You are on page 1of 16

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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

Available at www.sciencedirect.com

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

Evaluation of a combustion model for the simulation


of hydrogen spark-ignition engines using a CFD code

C.D. Rakopoulos a,*, G.M. Kosmadakis a, E.G. Pariotis b


a
Internal Combustion Engines Laboratory, Thermal Engineering Department, School of Mechanical Engineering, National Technical
University of Athens, 9 Heroon Polytechniou St., Zografou Campus, 15780 Athens, Greece
b
Laboratory of Naval Propulsion Systems, Section of Naval Architecture and Marine Engineering, Department of Naval Sciences, Hellenic
Naval Academy, End of Hatzikyriakou Ave., Hatzikyriakio, 18539 Piraeus, Greece

article info abstract

Article history: The present work deals with the evaluation of a combustion model that has been devel-
Received 12 July 2010 oped, in order to simulate the power cycle of hydrogen spark-ignition engines. The moti-
Received in revised form vation for the development of such a model is to obtain a simple combustion model with
1 September 2010 few calibration constants, applicable to a wide range of engine configurations, incorporated
Accepted 1 September 2010 in an in-house CFD code using the RNG ke3 turbulence model. The calculated cylinder
Available online 22 September 2010 pressure traces, gross heat release rate diagrams and exhaust nitric oxide (NO) emissions
are compared with the corresponding measured ones at various engine loads. The engine
Keywords: used is a Cooperative Fuel Research (CFR) engine fueled with hydrogen, operating at
Hydrogen a constant engine speed of 600 rpm. This model is composed of various sub-models used
Combustion for the simulation of combustion of conventional fuels in SI engines; it has been adjusted in
Spark-ignition the current study specifically for hydrogen combustion. The basic sub-model incorporated
CFD model for the calculation of the reaction rates is the characteristic conversion time-scale method,
Laminar meaning that a time-scale is used depending on the laminar conversion time and the
Turbulent turbulent mixing time, which dictates to what extent the combustible gas has reached its
chemical equilibrium during a predefined time step. Also, the laminar and turbulent
combustion velocity is used to track the flame development within the combustion
chamber, using two correlations for the laminar flame speed and the Zimont/Lipatnikov
approach for the modeling of the turbulent flame speed, whereas the (NO) emissions are
calculated according to the Zeldovich mechanism. From the evaluation conducted, it is
revealed that by using the developed hydrogen combustion model and after adjustment of
the unique model calibration constant, there is an adequate agreement with measured
data (regarding performance and emissions) for the investigated conditions. However,
there are a few more issues to be resolved dealing mainly with the ignition process and the
applicability of a reliable set of constants for the emission calculations.
ª 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.

Abbreviations: ABDC, after bottom dead center; ATDC, after top dead center; BBDC, before bottom dead center; BTDC, before top dead
center;  CA, degrees of crank angle; CFD, computational fluid dynamics; CR, compression ratio; EGR, exhaust gas recirculation; EVO,
exhaust valve opening; IT, ignition timing; IVC, inlet valve closure; NO, nitric oxide; NOx, nitrogen oxides; rpm, revolutions per minute;
SI, spark-ignition; TDC, top dead center.
* Corresponding author. Tel.: þ30 210 7723529; fax: þ30 210 7723531.
E-mail address: cdrakops@central.ntua.gr (C.D. Rakopoulos).
0360-3199/$ e see front matter ª 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2010.09.002
12546 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

Nomenclature V cylinder volume, m3


w velocity along z-axis, m/s
cp specific heat capacity under constant pressure,
Wgi gross indicated work, J
J kg1 K1
y distance of the computational node from the wall,
DT,u thermal diffusivity of the unburned mixture, m2/s
m
f residual gas mass fraction, kgi/kg
yþ non-dimensional distance from the wall, e
h enthalpy, J/kg
k turbulent kinetic energy (per unit mass), m2/s2 Greek symbols
Lt turbulent integral length scale, mm G4 diffusion coefficient, kg m1 s1
P pressure, N/m2 3 turbulent dissipation rate (per unit mass), m2/s3
Pr Prandtl number, e meff effective viscosity, kg m1 s1
qW wall heat flux, W/m2 n kinematic viscosity, m2/s
rk flame kernel radius, m r density, kg/m3
S4 source term rb burned gas density, kg/m3
Scr source term due to crevice flows ru unburned gas density, kg/m3
Sh source term of the enthalpy equation sh coefficient for the enthalpy equation, e
SRNG source term due to RNG turbulence model s3 coefficient for the dissipation of turbulent kinetic
Sct turbulent Schmidt number, e energy equation, e
t time, s sk coefficient for the turbulent kinetic energy
T temperature, K equation, e
TW wall temperature, K sW local wall shear stress, kg m1 s2
Tu unburned gas temperature, K sc characteristic conversion time, s
u velocity along x-axis, m/s sl laminar kinetics time, s
ul laminar flame speed, m/s st turbulent mixing time, s
ut turbulent flame speed, m/s Yi species mass fraction, kgi/kg
uT friction velocity, m/s Yi* species chemical equilibrium mass fraction, kgi/kg
u0 rms turbulent velocity, m/s f generalized variable
!u velocity vector, m/s 4 equivalence ratio, e
v velocity along y-axis, m/s

1. Introduction For the evaluation of this model, the calculated results


regarding engine performance and nitric oxide emissions are
The strict emission legislation has forced the automotive compared with the corresponding measured ones at various
industry to take into consideration the use of alternative fuels engine loads. From the analysis it is concluded that the
in internal combustion engines, in order to reduce exhaust presented combustion model manages to predict adequately
emissions with the least impact on engine’s performance. the engine performance within the range of examined
Towards this direction, bio-fuels gain an increasing attention equivalence ratios. As far as the nitric oxide exhaust emis-
lately, and research is conducted in order to evaluate the sions are concerned, the model manages to capture quali-
performance and emissions from internal combustion tatively correct the trend as equivalence ratio varies, while
engines using such promising fuels [1e4]. the absolute values present a small difference from the
Another alternative fuel that seems to be promising for measured ones.
spark-ignition engines is hydrogen, either used in dedicated
hydrogen engines [5] or blended with another fuel [6]. Its use in
such engines has been experimentally tested by many 2. Experimental data
researchers [7e9], and proper computational models have been
developed so far [10e13]. To gain a fundamental understanding The experimental data used in the present study, for the
of the hydrogen combustion mechanism in internal combus- evaluation of the developed combustion model, are provided
tion engines, detailed CFD simulation models should be used. by another research group [8]. These data concern measure-
The present study presents a hydrogen combustion model, ments on a Cooperative Fuel Research (CFR) engine, operating
which is incorporated in an in-house CFD code for the simu- at a constant engine speed equal to 600 rpm. Fuel is admitted
lation of a hydrogen spark-ignition engine. The combustion in the engine through a gas-carburetor in the air-inlet mani-
model uses recently developed correlations for the hydrogen fold and the fuel rate is controlled with a valve, in order to
laminar flame speed [11], while a more detailed description of adjust the equivalence ratio. This means that the load of the
the spatial flame propagation is accomplished by using local engine is altered with the variation of the equivalence ratio
gas properties. An advantage of this model is that only one (quantified load control). All measurements are conducted
constant is needed to be tuned within a quite small range, with a wide-open throttle (WOT). The main specifications of
contrary to what it is done in most existing multi-dimensional this engine can be seen in Table 1, where the valve timing
combustion models [14]. events are the ones referenced in [8].
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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0 12547

Table 1 e CFR engine specifications. Table 2 e Variables that represent the generalized
variable f in Eq. (1).
Engine model CFR engine, single cylinder,
and type naturally aspirated, Generalized Equation G4
four-stroke, water-cooled variable (f)

Bore 82.55 mm 1 Continuity 0


Stroke 114.2 mm u u-momentum meff
Swept volume 0.6117 l v v-momentum meff
Connecting rod length 254 mm w w-momentum meff
Compression ratio Variable (data provided h Enthalpy meff/sh
for CR: 7e9.5) k Turbulent kinetic energy meff/sk
Valves 2 (unshrouded) 3 Dissipation of turbulent meff/s3
Engine speed 600 rpm (constant) kinetic energy
Ignition timing Variable Yi Species (i) mass fraction meff/Sct
Number of piston rings 5
Valve Inlet valve opening 18 CA ATDC
timing Inlet valve closure 25 CA ABDC
the mesh generation model, as well as the detailed evaluation
Exhaust valve 34 CA BBDC
of the developed CFD model under motoring conditions, can
opening
Exhaust valve 7 CA ATDC be found in previous published papers by the authors [19,20].
closure

3.2. Heat transfer model

The available experimental data concern the inlet mass Hydrogen-fueled engines require special attention in the
flow rates of air and hydrogen, the cylinder pressure traces, description of the wall heat transfer, due to their thinner
together with the NOx emissions, when varying the thermal boundary layer (shorter quenching distance) and
compression ratio, equivalence ratio and ignition timing. their much higher laminar flame speed, compared to the ones
Further details concerning the implementation of the experi- of the conventional-fueled engines [8]. The heat transfer
mental facilities can be found in [8]. mechanism in this kind of engines has been enlightened by
some recent experimental studies, such as the ones in [21,22];
they have measured the wall heat fluxes in the cylinder of
3. Numerical model hydrogen spark-ignition engines, and concluded that the peak
heat flux is increased in this kind of engines as well as the
3.1. CFD model cumulative heat loss to the cylinder walls especially for stoi-
chiometric operation.
The CFD code developed by the authors can simulate three- In the present study the wall heat flux correlation devel-
dimensional curvilinear domains using the finite volume oped by the authors [23] is used for the simulation of the heat
method in a collocated grid. It incorporates the RNG ke3 transfer mechanism. Its formulation is based on a compress-
turbulence model with some slight modifications to introduce ible version of the standard law-of-the-wall [24] and includes
the compressibility of a fluid in generalized coordinates, as the unsteady pressure term, which has been also shown in
described in [15]. It solves the following transport equations [25] to give more accurate results. The wall heat flux (in W/m2),
for the conservation of mass, momentum and energy (Eq. (1)), which is inserted into the source term of the energy equation
as suggested in [16]. Sh for the boundary cells, is shown in Eq. (2). This expression is
used for every non-dimensional distance from the wall
vðrfÞ pffiffiffiffiffiffiffiffiffiffiffi pffiffiffi
þ V,ðr!
u fÞ ¼ V,ðGf VðfÞÞ þ Sf þ Scr (1) ( yþ ¼ yuT/n) with uT ¼ sW =r ¼ C1=4m ky=n the friction velocity.
vt
!
In Table 2, all the variables corresponding to the general  .  dP n yþ  40
ruT cp T ln TW T  þ 117:31
transport property f are shown. The source term Scr repre- dt uT 0:4767 þ 1=Pr
qW ¼     
sents the effect of the crevice model used [17]. Furthermore, 1 þ 1 1
ln y þ  ln 40 þ þ 10:2384
for the turbulence model the source terms of the turbulent 0:4767 0:4767Pr 0:4767Pr
kinetic energy and its dissipation are the same ones used in (2)
[17], where an extra term SRNG has been added to this version
Further details concerning the development, implementation
of turbulence model [18].
and validation of this heat transfer model used in the in-house
All the constants used in the implementation of the RNG
CFD code can be found in previous works of the authors [23].
ke3 turbulence model, the enthalpy equation and the calcu-
lation of the effective viscosity are the same as in [18]. More-
over, the volumetric heat of combustion, which is inserted in 3.3. Crevice model
the source term of the enthalpy equation Sh, takes also into
account the dissociation of the combustion products. For the simulation of the crevices a quite simpler phenome-
Further details of the in-house CFD model used, concern- nological model than the one found in [26] has been developed
ing the discretization process, the pressure-correction algo- and incorporated into the in-house CFD code. The method-
rithm used, the boundary conditions, the turbulence model, ology followed is in accordance to the one originally presented
12548 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

in [26], but for the calculation of the mass flow rates between formulas, as for example the one found in Ref. [39]. For this
the crevice regions simpler expressions are used. The ring reason, stretch-free expressions of the hydrogen laminar
motion is neglected (axial and twist), as well as the flow flame speed are not taken into consideration here.
through the ring grooves. For the calculation of the mass In the developed combustion model of the in-house CFD
flow rates in every inter-ring region, isentropic compressible code a correlation for the laminar flame speed has been
flow is assumed and an equivalent surface is introduced [27]. incorporated. The authors investigated the features of the
Therefore, the only input required is the number of the piston available ones in the literature, and selected for further
rings and the estimated/approximated dimensions of the investigation the correlations, which have been derived from
crevice regions, which are actually the equivalent discharge measurements under engine-like conditions (wide range of
surface and the volume of every inter-ring region. Fewer equivalence ratio, high pressure and temperature) [40].
dimensions are needed than in the original model presented A brief discussion of the correlations shown in Table 3
in [26], which makes the implementation of the present model follows. Beginning with the correlation of Milton and Keck
easier. Further details concerning the description, imple- [32], this is rarely used in engine simulations since it corre-
mentation and evaluation of this crevice model can be found sponds only to stoichiometric operation. The expression of Liu
in Ref. [17]. and MacFarlane [30] is also not suitable to be used in CFD
simulations, since it corresponds only to ambient pressure
3.4. Combustion model and, as it has been shown in Ref. [8], the magnitude of the
hydrogen burning velocity depends to some extent on the
In the next subsections the combustion model developed by mixture pressure. Therefore, their use will not be opted and
the authors will be presented, as applied for the simulation of the rest three studies shown in Table 3 are further examined
hydrogen-fueled, spark-ignition engines. There are various herein as being more suitable, although the correlation of
sub-models used in this combustion model to describe the Gerke et al. [34] is derived under conditions matching better
laminar flame speed, the ignition process etc., and for each the ones observed in real engines. From these experimental
one of them an analysis follows in order to justify its use. studies, correlations regarding the laminar flame speed are
obtained, which take into account the stretched flame. The
3.4.1. Hydrogen laminar burning velocity numerical formulas of these three correlations can be found
There are numerous experimental investigations concerning in the Appendix.
the hydrogen laminar burning velocity at atmospheric As discussed in Ref. [12], a correlation should adequately
conditions for flame speeds that are stretch-corrected [28,29] calculate the varying effect of all parameters involved in
or not [30,31]. On the other hand, measurements of the engines, namely the equivalence ratio, unburned gas
laminar burning velocities for elevated pressures and temperature, pressure, and residual gas mass fraction. Since
temperatures (relevant to engine conditions) are quite scarce, the correlation proposed by Iijima and Takeno [33] does not
due to issues concerning the flame stability in such conditions take into account the residual gas mass fraction, it has been
[8]. Table 3 shows a representative set of published experi- modified by adding the relevant term suggested by Verhelst
mental studies at high pressures and temperatures regarding [8]. It should be noted that this formula is the only reliable one,
the laminar flame speed [8,30,32e34]. concerning the effect of dilution on the hydrogen laminar
From the various experimental studies shown in Table 3, burning velocity that can be found in the literature [12].
algebraic correlations have been derived, which can be easily The effect of the equivalence ratio, the unburned gas
implemented in an engine simulation code [35]. Another temperature and pressure on hydrogen laminar flame speed is
approach for obtaining a correlation describing the hydrogen shown in Fig. 1aec. These parameters are varied in a range
laminar flame speed is through the use of one-dimensional similar to the one encountered in spark-ignition engines. The
chemical kinetics codes, incorporating one of the many residual gas mass fraction is kept constant and equal to 10%,
reaction mechanisms of hydrogen-air chemistry [36,37] that which is a realistic value for hydrogen-fueled engines in the
can be found in literature [38]. In that case, a stretch-free absence of external EGR.
laminar flame speed is calculated. The authors’ intention is to The correlation of Iijima and Takeno [33] has been derived
use in the expression of the turbulent flame speed (see Section when the flame is in the cellular region, and includes flame
3.4.2) a correlation for the laminar flame speed, which stretch effects, as described in Ref. [5]. Therefore, increased
includes stretch effects. In this way it will be avoided to values for the flame speed are expected [41], as shown in
include such effects externally, with the use of appropriate Fig. 1aec. On the other hand, the correlation of Verhelst [8]

Table 3 e Representative experimental studies concerning the hydrogen laminar flame speed.
Study Equivalence ratio range Temperature range (K) Pressure range

Milton and Keck [32] f¼1 300  Tu  550 0.5  P  7 atm


Liu and MacFarlane [30] 0.23  f  1.9 296  Tu  523 P ¼ 1 atm
Iijima and Takeno [33] 0.5  f  4 291  Tu  500 0.5  P  25 atm
Verhelst [8] 0.3  f  1 300  Tu  430 1  P  10 bar
Gerke et al. [34] 0.4  f  2.5 350  Tu  700 5  P  45 bar
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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0 12549

Correlations
Iijima and Takeno [33] (0.5<φ<4 , 291<Tu<500 K , 0.5<P<25 atm)
Gerke et al. [34] (0.4<φ<2.5 , 350<Tu<700 K , 10<P<45 bar)
16 Verhelst [8] (0.3<φ<1 , 300<Tu<430 K , 1<P<10 bar)
Hydrogen laminar flame speed (m/s)

14
a b c
P = 20 bar Tu = 800 K
12
φ = 0.6 φ = 0.6
Tu = 800 K f = 10% f = 10%
10 P = 20 bar
f = 10%
8

0
0.2 0.4 0.6 0.8 1 400 600 800 1000 1200 5 10 15 20 25 30 35 40
Equivalence ratio Unburned gas temperature (K) Pressure (bar)

Fig. 1 e Effect of (a): equivalence ratio, (b): unburned gas temperature, and (c): pressure, on the hydrogen laminar flame
speed calculated from various correlations under engine-relevant conditions.

predicts lower laminar flame speed as the equivalence ratio In Ref. [8], many turbulent flame speed formulas have been
(Fig. 1a) and pressure (Fig. 1c) varies, while this is also true for implemented in a quasi-dimensional code for the calculation
temperatures higher than approximately 500 K (Fig. 1b). In of the closed part of the engine’s cycle. Among the correla-
Fig. 1c, it is observed that beyond the lower pressure validity tions examined is the one proposed by Zimont/Lipatnikov and
limit of Gerke et al. [34] expression, a quite steep increase of Chomiak [47,48]. The same correlation is also examined in
the laminar flame speed occurs. Ref. [11], where it has been incorporated in a CFD code. In both
The authors decided not to use the correlation of Iijima these works the correlation of Zimont/Lipatnikov and Cho-
and Takeno [33], in the combustion model, since it does not miak is considered adequate to be used, since it is quite simple
take into consideration some recent findings concerning the with few uncertainties in its formulation. More specifically,
hydrogen laminar flame speed, such as the onset of cellu- this turbulent flame speed model has only one calibration
larity [42]. factor in its expression, making its implementation and use in
Concerning the last two correlations, in Section 4.3.1, a CFD code easier and less time consuming for its tuning.
a parametric investigation will take place concerning the Thus, the authors opted to use this correlation in the
simulation of the hydrogen-fueled engine, in order to better hydrogen combustion model developed. The expression
identify which correlation of the hydrogen laminar flame describing the turbulent flame speed is given by Eq. (3).
speed is more suitable to be used within the particular CFD
1=4 1=4
ut ¼ Aðu0 Þ
3=4
code developed. ul1=2 DT;u Lt þ ul (3)

where A is the only calibration constant with reported values


3.4.2. Hydrogen turbulent burning velocity from 0.40 to almost 1 [11,48], u’ the rms turbulent velocity
pffiffiffiffiffiffiffiffiffiffiffi
There are many expressions found in the literature concern- given by u0 ¼ 2k=3, ul the stretched laminar flame speed, DT,u
ing the turbulent burning velocity of a homogeneous mixture the thermal diffusivity of the unburned mixture, and Lt the
[43e48]. In their majority, they are not developed specifically turbulent integral length scale given by Lt ¼ 0:37ðu0 Þ3 =3.
for hydrogen flame propagation studies. Nevertheless, some Fig. 2a shows the dependence of the ratio of the turbulent
general findings of the aforementioned apply to the hydrogen to laminar flame speed (ut/ul) for different equivalence ratios
turbulent burning velocity as well. and turbulent intensities u0 under engine-relevant conditions,
Additionally, in order to decide which expression better including also the effect of the residual gas mass fraction f.
matches the specific needs of simulating hydrogen combus- According to Ref. [53], for near-to-stoichiometric mixtures this
tion in spark-ignition engines, one should have in mind the ratio takes quite low values as also depicted in Fig. 2a. On the
numerous relevant review works found in literature. Some of other hand, for lean or very lean mixtures the ratio of turbu-
those are of general purpose, such as [43,46,49e51], where lent to laminar flame speed increases significantly, and this
various turbulent flame speed correlations are compared with trend is also captured by the formula of the turbulent flame
experimental test cases, in order to identify which ones can speed (Eq. (3)) used in the present study.
qualitatively capture the effect of some crucial parameters, The same correlation concerning the turbulent flame
such as pressure, Lewis number, turbulence level u0 , etc. speed has been applied for different unburned gas tempera-
Specifically for hydrogen, more details can be found in Refs. tures, when varying the turbulent intensity. The calculated
[8,11,52]. results are shown in Fig. 2b. The effect of temperature is less
12550 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

Ratio of turbulent to laminar flame speed (-)


28 u' = 3 m/s a b φ = 0.4 c
u' = 2 m/s φ = 0.6
24 u' = 1 m/s Tu = 800 K
P = 20 bar f = 10%
u' = 0 m/s (ut=ul)
f = 10% A = 0.7 (Eq. (3))
20
Tu = 800 K A = 0.7 (Eq. (3))
16 P = 20 bar
f = 10%
12 A = 0.7 (Eq. (3))

0
0.4 0.6 0.8 1 400 600 800 1000 1200 5 10 15 20 25 30 35 40
Equivalence ratio Unburned gas temperature (K) Pressure (bar)

Fig. 2 e Effect of (a): equivalence ratio, (b): unburned gas temperature, (c): pressure and turbulent intensity, on the ratio of
turbulent to laminar flame speed under engine-relevant conditions.

pronounced than the one of the equivalence ratio. This minor


rt
dependence is due to the variation of the laminar flame speed rtþ1 ¼ dt ut utl þ rtk (5)
k
rb
with temperature (see. Fig. 1b) and of the mixture thermal
diffusivity included in the expression of the turbulent flame where (t þ 1) denotes values corresponding to the current time
speed (see Eq. (3)). step, while (t) to the previous one, and dt is the time step used.
The next parameter investigated is the effect of pressure The initial condition used for rk at the start of ignition (t ¼ tign)
on the turbulent flame speed, which has been pointed out by is equal to 0.5 mm (initial flame kernel diameter equal to
many researchers to play an important role on the absolute d0k ¼ 2r0k ¼ 1 mm [14]), which is approximately equal to the
value of the turbulent flame speed [49,53,54]. In Fig. 2c the spark-plug gaps more often used in spark-ignition engines.
effect of pressure on the ratio of turbulent to laminar flame The flame kernel radius is used, in order to track in which
speed is depicted corresponding to very lean mixtures (4 equal computational cells the propagating flame has reached. More
to 0.4), since for higher equivalence ratios this effect is negli- details about the process of cell tracking are presented in
gible as also reported in Ref. [53]. It is observed that for low Section 3.4.4, where the turbulent flame development is given.
turbulent intensity and elevated pressure the variation of the The end of the ignition process and the transition to the
ratio under investigation is negligible as well, whereas for turbulent combustion is determined by the time instant that
higher turbulence levels it becomes quite important. It is also the flame kernel radius exceeds a specific value (critical
concluded that for low pressure (independent of the turbu- radius). According to the two most widely used methods
lence level), the effect of pressure is much more pronounced, found in the literature dealing with the transition of the
according to the experimental findings in [53]. combustion process from laminar to turbulent mode, this
From the above it is concluded that Eq. (3) describes critical radius is either a constant value [11] or is related to the
adequately the effect of equivalence ratio, unburned gas integral length scale. In the latter case, it is usually taken to be
temperature and pressure, at least qualitatively. equal to two or three times the integral length scale [14,57],
which means that another calibration factor is added in the
3.4.3. Ignition phase simulation code. Nevertheless, the calculated values when
A simple approach has been selected for the simulation of the using both methods are within a small range.
ignition process. It has been shown that during the ignition In the present study, a steady value for the transition to the
period, the flame kernel propagates approximately with the turbulent combustion has been applied, since it seems that
laminar flame speed, since its length is considered to be this method is quite simpler and equally efficient. The value of
smaller than the prevailing turbulence eddies [11,55,56]. The the critical radius used is equal to rk,crit ¼ 4 mm [11]. According
laminar development of the flame kernel during the ignition to a parametric study that has been conducted, using the in-
phase is simulated here using Eq. (4), as proposed in Ref. [11]. house CFD code, it was revealed that by varying the critical
radius, from 3 to 5 mm, no significant influence on the overall
drk ru
¼ ul (4) combustion process, the performance and the heat release
dt rb
rate was observed.
where rk the radius of the flame kernel, ru the unburned gas The ignition energy supplied by the spark-plug was simu-
density, rb the burned gas density, and ul the laminar flame lated by adding energy to the computational cell located at the
speed. point where the spark-plug is placed, according to Refs.
The temporal discretization of Eq. (4) leads to Eq. (5), which is [58,59]. This energy deposition follows a specific pattern con-
used to describe the ignition phase of the developed CFD code. cerning its duration and profile.
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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0 12551

For the present simulation of the CFR engine, which has pressure, temperature, species mass fractions etc.). By doing
a constant low rotational speed (600 rpm), the spark duration so, the flame does not keep its exact initial spherical shape as
is equal to 5 CA (almost 1.5 ms). This value has been obtained the flame radius increases [57]. Nevertheless, as it will be
from similar spark-ignition systems widely used [39,56,60]. shown in a next subsection, the flame shape is close to
The supplied spark energy to the combustible mixture used is a spherical one for each case examined, as also noticed
10 mJ, being a quite low value, since it incorporates also the experimentally in Ref. [62].
heat energy loss to the electrodes; no additional model has Having determined the local flame propagation speed, it is
been implemented to simulate this process, due to the great possible to calculate the burned portion of each cell, defined as
uncertainty of the exact shape, temperature and transient the ratio of burned cell volume to the total cell volume. For
effects near the spark electrodes. Although the value used is a better understanding of the methodology used, Fig. 4 gives
not based on any measured value for the specific engine under schematically the main idea of this sub-grid approach
investigation, a sensitivity analysis that has been conducted implemented in the developed CFD code. The burned ratio of
revealed that the variation of this value does not have any the cells is identified in this figure. To simplify the calculation
significant influence on the engine’s performance, and the of the burned volume of the cell, the flame surface within the
influence on the heat release rate was actually negligible [60]. cell is supposed to be flat. This is not actually true, especially
Various spark power profiles have also been investigated in for cells near the spark-plug, because at those locations the
the present study, with the one finally used shown in Fig. 3, flame kernel radius is low. But for cells far away from the
which is a typical one resembling to some measured ones spark-plug, this is a fair approximation. This burned part of
[60,61]. the cell (corresponding to the calculated burned volume) is
taken into account in the source term of the species conser-
3.4.4. Turbulent flame development phase vation equations for the computational cells that are located
As was discussed in Section 3.4.3, as the initial flame kernel at the flame brush, instead of the whole cell’s volume. This is
expands and its radius reaches a critical value (in this case done, in order to make weaker the effect of grid size in the
equal to 4 mm), the ignition phase terminates and the main calculations and, at the same time, to enable the combustion
phase of flame propagation initiates. Then, the flame speed is model to calculate numerical solutions quite independent of
governed mainly by turbulent phenomena, since the turbu- the total number of computational cells used.
lent flame speed is quite higher than the laminar one. As mentioned earlier, the same methodology for flame
From that point and then, Eqs. (4) and (5) are not applied. tracking is also followed during the ignition phase. In that
Instead, Eq. (6) is used [11] incorporating the turbulent flame case, the calculated radius taking into consideration only the
speed ut, whose correlation was presented in a previous laminar flame speed is used to track in which cells the flame
subsection (see Eq. (3)). has reached.

drk ru
¼ ut (6) 3.4.5. Reaction rate calculation
dt rb
The basic sub-model incorporated in the developed CFD code
The temporal discretization of Eq. (6) leads to Eq. (7). is the characteristic conversion time-scale method. This sub-
model uses a time-scale, which depends on the laminar
rt conversion time and the turbulent mixing time, and dictates
rtþ1 ¼ dt ut utt þ rtk (7)
k
rb
to what extent the combustible gas of each computational cell
When applying Eq. (7), the radius of the flame is calculated in has reached its chemical equilibrium in a predefined time
order to track in which computational cells the flame exists. It step.
should be stressed that for each cell, which is located at the
outer part of the flame brush, a different turbulent flame
speed is calculated according to the local properties (e.g.
Cumulative spark energy (mJ)

10 10
Spark power
Spark energy
8 8
Spark power (W)

6 6

4 4

2 2
Total spark energy: 10 mJ
Spark duration: ~ 5 0CA
0 0
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Time after spark (ms)

Fig. 3 e Spark power and energy profile used during the Fig. 4 e Computational cells with unburned and burned
ignition phase. regions.
12552 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

The methodology followed originates from the work of (at medium load) is equal to 0.054, as also suggested in Refs.
Ref. [63], which has gone through some major improvements [14,58], while for leaner mixtures it is higher and for richer
ever since [14] and has also been extended to other types of lower. The delay coefficient fd found in the expression of the
engines. This combustion modeling methodology has been turbulent mixing time (Eq. (11)) is given by Eq. (12), as reported
successfully implemented in diesel engines [64,65], in HCCI in [14].
engines [66,67], as well as in spark-ignition engines [14,58].

Next, the combustion model is presented as was implemented fd ¼ 1  eðtts Þ=sd (12)
in the developed CFD code. where (t  ts) is the time after the spark, and sd ¼ Cm1Lt/ul is the
The species considered for the combustion of hydrogen in time required by the laminar flame to propagate by two to
a spark-ignition engine are: H2, O2, N2, H2O, H, O, N, OH, NO. At three times the turbulent integral length scale. The estimated
each time step, the reaction rates of each of these nine species value of the required times is expressed through the Cm1
is calculated from Eq. (8), similar to Ref. [63], except for the coefficient, and for the current study it is equal to a constant
nitric oxide one, which is kinetically controlled in engine value of 2.5 for all cases examined, as proposed in Ref. [63]. A
flows; the calculation of the latter is shown in Section 3.4.6. parametric investigation has revealed that by increasing the
 tþ1  t Cm1 coefficient, the timing of the peak pressure is advanced,
dYi ðYi Þt  Yi
¼ (8) while it has a weaker effect on the magnitude of the peak
dt sc
pressure. Nevertheless, no tuning of this coefficient has been
where (Yi)t is the species mass fraction at the previous time attempted here.
step, (Yi*)t is the species mass fraction corresponding to the
chemical equilibrium value at the previous time step, sc is the 3.4.6. NO formation modeling
characteristic conversion time, and (dYi/dt)tþ1 is the reaction In Section 3.4.5, it was discussed how the characteristic
rate of each species at the current time step, which is inserted conversion time needed for a species to reach its chemical
into the source term of the species transport equations for the equilibrium state is calculated. The aforementioned proce-
cells in which the flame has propagated. dure is followed for all species considered, except for the nitric
The explicit version of Eq. (8) has been chosen in order to oxide, for which it has been shown that its reaction rate is
calculate at the beginning of every time step the reaction kinetically controlled, according to three kinetic reactions,
rates, which are then held constant for the calculations during known as the extended Zeldovich mechanism [69].
this time step. Therefore, the computational time step during The methodology followed is the same as the one pre-
the combustion process should be substantially decreased, in sented in detail in Ref. [70]. The most usual set of constants for
order for the aforementioned methodology to produce reliable the reaction rates of the three chemical equations considered
results as is also the case followed here. is also used in the present study, which is the one provided by
The species concentrations corresponding to chemical Ref. [69]. Nevertheless, these constants were derived for
equilibrium Yi* are calculated using a developed subroutine hydrocarbon fuels, and their uncertainty (which is already
[68]. This subroutine has been greatly adjusted and partially quite high) increases when they are applied for the calculation
reformulated for the specific needs of hydrogen combustion of the nitric oxide reaction rates in the case of a fuel being
calculations, by omitting all species involved containing carbon-free [12]. The chemical reactions taken into consider-
carbon atoms, such as CO2 and CO. Its input is the local values ation for the calculation of the nitric oxide can be seen in
of pressure, temperature and species concentrations, while its Table 4, together with the constants used.
output is the chemical equilibrium values of all species
considered.
The characteristic conversion time sc, which is actually the 4. Results and discussion
time needed for the burning mixture to reach its chemical
equilibrium state, is the same for all nine species considered 4.1. Test cases examined
in this study; it is the sum of a laminar kinetics time sl and
a turbulent mixing time st. This characteristic time is calcu- The test cases examined in the current study are shown in
lated from Eq. (9), according to Ref. [63]. Table 5. Throughout the Cases 1e3, the equivalence ratio is
sc ¼ sl þ st (9) varied in order to evaluate the combustion model. On the

The laminar kinetics time is given by Eq. (10).

DT;u
sl ¼ (10) Table 4 e Chemical reactions and constants used for NO
u2l calculations [69].
while the turbulent one is given by Eq. (11). Chemical reactions Forward reaction rates:
kfi ¼ Afi Tbi eðEi =RTÞ ; i ¼ 1  3
f f

k
st ¼ cm;f fd (11) Af bf Ef
3
(mol m3 s1) (e) (J/mol)
The coefficient cm,f shown in Eq. (11) is a function of the local
1. N þ NO 4 N2 þ O 3.3  106 0 0
equivalence ratio, as is also explained in Ref. [14], in order to
2. N þ O2 4 NO þ O 6.4  103 1 26.2  103
capture better the combustion effects of variable mixture
3. N þ OH 4 NO þ H 4.1  107 0 0
compositions. In the present study, the reference value of cm,f
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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0 12553

Table 5 e Test cases examined in the present study. 60 Measured


Calculated (Correlation: [8])
Case Equivalence ratio (4) Compression ratio Calculated (Correlation: [34])

Cylinder pressure (bar)


50
1 0.71 8
CR=9.5
2 0.59 8
40 φ=0.59
3 0.50 8
IT=15 0CA BTDC
4 0.59 9.5
30

other hand, Case 4 (with higher compression ratio) is used to


20
comparatively evaluate the two alternative laminar flame
speed correlations [8,34], implemented into the combustion
10
model. All cases correspond to constant engine speed (equal
to 600 rpm) as well as ignition timing (equal to 15 CA BTDC).
0
120 150 180 210 240
4.2. Computational details Crank angle degrees (ABDC)

In the present study some initial and boundary conditions are Fig. 5 e Comparison of the calculated with the measured
provided from the experimental measurements, such as the pressure trace, using two laminar flame speed correlations
pressure at IVC, and the inlet air and hydrogen flow rates. For [8,34].
the rest of the required conditions, reliable estimations are
used. The wall temperature values are held constant
throughout the calculations and their values are estimated
a parametric study using the firing version of the developed
according to a recent work in the same CFR engine [21]. The
CFD code, have shown that the most reliable mesh, providing
initial flow field is calculated by considering an initial swirl
adequate accuracy and keeping the required CPU time relative
ratio equal to 1.5, since the inlet valve is unshrouded [66]. The
low, is the one having (40  40  40) grid lines along each axis
initial turbulence properties are the same as the ones derived
respectively. The methodology for adding or removing mesh
from a relevant work concerning the same engine [66], and are
layers is the same to the one described in Ref. [23].
held constant for every case examined here, since the engine
The computational time step used is kept low and equal to
speed is kept constant as well. More specifically, the initial u0
0.5 CA, since it has been shown that it is small enough to
at IVC is equal to 5.16 m/s, giving a rms turbulent velocity
provide time-step independent solutions [23], even for low
equal to around 1 m/s at TDC, although this value slightly
rotational speeds. During the combustion period, initiating
changes with engine load (from 0.9 m/s for the lower load to
from the ignition timing and afterwards, the time step is
1.2 m/s for the higher load). The initial gas temperature at IVC
further decreased and becomes equal to 0.1 CA, as in Ref. [11].
is held constant for all cases examined and equal to 370 K [8].
This occurs in order to increase the accuracy of the compu-
The estimation of the residual gas mass fraction plays an
tational results, but also to increase the stability of the code,
important role in the value of the laminar flame speed, as
since during each time step the variation of the cylinder
described in the correlations found in the Appendix. At IVC the
pressure and temperature is large.
initial mixture is composed by H2, O2, N2, and H2O, and the mass
content of each of these species is affected by the residual gas
mass fraction. Estimations concerning the residual gas and its
90 Experimental
trend are obtained from the works reported in Refs. [71,72]. No
Calculated (Correlation: [8])
overlapping period of the inlet and exhaust valve exists in the 80
Gross heat release rate (J/0CA)

Calculated (Correlation: [34])


specific engine, making the estimation of the residual gas quite 70
simpler. The exact value of the residual gas mass fraction, wall
60 CR=9.5
temperature and initial pressure, at each case examined, is
φ=0.59
shown in Table 6. 50 IT=15 oCA BTDC
Concerning the computational mesh used, previous rele- 40
vant studies of the present authors [17,23], together with
30

20

10
Table 6 e Initial and boundary conditions for the four
cases examined. 0

Case Wall Initial pressure Residual gas -10


temperature (K) at IVC (bar) mass fraction (%) 160 170 180 190 200
Crank angle degrees (ABDC)
1 430 1.079 12.04
2 425 1.054 9.25
Fig. 6 e Comparison of the calculated gross heat release
3 415 1.050 6.7
rates, using two laminar flame speed correlations, with the
4 430 1.059 7.5
experimental one.
12554 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

Fig. 6 shows a comparison of the predicted gross heat


Table 7 e Calculated NO emissions and gross indicated
work using the two correlations of the laminar flame release rates using the two alternative laminar flame speed
speed, and their comparison with the measured values. correlations with the corresponding experimental one. The
calculation of the experimental and numerical gross heat
NO emissions Gross indicated
(ppm) work (J) release rate is based on the cylinder pressure traces, using
a single-zone approach [21]. The CFD code using both corre-
Measured 2273 308.8
lations predicted similar gross heat release rates, although
Calculated (correlation: [34]) 2356 334.2
the calibration constant has very different values. In general,
Calculated (correlation: [8]) 2430 339.7
the predicted gross heat release rate is close to the experi-
mental one (qualitatively), while the timing of the peak gross
heat release is slightly advanced in comparison to the
4.3. Evaluation of the combustion model with measured one. Also, the calculated combustion duration is
comparison to experimental data well captured with both correlations used, although it is
slightly overestimated (approximately by 3 CA). The disad-
Prior to the evaluation of the combustion model, a compara- vantage observed when using the Verhelst correlation, is that
tive analysis of the two alternative laminar flame speed the combustion rate during the ignition process is greatly
correlations described previously is made. This is done in underestimated, increasing afterwards significantly during
order to decide which of these two correlations is more suit- the main flame propagation period, until approximately
able for our combustion model. 175 CA ABDC, due to the high value of the tuning constant
used. The change of the slope of the calculated gross heat
4.3.1. Investigation of laminar flame speed correlations release rate during the expansion stroke will be explained in
The two alternative laminar flame speed correlations Section 4.3.2.
considered are the ones proposed in Refs. [8,34]. To compar- Moreover, Table 7 shows the exhaust NO emissions and
atively evaluate these two expressions, the CFR engine is the gross indicated work when using the two correlations, and
examined with an increased compression ratio (Case 4), since their comparison with the measured values. The gross indi-
in that case the pressure and temperature reach higher values cated work is calculated during the closed part of the cycle
(compared to the rest of the available cases) and the demand (from the IVC until the EVO) and is given by Eq. (13).
for a reliable correlation is more intense. Fig. 5 shows the
predicted pressure traces using the two correlations, and their ZEVO
comparison with the measured one. Wgi ðJÞ ¼ pdV (13)
When using the Verhelst correlation, the calibration IVC

constant of the turbulent flame speed correlation (Eq. (3)) is where V the instantaneous cylinder volume.
set equal to 1.6, which is substantially higher than the corre- The calculated nitric oxide emissions using both correla-
sponding value when the Gerke et al. correlation is used (equal tions give a value very close to the measured one (discrepancy
to 0.8), in order to better match the measured in-cylinder smaller than 7%). On the other hand, using both correlations,
pressure trace. The calibration constant value used in the the calculated gross indicated work is higher than the
Verhelst correlation is also much higher than the reported measured one, since during the expansion stroke the cylinder
values from other researchers, which are held below unity. pressure is slightly over-predicted in comparison to the
Nevertheless, the calculated pressure history is well captured measured one (see Fig. 5). This may be attributed to the lower
using both correlations. heat loss calculated during this time period [23]. Nevertheless,

50
Measured
Cylinder pressure (bar)

Calculated
40

30
φ=0.71 φ=0.59 φ=0.50

20

10

0
120 150 180 210 240 120 150 180 210 240 120 150 180 210 240
Crank angle degrees (ABDC)
Fig. 7 e Comparison of calculated pressure traces with the measured ones at various engine loads (CR [ 8, IT [ 15 CA
BTDC).
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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0 12555

100
Experimental

Gross heat release rate (J/oCA)


Calculated
80

60 φ=0.71 φ=0.59 φ=0.50

40

20

160 170 180 190 200 160 170 180 190 200 160 170 180 190 200
Crank angle degrees (ABDC)

Fig. 8 e Comparison of calculated gross heat release rates with the experimental ones for three equivalence ratios (CR [ 8,
IT [ 15 CA BTDC).

this discrepancy is lower than 10%. It is to be understood in decreased rapidly, while the combustion duration is
this work that the calculated values concern nitric oxide (NO), increased. The latter is expected, owing to the decreased
while the available measured ones nitrogen oxides (NOx); hydrogen turbulent combustion velocity for lean mixtures.
however, as known, their difference at a spark-ignition engine The change of the slope of the calculated gross heat release
exhaust is extremely small. rate, during the expansion stroke, may be attributed to the
From the former extended evaluation, it has been revealed retarded combustion of the outer computational cells adja-
that both correlations can be used for reliable engine simu- cent to the cylinder liner. It has been observed that the
lations using the developed CFD code. Nevertheless, the use of propagating flame reaches these regions just before the end of
the correlation of Gerke et al. [34] is preferred for the combustion, where the prevailing unburned gas temperature
combustion model, since it is derived from conditions more and turbulence level is low. Thus, combustion of these specific
relevant to real engine operation, the calibration factor of the cells is delayed, leading to rapid ignition and combustion of all
turbulent flame speed formula is within the reported values these cells later on when the local conditions are proper. This
[11], and the ignition process is simulated better as shown causes an increase of the heat release for a few degrees CA.
previously. Nevertheless, further investigation is needed to understand
better the reason for the change of slope in the gross heat
4.3.2. Simulation of engine performance and emissions release rate.
In the current subsection, Cases 1e3 of Table 5 will be elabo- This comparison reveals the capability of the CFD code not
rated using the Gerke et al. correlation for the laminar burning only to capture the effect of equivalence ratio on the cylinder
velocity [34]. Fig. 7 shows the comparison of the calculated pressure history of a spark-ignition engine, but also to simu-
with the measured cylinder pressure traces at various engine late reliably the combustion processes taking place.
loads.
As shown, the peak pressure is well captured in all cases 6000
examined, as well as its timing, with the only exception for the Measured
richer mixture case (4 ¼ 0.71) where this timing is estimated Calculated
5000
almost 3 CA later. It should be stressed that the model
NO emissions (ppm)

manages to predict adequately the absolute value of the


4000
cylinder pressure at the IT, in all cases examined, without any CR=8
constant tuning. This is crucial for the proper estimation of IT=15 0CA BTDC
the cylinder pressure during the combustion phase, and is
3000
partly attributed to the further development of the CFD code
by including a crevice model and a detailed wall-function 2000
formulation [17,23]. An aspect that needs further investigation
is the slight pressure over-prediction during the expansion 1000
stroke, which becomes more pronounced for the richer
mixtures and, to some extent, can be attributed to the rela- 0
tively low wall heat flux observed during the expansion 0.45 0.5 0.55 0.6 0.65 0.7 0.75
stroke, as also described in Ref. [23].
Equivalence ratio
In Fig. 8, the calculated gross heat release rate is compared
with the experimental one for Cases 1e3. As the cylinder Fig. 9 e Calculated NO emissions and comparison with
charge becomes leaner, the peak gross heat release is measurements, when varying the equivalence ratio.
12556 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

Fig. 10 e Flame images for the three equivalence ratios examined (CR [ 8, IT [ 15 CA BTDC).

Fig. 9 shows the comparison of the calculated nitric oxide results would be more accurate, if a specially derived set of
emissions with the measured tailpipe ones for all cases exam- constants for carbon-free fuels were used for the Zeldovich
ined. As shown, the effect of varying the equivalence ratio is well mechanism, according to Ref. [12].
captured by the simulation model, although in absolute values The flame development images are shown in Fig. 10,
there is a small discrepancy, approximately 26% at high load and depicting how the flame propagates inside the cylinder after
50% at low load. However, in the latter case, the absolute value of the ignition timing for each case examined. It is interesting to
the measured nitric oxide emissions is very low. The calculated notice the difference of the flame development for the various
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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0 12557

Fig. 11 e NO mass fraction spatial distribution during the combustion period for the three equivalence ratios (Cases 1e3).

equivalence ratios. It should be noted that the spark-plug is plane crosses the spark-plug). The contour values are given as
offset from the cylinder axis, as one can guess from the mass fractions (kgNO/kgmixture) and their range is different for
location of the ignition kernel. every case, in order to better visualize the emissions produc-
Owing to the high burning velocity of hydrogen, the flame tion process.
has reached the liner for the rich mixture case (Case 1) at As shown in Fig. 11, as the equivalence ratio is increased
around TDC, while for the leaner cases (Cases 2 and 3) the NO emissions are radically increased. In all cases examined,
flame at that moment has also developed significantly. during the first stages of combustion, NO emissions are
Moreover, the flame has covered the whole combustion produced near the spark-plug where the local temperature is
chamber at around TDC for the richer mixture case (Case 1), at also higher [70]. However, as combustion continues, the core
10 CA ATDC for the mixture of Case 2, and at 15 CA ATDC for of the high NO concentration moves towards the center of the
the leaner mixture of Case 3. cylinder, primarily due to the cooling of the outer regions and
Fig. 11 shows the spatial distribution of the NO mass secondarily to the prevailing velocity field. This is an inter-
fraction inside the cylinder, at various time instants, for all esting finding revealing the importance of the proposed CFD
equivalence ratios examined at three planes (the top-right code, since in the simple quasi-dimensional or multi-zone
12558 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

models the total heat losses are equally distributed to the in- presented in the current work, can yield reliable results,
cylinder zones, irrespective of their location (close or far away without the need for recourse to a large set of calibration
from the cylinder walls), and also no convection phenomena constants. However, a more extended investigation should be
are considered there. conducted in the future covering a wider range of operating
conditions, to justify the validity of the model and its ability to
capture combustion phenomena without adjusting the value
of the calibration constant.
5. Summary and conclusions

In the present work a combustion model has been developed,


which has been incorporated in a CFD code for the simulation Acknowledgements
of a hydrogen-fueled spark-ignition engine. This combustion
model takes into consideration some recent findings con- The authors wish to thank Prof. S. Verhelst (Univ. of Gent,
cerning the hydrogen laminar flame speed, and it uses Belgium) for the provision of the experimental data. His
a robust turbulent flame speed correlation in order to track the contribution is greatly acknowledged. Also, G.M. Kosmadakis
flame-containing computational cells. In these cells the wishes to thank the Greek State Scholarships Foundation for
characteristic conversion time-scale method is applied, in granting him a Ph.D. research scholarship.
order to calculate the combustion rate. By doing so, the flame
propagation is obtained within the engine cylinder in order to
calculate the performance and NO emissions of the engine. Appendix.
The study showed that both hydrogen laminar flame speed
correlations investigated in detail, proposed by Verhelst [8] The correlations for the hydrogen laminar flame speed
and Gerke et al. [34], could be incorporated in the CFD code mentioned in the present study are given here. In every
for the simulation of the hydrogen engine, giving reasonably correlation the effect of the residual gas mass fraction has
acceptable results. However, the use of the latter correlation been added, according to Verhelst [8], where f is the residual
has been preferred, since it was based on a dataset in a wider gas mass fraction.
pressure and temperature range closer to real engine oper-
ating conditions. Also, using this correlation, the ignition 1) Iijima and Takeno correlation [33], valid for 0.5  f  4,
process is better captured. 291  Tu  500 K, 0.5  P  25 atm:
The combustion model has been applied to simulate the
power cycle of the same engine at various engine loads. The h i a
ul ¼ ul0 1 þ b log PP0 Tu
T0
ð1  gf Þ
calculated results (regarding cylinder pressure traces, gross
heat release rates, and exhaust nitric oxide emissions) have ul0 ¼ 2:98  ðf  1:7Þ2 þ0:32ðf  1:7Þ3
a ¼ 1:54 þ 0:026ðf  1Þ
been compared with the measured ones showing a good
b ¼ 0:43 þ 0:003ðf  1Þ
match. It should be noted that throughout this investigation, g ¼ 2:715  0:5f
the only calibration constant used in the model has preserved P0 ¼ 1 atm; T0 ¼ 291 K
its constancy of value.
However, some aspects need further investigation as
2) Verhelst correlation [8], valid for 0.3  f  1, 300  Tu 
follows:
430 K, 1  P  10 bar:

 The set of constants used for the calculation of the nitric


oxide emissions. a b
ul ¼ ul0 Tu
T0
P
P0
ð1  gf Þ
 The underestimated gross heat release rate at the beginning
ul0 ¼ 4:77f þ 8:65f2  0:394f  0:296
3
of combustion, and the change of the slope of the gross heat
a¼ 1:232
release rate diagram near the end of combustion. 2:9f3  6:69f2 þ 5:06f  1:16 f < 0:6

 The slightly overestimated cylinder pressure trace during 0:0246f þ 0:0781 f  0:6
the expansion stroke. g ¼ 2:715  0:5f
P0 ¼ 5 bar; T0 ¼ 365 K

In any case, the results obtained so far are encouraging, by


showing that the simulation of hydrogen-fueled engines using 3) Gerke et al. correlation [34], valid for 0.4  f  2.5, 350 
a detailed simulation numerical tool, such as the one Tu  700 K, 10  P  45 bar:

a b
ul ¼ ul0 TTu0 P
ð1  gf Þ
P0
0:25f  3:4774f5 þ 18:498f4  46:525f3 þ 52:317f2  13:976f þ 1:2994
6
0:4  f  2:5
ul0 ¼
.  1:272384ðf  2:5Þ f > 2:5
10:41096
a ¼ 0:0163 1 f þ 2:2937
.
b ¼ 0:2037 1 f  0:575
g ¼ 2:715  0:5f
P0 ¼ 20 bar; T0 ¼ 600 K
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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0 12559

references [19] Rakopoulos CD, Kosmadakis GM, Pariotis EG. Evaluation of


a new computational fluid dynamics model for internal
combustion engines using hydrogen under motoring
conditions. Energy 2009;34(12):2158e66.
[1] Agarwal AK. Biofuels (alcohols and biodiesel) applications as
[20] Rakopoulos CD, Kosmadakis GM, Pariotis EG. Investigation of
fuels for internal combustion engines. Prog Energy Combust
piston bowl geometry and speed effects in a motored HSDI
Sci 2007;33(3):233e71.
diesel engine using a CFD against a quasi-dimensional
[2] Rakopoulos CD, Michos CN. Generation of combustion
model. Energy Convers Manage 2010;51(3):470e84.
irreversibilities in a spark ignition engine under
[21] Demuynck J, Raes N, Zuliani M, De Paepe M, Sierens R,
biogasehydrogen mixtures fueling. Int J Hydrogen Energy
Verhelst S. Local heat flux measurements in a hydrogen and
2009;34(10):4422e37.
methane spark ignition engine with a thermopile sensor. Int
[3] Rakopoulos CD, Rakopoulos DC, Hountalas DT,
J Hydrogen Energy 2009;34(24):9857e68.
Giakoumis EG, Andritsakis EC. Performance and emissions of
[22] Shudo T, Nabetani S. Analysis of degree of constant volume
bus engine using blends of diesel fuel with bio-diesel of
and cooling loss in a hydrogen fuelled SI engine. Trans SAE J
sunflower or cottonseed oils derived from Greek feedstock.
Fuels Lubricants 2001;110:1911e7 [SAE Paper no. 2001-01-3561].
Fuel 2008;87(2):147e57.
[23] Rakopoulos CD, Kosmadakis GM, Pariotis EG. Critical
[4] Tsolakis A, Megaritis A, Yap D. Application of exhaust gas
evaluation of current heat transfer models used in CFD in-
fuel reforming in diesel and homogeneous charge
cylinder engine simulations and establishment of
compression ignition (HCCI) engines fuelled with biofuels.
a comprehensive wall-function formulation. Appl Energy
Energy 2008;33(3):462e70.
2010;87(5):1612e30.
[5] Verhelst S, Wallner T. Hydrogen-fueled internal combustion
[24] Launder BE, Spalding DB. The numerical computation of
engines. Prog Energy Combust Sci 2009;35(6):490e527.
turbulent flows. Comput Methods Appl Mech Eng 1974;3(2):
[6] Dimopoulos P, Bach C, Soltic P, Boulouchos K.
269e89.
Hydrogenenatural gas blends fuelling passenger car engines:
[25] Han Z, Reitz RD. A temperature wall function formulation for
combustion, emissions and well-to-wheels assessment. Int J
variable-density turbulent flows with application to engine
Hydrogen Energy 2008;33(23):7224e36.
convective heat transfer modeling. Int J Heat Mass Transfer
[7] Kirchweger W, Haslacher R, Hallmannsegger M, Gerke U.
1997;40(3):613e25.
Applications of the LIF method for the diagnostics of the
[26] Namazian M, Heywood JB. Flow in the pistonecylinderering
combustion process of gas-IC-engines. Exp Fluids 2007;43
crevices of a spark-ignition engine: effect on hydrocarbon
(2e3):329e40.
emissions, efficiency and power. Trans SAE (Section 1) 1982;
[8] Verhelst S. A study of the combustion in hydrogen-fuelled
91:261e88 [SAE Paper no. 820088].
internal combustion engines. Ph.D. thesis, Ghent University,
[27] Keribar R, Dursunkaya Z, Flemming MF. An integrated model
Ghent, Belgium; 2005.
of ring pack performance. Trans ASME J Eng Gas Turbines
[9] Rottengruber H, Berckmueller M, Elsaesser G, Brehm N,
Power 1991;113:382e9.
Schwarz C. Direct-injection hydrogen SI-engine operation
[28] Vagelopoulos CM, Egolfopoulos FN, Law CK. Further
strategy and power density potentials. Trans SAE J Fuels
considerations on the determination of laminar flame
Lubricants 2004;113:1749e61 [SAE Paper no. 2004-01-2927].
speeds with the counterflow twin-flame technique. Symp
[10] Dimopoulos P, Rechsteiner C, Soltic P, Laemmle C,
(Int) Combust [Proc] 1994;25(1):1341e7.
Boulouchos K. Increase of passenger car engine efficiency
[29] Kwon OC, Faeth GM. Flame/stretch interactions of premixed
with low engine-out emissions using hydrogenenatural gas
hydrogen-fueled flames: measurements and predictions.
mixtures: a thermodynamic analysis. Int J Hydrogen Energy
Combust Flame 2001;124(4):590e610.
2007;32(14):3073e83.
[30] Liu DDS, MacFarlane R. Laminar burning velocities of
[11] Gerke U. Numerical analysis of mixture formation and
hydrogeneair and hydrogeneairesteam flames. Combust
combustion in a hydrogen direct-injection internal
Flame 1983;49(1e3):59e71.
combustion engine. Ph.D. thesis, Diss. ETH no. 17477,
[31] Koroll GW, Kumar RK, Bowles EM. Burning velocities of
Cuvillier Göttingen, ETH Zurich, Switzerland; 2007.
hydrogeneair mixtures. Combust Flame 1993;94(3):330e40.
[12] Knop V, Benkenida A, Jay S, Colin O. Modelling of combustion
[32] Milton B, Keck J. Laminar burning velocities in stoichiometric
and nitrogen oxide formation in hydrogen-fuelled internal
hydrogen and hydrogenehydrocarbon gas mixtures.
combustion engines within a 3D CFD code. Int J Hydrogen
Combust Flame 1984;58(1):13e22.
Energy 2008;33(19):5083e97.
[33] Iijima T, Takeno T. Effects of temperature and pressure on
[13] Verhelst S, Sierens R. A quasi-dimensional model for the
burning velocity. Combust Flame 1986;65(1):35e43.
power cycle of a hydrogen-fuelled ICE. Int J Hydrogen Energy
[34] Gerke U, Steurs K, Rebecchi P, Boulouchos K. Derivation of
2007;32(15):3545e54.
burning velocities of premixed hydrogen/air flames at
[14] Fan L, Reitz RD. Development of an ignition and combustion
engine-relevant conditions using a single-cylinder
model for spark-ignition engines. Trans SAE J Engines 2000;
compression machine with optical access. Int J Hydrogen
109:1977e89 [SAE Paper no. 2000-01-2809].
Energy 2010;35(6):2566e77.
[15] Thakur S, Wright JR. A multiblock operator-splitting
[35] Verhelst S, Woolley R, Lawes M, Sierens R. Laminar and
algorithm for unsteady flows at all speeds in complex
unstable burning velocities and Markstein lengths of
geometries. Int J Numer Methods Fluids 2004;46(4):383e413.
hydrogeneair mixtures at engine-like conditions. Proc
[16] Shyy W. Computational modeling for fluid flow and
Combust Inst 2005;30(1):209e16.
interfacial transport. Amsterdam: Elsevier Science; 1994.
[36] Marinov NM, Curran HJ, Pitz WJ, Westbrook CK. Chemical
[17] Rakopoulos CD, Kosmadakis GM, Dimaratos AM, Pariotis EG.
kinetic modeling of hydrogen under conditions found in
Investigating the effect of crevice flow on internal
internal combustion engines. Energy Fuels 1998;12:78e82.
combustion engines using a new simple crevice model
[37] O’Conaire M, Curran HJ, Simmie JM, Pitz WJ, Westbrook CK. A
implemented in a CFD code. Appl Energy; 2010;. doi:10.1016/j.
comprehensive modeling study of hydrogen oxidation. Int J
apenergy.2010.07.012.
Chem Kinetics 2004;36:603e22.
[18] Han Z, Reitz RD. Turbulence modeling of internal
[38] Konnov AA. Remaining uncertainties in the kinetic mechanism
combustion engines using RNG ke3 models. Combust Sci
of hydrogen combustion. Combust Flame 2008;152(4):507e28.
Technol 1995;106(4):267e95.
12560 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 5 ( 2 0 1 0 ) 1 2 5 4 5 e1 2 5 6 0

[39] Herweg R, Maly RR. A fundamental model for flame kernel [57] Tan Z, Reitz RD. An ignition and combustion model based
formation in S.I engines. SAE paper no. 922243; 1992. on the level-set method for spark-ignition engine
[40] Hu E, Huang Z, He J, Miao H. Experimental and numerical multidimensional modeling. Combust Flame 2006;145(1e2):
study on laminar burning velocities and flame instabilities of 1e15.
hydrogeneair mixtures at elevated pressures and [58] Kuo T-W. Multidimensional port-and-cylinder gas flow, fuel
temperatures. Int J Hydrogen Energy 2009;34(20):8741e55. spray, and combustion calculations for a port-fuel-injection
[41] Chen JH, Im HG. Stretch effects on the burning velocity of engine. SAE paper no. 920515; 1992.
turbulent premixed hydrogen/air flames. Proc Combust Inst [59] Yossefi D, Maskell SJ, Ashcroft SJ, Belmont MR. Ignition
2000;28(1):211e8. source characteristics for natural-gas-burning vehicle
[42] Bradley D, Lawes M, Liu K, Verhelst S, Woolley R. Laminar engines. Proc Instn Mech Engrs Part D J Automob Eng 2000;
burning velocities of lean hydrogeneair mixtures at 214(2):171e80.
pressures up to 1.0 MPa. Combust Flame 2007;149(1e2): [60] Arcoumanis C, Bae C- S. Correlation between spark ignition
162e72. characteristics and flame development in a constant-volume
[43] Duclos JM, Veynante D, Poinsot T. A comparison of flamelet combustion chamber. Trans SAE J Engines 1992;101:556e70
models for premixed turbulent combustion. Combust Flame [SAE Paper no. 920413].
1993;95(1e2):101e17. [61] Aleiferis PG, Taylor AMKP, Ishii K, Urata Y. The relative
[44] Weller HG, Uslu S, Gosman AD, Maly RR, Herweg R, Heel B. effects of fuel concentration, residual-gas fraction, gas
Prediction of combustion in homogeneous-charge spark- motion, spark energy and heat losses to the electrodes on
ignition engines. In: Proceedings of 3rd international flame-kernel development in a lean-burn spark ignition
symposium on “Diagnostics and modelling of combustion in engine. Proc Inst Mech Engrs Part D J Automob Eng 2004;218
internal combustion engines (COMODIA 1994)”, Yokohama, (4):411e25.
Japan; July 1994. p. 163e9. [62] Heywood JB, Vilchis FR. Comparison of flame development in
[45] Peters N. The turbulent burning velocity for large-scale and a spark-ignition engine fueled with propane and hydrogen.
small-scale turbulence. J Fluid Mechanics 1999;384(1): Combust Sci Technol 1984;38(5):313e24.
107e32. [63] Abraham J, Bracco FV, Reitz RD. Comparison of computed
[46] Guelder OL. Turbulent premixed flame propagation models and measured premixed charge engine combustion.
for different combustion regimes. Symp (Int) Combust [Proc] Combust Flame 1985;60(3):309e22.
1990;23(1):743e50. [64] Kong S-C, Han Z, Reitz RD. The development and application
[47] Zimont VL. Gas premixed combustion at high turbulence. of a diesel ignition and combustion model for
Turbulent flame closure combustion model. Exp Therm Fluid multidimensional engine simulations. Trans SAE J Engines
Sci 2000;21(1e3):179e86. 1995;104:502e18 [SAE Paper no. 950278].
[48] Lipatnikov AN, Chomiak J. A simple model of unsteady [65] Masood M, Ishrat MM, Reddy AS. Computational combustion
turbulent flame propagation. Trans SAE J Engines 1997;106: and emission analysis of hydrogen-diesel blends with
2441e52 [SAE Paper no. 972993]. experimental verification. Int J Hydrogen Energy 2007;32(13):
[49] Lipatnikov AN, Chomiak J. Modeling of pressure and non- 2539e47.
stationary effects in spark ignition engine combustion: A [66] Kong S-C, Ayoub N, Reitz RD. Modeling combustion in
comparison of different approaches. Trans SAE J Engines compression ignition homogeneous charge engines. Trans
2000;109:1833e50 [SAE Paper no. 2000-01-2034]. SAE J Engines 1992;101:896e911 [SAE Paper no. 920512].
[50] Dinkelacker F, Hoelzler S. Investigation of a turbulent flame [67] Kim S, Ito K, Yoshihara D, Wakisaka T. Application of
speed closure approach for premixed flame calculations. a genetic algorithm to the optimization of rate constants in
Combust Sci Technol 2000;158(1):321e40. chemical reaction submodels for engine combustion
[51] Lipatnikov AN. Testing premixed turbulent combustion simulation. In: Proceedings of 6th international symposium
models by studying flame dynamics. Int J Spray Combust on “Diagnostics and modelling of combustion in internal
Dynamics 2009;1(1):39e66. combustion engines (COMODIA 2004)”, Yokohama, Japan,
[52] Abdel-Gayed RG, Bradley D. Dependence of turbulent August 2004, pp. 43e51.
burning velocity on turbulent Reynolds number and ratio of [68] Rakopoulos CD, Hountalas DT, Tzanos EI, Taklis GN. A fast
laminar burning velocity to r.m.s. turbulent velocity. Symp algorithm for calculating the composition of diesel
(Int) Combust [Proc] 1976;16:1725e35. combustion products using 11 species chemical equilibrium
[53] Kitagawa T, Nakahara T, Maruyama K, Kado K, Hayakawa A, scheme. Adv Eng Software 1994;19(2):109e19.
Kobayashi S. Turbulent burning velocity of hydrogen-air [69] Lavoie GA, Heywood JB, Keck JC. Experimental and
premixed propagating flames at elevated pressures. Int J theoretical study of nitric oxide formation in internal
Hydrogen Energy 2008;33(20):5842e9. combustion engines. Combust Sci Technol 1970;1(4):313e26.
[54] Brandl A, Pfitzner M, Mooney JD, Durst B, Kern W. [70] Rakopoulos CD, Michos CN, Giakoumis EG. Thermodynamic
Comparison of combustion models and assessment of their analysis of SI engine operation on variable composition
applicability to the simulation of premixed turbulent biogasehydrogen blends using a quasi-dimensional, multi-
combustion in IC-engines. Flow Turbul Combust 2005;75 zone combustion model. SAE Int J Engines 2009;2(1):880e910
(1e4):335e50. [SAE Paper no. 2009-01-0931].
[55] Willems H, Sierens R. Modeling the initial growth of the [71] Fox JW, Cheng WK, Heywood JB. A model for predicting
plasma and flame kernel in SI engines. Trans ASME J Eng Gas residual gas fraction in spark-ignition engines. Trans SAE J
Turb Power 2003;125(2):479e84. Engines 1993;102:1538e44 [SAE Paper no. 931025].
[56] Pischinger S, Heywood JB. A model for flame kernel [72] Senecal PK, Xin J, Reitz RD. Predictions of residual gas
development in a spark-ignition engine. Symp (Int) Combust fraction in IC engines. Trans SAE J Engines 1996;105:2243e54
[Proc] 1991;23(1):1033e40. [SAE Paper no. 962052].

You might also like