;
1977. SAE paper no 770647.
[30] Blizard NC, Keck JC. Experimental and theoretical
investigation of turbulent burning model for internal
combustion engines; 1974. SAE paper no 740191.

9901

[31] Sarli VDi, Benedetto ADi. Laminar burning velocity of


hydrogen-methane/air premixed flames. Int J Hydrogen
Energy 2007;32:637e46.
[32] Gottgens J, Mauss F, Peters N. Analytic approximations of
burning velocities and flame thicknesses of lean hydrogen,
methane, ethylene, ethane, acetylence, and propane flames.
Twenty-fourth symposium on Combustion/The Combustion
Institute. 1992:129e135.

You might also like