Вы находитесь на странице: 1из 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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

Available online at www.sciencedirect.com

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

Development and validation of a laminar flame


speed correlation for the CFD simulation of
hydrogen-enriched gasoline engines
Changwei Ji*, Xiaolong Liu, Shuofeng Wang, Binbin Gao, Jinxin Yang
College of Environmental and Energy Engineering, Beijing University of Technology, Beijing 100124, China

article info

abstract

Article history:

In this paper, a laminar flame speed correlation was developed and validated for the

Received 25 September 2012

computational fluid dynamics (CFD) simulation of hydrogen-enriched gasoline engines.

Received in revised form

This correlation was derived through the tabulated data which was determined by a self-

26 November 2012

developed calculation program according to the flame temperature-based mixing rule.

Accepted 28 November 2012

Wide ranges of hydrogen volume fractions (0e10%), equivalence ratios (0.6e1.5), unburned

Available online 21 December 2012

gas temperatures (300e2500 K), pressures (1e50 bar) and residual gas mass fractions (0
e20%) were simultaneously considered in this correlation to cover the burning conditions

Keywords:

encountered in SI engines. The estimated values of the new correlation were found to be in

Hydrogen

satisfying agreement with the experimental data under normal burning conditions.

Gasoline

Moreover, the new correlation was implemented in the extended coherent flame model to

Laminar flame speed

evaluate its suitability for CFD simulation. Satisfying agreement between the experimental

SI engines

and calculated results was observed under all examined hydrogen addition levels. This

CFD simulation

indicated that the new correlation was suitable for the CFD simulation of hydrogenenriched gasoline engines.
Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.

1.

Introduction

Facing the severe energy and environmental problems, more


and more researches have been dedicated to explore alternative fuels for internal combustion (IC) engines. Because of
the attractive features of hydrogen as an energy carrier such
as potential for zero CO2 and ultra low emissions, hydrogen
has been widely considered as a promising alternative fuel for
future spark-ignited (SI) engines [1e3]. The addition of
hydrogen could heighten the thermal efficiency, reduce the
toxic emissions and ease the cyclic variation of tradition
gasoline engines [4e7]. Thus, the hydrogen-enriched gasoline
engines have been considered as a feasible and effective
approach for applying hydrogen to SI engines.

In recent years, the in-depth investigation and optimization of alternative fuel-powered engines are increasingly
performed using the CFD simulation. The predictive CFD
combustion models for hydrogen-gasoline blends could
greatly facilitate the development of hydrogen-enriched
gasoline engines. Laminar flame speed is one of the most
fundamental properties of combustive mixtures. Most of the
CFD combustion models, such as coherent flame model (CFM),
stress the importance of the laminar flame speed data at
engine-relevant conditions [8]. In these models, the determination of burning rate highly relies on the laminar flame speed
at instantaneous cylinder temperature, pressure, and mixture
composition. Therefore, it is of strong necessity to obtain the
accurate data of laminar flame speed at engine-relevant

* Corresponding author. Tel./fax: 86 1067392126.


E-mail address: chwji@bjut.edu.cn (C. Ji).
0360-3199/$ e see front matter Copyright 2012, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijhydene.2012.11.139

1998

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

conditions when the CFD combustion models are adapted to


alternative fuels.
Since correlations can be more efficiently applied in CFD
codes than tabulated data, many studies have been concentrated on developing laminar flame speed correlations of
alternative fuels for CFD simulation. Verhelst et al. [9] derived
a hydrogen laminar flame speed correlation for use in engine
simulation through the one-dimensional chemical kinetic
calculations. Knop et al. [10] proposed a laminar flame speed
correlation of hydrogen-air mixture according to literature
data, accounting for the influences of equivalence ratio,
temperature, pressure and residual gas mass fraction. By
implementing this correlation in the extended coherent flame
model (ECFM), they found that the modified CFD model could
accurately capture the combustion behavior in the hydrogenfueled port fuel injection and direct injection engines. Rakopoulos et al. [11] evaluated two correlations of hydrogen
laminar flame speed reported in Refs. [12] and [13] for the use
in CFD simulation. The results showed that the correlation of
Gerke et al. [13] was preferred since it could more accurately
predict the heat release during the early stage of combustion.
Bougrine et al. [14] applied GRI 3.0 mechanism to derive the
laminar flame speed data of hydrogen-methane blends under
wide range of conditions. A correlation is then developed
accordingly through optimization algorithm. However, there
are still few available correlations that can be applied in the
CFD simulation of hydrogen-enriched gasoline engines.
In this paper, a new laminar flame speed correlation of
hydrogen-gasoline blends is developed for the CFD simulation
according to the flame temperature-based mixing rule [15,16].
Extended burning conditions are considered in this correlation with the hydrogen volume fraction in the total intake gas
varying from 0% to 10%, equivalence ratio varying from 0.6 to
1.5, unburned gas temperature varying from 300 K to 2500 K,
pressure varying from 1 to 50 bar, and residual gas mass
fraction varying from 0% to 20%. The values of the new
correlation are compared with the literature data under
different burning conditions. Moreover, in order to evaluate
the new correlation for the use in CFD simulation, the new
correlation and the Le Chateliers Rule-like formula [17] are
respectively implemented in the ECFM to calculate the
combustion process of a hydrogen-enriched gasoline engine
at different hydrogen addition levels. The predictive accuracy
is then assessed by comparing the calculated engine
combustion parameters with the experimentally determined
data.

2.
Review of the literature data and feasible
approaches
2.1.

Literature data

Up to now, there is still very limited laminar flame speed data


available for hydrogen-gasoline blends. Through the experimental measurement, Mandilas et al. [18] observed that the
hydrogen addition could enhance the laminar flame speed of
iso-octane-air flames. Their investigation was performed at
the unburned gas temperature of 360 K and the initial pressure at 5 bar. The hydrogen addition level was fixed at

a hydrogen mass fraction in the total fuel of 5%. However, no


more data was provided under other burning conditions and
hydrogen addition levels. Tahtouh et al. [19] experimentally
measured the laminar flame speed of diluted hydrogen-isooctane-air flames. The burning condition was limited at the
unburned gas temperature of 300 K, the pressure of 1 bar, and
stoichiometric conditions. A correlation was developed
according to the experimental results, which could take
account for the effects of hydrogen addition level and dilution
on the laminar flame speed. However, since this correlation
was failed to cover elevated temperatures and pressures, it
cannot be applied in engine simulation. In this paper, the
correlation of Tahtouh et al. [19] is used as a reference to
assess the predictive accuracy of the laminar flame speed
correlation of hydrogen-gasoline blends at normal burning
conditions.

2.2.
Feasible approaches for estimating laminar flame
speed of hydrogen-gasoline blends
Since the combustion reaction process is highly nonlinear, the
laminar flame speed of binary fuel blends cannot be determined by linearly combining the laminar flame speed of each
individual fuel component. Aiming at efficiently estimating
the laminar flame speed of binary fuel blends with satisfying
accuracy, many researches have been dedicated to develop
predictive laminar flame speed models for binary fuel blends.
Law and coauthors [20,21] observed that the laminar flame
speed of hydrogen-hydrocarbon blends was increased almost
linearly with the increase of the hydrogen addition parameter
RH. A linear correlation was developed as follows:
SL F; RH SL F; 0 kF$RH

(1)

in Eq. (1), F means the equivalence ratio, while SL (F,0) and SL


(F,RH) are the laminar flame speed of the pure hydrocarbon
fuel and that of the hydrogen-hydrocarbon blends at the
hydrogen addition parameter of RH, respectively. The coefficient k(F) is derived from linear fitting and varied with
hydrocarbon fuel type and equivalence ratio. However, concerning that it is difficult to determine the coefficient k(F) for
hydrogen-gasoline blends under engine-relevant burning
conditions, this linear correlation cannot be directly used in
the CFD code to provide the laminar flame speed data. Moreover, as it is found by Chen et al. [22], this linear correlation
gives accurate prediction only before RH exceeds 50% and it
cannot work for further elevated hydrogen addition levels.
Chen et al. [22] theoretically developed a squared model for
laminar flame speed of binary fuel blends:


m2 cY1;u m21 Y2;u m22 cY1;u Y2;u

(2)

In Eq. (2), mi ru,iSL,i with ru,i and SL,i being the density of the
unburned mixture and laminar flame speed of the ith fuel
component, respectively. Yi,u is the mass fraction of the ith
fuel component in the unburned mixture. The coefficient c is
a free parameter. The laminar flame speed of binary fuel
blends can be obtained as SL m/ru. The proposed model was
found to be predictive for hydrogen-methane and DMEmethane blends under normal burning conditions. However,
the squared model of Chen et al. [22] is not parameter free.

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

1999

Besides the laminar flame speed of each neat fuel component,


the laminar flame speed data at one selected blending ratio is
also required to determine the coefficient c. Thus, this model
also cannot be applied to estimate the laminar flame speed of
hydrogen-gasoline blends for CFD simulation.
Di Sarli et al. [17] predicted the laminar flame speed of
hydrogen-methane blends through a Le Chateliers Rule-like
formula as follows:

estimating the laminar flame speed of hydrogen-gasoline


blends with satisfying accuracy. Thus, a laminar flame
speed correlation of hydrogen-gasoline blends is developed
according to this mixing rule. The mixing rule, the calculation
step and laminar flame speed database, and the proposed
correlation are respectively introduced in the following
subsections.


1
SL;hydrogenmethane XH2 =SL;hydrogen 1  XH2 =SL;methane

3.1.
Flame temperature-based mixing rule for laminar
flame speed

(3)

where the SL,hydrogenmethane, SL,hydrogen and SL,methane represent


the laminar flame speeds of hydrogen-methane blends, pure
hydrogen and pure methane, respectively. The XH2 means the
hydrogen mole fraction in the total fuel. Satisfying agreement
between the estimated data and the chemical kinetic calculation results was observed except for rich mixtures at high
hydrogen addition level. Unlike the linear correlation of Law
et al. [20,21] and the squared model proposed by Chen et al. [22],
the Le Chateliers Rule-like formula [17] is parameter free and
consequently can be directly used to calculate the laminar
flame speed of hydrogen-gasoline blends for CFD calculation.
However, the suitability of this formula for hydrogen-gasoline
blends at engine-relevant burning conditions needs to be
further examined.
Hirasawa et al. [15] proposed a semiempirical flame
temperature-based mixing rule for the laminar flame speed.
This mixing rule accurately predicted the laminar flame
speeds of binary fuel blends of ethylene, n-butane and
toluene. Ji et al. [16] experimentally and numerically investigated the laminar flame propagation of binary fuel blends of
n-dodecane, methylcyclohexane and toluene under an
extended range of equivalence ratios. The flame temperaturebased mixing rule was also confirmed to be predictive for the
considered fuel blends. This can be attributed to the fact that,
the laminar flame speed of the above fuel blends is mostly
sensitive to the flame temperature through its influence on
the main branching reaction H O2 / OH O, while the
kinetic couplings tend to have minor effect. Furthermore,
considering that the reaction H O2 / OH O is dominating
for the laminar flame speed of hydrogen-hydrocarbon-air
flames [14], this mixing rule tends to be also capable of estimating the laminar flame speeds of hydrogen-gasoline blends
with satisfying accuracy.
Thus, in this paper, a laminar flame speed correlation of
hydrogen-gasoline blends is developed according to the flame
temperature-based mixing rule [15,16]. The predictive accuracies of the new correlation and the Le Chateliers Rule-like
formula [17] are assessed through the comparison with the
experimentally derived correlation of Tahtouh et al. [19] under
normal burning conditions. Moreover, the suitability of
the new correlation for use in CFD simulation is then evaluated in comparison with that of the Le Chateliers Rule-like
formula [17].

In the present study, the laminar flame speed data of


hydrogen-gasoline blends is obtained through the flame
temperature-based mixing rule proposed by Hirasawa [15] and
Ji [16]:
"
SL;M exp

Description of the approach

As it is analyzed in the above section, the flame temperaturebased mixing rule [15,16] is a potential way for effectively

 
si $ln SL;i

#
(4)

i1

In Eq. (4), SL,m represents the laminar flame speed of binary


fuel blends [cm/s]. SL,i symbolizes the laminar flame speed of
the ith neat fuel component in the fuel blends [cm/s]. si is
a dimensionless number that reflects the contribution of the
ith fuel component to the adiabatic flame temperature of the
fuel blends. si is defined as:
si



Xi $Ni
$
$Tad;i
Tad;m
Nm
1

(5)

In Eq. (5), Xi is the mole fraction of the ith fuel component in


P
the binary fuel mixture 2i1 Xi 1. Nm symbolizes the total
mole numbers of the combustion products when 1 mol binary
fuel mixture is combusted in the air [mol], while Ni represents
the total mole numbers of the combustion products of 1 mol
the ith neat fuel component [mol]. Tad,m and Tad,i are the
adiabatic flame temperatures of the binary fuel mixture and
its ith fuel component [K], respectively. Nm, Tad,i and Tad,m can
be determined as follows:
Nm

2
X
Xi $Ni

(6)

i1

Tad;i Tu

Tad;m

Qi
Ni $Cp

2 
X
Xi $Ni
i1

Nm

(7)

$Tad;i

(8)

In Eq. (7), Tu represents the unburned gas temperature [K]. Qi


is the heat release per mole of the ith fuel component [J]. Cp
symbolizes the mean molar specific heat of the combustion
products [J/(mol K)].

3.2.

3.

2
X

Calculation step and database

According to the flame-temperature-based mixing rule introduced above, a calculation program is specially developed to
calculate the laminar flame speeds of hydrogen-gasoline
blends at a wide range of burning conditions.

2000

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

The calculation steps are as follows. Firstly, the species


concentrations of the combustion reactants and products are
calculated for the hydrogen-air and gasoline-air mixtures. The
reactants species are composed of fuel (H2 or C7.76H13.10 [23]),
O2 and N2, while the species including O2, N2, H2O, CO2, CO and
H2 are treated as the main combustion products. For fuel-rich
mixtures, the water gas shift reaction is solved with the
species equilibrium constant for formation data as a function
of the adiabatic flame temperature. The parameter Ni is
deduced accordingly for different equivalence ratios and
adiabatic flame temperatures. Meanwhile, according to the
derived species concentrations in the combustion products,
the mean molar specific heat of the combustion products Cp is
calculated with the species molar specific heat data as
a function of the average of the unburned gas temperature
and the adiabatic flame temperature [24]. The heat release per
mole of fuel Qi is derived by calculating the difference between
the total enthalpies of the reactants and the products at the
unburned gas temperature. The total enthalpies of
the combustion reactants and products are determined by the
absolute enthalpy databases and the molar concentrations of
each species. Subsequently, an iterative calculation of the
adiabatic flame temperature is conducted with an initial value
of 2000 K. The convergence criterion is jTad;n  Tad;n1 jh1K,
where the Tad,n means the calculated adiabatic flame

3 2
SL;0
a0
4 aT 5 4 b0
bp
c0
2

a1
b1
c1

2500 K. The pressure is varied from 1 to 50 bar. The database


totally contains 14,520 data points, which can cover the
burning conditions encountered in naturally aspirated SI
engines.

3.3.

As correlations can be more efficiently implemented in the


CFD combustion models than the tabulated data, a new
laminar flame speed correlation for hydrogen-gasoline blends
is proposed according to the above database. The correlation
is in the following form, which is similar to that presented by
Metghalchi and Keck [27]:


SL;hydrogengasoline aH2 ; F; Tu ; p; Ydil

a 
bp
Tu T
P
$
$1  2:1Ydil
SL;0 $
Pref
Tref

where the VH2 and Vair represent the volumes of hydrogen and
air in the intake charge [dm3], respectively. In the database,
the hydrogen volume fraction in the total intake gas is varied
from 0% to 10%. The equivalence ratio is varied from 0.6 to
1.5. The unburned gas temperature is varied from 300 to

(10)

In Eq. (10), the laminar flame speed of hydrogen-gasoline


blends SL,hydrogengasoline [cm/s] is treated as a multivariate
function of hydrogen volume fraction in the total intake gas aH2 ,
equivalence ratio F, unburned gas temperature Tu [K], pressure
p [bar], and residual gas mass fraction Ydil. The reference
temperature Tref and reference pressure pref are equal to 298 K
and 1 bar, respectively. The SL,0, aT and bp are functions of aH2
and F, which can be determined through Eq. (11):

3
.a8

T
.b8 5$ 1 F F2 aH2 a2H2 FaH2 F2 aH2 Fa2H2 F2 a2H2
.c8

temperature after the nth iteration. At a specified burning


condition and a given fuel blending ratio Xi, the dimensionless
number si is deduced by Eq. (5) with the calculated values of Ni
and Tad,i from the iteration. According to the calculated si and
SL,i, the laminar flame speeds of hydrogen-gasoline blends are
computed through Eq. (4) at different unburned gas temperatures, pressures, equivalence ratios and hydrogen volume
fractions in the total intake gas. In this program, the equilibrium constant for formation, molar specific heat and absolute
enthalpy databases for each species are derived from the
JANAF tables [25]. The laminar flame speeds of hydrogen-air
and gasoline-air mixtures are derived from the correlation
reported in Refs. [10] and [26], respectively.
A laminar flame speed database of hydrogen-gasoline
blends is then generated from the above calculation
program. In order to facilitate the CFD simulation of
hydrogen-enriched gasoline engines, the hydrogen volume
fraction in the total intake gas aH2 is selected to represent the
hydrogen addition level, which is defined as:


(9)
aH2 VH2 = VH2 Vair

Laminar flame speed correlation

(11)

Table 1 lists the fit coefficients a0 to c8 in Eq. (11). The fit


coefficients are obtained through the LevenbergeMarquardt
method [28,29]. Meanwhile, since there is no available data
on the laminar flame speeds of diluted hydrogen-gasolineair mixtures at elevated temperatures and pressures, the
part of (1e2.1Ydil), which has been widely used in the
previous correlations [23], is retained in the new correlation
to represent the influence of the residual gas. This correlation is suitable for extended burning conditions with
the hydrogen volume fraction in the total intake gas varying
from 0% to 10%, equivalence ratio varying from 0.6 to
1.5, unburned gas temperature varying from 300 K to 2500 K,
pressure varying from 1 to 50 bar, and residual gas
mass fraction varying from 0% to 20%, which could cover
the burning conditions encountered in traditional SI
engines.
Fig. 1 compares the laminar flame speeds predicted by
the correlation with those in the database for all the fitted
data points (14,520 in total). The relative error is less than
5% for all the data points, and more than 98.46% data
points are within the error limit of 2%. The coefficient of
determination R2 is shown to be 0.99984. A satisfying
agreement is observed, suggesting that the influences of aH2 ,
F, Tu and p on the laminar flame speed calculated from the
self-developed program are well reproduced in the new
correlation.

2001

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

Table 1 e Fit coefficients of the new laminar flame speed correlation.


Parameter
a0
a1
a2
a3
a4
a5
a6
a7
a8

Value
49.8508
133.9972
55.4971
157.4988
1516.7802
545.4127
241.5668
646.0548
10.4278

Parameter
b0
b1
b2
b3
b4
b5
b6
b7
b8

(F 
(F 
(F 
(F 
(F 
(F 
(F 
(F 
(F 

4.

Results and discussions

4.1.

Comparison with literature data

1:0)
1:0)
1:0)
1:0)
1:0)
1:0)
1:0)
1:0)
1:0)

Value
2.9486
0.7494
0.0264
20.1404
13.3876
33.0797
15.6721
42.9936
28.0143

In this section, the estimated values form the new correlation


and the Le Chateliers Rule-like formula [17] are firstly
compared with the experimental data of Tahtouh et al. [19]
under different hydrogen volume fractions and residual gas
mass fractions at normal burning condition. Subsequently,
comparisons are conducted to analyze the difference of the
new correlation and the Le Chateliers Rule-like formula [17]
under engine-relevant conditions.
Fig. 2 illustrates the effects of hydrogen volume fraction
and residual gas mass fraction on the laminar flame speed
calculated from various correlations under normal burning
condition (Tu 300 K, p 1 bar, F 1.0). It can be found in
Fig. 2a that in comparison with the experimentally derived
correlation of Tahtouh et al. [19], the Le Chateliers Rule-like
formula [17] significantly overestimates the laminar flame
speed with hydrogen addition. Comparatively, the new
correlation tends to be maintained at satisfying accuracy
under high hydrogen addition levels. The relevant error of the
new correlation is less than 10% before the hydrogen volume
fraction exceeds 8%, while that of the Le Chateliers Rule-like
formula [17] researches 10% at the hydrogen volume fraction

Fig. 1 e Comparison of the laminar flame speeds estimated


by the new correlation with those in the database for all
the fitted data points.

Parameter
b0
b1
b2
b3
b4
b5
b6
b7
b8

(F > 1:0)
(F > 1:0)
(F > 1:0)
(F > 1:0)
(F > 1:0)
(F > 1:0)
(F > 1:0)
(F > 1:0)
(F > 1:0)

Value

Parameter

Value

2.9729
0.8017
0.0018
6.5013
12.5686
4.5552
0.7006
16.3882
5.5966

c0
c1
c2
c3
c4
c5
c6
c7
c8

0.3781
0.2096
0.0048
3.3960
1.7898
3.9679
1.3055
1.3144
0.1638

of only 0.3%. The relevant errors of the new correlation and


the Le Chateliers Rule-like formula [17] at the hydrogen
volume fraction of 10% are 13.7% and 167.3%, respectively.
Meanwhile, in Fig. 2b, the new correlation is also observed in
better agreement with the experimental data of Tahtouh et al.
[19] under different residual gas mass fractions, compared
with the Le Chateliers Rule-like formula [17]. When the
residual gas mass fraction is 5%, the errors of the new

Fig. 2 e a: Effect of hydrogen volume fraction on the


laminar flame speed calculated from various correlations
under normal burning condition, b: Effect of residual gas
mass fraction on the laminar flame speed calculated from
various correlations under normal burning condition.

2002

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

Fig. 3 e a: Effect of unburned gas temperature on the


laminar flame speed calculated from various correlations
under engine-relevant condition, b: Effect of pressure on
the laminar flame speed calculated from various
correlations under engine-relevant condition.

Table 2 e Specifications of the modeled engine.


Engine type
Number of valves
Bore
Stroke
Compression ratio
Displacement
Connecting rod
length
Intake valve
opening
Intake valve
closure
Exhaust valve
opening
Exhaust valve
closure

SI engine, 4-cylinder, 4-stroke,


naturally aspirated, port fuel injection
4 per cylinder
77.4 mm
85.0 mm
10.0:1
1.599 L
141 mm
45  CA BTDC
135  CA BTDC
135  CA ATDC
45  CA ATDC

Fig. 4 e a: Comparison of the experimental determined and


calculated MFB traces using different laminar flame speed
correlations at the hydrogen volume fraction of 0%, b:
Comparison of the experimental determined and
calculated MFB traces using different laminar flame speed
correlations at the hydrogen volume fraction of 3%, c:
Comparison of the experimental determined and
calculated MFB traces using different laminar flame speed
correlations at the hydrogen volume fraction of 6%.

2003

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

correlation and the Le Chateliers Rule-like formula [17] in


comparison with the experimental data of Tahtouh et al. [19]
are 2.3 cm/s and 28.6 cm/s at the hydrogen volume fraction
of 3%, respectively, while those at the hydrogen volume
fraction of 10% are 6.6 cm/s and 70.4 cm/s, respectively.
Moreover, at the hydrogen volume fraction of 10%, when the
residual gas mass fraction is elevated from 0% to 20%, the
laminar flame speeds calculated from the correlation of Tahtouh et al. [19] and the new correlation are decreased by 21.5%
and 22.1%, respectively. This means that before the residual
gas mass fraction exceeds 20%, the effect of dilution on
reducing the laminar flame speed of hydrogen-gasoline
blends can be well reproduced in the new correlation.
Therefore, it can be concluded from Fig. 2 that, the new
correlation is able to reproduce the effects of hydrogen addition level and dilution on the laminar flame speed of
hydrogen-gasoline blends with satisfying accuracy at the
normal burning conditions. However, the values calculated
from the Le Chateliers Rule-like formula [17] are significantly
overestimated, especially at high hydrogen addition levels.
Fig. 3 compares the laminar flame speed calculated from
the new correlation and the Le Chateliers Rule-like formula
[17] under engine-relevant burning conditions. The selected
temperatures and pressures are in the ranges which are
similar to those encountered in an SI engine. The residual gas
mass fraction of 5% is chosen since it is a representative value
for the hydrogen-enriched gasoline engine without external
EGR. As it is seen in Fig. 3, compared with the new correlation,
the laminar flame speed is significantly overestimated by the
Le Chateliers Rule-like formula [17] under engine-relevant
burning conditions. The difference is more pronounced at
high hydrogen addition levels than at low hydrogen addition
levels. In Fig. 3a, when the unburned gas temperature is
elevated from 1500 to 2000 K, the laminar flame speed at
hydrogen volume fraction of 3% estimated from the Le Chateliers Rule-like formula [17] rises from 1017.1 cm/s to
1702.2 cm/s, whereas those calculated from the new correlation increases from 640.5 cm/s to 1168.4 cm/s. Moreover, more
significant difference is observed at the hydrogen volume
fraction of 10%. At the unburned gas temperature of 2000 K,
the laminar flame speed estimated from the Le Chateliers
Rule-like formula [17] is 740.5 cm/s higher than that calculated
from the new correlation. Meanwhile, it can be found in Fig. 3b

that, the similar difference between these two correlations is


also can be observed under different pressures. Under the
pressure of 20 bar, the laminar flame speeds calculated from
the Le Chateliers Rule-like formula [17] is 389.8 and 691.1 cm/s
higher than those estimated by the new correlation at the
hydrogen volume fractions of 3% and 10%, respectively. Thus,
at the engine-relevant burning conditions considered in Fig. 3,
the values calculated from the Le Chateliers Rule-like formula
[17] are significantly higher than those estimated from the
new correlation, especially at high hydrogen addition levels.

4.2.

CFD results

In order to evaluate the suitability of the new correlation and


the Le Chateliers Rule-like formula [17] for use in CFD simulation, the two correlations are respectively implemented in
the ECFM to calculate the combustion process of a hydrogenenriched gasoline engine at different hydrogen addition
levels. In this section, the CFD model and calculation results in
comparison with experimental data are introduced.

4.2.1.

CFD model

The calculations are performed using the CFD code AVL FIRE.
The k-z-f model [30] is applied to take account of the turbulence effect. The combustion is described by the ECFM [31,32].
The spherical model [31] is used to capture the flame kernel
formation. The flame thickness is determined through the
Blint correlation [33]. Meanwhile, the near-wall treatment
follows the description proposed by Popovac et al. [34].
The modeled engine is a 1.6L four-cylinder SI engine
manufactured by Beijing Hyundai Motors. The detailed engine
specifications are listed in Table 2. The original engine is
modified so that the hydrogen and gasoline can be simultaneously injected in the intake manifolds and mixed with air at
the end of the intake stroke. The calculations are conducted at
an engine speed of 1400 rpm and a MAP of 61.5 kPa, which
could represent one of the typical city driving conditions. The
global equivalence ratio of hydrogen-gasoline-air mixtures is
fixed at 0.8 and the spark timing is kept at 22  CA ATDC for all
the test cases. Three hydrogen volume fractions in the total
intake gas of 0%, 3% and 6% are selected to explore the
combustion process in the hydrogen-enriched gasoline
engine. The corresponding experimental results have been

Table 3 e Comparison of calculated and experimental results.




Pmax/bar qPmax/ CA pmax/% CA0-10/ CA CA0-10/% CA10-90/ CA CA10-90/% CA50/ CA


aH2 0% (experiment)
aH2 0% (CFD, Le Chateliers Rule-like
formula [17])
aH2 0% (CFD, new correlation)
aH2 3% (experiment)
aH2 3% (CFD, Le Chateliers Rule-like
formula [17])
aH2 3% (CFD, new correlation)
aH2 6% (experiment)
aH2 6% (CFD, Le Chateliers Rule-like
formula [17])
aH2 6% (CFD, new correlation)

16.96
16.61

23.1
21.7

e
2.06

27.4
28.0

e
2.19

38.8
37.8

e
2.58

23.2
22.2

16.87
25.20
36.09

21.8
17.2
6.5

0.53
e
43.21

27.7
23.0
15.5

1.09
e
32.61

36.7
21.6
12.2

5.41
e
43.52

21.7
10.6
0.2

25.62
31.58
37.94

17.1
12.5
5.5

1.67
e
20.14

22.6
18.2
14.0

1.74
e
23.08

21.0
16.7
11.8

2.78
e
29.34

10.9
5.6
2.1

31.56

12.2

0.06

19.2

5.49

17.0

1.80

5.6

2004

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

published previously [6]. The boundary and initial conditions


used in the CFD calculation are specified according to the
experimental measurements as well as the reliable estimations. The calculations begin at the time of intake valve
closure (IVC) and end at the time of exhaust valve opening
(EVO). The time step is kept at 0.5  CA before the spark
discharge at 22  CA ATDC and then further decreased to
0.1  CA afterwards to track the combustion period.
During the calculation, the new correlation and the Le
Chateliers Rule-like formula [17] are respectively implemented in the CFD code to calculate the laminar flame speed
data for each computational cell according to the local
burning condition and hydrogen addition level. The predictive
accuracy is examined without case-specific parameter tuning.
For all the test cases, the stretch factor and consumption
factor in the ECFM are kept at 1.2 and 1.0, respectively. The
CFD code AVL FIRE is executed in parallel on an 8-core server.
The calculation consumes about 54 h for each test case.

4.2.2.

Comparison with experimental data

In this section, the CFD calculation results using the new


correlation and the Le Chateliers Rule-like formula [17] are
respectively compared with the experimental data under
various hydrogen addition levels. The uncertainties of all the
equipments in the experimental system have been listed in
Ref. [6] in detail.
Fig. 4 shows the experimentally determined and calculated
mass fraction burned (MFB) traces using different laminar
flame speed correlations under the hydrogen volume fractions of 0%, 3% and 6%. It can be observed that the calculated
MFB traces using the new correlation are approximately
constant with the experimental results under all hydrogen
addition levels, while those calculated using the Le Chateliers
Rule-like formula [17] is significantly advanced and shortened
at hydrogen volume fractions of 3% and 6%. The relevant error
on the flame propagation period (CA10-90) is 5.41%, 2.78% and
1.80% at hydrogen volume fractions of 0%, 3% and 6% when
the new correlation is used in CFD calculation. However, with
the implementation of the Le Chateliers Rule-like formula
[17], the CFD results of flame propagation period (CA10-90) are
2.58%, 43.52%, and 29.34% shorter than the experimental data
(as it is seen in Table 3). This can be explained by the fact that,
in ECFM, the laminar flame speed not only directly influences
the reaction rate, but also affects the flame surface production
through its influence on the flame thickness. Since the
laminar flame speed of hydrogen-gasoline blends is overestimated by the Le Chateliers Rule-like formula [17] at
engine-relevant burning conditions (as it is demonstrated in
Fig. 3), the degree of flame wrinkling and local combustion
reaction rate are both enhanced. This leads to the advanced
and shortened heat release when the Le Chateliers Rule-like
formula [17] is used in CFD calculation.
Fig. 5 compares the measured and calculated mean incylinder pressure traces using different laminar flame speed
correlations at the hydrogen volume fractions of 0%, 3% and
6%. It is found that the in-cylinder pressure trace calculated
using the new correlation is in good agreement with the
experimental results. However, with the implementation of
the Le Chateliers Rule-like formula [17], the peak pressure
(Pmax) is overestimated and the relevant crank angle for the

Fig. 5 e a: Comparison of the measured and calculated


mean in-cylinder pressure traces using different laminar
flame speed correlations at the hydrogen volume fraction
of 0%, b: Comparison of the measured and calculated mean
in-cylinder pressure traces using different laminar flame
speed correlations at the hydrogen volume fraction of 3%,
c: Comparison of the measured and calculated mean incylinder pressure traces using different laminar flame
speed correlations at the hydrogen volume fraction of 6%.

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

peak pressure (qPmax) is advanced compared with the experimental data at the hydrogen volume fractions of 3% and 6%.
When the CFD calculation is conducted using the new correlation, the relevant error on the peak pressure (pmax) is 0.53%,
1.67% and 0.06% at the hydrogen volume fractions of 0%, 3%
and 6%. The error on the relevant crank angle for the peak
pressure at the hydrogen volume fractions of 3% and 6% is
only 0.1 and 0.3  CA, respectively. Comparatively, with the
implementation of the Le Chateliers Rule-like formula [17],
the pmax is 2.06%, 43.21% and 20.14% at the hydrogen volume
fractions of 0%, 3% and 6%. Moreover, the error on the relevant
crank angle for the peak pressure is 10.7 and 7.0  CA at the
hydrogen volume fractions of 3% and 6%, respectively. This is
because the overestimated laminar flame speed of the Le
Chateliers Rule-like formula [17] leads to the advanced and
shortened heat release of the CFD calculation, which results in
the disagreement with the experimental results after the
addition of hydrogen.
The detailed comparison data is listed in Table 3, including
the peak in-cyinder pressure (Pmax), relevant crank angle for
the peak pressure (qPmax), relevant error on the peak pressure
(pmax), flame development period (CA0-10), relevant error on
the flame development period (CA0-10), flame propagation
period (CA10-90), relevant error on the flame propagation
period (CA10-90), and relevant crank angle for 50% mass fraction burned (CA50). The calculated results at hydrogen volume
fraction of 0% are in good agreement with the experimental
data for both correlations. For the high hydrogen addition
levels, the CFD results using the new correlation are much
closer to the experimental data than those calculated using
the Le Chateliers Rule-like formula [17]. This indicates that
the new correlation is more suitable for the CFD calculation on
the combustion process in hydrogen-enriched gasoline
engines.

5.

Conclusions

This paper developed and validated a laminar flame speed


correlation for use in CFD simulation of hydrogen-enriched
gasoline engines. The influences of hydrogen volume fractions, equivalence ratios, unburned gas temperatures, pressures and residual gas mass fractions on the laminar flame
speed of hydrogen-gasoline blends are simultaneously
considered in this correlation. The estimated values of the
new correlation are found to be in satisfying agreement with
the experimental data reported in Ref. [19] under normal
burning conditions. Moreover, in order to evaluate the suitability of the new correlation for use in CFD simulation, the
new correlation and the Le Chateliers Rule-like formula [17]
are respectively implemented in ECFM to calculate the
combustion process of a hydrogen-enriched gasoline engine
at different hydrogen addition levels. With the implementation of the new correlation, satisfying agreement between the
experimental and calculated results is observed at all the
examined hydrogen addition levels. Comparatively, the CFD
model using the Le Chateliers Rule-like formula [17] is failed
to capture the combustion characteristics after the addition of
hydrogen. Therefore, the new correlation is preferred to be
applied in the CFD simulation.

2005

Acknowledgments
This work was supported by National Program on Key Basic
Research Project (973 Program) (Grant No.2013CB228403), Key
Program of Sci & Tech Project of Beijing Municipal Commission of Education (Grant No.KZ201210005002), Ph.D. Programs
Foundation of Ministry of Education of China (Grant
No.20111103110010) and Beijing Municipal Natural Science
Foundation (Grant No. 3122006). The authors would also like
to express their gratitude towards the technical support
received from Dr. P. Priesching and Mr. M. Suffa in AVL List
GmbH, Graz, Austria.

references

[1] Verhelst S, Wallner T. Hydrogen-fueled internal combustion


engines. Prog Energy Combust Sci 2009;35:490e527.
[2] Verhelst S, Woolley R, Lawes M, Sierens R. Laminar and
unstable burning velocities and Markstein lengths of
hydrogeneair mixtures at engine-like conditions. Proc
Combust Inst 2005;30:209e16.
[3] Verhelst S, Maesschalck P, Rombaut N, Sierens R. Efficiency
comparison between hydrogen and gasoline, on a bi-fuel
hydrogen/gasoline engine. Int J Hydrogen Energy 2009;34:
2504e10.
[4] Ji C, Wang S. Experimental study on combustion and
emissions performance of a hybrid hydrogenegasoline engine
at lean burn limits. Int J Hydrogen Energy 2010;35:1453e62.
[5] Ji C, Wang S. Effect of hydrogen addition on the idle
performance of a spark ignited gasoline engine at
stoichiometric condition. Int J Hydrogen Energy 2009;34:
3546e56.
[6] Ji C, Wang S. Effect of hydrogen addition on combustion and
emissions performance of a spark ignition gasoline engine at
lean conditions. Int J Hydrogen Energy 2009;34:7823e34.
[7] Lucas GG, Richards WL. The hydrogen/petrol engine e the
means to give good part-load thermal efficiency. SAE Paper
No.820315 1982.
[8] Veynante D, Vervisch L. Turbulent combustion modeling.
Prog Energy Combust Sci 2002;28:193e266.
[9] Verhelst S, TJoen C, Vancoillie J, Demuynck J. A correlation
for the laminar burning velocity for use in hydrogen spark
ignition engine simulation. Int J Hydrogen Energy 2011;36:
957e74.
[10] Knop V, Benkenida A, Jay S, Colin O. Modelling of combustion
and nitrogen oxide formation in hydrogen-fuelled internal
combustion engines within a 3D CFD code. Int J Hydrogen
Energy 2008;33:5083e97.
[11] Rakopoulos CD, Kosmadakis GM, Pariotis EG. Evaluation of
a combustion model for the simulation of hydrogen sparkignition engines using a CFD code. Int J Hydrogen Energy
2010;35:12545e60.
[12] Verhelst S. A study of the combustion in hydrogen-fuelled
internal combustion engines. Ph.D. thesis. Ghent, Belgium:
Ghent University; 2005.
[13] Gerke U, Steurs K, Rebecchi P, Boulouchos K. Derivation of
burning velocities of premixed hydrogen/air flames
atengine-relevant conditions using a single-cylinder
compression machine with optical access. Int J Hydrogen
Energy 2010;35:2566e77.
[14] Bougrine S, Richard S, Nicolle A, Veynante D. Numerical
study of laminar flame properties of diluted methanehydrogen-air flames at high pressure and temperature using
detailed chemistry. Int J Hydrogen Energy 2011;36:12035e47.

2006

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 8 ( 2 0 1 3 ) 1 9 9 7 e2 0 0 6

[15] Hirasawa T, Sung CJ, Joshi A, Yang Z, Wang H, Law CK.


Determination of laminar flame speeds using digital particle
image velocimetry: binary fuel blends of ethylene, n-butane,
and toluene. Proc Combust Inst 2002;29:1427e34.
[16] Ji C, Egolfopoulos FN. Flame propagation of mixtures of air
with binary liquid fuel mixtures. Proc Combust Inst 2011;33:
955e61.
[17] Di Sarli V, Di Benedetto A. Laminar burning velocity of
hydrogen-methane/air premixed flames. Int J Hydrogen
Energy 2007;32:637e46.
[18] Mandilas C, Ormsby MP, Sheppard CGW, Woolley R. Effects
of hydrogen addition on laminar and turbulent premixed
methane and iso-octane-air flames. Proc Combust Inst 2007;
31:1443e50.
[19] Tahtouh T, Halter F, Mounam-Rousselle C. Laminar
premixed flame characteristics of hydrogen blended isooctane-air-nitrogen mixtures. Int J Hydrogen Energy 2011;36:
985e91.
[20] Yu G, Law CK, Wu CK. Laminar flame speed of
hydrocarbonair mixtures with hydrogen blending.
Combust Flame 1986;63:339e47.
[21] Tang C, Huang Z, Law CK. Determination, correlation, and
mechanistic interpretation of effects of hydrogen addition
on laminar flame speeds of hydrocarbon-air mixtures. Proc
Combust Inst 2011;33:921e8.
[22] Chen Z, Dai P, Chen S. A model for the laminar flame speed
of binary fuel blends and its application to methane/
hydrogen mixtures. Int J Hydrogen Energy 2012;37:10390e6.

[23] Turns SR. An introduction to combustion: concepts and


applications. 2nd ed. New York: McGraw-Hill; 2000.
[24] Wark K. Thermodynamics. 6th ed. New York: McGraw-Hill;
1999.
[25] NIST-JANAF. Thermochemical tables, http://kinetics.nist.
gov/janaf/.
[26] Heywood JB. Internal combustion engine fundamentals. New
York: McGraw-Hill; 1988.
[27] Metghalchi M, Keck JC. Laminar burning velocity of propaneair mixtures at high temperature and pressure. Combust
Flame 1980;38:143e54.
[28] Levenberg K. A method for the solution of certain
nonlinear problems in least squares. Quart Appl Math
1944;2:164e8.
[29] Marquardt DW. An algorithm for least-squares estimation of
nonlinear inequalities. SIAM J Appl Math 1963;11:431e41.
 K, Popovac M, Hadz
 M. A robust near-wall
iabdic
[30] Hanjalic
elliptic-relaxation eddy-viscosity turbulence model for CFD.
Int J Heat Fluid Flow 2004;25:1047e51.
[31] AVL FIRE users guide version. AVL LIST GmbH; 2010.
[32] Colin O, Benkenida A, Angelberger C. 3D modeling of mixing,
ignition and combustion phenomena in highly stratified
gasoline engines. Oil Gas Sci Technol 2003;58:47e62.
[33] Blint RJ. The relationship of the laminar flame width to flame
speed. Combust Sci Technol 1986;49:79e92.
 K. Compound wall treatment for RANS
[34] Popovac M, Hanjalic
computation of complex turbulent flows and heat transfer.
Flow Turbul Combust 2007;78:177e202.

Вам также может понравиться