Вы находитесь на странице: 1из 19

A NUMERICAL STUDY OF EFFECTS OF COMBUSTION CHAMBER

SHAPE ON PRE-CHAMBER COMBUSTION

S. Matsuo, Y. Kawabata, K. Okamoto and T. Amano


Tokyo Gas Co., Ltd.
Japan

ABSTRACT

To accomplish a 3-D simulation of pre-chamber combustion, the original turbulent


combustion model was developed. Two experiments, the performance test and the
combustion visualization one, were performed using a single-cylinder pre-chamber engine to
verify the applicability of the model. The history of the main chamber pressure, the flame jet
behavior and the flame propagation process were compared between the numerical and the
experimental results, and consequently, a good agreement was obtained.

Then, the effects of cavity area ratio on pre-chamber combustion were studied using the
two model engines with a bathtub type cavity. As a result, it is found that combustion in the
main chamber was promoted when the engine with the smaller cavity area ratio was used. In
addition, the effects of piston cavity type were also examined using the model engines with a
bathtub and a toroidal type cavity. As a consequence, it is found that combustion in the main
chamber was promoted for the toroidal type cavity compared with for the bathtub type one.
The causes of such combustion promotion could be explained with the differences of the
flame jet behavior and the flame propagation process induced by those of combustion
chamber shape factors.

1
INTRODUCTION

Recently, interest on the energy saving and the less environmental harm of exhaust
are increasing. Thus, gas engines for co-generation systems are required to have higher
thermal efficiency and less harmful emissions such as NOx in their exhaust. For lean
burn method, the theoretical thermal efficiency is higher and NOx is less exhausted than
for the stoichiometric burn method because the ratio of the specific heat of the mixture
is higher and the flame temperature is lower, respectively. Hence, the lean burn method
is regarded as one of effective ways to meet these requirements.

Lean burn gas engines are categorized into two types, i.e. open chamber type and
pre-chamber one. For the lean burn gas engines with a relatively large bore diameter,
which is larger than 150 mm in general, the latter type is adopted. The lean burn method
has the advantages mentioned above, however, it also has the weak points of low
ignitability and slow flame propagation. A pre-chamber engine can overcome these
problems because the flame jets with high energy play a role of ignition source of the
mixture in the main chamber and the turbulence produced by the flame jets eruption
promotes the flame propagation. Several researchers have investigated the operating
characteristics of pre-chamber lean burn gas engines[1]-[3].

For pre-chamber engines, combustion in the main chamber, which affects the
performance of the engine, depends strongly on the behavior of the flame jets and the
flame propagation process in the main chamber. In addition, those features are deeply
affected by the combustion chamber shape. Therefore, in order to achieve higher
performance of the engines, the combustion chamber shape needs to be optimized, and
relating parametric studies should be carried out. It takes much cost and time for the
experimental tests because the combustion chamber has many shape factors such as the
number of jet nozzles, diameter of the jet nozzle, cavity area ratio, piston cavity type
etc. Thus, numerical simulations are expected to be applied to such parametric studies to
perform an effective development of the engines.

However, it is difficult to simulate combustion by a flame jet and consequently, only


a few results of the simulations of pre-chamber combustion have been reported. While
flame jet is erupting with high velocity, extinction by flame stretch mechanism probably
occurs at the boundary between the flame jet and the surrounding mixture in the main

2
chamber and consequently, the flame can not propagate to the directions perpendicular
to the flame jet axis. It is considered that one of the causes that the existing combustion
models can not simulate combustion by a flame jet is that they do not consider the
extinction phenomenon.

In this study, in order to simulate pre-chamber combustion, we developed the


original turbulent combustion model (Turbulent Flame Speed Closure Model) in which
the extinction by flame stretch is considered, and tried to simulate the pre-chamber
combustion using the new model. The results of the present study could simulate well
the corresponding experimental results. Thus, we investigated the effects of cavity area
ratio and piston cavity type on pre-chamber combustion using the present combustion
model.

TURBULENT FLAME SPEED CLOSURE MODEL

Summary
In this combustion model, the fuel reaction rate can be determined using two
different mechanisms, auto-ignition and flame propagation schemes. The larger reaction
rate of these two mechanisms is the dominant one. Hence, the fuel reaction rate ωfuel
can be described using a maximum operator via:

ωfuel = max { Auto − ignition ω AI , Flame Pr opagation ω FP } (1)

The first scheme is built on the database of the reaction rate. The database is made by
using CHEMKIN Ⅱ(SENKIN). The auto-ignition reaction rate ωAI can be written as:

 a 
ω AI = a 1 ρ a 2 y afuel
3
y aO42 T a 5 exp − 6  (2)
 T

where a1 to a6 are empirical coefficients, ρ is the gas density, yfuel and yo2 are the
fuel and oxygen mass fractions, and T is the temperature, respectively.

The reaction rate of the flame propagation mechanism ωFP, the second one, can be
written as:

ωFP = ρ ST ∇c f st (3)

3
where ST represents the turbulent burning velocity, c the reaction progress variable, which
is equivalent to the mass of combustion products divided by fully burnt (theoretical
maximum) mass of combustion products, and fst stoichiometric mixture fraction.

The turbulent burning velocity ST [4] can be written as:

  
0.38

ST =  2.9 ×  p  u '  × SL for 0 < Ka ≤ Ka th (4)
 p 0  S L 
 0 for Ka th < Ka

where SL represents the laminar burning velocity, p the pressure, p0 the datum
(=0.1MPa), u’ the turbulence intensity, Ka the Karlovitz number, and Kath the threshold
of Ka, respectively. The above formula indicates that the flame is extinguished by the
effect of flame stretch when the Karlovitz number is larger than the threshold. The
Karlovitz number can be written as:

b2 b3
 u'  δ 
Ka = b1 ×   ×  L  (5)
 SL   LT 

where b1 to b3 are empirical coefficients, δL is the laminar flame thickness, and LT is the
turbulent length scale, respectively.

The laminar burning velocity SL and the laminar flame thickness δL can be
expressed as follows:

SL 
δ L
(
 = c1 + c 2 λ + c 3 λ + c 4 λ + c 5 λ
2 3 4
)
(6)
 c7 c9    c 
×  c 6 + +  ×  c + c13 exp 14  
2   12
 c 8 + p c10 + c11 p + p    T 

where c1 to c14 are empirical coefficients and differ in their individual parameters for each. The
above formula is built on the database which was made by using CHEMKIN Ⅱ(PREMIX).

4
The turbulent length scale LT is determined using the following formulation via:

3 k 1.5
LT = cµ 4 × (7)
ε

with k as the turbulence kinetic energy ε as its dissipation rate and cμ as constant,
respectively.

Verification of TFSCM
The applicability of TFSCM to a 3-D simulation of pre-chamber combustion was
verified by comparing the numerical results with the experimental ones. Two
experiments were performed using a single-cylinder pre-chamber engine. One was
measuring the performance of the engine and another was observing the combustion
process. The history of the main chamber pressure, the behavior of the flame jets and
the flame propagation process for each result were compared.

Table 1 shows the specifications of the engine used in the experiments. The piston had a
bathtub type cavity with its area ratio of 0.33. By changing the piston, both experiments could
be conducted with the identical engine. In the visualization experiment, a piston with a glass
which enabled us to observe combustion in the main chamber from the bottom was
employed. The details of the experimental apparatus and the method have been reported
elsewhere[5]. The experimental conditions are listed in table 2.

Table 1. Engine Specifications

Engine Type 4-stroke, 4-valve, OHV


Bore × Stroke [mm] 150 × 150
Displacement Volume [cm3] 2649
Compression Ratio 13
Swirl Ratio 2
Number of Nozzles 4

5
Table 2. Experimental Conditions

Engine Speed [rpm] 1000


Mixture Temperature [K] 313
Mixture Pressure [kPa] 157
Ignition Timing [deg. aTDC] -11
Fuel Natural Gas
Total Excess Air Ratio 1.9

Figure 1 shows the computational grid of the main chamber of the corresponding engine.
The simulation was performed using a CFD-code FIRE v7.3 which was developed by AVL
LIST GmbH. The calculation conditions except the ignition timing were the same as the
experimental ones. The ignition timing was adjusted so that the first peak in the pre-chamber
pressure history of the numerical result would agree with that of the experimental one. As a
result, the ignition timing used in the simulation was -9 deg. aTDC. This adjustment was
reasonable because the ignition timing was found to fluctuate by around one degree from the set
value in the experiments and spark ignition phenomenon was not modeled strictly in the
calculation. The turbulent model used in this calculation was the standard k-εmodel. The
auto-ignition reaction rate ωAI was set to zero constantly, that is the auto-ignition would not
occur because the present auto-ignition model in which the ignition delay was not considered
was insufficient and consequently could not simulate the phenomenon.

Figure 1. Computational Grid of Main Chamber (A.R.=0.33 Bathtub Type)

Figure 2 shows the comparison of the history of the main chamber pressure
between the experimental and numerical results. The numerical result shows a good
agreement with the experimental one. However, this result is insufficient to verify the
applicability of TFSCM to a 3-D simulation of pre-chamber combustion because that
can be also achieved by assuming that the history of the heat release rate agrees even
though the behavior of flame jets and the flame propagation process do not agree. For
pre-chamber combustion, differences of combustion chamber shape cause changes of

6
behavior of flame jets and flame propagation process, furthermore, those induce
differences of the overall combustion characteristics such as the heat release rate and the
main chamber pressure etc. Therefore, a combustion model for pre-chamber combustion
must be able to simulate flame jet behavior and flame propagation process well to
evaluate effects of combustion chamber shape on pre-chamber combustion.
Accordingly, the comparison of the features between the results of the visualization
experiments and those of the numerical simulations was also made.

12
C alculation
Experim ent
10
M ain cham ber pressure, M P a

0
-20 -10 0 10 20 30 40
C rank angle, deg. aT D C

Figure 2. Comparison of History of Main Chamber Pressure between Calculation and


Experiment (A.R.=0.33 Bathtub Type)

Figure 3 indicates the combustion process in the main chamber observed in the
visualization experiment. The self-luminescence integrated along the cylinder axis appears
in the images, which were captured from the bottom of the engine. The circles shown in the
images indicate the edge of the observation window with a diameter of 87 mm, and mostly
coincide with the side wall of the piston cavity, where the flame jets impinge.

After the spark ignition in the pre-chamber whose timing was set to -11 deg. aTDC,
the mixture in the pre-chamber burns and the reacting flow with high energy, i.e. the
flame jets erupt to the main chamber from the pre-chamber. They appear in the main
chamber at -4.3 deg. aTDC and go toward the piston cavity wall retaining their nearly
straight-line structure without being bent by the clockwise swirl in the figure. They have
strong penetration. Next, they impinge intensely on the wall and split, then flow along
the wall circumferentially. The eruption of the flame jets ends at around the TDC. The
flame dispersed in the cavity by the impingement plays a role of the ignition source of
the mixture in the main chamber and flame propagation in the main chamber start from

7
that. The propagating flame seems to burn the mixture near the side wall first, then that
in the zone between adjacent flame jets and that in the center of the cavity last while
being affected by the swirl.

-4.3 deg.
aTDC TDC

-3.6 deg. 2.4 deg.


aTDC aTDC

-2.9 deg. 5.1 deg.


aTDC aTDC

-2.2 deg. 7.8 deg.


aTDC aTDC

-1.5 deg. 9.8 deg.


aTDC aTDC

-0.7 deg. 15.1 deg.


aTDC aTDC

Figure 3. Combustion Process (Experiment, A.R.=0.33 Bathtub Type)

Figure 4 shows the calculated distribution of the products mass fraction indicating
the combustion process. Though the flame jets begin to spread at an earlier time in the
simulation than in the experiment, the flame jets impinge on the cavity wall and split
and flow along the wall circumferentially as shown in the experimental result. Figure 5
shows the calculated distribution of the velocity in the x direction around a nozzle. The
flame jet eruption ends between -2.5 and -2.0 deg. aTDC. Accordingly, the period of the
flame jet eruption is about 3 deg. C.A., which is shorter than that of the experimental
result. This figure also indicates that after the finish of the eruption, the mixture begins
to flow from the main chamber to the pre-chamber reversely. A phenomena probably
related with this reverse flow was observed in the visualization experiment. The detail
of the phenomenon is described later. Figure 4, in addition, indicates that the impinged
flame jets split not only circumferentially but also up and down, and then the flame
from the downward split flame jets propagate to the center of the cavity after reaching
the bottom. The mixture in the lower zone of the center of the cavity is burned by also
the flame propagating from above simultaneously. Flame from the circumferentially

8
split flame jets and the flame jets remaining on the nozzle axes propagate to the zone
between the adjacent jets while being affected by the swirl.

Products mass fraction [-]

0 0.074 0.148

-5 deg.
aTDC

-4 deg.
aTDC

-2 deg.
aTDC

TDC

2 deg.
aTDC

4 deg.
aTDC

6 deg.
aTDC

8 deg.
aTDC

10 deg.
aTDC

15 deg.
aTDC

Figure 4. Combustion Process (Calculation, A.R.=0.33 Bathtub Type)

9
Velocity [m/s]

-100 0 100
x

-2.5 deg.
aTDC

-2.0 deg.
aTDC

-1.5 deg.
aTDC

Figure 5. Reverse Flow of Main Chamber Mixture

As discussed above, although the room for improvement about a simulation of flame
jet eruption is remained, the combustion process is also well simulated. Thus, we can
conclude that TFSCM is applicable to a 3-D simulation of pre-chamber combustion.

RESULTS AND DISCUSSION

Effects of Cavity Area Ratio

Effects of cavity area ratio on pre-chamber combustion were studied using two model
engines with a bathtub type cavity. Only the difference in the specifications of the two
engines is cavity area ratio. Cavity area ratios of the two engines are 0.33 and 0.50. The
computational grid of the main chamber having a cavity with its area ratio of 0.33 (the
0.33-bathtub-type engine) was already shown in figure 1 and the calculation conditions used
and the results were described above. Figure 6 shows the computational grid of the other i.e.
the main chamber having a cavity with its area ratio of 0.50 (the 0.50-bathtub-type engine).

10
Because all the specifications except cavity area ratio of the two engines are identical, a
cavity depth decreases and the zone between the adjacent flame jets enlarges with an
increase in cavity area ratio. The calculation for the 0.50-bathtub-type engine was
performed under the same conditions used in that of the 0.33-bathtub-type one.

Figure 6. Computational Grid of Main Chamber (A.R.=0.50 Bathtub Type)

Figure 7 shows the calculated distribution of the products mass fraction indicating
the combustion process for the 0.50-bathtub-type engine. The flame jets erupted to the
main chamber impinge on the side wall of the cavity and split and then, flow
circumferentially and up and down along the wall as shown in the results of the
0.33-bathtub-type engine. However, the downward split flow reaches the bottom and the
propagating flame from the flame jets remaining on the nozzle axes reaches the bottom
and the cylinder head at around the TDC which is an early stage of combustion because
the cavity depth is small. Then, flame propagates mainly to the zone between adjacent
flame jets. On the other hand, for the 0.33-bathtub-type engine, flame from the
dispersed ignition sources by the impingement propagates to many directions
simultaneously as shown in figure 4. Therefore, it is considered that the flame front area
for the 0.33-bathtub-type engine is larger than that for 0.50-bathtub-type one during a
main combustion period, as a consequence, combustion is promoted for the former
compared with for the latter.

Then, the history of the heat release rate for each result was compared as shown in
figure 8. Both have a small peak around the -5 deg. aTDC. As shown in figure 4 and 8,
the mixture in the main chamber does not burn before -5 deg. aTDC. Thus, this first
peak is caused by combustion in the pre-chamber and the rise after that by combustion
in the main chamber. Each value of X10-90% representing overall burning velocity of
mixture in the main chamber during the main combustion period is 10.34 deg. C.A. for
the 0.33-bathtub-type engine and 11.86 deg. C.A. for the 0.50-bathtub-type one. This
indicates that combustion in the main chamber is promoted for 0.33-bathtub-type engine
compared with for 0.50-bathtub-type one. Accordingly, the maximum value in the

11
history of the main chamber pressure is larger and the timing is earlier for the
0.33-bathtub-type engine than for the 0.50-bathtub-type engine as shown in figure 9.

Products mass fraction [-]

0 0.074 0.148

-5 deg.
aTDC

-4 deg.
aTDC

-2 deg.
aTDC

TDC

2 deg.
aTDC

4 deg.
aTDC

6 deg.
aTDC

8 deg.
aTDC

10 deg.
aTDC

15 deg.
aTDC

Figure 7. Combustion Process (A.R.=0.50 Bathtub Type)

12
800
0.33-bathtub-type
700 0.50-bathtub-type

H eat release rate, J/deg.


600

500

400

300

200

100

0
-20 -10 0 10 20 30 40
C rank angle, deg. aT D C

Figure 8. Comparison of History of Heat Release Rate (A.R.=0.33 Bathtub Type,


A.R.=0.50 Bathtub Type)

12

10
M ain cham ber pressure, M P a

2
0.33-bathtub-type
0.50-bathtub-type
0
-20 -10 0 10 20 30 40
C rank angle, deg. aT D C

Figure 9. Comparison of History of Main Chamber Pressure (A.R.=0.33 Bathtub Type,


A.R.=0.50 Bathtub Type)

Thus, the difference of cavity area ratio causes that of the combustion process and
furthermore, overall combustion characteristics as the histories of heat release rate and
the main chamber pressure.

Effects of Piston Cavity Type


The calculations for two model engines were performed to study the effects of
piston cavity type on pre-chamber combustion. The cavity area ratio of both engines is
0.50 and cavity types are bathtub type (the 0.50-bathtub-type engine) and toroidal type
(the 0.50-toroidal-type engine). The calculation for the 0.50-bathtub-type engine was
described above.

13
Figure 10 shows the computational grid of the main chamber of the
0.50-toroidal-type engine. All the specifications of the two engines are the same except
for piston cavity type. The calculation for 0.50-toroidal-type engine was performed
under the same conditions for 0.50-bathtub-type engine.

Figure 10. Computational Grid of Main Chamber (A.R.=0.50 Toroidal Type)

Figure 11 shows the calculated distribution of the products mass fraction indicating
the combustion process for the 0.50-toroidal-type engine. The flame jets impinge on the
cavity wall and split and then, flow circumferentially and up and down along the wall.
For the 0.50-toroidal-type engine, the flame jets spread more widely at the upper zone
of the cavity, on the contrary, less at the lower zone than for the 0.50-bathtub-type one.
This is probably due to the difference of the impingement angles of the flame jet to the
tangential plane on the point where the flame jet impinges. It is considered that much
time is needed for the flame to cover all the lower zone of the cavity because the jet
spreads less widely there. However, that can not be seen because the circumferential
interval between the adjacent jets is smaller at lower zone of the cavity. The upper zone
of the cavity, in particular near the wall, is covered more rapidly owing to the more
widely spread flame jets by the impingement. Accordingly, combustion is promoted for
the 0.50-toroidal-type engine than for the 0.50-bathtub-type one.

For the 0.50-bathtub-type engine, the lower zone of the center of the cavity is
difficult to be covered by flame although the cavity depth is small as shown in figure 8.
For the 0.50-toroidal-type engine, this zone is removed and the corresponding volume is
added to the zone where the flame reaches easily. This also probably causes the
promotion of overall combustion.

Thus, the changing the cavity shape from bathtub type to toroidal one causes the
change of the combustion process and reallocation of the zone which the flame is
difficult to reach to the zone where the flame reaches easily. As a result, combustion is
promoted.

14
Products mass fraction [-]

0 0.074 0.148

-5 deg.
aTDC

-4 deg.
aTDC

-2 deg.
aTDC

TDC

2 deg.
aTDC

4 deg.
aTDC

6 deg.
aTDC

8 deg.
aTDC

10 deg.
aTDC

15 deg.
aTDC

Figure 11. Combustion Process (A.R.=0.50 Toroidal Type)

Figure 12 shows the comparison of the history of the heat release rate between the
0.50-bathtub-type engine and 0.50-toroidal-type one. It is clear that combustion is
promoted for the 0.50-toroidal-type engine compared with for the 0.50-bathtub-type
one. Each value of X10-90 is 10.28 and 11.86 degree C.A., for 0.50-bathtub-type and

15
for 0.50-toroidal-type, respectively. On the basis of this result, it is easy to understand
the history of the main chamber pressure as shown in figure 13.

800
0.50-bathtub-type
700 0.50-toroidal-type

H eat release rate, J /deg.


600

500

400

300

200

100

0
-20 -10 0 10 20 30 40
C rank angle, deg. aTD C

Figure 12. Comparison of History of Heat Release Rate (A.R.=0.50 Bathtub Type,
A.R.=0.50 Toroidal Type)

12

10
M ain cham ber pressure, M P a

2 0.50-bathtub-type
0.50-toroidal-type
0
-20 -10 0 10 20 30 40
C rank angle, deg. aTD C

Figure 13. Comparison of History of Main Chamber Pressure (A.R.=0.50 Bathtub Type,
A.R.=0.50 Toroidal Type)

Reverse Flow Phenomenon of the Main Chamber Mixture


As shown in figure 5, the mixture in the main chamber flows to the pre-chamber after the
finish of the flame jet eruption. Figure 14 shows the calculated distribution of the unburned
fuel mass fraction. The reverse flow is found to include the unburned fuel. By an analysis of
the velocity distribution, this reverse flow was observed from -2 to 10 deg. aTDC.

16
The pressure difference between the main chamber and pre-chamber is studied to
clarify the cause of the reverse flow phenomenon. It is found that the pre-chamber
pressure is lower than that in the main chamber from about –3 to 7.5 degree aTDC. This
is due to rapid pressure rise in the main chamber in the post-flame jet eruption period
caused by combustion in the main chamber. In the pre-chamber, in contrast, such
pressure rise does not occur since combustion of the mixture in the pre-chamber is
already completed and emitted as flame jets. Thus, pressure is higher in the main
chamber than in the pre-chamber after flame jets are seen. This causes the reverse flow
of the unburned mixture to the main chamber.

Unburned fuel mass fraction [-]

0 0.0145 0 .029

-2.5 deg.
aTDC

-2.0 deg.
aTDC

-1.5 deg.
aTDC

Figure 14. Reverse Flow of Unburned Fuel to Pre-chamber

A phenomenon that is probably caused by this reverse flow including the unburned
fuel was observed in the visualization experiment. The burned gas with faint light
emitting flows out from the pre-chamber at the late stage of combustion by which it is
completed as shown in figure 15. Unlike the flame jets associated with combustion in

17
the pre-chamber right after spark ignition, the gas flow has little penetration and is
diffused near the pre-chamber nozzle largely by the swirl flow.

33.2 deg.
aTDC

40.6 deg.
aTDC

48.0 deg.
aTDC

55.3 deg.
aTDC

Figure 15. Flow Out Phenomenon of Reversed Main Chamber Mixture

The gas flow with faint light emitting was observed or not observed even if the
experiments were performed under the same conditions. It is considered that when the
light emitting is observed, the reversed unburned fuel is ignited by certain causes and
the burned gas flows out as the main chamber pressure decreases. On the contrary, when
that is not observed, the unburned fuel is exhausted without reacting. This can lead
efficiency to decrease and emission of unburned hydrocarbons to increase.

CONCLUSION

The original turbulent combustion model (TFSCM) was developed to simulate


pre-chamber combustion. The applicability of the model to pre-chamber combustion
was verified by comparing the calculated results with the experimental ones, which are
for the performance test and for the combustion visualization test. Furthermore, the effects
of cavity area ratio and piston cavity type on pre-chamber combustion were studied by the
simulations with the combustion model. The major findings are as follows:

1. The original turbulent combustion model we developed is applicable to a simulation


of pre-chamber combustion.

18
2. By changing the cavity area ratio, the flame jet behavior and the flame propagation
process change and consequently, the pattern of the heat release varies.
3. It is suggested that combustion in a main chamber is promoted for smaller cavity
area ratio compared with for larger one.
4. By changing the piston cavity type, the flame jet behavior and the flame
propagation process change and consequently, the pattern for the heat release varies.
5. It is suggested that combustion in a main chamber is promoted for a toroidal type
cavity compared with for a bathtub type one.
6. It is indicated that the unburned mixture flow from a main chamber to a pre-chamber
after the finish of the flame jet eruption. The mixture sometimes flows back to a main
chamber without reacting in a pre-chamber. This can cause efficiency decrease and
increase in concentration of the unburned hydrocarbons in emission.

ACKNOWLEDGMENTS

The support of AVL LIST GmbH is gratefully acknowledged.

REFERENCES CITED

1. Nakazono et al., “A study for lean burn gas engine 1st report“, the 9th Internal
Combustion Engine Symposium, Japan, pp.313-318, 1991
2. Nakagawa et al., “High Efficiency, Low NOx, Ultra Lean Burn SI Prechamber
Gas Engine “, the 72nd JSME Annual Meeting, No940-30, pp.356-358, 1994
3. Fujiwaka et al., ”Improvement to the Combustion Characteristics of a Lean-Burn
Gas Engine with Prechamber”, the 14th Internal Combustion Engine
Symposium, Japan, pp.91-96, 1997
4. Kobayashi et al., “Experimental Study on General Correlation of Turbulent
Burning Velocity at High Pressure”, the 27th Symposium (International) on
Combustion, pp.941-948, 1998
5. Kawabata et al., “Combustion Diagnostics & Improvement of a Prechamber
Lean-Burn Natural Gas Engine”, SAE Technical Paper Series,
No.2004-01-0979,2004

19

Вам также может понравиться