You are on page 1of 15

Effect of heat transfer and crevice flow on the engine performance and NO

emissions in a hydrogen-fueled spark-ignition engine


E.G. Pariotis 1,*, G.M. Kosmadakis 2, and C.D. Rakopoulos 2
1
Laboratory of Naval Propulsion Systems, Section of Naval Architecture and Marine Engineering, Department
of Naval Sciences, Hellenic Naval Academy, Piraeus, Greece
2
Internal Combustion Engines Laboratory, Thermal Engineering Department, School of Mechanical
Engineering, National Technical University of Athens, Athens, Greece

Abstract: The motivation of the present work is to investigate the effect of heat and mass
transfer on the combustion process of a hydrogen-fueled spark-ignition engine, operating at
various loads and constant engine speed. The scope is to compare the calculated heat fluxes
with the available measured ones at specific locations on the cylinder walls of the engine
using three heat transfer models of increasing complexity (two existing and one proposed by
the authors). Apart from that, the effect of mass transfer through the crevice regions is also
investigated, using a phenomenological crevice model. The aforementioned models together
with the hydrogen combustion model are all included in an in-house 3D-CFD engine
simulation model, which has been extensively validated both at motoring and firing
conditions.

The calculated results have been compared to available measurements from a hydrogen-fueled
spark-ignition engine. This comparison actually concerns the cylinder pressure traces, local
heat fluxes and NO exhaust emissions. Comparing the results obtained with the various heat
transfer models considered, it has been revealed that the standard wall function formulation
under-predicts the local heat fluxes, while the robust heat transfer model, developed by the
authors, is proved to be much more accurate. Moreover, it has been revealed the importance
of using a crevice model, in order to calculate in better terms the in-cylinder pressure during
the combustion phase and expansion stroke, as well as the NO emissions.

Keywords: computational fluid dynamics, heat transfer, crevices, hydrogen, spark-ignition


engine.

*
Corresponding author. Tel.: +30 210 4581663; Fax: +30 211 7701574.
E-mail addresses: pariotis@snd.edu.gr (E.G. Pariotis), gkosmad@central.ntua.gr (G.M. Kosmadakis),
cdrakops@central.ntua.gr (C.D. Rakopoulos).

1
1. Introduction

One important research aspect of the automotive industry is the improvement of the engine
efficiency and the exploitation of the use of alternative fuels, such as hydrogen. Extended
experimental studies have already been made to investigate the combustion process and
performance of the spark-ignition engines running on hydrogen under different operating
conditions and strategies [1-4]. For this task many important parameters are measured, such
as the cylinder pressure, exhaust emissions, local wall temperatures, and processed to evaluate
the engine performance [5,6].

Apart from the experimental investigations of hydrogen-fueled spark-ignition engines,


numerical tools are also used [4,7,8]. Computational fluid dynamics (CFD) codes [4], not
only predict the engine performance, but they provide detailed and fundamental
understanding of the physical and chemical processes that take place inside the combustion
chamber. Thus, reliable predictions can be obtained regarding the flame kernel development
[8] and the in-cylinder heat and mass transfer.

In the present study an already validated in-house CFD code has been applied to simulate a
hydrogen-fueled, spark ignition engine [8], under various operating conditions, focusing on
the heat and mass transfer. In order to do so, different heat transfer models have been used (all
of them based on wall functions formulations), while, the effect of a crevice model is also
examined. The calculated results are compared with the corresponding measured ones, which
concern the in-cylinder pressure, the local heat fluxes at certain locations on the cylinder
walls and the exhaust NO emissions at various engine loads.

2. Available experimental data

The experimental data used in the present study concern measurements on a CFR engine,
operating at a constant engine speed equal to 600 rpm [2]. It is equipped with port fuel
injection (PFI) and has a variable compression ratio. The load of the engine is altered with the
variation of the equivalence ratio (quality load control), by controlling the fuel injection
timing / duration. All measurements have been conducted at wide-open throttle (WOT), due
to the specific characteristics of hydrogen [1], reducing at the same time the pumping losses.
The main specifications of this engine are given in Table 1.

2
Table 1. CFR engine specifications

CFR engine, single cylinder, naturally


Engine model and type
aspirated, four-stroke, water-cooled
Bore 82.55 mm
Stroke 114.2 mm
Swept volume 0.6117 lit
Connecting rod length 254 mm
Compression ratio Variable (CR: 8 and 9)
Valves 2 (unshrouded)
Ignition timing ΜΒΤ
Number of piston rings 5
Valve timing Inlet valve opening 17 oCA ATDC
events Inlet valve closure 26 oCA ABDC
Exhaust valve opening 32 oCA BBDC
Exhaust valve closure 6 oCA ATDC

The measured data concern the inlet mass flow rates of air and hydrogen, the cylinder
pressure traces, the NO/NOx emissions, the inlet and exhaust gas temperatures and the wall
temperatures and heat flux rates at three locations on the cylinder walls [2]. These data are
measured for different operating conditions, varying the compression ratio, and the fuel-to-air
equivalence ratio φ (i.e. the actual fuel-to-air ratio divided by its stoichiometric value). A
more detailed description of the measuring equipment can be found in Ref. [2], while a
general overview of the test setup with the notation of the measurement positions for local
temperature and heat flux and the location of the spark-plug is shown in Fig. 1.

Fig. 1. Axial and Cross-sections of the CFR engine, P1: spark plug, P2–P4: Temperature
sensor positions, IV: intake valve, EV: exhaust valve (sensors are located 9 mm lower than
the cylinder head)

3. Numerical model

3.1. CFD model

3
The firing version of the CFD code developed by the authors can simulate three-dimensional
curvilinear domains, using the finite volume method in a collocated grid. It has been validated
in a previous published work [9] in comparison to measured data, showing that it is capable of
adequately simulating the power cycle of hydrogen-fueled spark-ignition engines. Further
details of the in-house CFD model used, as well as a detailed evaluation of the developed
CFD model under both motoring and firing conditions, can be found in Refs. [9-12].

3.2. Heat transfer models considered

Various heat transfer models can be found in the literature, which are based on the wall-
function formulation [11]. This method is widely used in multidimensional simulations, since
it provides a robust way of solving the thermal boundary layer, without increasing the
computational nodes that are placed within this layer.

Three representative models have been selected and incorporated in the CFD code presented
in this study, to identify their relative performance in comparison to measured local wall heat
fluxes. Two of them are widely used in commercial and research codes to calculate the heat
transfer mechanism in all kind of engines. These are the model of Launder and Spalding [13]
and the model of Angelberger et al. [14], where the latter takes into account the variation of
the gas density and viscosity within the boundary layer and is actually a compressible version
of the former model. The third model considered has been developed by the authors [11] and
its formulation is based on a compressible version of the standard law-of-the-wall [13], while
it also includes an unsteady pressure term (dP/dt) and a combustion term ( Q c ), which has
been also shown in [11,15] to give more accurate results.

The wall heat flux, qW (in W/m2), which is inserted into the source term of the energy
equation (Sh) for the boundary cells, is shown in Table 2 for the three heat transfer models
considered. It should be mentioned that in these equations, the following parameters are used:
the non-dimensional distance from the wall, y   yuT  , the friction velocity,

uT   w   C 1/ 4 k y  with Cμ equal to 0.0845, according to the RNG k–ε turbulence

model [15], the local wall shear stress, τW, the viscous sublayer thermal resistance factor,
0.5 0.25
PPr Prt    4 sin  4  A   Pr Prt  1Pr Prt  , the laminar and turbulent Prandtl
numbers, Pr, Prt, while κ, Ε, A are coefficients with values 0.4187, 9.79 and 26, respectively
[13]. Further details concerning the use of different heat transfer models in the in-house CFD
code, and the validation of the one developed by the authors, can be found in Refs. [11,16].

4
Table 2. Wall heat flux (qW) corresponding to each one of the three models considered

Model Wall heat flux y+ valid


uT c p TW  T 
qW  y+≤11.63
Pr y 
Launder and
Spalding uT c p TW  T 
qW 
[13] 1  Pr  y+>11.63
 
Prt  ln Ey   P 

   Prt  
T
uT c pT ln  W T 
qW    y+≤11.63

Pr y
Angelberger
et al. [14] uT c pT ln TW 
qW   T
1  Pr  y+>11.63
 
Prt  ln Ey   P 

   Prt  
 y   40 
T  dP   
 uT c pT ln  w T    
 Qc  117.31
Rakopoulos    dt  0.4767  1
 uT 
qW   Pr  All y+
et al. [11] 1    1   1 
ln  y    ln  40    10.2384
0.4767   0.4767 Pr   0.4767 Pr 

3.3. Crevice model

The phenomenological crevice model developed by the authors [10] requires fewer input data
than the model of Namazian and Heywood [17], making its implementation easier. With the
use of this crevice model, the variation of the in-cylinder mass (due to the mass transfer
through the crevice rings) is taken into account during the closed engine cycle. Further details
concerning its development and implementation, can be found in previous works [11,17].

3.4. Combustion model

The detailed presentation and evaluation of the developed hydrogen combustion model is
provided in [9]. This combustion model can track the flame during its propagation, by
calculating the flame radius at each time step [4]. For the calculation of the reaction rates, the
characteristic time-scale method is used [18], while the NO reaction rates are calculated using
the extended Zeldovich mechanism [8] with the constants proposed by Lavoie et al. [19].

4. Results and discussion

4.1. Test cases examined

5
The operating conditions of the CFR engine investigated in the current study concern the
variation of the compression ratio (CR=8 and 9) and the equivalence ratio (φ=0.5, 0.75 and 1).
All cases examined were under wide-open throttle (WOT) operation, and the spark-timing
was set to its maximum brake torque (MBT) value, while the engine speed was kept constant
at 600 rpm.

4.2. Initial and Boundary Conditions - Computational details

In the present study some initial and boundary conditions are provided from the experimental
data, such as the temperature and pressure at IVC, the inlet air and hydrogen flow rates, and
the wall temperatures. For the rest of the required conditions, reliable estimations are used.
The initial gas temperature at IVC is varied from 370 K to 400 K, as the mixture becomes
richer (for both compression ratios), while further details, concerning the initial and boundary
conditions, the computational mesh, as well as the calculation of the residual gas fraction, can
be found in [9,16]. It should be mentioned that the only calibration constant used in the in-
house CFD code exists in the expression of the turbulent flame speed [9], and retains the
constant value A=0.8 for all cases examined.

4.3. Effect of heat transfer model

The investigation begins with the effect of the heat transfer model on the local wall heat
fluxes, the engine performance and NO emissions. This investigation is conducted for
compression ratio (CR) equal to 8 and for the whole range of equivalence ratio (0.5 – 1.0),
since similar conclusions are derived for compression ratio equal to 9.

4.3.1. Local wall heat fluxes

The local wall heat fluxes have been calculated using three different heat transfer models, as
already mentioned. In Fig. 2 is shown, the comparison of the predicted with the measured heat
fluxes at points named P2 and P3 (near the cylinder head) during the closed engine cycle.
These two points are located at the cylinder liner at different distances from the spark-plug.
From Fig. 2 it is observed that the heat transfer model developed by the authors [11] can
better predict the local heat fluxes, at both points compared with the other two heat transfer
models, while the matching with the measured one is better at location P2. This is mainly
attributed to the pressure and combustion term used in its formulation (see Table 2) [16]. The
predictions using the other two wall functions (Angelberger et al. [14] and Launder and
Spalding [13]) show large deviations from the measured data for all cases. The peak heat

6
fluxes are greatly under-predicted, while during the compression and expansion stroke the
calculated fluxes are significantly lower than the measured ones.

5
Local wall-heat flux (MW/m2)

P2 location - CR: 8
Measured
4 φ: 1
Calculated (Rakopoulos et al.)
IT: 6 oCA ATDC
Calculated (Angelberger et al.)
3 Calculated (Launder and Spalding)

2 φ: 0.75 φ: 0.50
IT: 2 oCA ATDC IT: 4 oCA BTDC

0
170 180 190 200 210 220 230 240 170 180 190 200 210 220 230 240 170 180 190 200 210 220 230 240

5 Crank angle degrees (ABDC)


Local wall-heat flux (MW/m2)

P3 location - CR: 8
Measured
4
Calculated (Rakopoulos et al.)
φ: 1 Calculated (Angelberger et al.)
3 IT: 6 oCA ATDC Calculated (Launder and Spalding)
φ: 0.75
2 IT: 2 oCA ATDC φ: 0.50
IT: 4 oCA BTDC
1

0
170 180 190 200 210 220 230 240 170 180 190 200 210 220 230 240 170 180 190 200 210 220 230 240
Crank angle degrees (ABDC)

Fig. 2. Comparison of calculated local wall heat fluxes using three heat transfer models with
the measured ones at locations P2 and P3, for compression ratio equal to 8

It is interesting to notice that the model of Angelberger et al. [14], which showed good
performance at various operating conditions and engines at motoring conditions [11], has a
poor performance at firing conditions, although much better than the model of Launder and
Spalding [13].

4.3.2. Engine performance

The adequate prediction of the heat transfer process is a prerequisite for the proper calculation
of the engine performance. Taking into account the significant differences observed in Fig.3,
regarding the measured and the calculated heat fluxes at certain points on the cylinder linear,
similar deviations concerning the cylinder pressure traces are expected. In Fig. 3 is shown the
calculated cylinder pressure traces using the three heat transfer models and their comparison
with the available measured data, for compression ratio equal to 8 and equivalence ratios
ranging from 0.5 to 1.

7
50
CR: 8 Measured
φ: 1 Calculated (Rakopoulos et al.)
IT: 6 0CA ATDC Calculated (Angelberger et al.)
Cylinder pressure (bar)
40
CR: 8 Calculated (Launder and Spalding)
φ: 0.75
IT: 2 0CA ATDC
30 CR: 8
φ: 0.50
IT: 4 0CA BTDC
20

10

0
140 160 180 200 220 240 140 160 180 200 220 240 140 160 180 200 220 240
Crank angle degrees (ABDC)

Fig. 3. Comparison of calculated cylinder pressure using three heat transfer models with the
measured data, for compression ratio equal to 8 and equivalence ratios φ=0.5, 0.75 and 1.0

It can be observed that the calculated cylinder pressure using the models of Launder and
Spalding [13] and Angelberger et al. [14] is higher compared to the measured one, not only
during the combustion phase, but also during the compression and late expansion stroke
(over-predicting the peak pressure by approx. 2 bar). On the contrary, when using the heat
transfer model developed by the authors [11], the cylinder pressure almost coincide with the
measured one at medium and high load, while at low load a small discrepancy is observed at
the initial phase of the combustion process. As a result in this case, the indicated mean
effective pressure (IMEP) and indicated efficiency are better predicted. The calculated values
of these parameters are depicted in Fig. 4 and compared with the measured ones, using the
three heat transfer models examined in this study.
Indicated mean effective pressure (bar)

9 Measured 36
Calculated (Rakopoulos et al.)
Indicated efficiency (%)

Calculated (Angelberger et al.)


8
Calculated (Launder and Spalding) 32

7
28
6

24
5 CR: 8 CR: 8
IT: MBT IT: MBT

4 20
0.5 0.75 1 0.5 0.75 1
Equivalence ratio (-) Equivalence ratio (-)

Fig. 4. Measured and Calculated Engine IMEP and indicated efficiency using the three heat
transfer models for compression ratio equal to 8 and equivalence ratios φ=0.5, 0.75 and 1.0

8
With the use of the heat transfer models of Angelberger et al. [14] and Launder and Spalding
[13] the performance values of the hydrogen-fuelled spark-ignition engine are significantly
over-predicted by a factor of appox. equal to 16-18% for all engine loads (around 1 bar). This
is an important disadvantage of these two heat transfer models, although the trend of IMEP
and indicated efficiency, as the engine load varies, is correctly captured. On the other hand,
the use of the heat transfer model developed in [11] provides much more reliable results
(difference lower than 2% in comparison with the measured values).

4.3.3. Exhaust NO emissions

In Fig. 5 are shown the predicted NO emissions using the three heat transfer models and are
compared with the corresponding measured values. It is observed that the heat transfer model
developed by the authors, manages to predict better the absolute values of the NO emissions
at all operating conditions, while all models predict adequately the trend. The relative error
between the predicted and the measured value of NO is approx. 15% when using the proposed
heat transfer model, with the higher differences observed at medium and high load.

7000
Measured
Exhaust NO emissions (ppm)

Calculated (Rakopoulos et al.)


6000
Calculated (Angelberger et al.)
Calculated (Launder and Spalding)
5000

4000

3000

2000
CR: 8
1000 IT: MBT

0
0.5 0.75 1
Equivalence ratio

Fig. 5. Measured and Calculated NO emissions using the three heat transfer models for
compression ratio equal to 8 and equivalence ratios φ=0.5, 0.75 and 1.0

To better explain the differences on the predicted values of NO emissions using the three heat
transfer models examined, in Fig.6 are presented the mass-weighted in-cylinder gas
temperatures of the burnt and un-burnt mixture [16], during the close part of engine cycle, as
equivalence ratio ranges from 0.5 to 1.0, using the three heat transfer models. NO emissions
are highly affected by the local in-cylinder temperature distribution, which in turn is affected
by the heat transfer model used.

9
Burned/unburned gas temperatures
3000
CR: 8 Rakopoulos et al.
Mass-weighted temperatures (K)

φ: 1 Angelberger et al.
IT: 6 oCA ATDC CR: 8
2500 Launder and Spalding
φ: 0.50
IT: 4 oCA BTDC
CR: 8
2000 φ: 0.75
IT: 2 oCA ATDC

1500 Burned
Unburned Unburned
Burned
Unburned Burned
1000

500
140 160 180 200 220 240 140 160 180 200 220 240 140 160 180 200 220 240

Crank angle degrees (ABDC)

Fig. 6. Calculated burned and unburned gas temperatures using the three heat transfer models,
for compression ratio equal to 8 and equivalence ratios φ=0.5, 0.75 and 1.0

It is observed that even the calculated unburned gas temperature during the compression
stroke and the early combustion period has some differences, when using different heat
transfer model. But the most critical one is the variation of the burned gas temperature, where
the resulting differences are important (including its peak value). The discrepancy of the peak
value is increased as the engine load decreases, while during the whole expansion stroke very
large temperature differences are observed (around 200-300 K) for all engine loads, when
using different heat transfer model. In all cases the predicted temperatures using the heat
transfer model proposed by the authors are lower compared with the other two heat transfer
models, which is in accordance with the corresponding trend observed regarding the NO
emissions (see Fig. 5).

4.4. Effect of crevice model

In the present study, apart from the heat transfer models, the effect of the crevice flow on the
predicted engine performance and NO emissions are examined. To this scope, the simulation
model developed by the authors is used in two versions: with and without the crevice model.
Three operating conditions are selected to evaluate the effect of the crevice model on engine
performance and NO emissions. In Table 3 are shown the test cases examined.

Table 3. Test cases examined for the effect of the crevice model

Case Equivalence ratio (φ) Compression ratio MBT timing (oCA ATDC)

10
1 0.75 8 2
2 0.75 9 8
3 1 9 14

4.4.1. Engine performance

The engine performance is investigated first. The heat transfer model developed in [11] has
been used in all the simulations. In Fig. 7 are shown the calculated cylinder pressure traces for
the three test cases shown in Table 3, with and without the use of the crevice model, and
compared with the corresponding measured values.

50
Test Case 1 Measured
Cylinder pressure (bar)

CR: 8 Calculated (with crevices)


40 Test Case 3
IT: 2 oCA ATDC Calculated (without crevices)
φ: 0.75 CR: 9
30 Test Case 2 IT: 14 oCA ATDC
CR: 9 φ: 1
IT: 8 oCA ATDC
20 φ: 0.75

10

0
140 160 180 200 220 240 140 160 180 200 220 240 140 160 180 200 220 240
Crank angle degrees (ABDC)

Fig. 7. Calculated cylinder pressure traces with and without the use of the crevice model and
comparison with the corresponding measured data for the three test cases examined

It should be noticed that in all test cases examined the spark-timing is after TDC, which helps
to identify more clearly the effect of the crevice model during the compression and
combustion phase. As observed in Fig. 7, when the crevice model is de-activated, the cylinder
pressure is over-predicted from the end of the compression stroke (near TDC) and afterwards.
The calculated peak pressure is significantly higher than the measured one for all test cases
examined (by around 3 bar), while during the expansion stroke their difference is almost
constant (around 1 bar) [10]. On the other hand, when activating the crevice model, the
cylinder pressure is better predicted and minor differences can be noticed in comparison to the
measured pressure.

The corresponding effect of the crevice flow on the IMEP and indicated efficiency is shown
in Fig. 8.

11
Indicated mean effective pressure (bar)
7.5 Measured 32
Calculated (with crevices)
Calculated (without crevices)

Indicated efficiency (%)


7 30

6.5 28

6 26

5.5 24
IT: MBT IT: MBT
5 22
1 2 3 1 2 3
Test cases Test cases

Fig. 8. Calculated IMEP and indicated efficiency with and without the use of the crevice
model and comparison with the measured data for the three test cases

The over-estimation of the calculated IMEP and indicated efficiency, when the crevice model
is not used, is significant, being approx. 10-15% higher than the measured value. On the other
hand with the use of the crevice model, the calculations are more reliable, since the
performance parameters almost match the measured values, with a small discrepancy
observed only in case 1.

4.4.2. Exhaust NO emissions

To investigate the effect of the crevice flow on the predicted NO emissions, in Fig. 9 are
compared the calculated NO emissions with/without the usage of the crevice model and with
the corresponding measured ones, at the three test cases examined.

5000 Measured
Exhaust NO emissions (ppm)

Calculated (with crevices)


Calculated (without crevices)
4000

3000

2000
IT: MBT

1000
1 2 3
Test case

Fig. 9. Calculated NO emissions with and without the use of the crevice model and
comparison with the measured data for the three test cases

12
As observed the use of the crevice model has a minor effect on the calculated NO emissions,
which is less than the corresponding effect of the heat transfer model already examined (see
Fig. 5). This is attributed to the fact that the crevice model affects the calculated in-cylinder
trapped mass at each crank angle, which has a major effect on the in-cylinder pressure, but
just a minor on the gas temperature, as observed in Fig. 10, where the mass-weighted burned
and unburned gas temperatures are presented for the three test cases examined.

3000 Burned/unburned gas temperatures


Mass-weighted temperatures (K)

Test case 1 With crevices


CR: 8 Without crevices
2500 φ: 0.75 Test case 3
IT: 2 oCA ATDC CR: 9
Test case 2 φ: 1
2000 CR: 9 IT: 14 oCA ATDC
φ: 0.75
IT: 8 oCA ATDC
1500
Burned
Unburned Unburned
Burned
Unburned Burned
1000

500
140 160 180 200 220 240 140 160 180 200 220 240 140 160 180 200 220 240

Crank angle degrees (ABDC)

Fig. 10. Calculated burned and unburned gas mass-weighted temperatures with and without
the use of the crevice model for the three test cases examined

This is in accordance with what has already been observed in similar simulations conducted
by the authors at motoring conditions [10]. More specifically, the unburned gas temperature
remains unaffected, while as far as the burned gas temperature is concerned, just a minor
difference is noticed during the expansion stroke and especially for the stoichiometric mixture
(test case 3), when cold gas is entering the cylinder from the crevice regions, because of the
decrease of the in-cylinder pressure [10].

5. Summary and conclusions

In the present study, the effect of the heat transfer model and the use of a crevice model on the
predicted performance and NO emissions of a hydrogen-fueled spark-ignition engine at
various engine loads have been investigated. As observed, the heat transfer model used affects
significantly the predicted performance and NO emissions. Using two well-known wall-
function formulations for the heat transfer modeling, the local heat fluxes are under-estimated,
leading to an important over-prediction of the in-cylinder pressure, the gas temperatures and
the NO emissions. However, the heat transfer model proposed by the authors seems to

13
perform much better predicting well the engine performance and NO emissions at all test
cases examined. On the other hand, the use of a crevice model influences to a smaller degree
the performance of the engine, while it has a minor effect on the prediction of the NO
emissions. Concluding, through this study it has been demonstrated the importance of using a
reliable heat transfer model for the proper prediction of engine performance and emissions,
while the effect of the crevice model seems to be less intense, but equally important for an
accurate simulation of the physical processes taking place inside the cylinder.

Acknowledgements

The authors wish to thank Prof. S. Verhelst (Univ. of Gent, Belgium) for the provision of the
experimental data.

References:

1. Verhelst, S. A study of the combustion in hydrogen-fuelled internal combustion engines,


Ph.D. Thesis, Ghent University, 2005. (URL: http://hdl.handle.net/1854/3378).
2. Demuynck, J., Raes, N., Zuliani, M., De Paepe, M., Sierens, R., and Verhelst, S. Local
heat flux measurements in a hydrogen and methane spark ignition engine with a
thermopile sensor. Int. J. Hydrogen Energy, 2009, 34(24), 9857–68.
3. Rottengruber, H., Berckmueller, M., Elsaesser, G., Brehm, N., and Schwarz, C. Direct-
injection hydrogen SI-engine operation strategy and power density potentials. Trans.
SAE, J. Fuels Lubricants, 2004, 113, 1749–61 [SAE Paper no. 2004-01-2927].
4. Gerke, U. Numerical analysis of mixture formation and combustion in a hydrogen direct-
injection internal combustion engine. Ph.D. Thesis, ETH Zurich, Switzerland, 2007.
(URL: http://e-collection.ethbib.ethz.ch/eserv/eth:30102/eth-30102-02.pdf).
5. Shudo, T. Improving thermal efficiency by reducing cooling losses in hydrogen
combustion engines. Int. J. Hydrogen Energy, 2007, 32(17), 4285–93.
6. Mohammadi, A., Shioji, M., Nakai, Y., Ishikura, W., and Tabo, E. Performance and
combustion characteristics of a direct-injection SI hydrogen engine. Int. J. Hydrogen
Energy, 2007, 32(2), 296–304.
7. Verhelst, S., and Sierens, R. A quasi-dimensional model for the power cycle of a
hydrogen-fuelled ICE. Int. J. Hydrogen Energy, 2007, 32(15), 3545–54.

14
8. Knop, V., Benkenida, A., Jay, S., and 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(19), 5083–97.
9. Rakopoulos, C.D., Kosmadakis, G.M., and Pariotis, E.G. Evaluation of a combustion
model for the simulation of hydrogen spark-ignition engines using a CFD code. Int. J.
Hydrogen Energy, 2010, 35(22), 12545–60.
10. Rakopoulos, C.D., Kosmadakis, G.M., Dimaratos, A.M., and Pariotis, E.G. Investigating
the effect of crevice flow on internal combustion engines using a new simple crevice
model implemented in a CFD code. Appl. Energy, 2011, 88(1), 111–26.
11. Rakopoulos, C.D., Kosmadakis, G.M., and Pariotis, E.G. Critical evaluation of current
heat transfer models used in CFD in-cylinder engine simulations and establishment of a
comprehensive wall-function formulation. Appl. Energy, 2010, 87(5), 1612–30.
12. Rakopoulos, C.D., Kosmadakis, G.M., and Pariotis, E.G. Evaluation of a new
computational fluid dynamics model for internal combustion engines using hydrogen
under motoring conditions. Energy, 2009, 34(12), 2158–66.
13. Launder, B.E., and Spalding, D.B. The numerical computation of turbulent flows.
Comput. Methods Appl. Mech. Engrg., 1974, 3(2), 269–89.
14. Angelberger, C., Poinsot, T., and Delhaye, B. Improving near-wall combustion and wall
heat transfer modeling in SI engine computations. SAE Paper no. 972881, 1997.
15. Han, Z., and Reitz, R.D. A temperature wall function formulation for variable-density
turbulent flows with application to engine convective heat transfer modeling. Int. J. Heat
Mass Transfer 1997, 40(3), 613–25.
16. Rakopoulos, C.D., Kosmadakis, G.M., Demuynck, J., De Paepe, M., and Verhelst, S. A
combined experimental and numerical study of thermal processes, performance and nitric
oxide emissions in a hydrogen-fueled spark-ignition engine. Int. J. Hydrogen Energy,
2011, 36(8), 5163–80.
17. Namazian, M., and Heywood, J.B. Flow in the piston-cylinder-ring crevices of a spark-
ignition engine: effect on hydrocarbon emissions, efficiency and power. Trans. SAE
(Section 1), 1982, 91, 261-88 [SAE Paper no. 820088].
18. Abraham, J., Bracco, F.V., and Reitz, R.D. Comparison of computed and measured
premixed charge engine combustion. Combust. Flame, 1985, 60(3), 309–22.
19. Lavoie, G.A., Heywood, J.B., and Keck, J.C. Experimental and theoretical study of nitric
oxide formation in internal combustion engines. Combust. Sci. Technol., 1970, 1, 313–26.

15

You might also like