You are on page 1of 13

Applications in Energy and Combustion Science 15 (2023) 100192

Contents lists available at ScienceDirect

Applications in Energy and Combustion Science


journal homepage: www.sciencedirect.com/journal/applications-in-energy-and-combustion-science

Computational investigation of methanol pre-chamber combustion in a


heavy-duty engine
Xinlei Liu a, *, Priybrat Sharma a, Mickael Silva a, Abdullah S. AlRamadan b, Emre Cenker b,
Qinglong Tang c, Gaetano Magnotti a, Hong G. Im a
a
CCRC, Physical Science and Engineering Division, King Abdullah University of Science and Technology (KAUST), Thuwal, Saudi Arabia
b
Transport Technologies Division, R&DC, Saudi Aramco, Dhahran, Saudi Arabia
c
State Key Laboratory of Engines, Tianjin University, Tianjin, China

A R T I C L E I N F O A B S T R A C T

Keywords: This work explored the potential of methanol pre-chamber combustion (PCC) for heavy-duty engine applications.
Methanol An optical engine experiment was conducted to visualize the jet flame development. The measured pressure
Pre-chamber engine traces and natural flame luminosity images were also used for the validation of three-dimensional computational
Lean-burn
fluid dynamics simulations. It was demonstrated that the main chamber (MC) combustion was successfully
Jet flame
established by the reactive jet issued from the pre-chamber. Compared to methane PCC in our previous study, the
Optical diagnostics
Dual fuel distributed reacting jets were significantly thinner, in particular at the learner condition. The active PCC mode,
which comprises enrichment of the mixture in the pre-chamber (PC) by means of direct methane injection, was
effective in improving the engine performance. However, excessive PC fueling ratio (PCFR) resulted in lower
thermal efficiency due to the higher wall heat transfer and combustion losses. In addition, the effects of various
PC and piston geometries on the methanol/methane PC combustion were evaluated. The combination of an
optimized PC and a flat piston yielded the highest thermal efficiency owing to the relatively lower combustion
and wall heat transfer losses. At engine loads higher than 12.5 bar indicated mean effective pressure, exhaust gas
recirculation must be implemented to avoid end-gas autoignition and reduce nitric oxides (NOx) emissions. As
expected, the increase in (CR) further promoted engine work because of the higher expansion ratio. With CR of
13 and 14, higher thermal efficiency and lower NOx emission were simultaneously achieved under both inter­
mediate and high loads when the engine was operating at the pure methanol PC combustion mode.

major modifications. Therefore, e-methanol is considered a viable


1. Introduction renewable fuel in engine combustion [6]. In addition, methanol is
characterized by high octane number and heat of vaporization, so the
Advanced internal combustion engines (ICEs) will play a significant combustion is less likely to lead to knocking combustion in spark igni­
role in achieving sustainable mobility [1]. In particular, for heavy-duty tion (SI) engines, making it a good alternative fuel for gasoline [7]. A
applications such as long-haul trucks, agricultural machines, and marine relatively high compression ratio (CR) can be implemented when uti­
ships, ICE is nearly irreplaceable and will continue to dominate the lizing the direct-injection stratified-charge method with an overall-lean
market in the foreseeable future, in favor of high power density, low mixture, so a higher engine thermal efficiency is expected compared to
cost, and longevity compared to the emerging alternative powertrains. conventional light-duty gasoline SI engines [8]. On the other hand,
Moreover, the stringent emission regulations and decarbonization goals methanol can also be used in heavy-duty compression ignition (CI) en­
can also be addressed by utilizing renewable biofuels or electro fuels gines [9–13]. Due to the high autoignition temperature of methanol,
(e-fuels) in ICEs [2,3]. however, steady engine operation depends highly on auxiliary means to
Methanol is a promising carrier for hydrogen (H2), which can be promote fuel reactivity, such as intake heating or the addition of a ce­
sustainably produced from captured CO2 and renewable H2 in mass scale tane improver [14]. The challenge in ignition becomes more severe at
[4,5]. In addition, because methanol is liquid at room temperature, it is idle and low-load conditions when the chamber temperature is low.
possible to utilize the current distribution system for petroleum without Pre-chamber combustion (PCC) has gained a lot of attention in recent

* Corresponding author.
E-mail address: xinlei.liu@kaust.edu.sa (X. Liu).

https://doi.org/10.1016/j.jaecs.2023.100192

Available online 19 August 2023


2666-352X/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

Nomenclature ITE indicated thermal efficiency


λ lambda
AHRR apparent heat release rate MC main chamber
AMR adaptive mesh refinement MFB mass fraction of fuel burnt
aTDC after the top dead center MFR mass flow rate
CAD, crank angle degree NFL natural flame luminosity
CFD computational fluid dynamics NOx nitric oxide
CH2O formaldehyde Pin intake pressure
CI compression ignition PC pre-chamber
CO2 carbon dioxide PCC pre-chamber combustion
CR compression ratio PCFR fuel mass ratio within the PC
DoE design of experiment RNG renormalization group
E-fuel electro fuel SI spark-ignition
EGR exhaust gas recirculation ST spark timing
EVO exhaust valve opening T temperature
H2 hydrogen Tavg average temperature
H2O water V velocity
HT heat transfer WSR well-stirred reactor
HTR heat transfer rate 0D zero-dimensional
ICE internal combustion engine 3D three-dimensional

years [15–20] as an enabler of lean-burn high-efficiency engines. It is


considered in this study to solve the ignition issue for methanol engines
under low-load conditions. Compared to the conventional SI, the com­
bustion duration is shortened at the PCC mode, owing to the generated
distributed jet flames from the pre-chamber (PC) [21,22]. In particular,
when the engine is operated at the active PCC mode, more intense
reacting jets are generated, further promoting the combustion speed in
the main chamber (MC) [23]. As a result, the lean-burn limit can be
further extended without the use of the throttle, resulting in higher
thermal efficiency.
Extensive efforts have been performed to provide a fundamental
understanding of the turbulent reacting jet dynamics and optimize the
PC engine performance [20,24–29]. However, most of the previous
works were primarily focused on the applications of fossil fuels such as
gasoline [30,31] or natural gas [32–34]. In recent years, investigations
on renewable oxygenated fuels such as methanol [35,36] and ethanol
[37] are gaining attention because of the imminent net-zero carbon
policy. It has been demonstrated that methanol exhibited a wider
lean-burn limit and higher ITE in contrast to ethanol or gasoline, pri­
marily owing to the higher anti-knocking resistance [36,38]. Therefore,
methanol PC engine should be a promising solution for future
transportation.
Note that for the practical application of the methanol PCC concept
on heavy-duty engines, some key challenges must be addressed, Fig. 1. Schematic of the optical engine configuration.
including the issues of low-load ignition and knock limitation of
maximum torque. Our previous experimental studies [38,39] reported recirculation (EGR) rate, and CR on the combustion and emission
high incomplete combustion losses were generated under ultra-lean characteristics were evaluated. It is anticipated that the results of this
conditions even at the active PCC mode. Although elevating the study will serve as a guide to future powertrain solutions using renew­
compression ratio (CR) may effectively improve the engine performance able fuels.
at low-load operation, a high-load operation is restricted by the
increased tendency of knocking combustion. Consequently, an optimi­ 2. Methodology
zation study is required to compromise various trade-offs, so as to ach­
ieve the best engine performance. 2.1. Optical engine experiment
The present work explored the potential benefits and disadvantages
of PCC on a heavy-duty engine that is operated with methanol fuel over a A new methanol PCC experiment was conducted on a single-cylinder
wide load range. The three-dimensional (3D) computational fluid dy­ optical engine. The measured multi-cycle pressure profiles and averaged
namics (CFD) method was employed for parametric investigation. As natural flame luminosity (NFL) images were utilized for computational
part of the validation process, an optical engine experiment was per­ validations. Schematic of the engine configuration is presented in Fig. 1.
formed to visualize the combustion process. The jet flame dynamics of The engine specification and experimental conditions are summarized in
methanol PCC at both the passive and active modes were examined. Table 1. Further details of the engine are available in [27]. Note that the
Extensive parametric simulations were conducted and the effects of single-cylinder optical engine was modified from a six-cylinder metal
various piston and PC geometries, PCFR values, overall λ, exhaust gas diesel engine, where five cylinders were deactivated by removing their

2
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

Table 1 indicating no auxiliary fuel was supplied within the PC. Methanol was
Engine specification and operating conditions. injected into the intake port with two injectors. Since the port injectors
Type Volvo, four-stroke were originally designed for gasoline injection and the lower heating
value of methanol was almost half of the gasoline, the higher mass flow
Bore/stroke (mm) 131/158
Connecting rod length (mm) 255 rate was required to maintain the same engine load. The cyclic total
Displacement volume (L) 2.1 injected mass was controlled at 250 mg/cycle, resulting in an overall λ of
Geometric CR 11.4 1.1. The spark timing (ST) was fixed at − 15◦ after the top dead center
PC throat diameter (mm) 3.3 (aTDC). In addition, two piezoelectric pressure sensors were mounted in
PC nozzle diameter (mm) 1.5/1.42*
PC nozzle layer number 2/1*
the PC and MC to detect the pressure signals with a resolution of 0.2
PC nozzle number 12/8* crank angle degrees (CADs). The engine was operated for every 20
PC nozzle angle (◦ ) 134/160* consecutive cycles followed by 10 misfire cycles to protect the optical
Engine speed (rpm) 1200 piston. The NFL images were also recorded with a resolution of 0.2
Intake pressure (bar) 1
CADs.
Intake temperature (K) 313
Overall λ 1.1
Spark timing (aTDC) -15 2.2. Computational setup
*
Normal/bold fonts indicate the baseline and optimized PCs. Their
detailed geometries are shown in Fig. 2. Full-cycle CFD simulations were performed utilizing the CONVERGE
code [40]. Each simulation began from the exhaust valve opening (EVO)
valves. As presented in Fig. 1, the original metal piston was replaced by timing and lasted for 720◦ to minimize the impact of internal gas ex­
an extended flat optical piston with a long sleeve, providing optical change on computational uncertainties. The measured intake and
access from the bottom. A high-speed camera was used to record the exhaust pressure traces were imposed on the inlet and outlet boundaries,
combustion process. In addition, to achieve the PCC mode, a PC as­ respectively. A premixed homogenous mixture distribution was assumed
sembly that accommodates the spark plug, auxiliary fuel channel, and a at the inlet for simplifications. The PISO algorithm was applied to solve
piezoelectric pressure sensor was implemented at the original diesel the compressible conservation equations. A variable time-step scheme
injector socket. Owing to the long and narrow throat design of the PC was employed. A unity convective CFL number was utilized. Moreover,
[34], no complex modification of the cylinder head was needed. the maximum Mach CFL number was set to 3 during the jet issuing
In the experiment, the engine speed was fixed at 1200 rpm with an process to ensure computational accuracy.
electric dynamometer. The intake pressure was kept at 1 bar. The intake In addition, the Redlich-Kwong equation of state [41] was applied to
temperature was elevated to 313 K with a heater to attenuate cycle-to- compute gas density. The renormalization group (RNG) k-ε model was
cycle variation. The engine was operating at the passive PCC mode, used to predict turbulence [42]. The O’Rourke heat transfer (HT) model
was utilized to calculate wall HT loss [43]. The well-stirred reactor

Fig. 2. Schematics of the (a) computational domain, (b) PC geometries, and (c) piston profiles for metal engine modeling.

3
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

significantly shorter crevice. The mesh is illustrated in Fig. S1 of the


Supplementary Material. A base grid size of 4.0 mm was used. Fixed
embeddings were imposed in both the PC and nozzle regions. The PC
and nozzle downstream regions were refined with a grid size of 0.25
mm. The spark plug and nozzle surrounding regions were refined with a
grid size of 0.125 mm. Furthermore, the adaptive mesh refinement
(AMR) model was activated. A refined scale of 4 and 3 was implemented
in the MC during the SI stage and subsequently, generating minimum
mesh sizes of 0.25 and 0.5 mm, respectively. As seen in Fig. S1, the jet
flame development was reasonably tracked with AMR. In our previous
study [20], mesh convergence had been achieved with a similar mesh
Fig. 3. Geometry details of the three pistons. configuration.
Besides the baseline narrow-throat PC with two layers of nozzles, the
design of experiment (DoE)-optimized PC with one layer of nozzles was
also applied for comparative analyses. In the numerical work of Silva
et al. [51], the DoE-optimized PC yielded the highest ITE among 300
different PC designs for methane PC engine combustion. The specifica­
tion and geometries for these two PCs are presented in Table 1 and Fig. 2
(b), respectively. The optimized PC has a larger throat diameter and only
one layer of nozzles. In addition, the nozzle diameter is relatively
smaller but the included angle is slightly wider in comparison to the
baseline case. Furthermore, to assess the effect of piston geometry on the
methanol PCC, the ω and U pistons were also simulated, as depicted in
Fig. 3. The ω piston was utilized on our metal PC engine [38] and the U
piston was found to yield the highest ITE among seven different pistons
for methane PCC [52]. Note that the optimized PC and piston that
worked well for methane PCC might not be the optimal choice for
methanol PCC. Therefore, further evaluations were still required to
identify the optimal combination for methanol PCC.

Fig. 4. Comparison of the measured and predicted pressure traces in the (a) PC 2.3. Computational validations
and (b) MC. Gray lines represent the multiple pressure traces.
Fig. 4 compares the measured and predicted pressure traces within
(WSR) combustion solver SAGE [44] coupled with a reduced methanol the PC and MC. The measured multi-pressure traces were also added to
chemistry mechanism [45] was adopted to model the combustion pro­ show the cycle-to-cycle variation. An overall reasonable agreement with
cess. NOx emissions were predicted with an extended Zel’dovich the experiment was obtained. Although the modeling gave slight over­
mechanism [46], which has been calibrated in our previous work for predictions of the peak combustion pressure in both the PC and MC, the
methane PCC modeling [20]. Note that the SAGE combustion solver is discrepancies were still acceptable considering the cyclic variation. Two
not only able to adequately solve the CI problems [47], but it is also able peaks were seen for the PC pressure trace. The first peak was due to the
to provide reasonable predictions for the SI problems [48,49], where the flame propagation and rapid pressure buildup within the PC, while the
turbulence-chemistry interaction is important. Pomraning et al. [50] second one was due to the subsequent flame propagation and pressure
reported that the numerical and turbulence-chemistry interaction errors rise within the MC. Fig. 5 presents the measured averaged NFL images
are attenuated utilizing a fine mesh resolution for RANS simulations. and predicted projected density of hydroxyl radical (OH). Our previous
Consequently, in this work, the SAGE combustion solver was employed work [29] has demonstrated that the line-of-sight OH distribution was
in all simulations with a reasonably fine mesh setup to model the able to qualitatively represent the jet flame development. As seen in
methanol PCC. Fig. 4, the jet flame was reasonably captured by the modeling work,
Fig. 2(a) presents the computational domain for metal engine despite a slight lag-behind was seen. Given the NFL images were
modeling. The only major difference between the metal and optical 20-cycle averaged results, the 1 CAD discrepancy still made sense. Note
engine setups lies in their crevice height, with the former case having a that reacting jets for methanol PCC were thinner compared to the active

Fig. 5. Comparison of the measured NFL images and predicted projected density of OH.

4
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

3. Results and discussion

In this work, the effect of overall λ on jet flame dynamics was first
evaluated on the optical engine. Subsequently, an extensive modeling
investigation was performed to optimize the performance of the metal
engine. Multi-design parameters including the PC and piston geome­
tries, PCFR values, EGR ratio, and CR were evaluated. Three engine
loads ranging from low to high load were considered, corresponding to
an indicated mean effective pressure (IMEP) at about 5.5, 12.5, and 19.5
bar, respectively.

3.1. Effects of overall λ


Fig. 6. Predicted pressure traces with the variation of λ.
Before performing the optimization study, the effect of overall λ on
the passive methanol PCC was evaluated on the optical engine. For
comparative analyses, two additional cases with λ = 1.5 and 2.0 were
conducted based on the validation case with λ = 1.1. These three cases
(λ = 1.1, 1.5, and 2.0) had a cyclic fuel mass at 245.9, 189.9, and 145.5
mg, corresponding to an IMEP at about 9.0, 7.0, and 5.3 bar with an
assumed ITE of 40%. Fig. 6 presents the predicted pressure traces for
both the MC and PC with the variation of λ. As expected, a leaner
mixture yielded a lower pressure build-up within the PC and hence the
postponed combustion phasing of the MC. Note that a misfire occurred
at λ = 2.0, indicating the active PCC mode may be needed under low-
load conditions.
To further clarify the jet flame dynamics with the variation of λ, the
predicted traces of mass flow rate (MFR) between the PC and MC and
apparent heat release rate (AHRR) were presented in Fig. 7. Two jet
stages (cold and hot) were identified based on the variations of MFR and
temperature (T). As shown in Figure S2, the cold jet stage started with a
positive MFR from the PC to the MC until a high-temperature jump
occurred, followed closely by the hot jet stage that lasted until when the
MFR became negative. Note that, similar to the methane PCC [29], the
upper layer of nozzles always yielded a significantly lower MFR than the
lower layer of nozzles, indicating no jet flame was distributed from
there. This finding was further confirmed by Figure S3, showing the
predicted projected density of OH with the variation of λ. As expected,
Fig. 7. Predicted traces of mass flow rate between the PC and MC and AHRR in
only six reacting jets were generated for both cases. Note that the OH
the PC and MC at (a) λ = 1.1 and (b) 1.5.
formation was significantly weaker at λ = 1.5 for the less intense heat
release.
methane PCC [29], primarily owing to the significantly lower heating Based on the traces of AHRR in the PC and MC, three different
value of methanol. As a result, the flame was seen to propagate rapidly combustion stages were identified, including the PC flame propagation-
only after the jet flame-piston interaction. Because of the dominated stage, mixed mode stage, and MC flame propagation-
near-stoichiometric condition, a fast MC flame propagation process was dominated stage. As presented in Fig. 7, the PC flame propagation
generated. Within just five CADs, the flame has almost filled up the stage started from the ST until when AHRR in the PC was exceeded by
whole combustion chamber. the MC, covering the whole cold jet stage and part of the hot jet stage.
The mixed mode indicated the MC combustion heat release was domi­
nated by both the jet flame and MC flame propagation process. In the

Fig. 8. Predicted distributions of CH2O and OH at (a) λ = 1.1 and (b) 1.5.

5
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

Fig. 9. Predicted distributions of CH2O and V with λ = 1.1 and 1.5. The red line represents T = 1000 K.

Fig. 10. Predicted T-ϕ-V/CH2O distributions with (a) λ = 1.1 and (b) 1.5 at various timings.

present work, the end of the hot jet stage was selected as the start when calculation. At both λ, a negligible amount of CH2O was yielded when T
the MC flame propagation dominated. In contrast to the case with λ = < 1200 K because of the long ignition delay. Tremendous CH2O was
1.1, a longer PC flame propagation stage was yielded with λ = 1.5, owing accumulated when T > 1250 K. But when the temperature was overly
to the slower flame propagation process. For the same reason, the high (> 1450 K), CH2O was instantly consumed, generating abundant
significantly postponed MC flame propagation stage was generated. OH and a drastic heat release. As expected, the case with λ = 1.5 yielded
Fig. 8 shows additional details of the flame development at two λ. a lower peak concentration of CH2O owing to the leaner condition.
The predicted distributions of formaldehyde (CH2O) and OH were uti­ The zero-dimensional (0D) simulation results agreed with the pre­
lized to clarify the flame structure. The reacting jet impingement dicted T-V-CH2O distributions from CFD simulations as presented in
occurred just within about 1 CAD after the ejection of reacting jets. Fig. 10. On the other hand, in CFD modeling cases, high concentrations
Owing to the slower PC flame propagation process and hence lower of CH2O also were generated when T > 1450 K. It was because the
pressure build-up within the PC at λ = 1.5, the discharge of the reacting convection term also acted a significant role in species transport. Due to
jets was about 2 CADs later than the case at λ = 1.1. In addition, the the super-high jet velocity, the generated CH2O in the upstream jet was
leaner case yielded a significantly lower OH concentration on both the fully oxidized and immediately transported downstream. As evidenced
reacting jets and flame fronts. As expected, for both cases, CH2O was in Fig. 9 at − 8.7◦ with λ = 1.1 and − 6.4◦ with λ = 1.5, some CH2O was
primarily formed in the flame surroundings, where the temperature was formed downstream of the jet tip, which was primarily transported from
relatively lower. On the other hand, OH was mainly generated on or the upstream rather than generated by autoignition reactions, because
within the high-temperature flame surfaces. In contrast, the leaner case the local temperature was relatively low.
yielded significantly less OH due to the lower combustion temperature.
It is of interest to find whether the MC combustion was initiated by 3.2. Low-load operation
the jet flame or reactive intermediate species. To clarify this, the pre­
dicted distributions of CH2O and velocity (V) around the jet flame-piston The low-load operation has been a challenge for conventional SI
interaction instant at different λ were presented in Fig. 9. The isolines engines. Because a small throttle opening angle is required to enrich the
with T = 1000 K were depicted to show the high-temperature reacting fuel-air mixture, high pumping loss, and low air exchange rate were
regions. For both cases, CH2O was primarily located within the high- generated, significantly deteriorating the fuel economy. In comparison,
temperature pockets. Note that when the reacting jet was issued (at due to the distributed jet flame feature, the PCC mode can operate
− 8.9◦ with λ = 1.1 and − 6.6◦ with λ = 1.5), the CH2O from both PCs had smoothly even at an ultra-lean condition, reliving the use of the throttle.
already been depleted. This implies that the initial formation of CH2O Therefore, the PCC mode is expected to yield better performance
within the MC originated from the instantaneous high-temperature re­ compared to the SI mode at low loads.
action, i.e. the MC combustion was initiated by the high temperature This subsection intends to optimize the performance of the methanol
rather than intermediate species. To support this finding, Figure S4 PC engine operated at an IMEP of about 5.5 bar. The previous case with
presents the predicted MSF of CH2O with the variation of initial T with a λ = 2.0 was taken as the baseline case. Note that since the misfire
residence time of 0.1 deg at 1200 rpm and ambient pressure of 20 bar. occurred at λ = 2.0, the active PCC mode must be applied. Considering
The SENKIN code within the CHEMKIN package [53] was used for the the good mixing quality, gaseous methane was supplied as an auxiliary

6
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

fuel burnt (MFB), energy balance, and NOx emission with various PCs
and PCFR values. Similar to our previous work on methane PCC [52],
the best DoE PC with a small PCFR value also yielded the highest ITE for
methanol PCC. Comparatively, the cases with the baseline
narrow-throat PC yielded a higher wall HT loss than the best DoE PC,
primarily owing to the overall near-stoichiometric mixture distribution
within the PC (Fig. 12), leading to the higher pressure build-up within
the PC and more intense jet flame-piston interaction. For the same
reason, more advanced combustion phasing and higher peak combus­
tion temperature were generated, resulting in more complete combus­
tion but more NOx emission than the cases with the best DoE PC. In
addition, the NOx emission followed the same trend as ITE with the
variation of PCFR value. The cases with PCFR = 5% yielded higher NOx
emissions owing to the higher peak combustion temperature within the
PC.
Compared to the cases utilizing the baseline PC, various PCFR values
exhibited more impact on the pressure rise within the PC utilizing the
best DoE PC. Fig. 13 shows that a larger amount of methane was trapped
within the PC with the best DoE PC. As such, the overall richer mixture
distribution and smaller pressure build-up were generated compared to
the cases with the baseline PC. Note that due to the wide-throat design,

Fig. 11. Predicted (a) traces of pressure and MFB and (b) energy balance and
NOx emission with various PCs and PCFR values.

fuel within the PC at − 360◦ aTDC. Three PCFR values (5, 10, and 15%)
were used and the injection profiles were referenced from our previous
work [54]. The PCFR value was defined as the auxiliary fuel energy of
methane divided by the total fuel energy. The effects of PC (baseline and
best DoE) and piston (flat, ω, and U) geometries on the combustion and
emission characteristics were also assessed.

3.2.1. Effects of PCFR value and PC geometry Fig. 13. Predicted traces of methane mass in PC with various PCs and
Fig. 11 shows the predicted traces of pressure and mass fraction of PCFR values.

Fig. 12. Predicted distributions of λ, PDF of λ, and distributions of T (isosurface at 1600 K) and TKE with various PCs and PCFR values.

7
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

the flow reversal from the MC to the PC is directed toward the spark plug
without sufficient mixing, as shown in Fig. 12. This transient and un­
predictable flow contributed to uncertainties in the SI process. There­
fore, the PC combustion phasing and jet flame feature were more easily
affected by the PCFR value using the DoE-optimized PC. On the other
hand, although the case with a smaller PCFR value resulted in the overall
leaner mixture distribution and hence the faster flame propagation
within the PC, the fuel energy within the PC was reduced accordingly.
These two trade-off factors eventually led to similar pressure rise levels
and jet flame features with the variation of PCFR value.
Fig. 14 compares the predicted distributions of T on the liner, piston,
and central cut plane at 0◦ and 10◦ aTDC with two different PCs. The
predicted traces of HTR were presented in Figure S5. Owing to the
intense reacting jet issuing process, the case with the baseline PC yielded
a significantly wider flame surface and thus the higher wall HT loss
compared to the best DoE PC. At 10◦ aTDC, the flame in the former case
had already propagated to cover a wider crevice region. In contrast, a
large area of low-temperature pockets was generated in the best DoE PC,
resulting in higher incomplete combustion loss eventually. Furthermore,
as shown in Fig. 11(b), the higher PCFR value tended to generate a
higher incomplete combustion loss, in particular at PCFR = 15%. This
Fig. 14. Predicted distributions of T on the liner, piston, and central cut plane was because λ of the bulk mixture was elevated in order to keep the
at 0◦ and 10◦ aTDC with various PCs. overall λ unchanged, slowing down the oxidization process.

3.2.2. Effects of piston geometry


The effects of three different pistons were further assessed. The ge­
ometry details of these three pistons are shown in Fig. 3. The previous
two cases with different PCs at PCFR = 5% were taken as the reference
cases. Fig. 15(a) presents the predicted traces of pressure and MFB. The
piston geometry is seen to affect the combustion processes in both the PC
and MC. The flat and ω pistons resulted in a more advanced combustion
phasing within the PC. Fig. 16 reveals that the flow field and mixture
distribution were affected by the piston geometry, although the PC ge­
ometry exhibited a more significant impact. In comparison, various
piston geometries exhibited the more significant influence for the three
cases with the best DoE PC. As shown by the predicted distribution
contours of λ for various cases, the cases with the baseline PC exhibited
the overall leaner and relatively more homogeneous mixture distribu­
tions than the three best DoE PC cases. Therefore, a similarly high-
pressure rise and intense jet flame were generated. In contrast, the
cases with the best DoE PC were very sensitive to the flow field, in
particular near the spark plug. The U piston case yielded a higher pro­
portion of lean mixture, resulting in the most postponed pressure build-
up and jet issuing processes.
Fig. 15(b) shows the predicted energy balance and NOx emission. As
expected, due to the higher wall HT loss, the cases with the baseline PC
yielded the overall lower ITE compared to the cases with the best DoE
PC. For both sets of cases, the highest but lowest ITE was generated
utilizing the U and ω piston, respectively. The ω piston coupled to the
baseline PC resulted in the highest incomplete combustion loss, followed
by the U piston coupled to the best DoE PC. For both sets of cases, the
Fig. 15. Predicted (a) traces of pressure and MFB and (b) energy balance and
NOx emission with various PCs and pistons.
lowest NOx emission was generated utilizing the flat piston. For eluci­
dations, Fig. 17 presents the predicted temperature distributions on the
piston and liner walls, as well as the central cut plane. Note that the jet

Fig. 16. Predicted distributions of λ with various PCs and pistons.

8
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

Fig. 17. Predicted distributions of T on the liner, piston, and central cut plane at 0◦ and 10◦ aTDC with various PCs and pistons.

flame-piston interaction was significantly affected by the different pis­


ton designs. For the three cases with the baseline PC, most of the jet
flame was effectively confined within the inner piston using the ω pis­
ton, leading to its lowest wall HT loss. On the other hand, oxidization of
the mixture within the crevice region was significantly slowed down,
resulting in the highest incomplete loss and lowest ITE eventually. Since
the jet flame was partially confined within the inner piston, the U piston
yielded the lower wall HT loss and hence higher ITE and NOx emission
than the flat piston.
On the other hand, for the three cases utilizing the best DoE PC, the
highest wall HT loss and thus lowest ITE were obtained with the ω
piston. This was because its reacting jets were directly impinged on the
inner piston edge, significantly enhancing the HTR from the piston.
However, the flame propagation within the squish region was enhanced,
promoting the oxidization of the mixture in the crevice and reducing the
incomplete loss compared to the case with the baseline PC. As expected,
in contrast to the flat piston, the U piston was able to partially confine
the reacting jets within the piston, expediting the overall combustion
process and NOx formation but reducing the wall HT loss from the cyl­
inder head. Nevertheless, for this same reason, the flame propagation
within the squish and crevice regions was slowed down and hence the
higher incomplete loss was generated.

3.3. Mid-load operation

The mid-load operation was investigated at two overall λ, i.e. 1.5 and
Fig. 18. Predicted (a) traces of pressure and MFB and (b) energy balance and
2.0. The previous case using the flat piston and best DoE PC was taken as
NOx emission with various λ and STs.
the baseline case due to the highest ITE and reasonably low incomplete
combustion loss. For both λ, the intake pressure was elevated accord­
ingly to maintain the same total fuel energy and hence achieve the same

9
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

Fig. 19. Predicted distributions of CH2O and OH at various timings with (a) λ = 1.5 and (b) 2.0.

Table 2
Parameter details.
Cases EGR [%] Pin ST
[bar] [aTDC]

CR=11 25 2.75 − 15
CR=12 28 2.82 − 15
CR=13 30 2.88 − 10
CR=14 30 2.88 − 5

the predicted distributions of CH2O and OH at various CADs. Compared


to the results in Fig. 8(b), the active PCC mode effectively enhanced the
generation of reacting jets, as evidenced by the continuously wider jet
flame feature. Nevertheless, the peak concentrations of CH2O and OH
were reduced because of the implementation of EGR resulting in a lower
combustion temperature. The combustion temperature was further
reduced at λ = 2.0, with a lower concentration of OH and higher CH2O
on the jet flame compared to λ = 1.5. To further explain why the leaner
case yielded the higher wall HT loss as shown in Fig. 18(b), the predicted
distributions of T on the liner, piston, and central cut plane were pre­
sented in Fig. 20. Due to the more advanced ST, the flame surface area at
λ = 2.0 was significantly larger than λ = 1.5. As a result, despite the

Fig. 20. Predicted distributions of T on the liner, piston, and central cut plane
at 0◦ and 10◦ aTDC with various λ.

IMEP at about 12.5 bar. Note that in the low-load baseline case, the NOx
emission (0.42 g/kW-h) was already below the EURO VI emission
regulation limit (0.46 g/kW-h). Therefore, EGR must be employed to
inhibit NOx formation at the mid-load operation. Considering that the
larger EGR rate would result in higher incomplete combustion loss, the
EGR rate for both cases was kept at less than 30%. Furthermore, the ST
was also swept to identify the optimal operating point that yielded the
highest ITE but relatively low NOx emission.
Fig. 18 presents the predicted traces of pressure and MFB, energy
balance, and NOx emission. Due to the lower EGR rate at λ = 2.0, the
more advanced combustion phasing and thus higher ITE were generated
compared to λ = 1.5, despite the promoted wall HT loss. In addition,
lower NOx emissions were yielded with the leaner cases, mainly owing
to the lower combustion temperature. The predicted ITE exhibited a
monotonically growing trend with the advancement of ST at λ = 1.5.
However, restricted by the increased NOx emission, the optimal ST was
limited to − 15◦ aTDC, yielding an ITE of 43.9%. On the other hand, the
predicted ITE at λ = 2.0 saw a first growing and then a declining trend.
The optimal operating point with the highest ITE of 45.5% was obtained
at ST = − 20◦ aTDC. Further advancement of ST resulted in the promoted
wall HT loss and reduced ITE.
The jet flame structures at two λ were further analyzed. Fig. 19 shows Fig. 21. Predicted (a) traces of pressure and MFB and (b) energy balance and
NOx emission with the variation of CR.

10
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

Fig. 22. Predicted distributions of T on the liner, piston, and central cut plane at 0◦ to30◦ aTDC with various CRs.

lower peak combustion temperature, the higher wall HT loss was high-load study with an IMEP of about 19.5 bar. To avoid knocking and
generated at λ = 2.0 eventually. end-gas autoignition issues, a passive PCC mode was applied. The effects
of four CRs ranging from 11 to 14 were evaluated. At each CR, the
corresponding EGR rate, intake pressure (Pin), and ST were adjusted
3.4. High-load operation jointly. As shown in Table 2, with the growth in CR, the higher EGR rate
and Pin and postponed ST were adopted to attenuate end-gas auto­
High-load operation is another significant challenge for heavy-duty ignition and NOx emission.
applications, in particular for a PC engine. To ensure the engine is Fig. 21 presents the predicted traces of pressure and MFB, energy
operating at the lean-burn condition, a high intake pressure must be balance, and NOx emission with the variation of CR. As expected, the
utilized. However, the highest achievable intake pressure and hence overall higher ITE was generated at a larger CR owing to the promoted
lean-burn limit are restricted by the capability of the turbocharger. In expansion ratio and thus reduced exhaust loss, despite the increased
addition, due to the higher combustion temperature and pressure, the wall HT loss. Note that the ITE was reduced at an over-high CR (14)
risk of end-gas autoignition and NOx emission is also increased. There­ because a postponed ST was applied to inhibit end-gas autoignition. In
fore, the EGR rate and ST should be adjusted simultaneously to achieve addition, NOx emission and incomplete combustion loss both saw a
optimal engine performance. monotonically declining trend with the growth in CR.
In the current work, considering the limit of intake pressure on a For clarification, Fig. 22 compares the predicted distributions of T on
practical engine, an overall λ of 1.5 instead of 2.0 was utilized for the the liner, piston, and central cut plane at 0◦ to 30◦ aTDC. The predicted
traces of HTR, Tavg, NOx, and THC are presented in Figure S6. Due to the
Table 3 increased EGR rate and postponed ST at a larger CR, the combustion
Parameter details at various loads. temperature was generally reduced, resulting in lower NOx emission.
Load λ PCFR [%] EGR [%] Pin ST However, since the squish height was minimized, the jet flame-wall
[bar] [aTDC] interactions were promoted, enhancing the wall HT loss. On the other
Low 2.0/2.0 5/5 18/20 1.0/1.0 − 15/− 12 hand, the reduced squish height promoted flame propagation within the
Mid 1.5/1.5 0/0 26.5/27.5 1.86/1.86 − 13/− 8 crevice region during the post-combustion stage (after 20◦ aTDC),
High 1.5/1.5 0/0 30/30 2.88/2.88 − 10/− 5 leading to fewer low-temperature pockets and more complete
Note: normal and bold fonts represent CR at 13 and 14, respectively. combustion.

11
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

as follows:

(1) Relatively efficient and clean combustion process was achieved


with methanol PCC at three different engine loads. Under the
conditions considered, the optimal engine performance was ob­
tained at CR = 14. The implementation of an even higher CR was
restricted by the increased risk of end-gas autoignition in the
squish region.
(2) Due to the smaller heating value of methanol compared to
methane, methanol PCC generated significantly thinner reacting
jets than methane PCC. Therefore, to stabilize the low-load
operation and reduce incomplete combustion loss, the active
methanol PCC mode was necessary.
(3) Similar to the pure methane PCC mode, the highest ITE was
achieved by utilizing the best DoE PC with a small PCFR value.
However, an excessive PCFR value increased incomplete com­
bustion loss because of the over-lean bulk mixture.
(4) Although a relatively high thermal efficiency was obtained with
the U piston, a high incomplete combustion loss was yielded. This
was due to the intense jet flame impingement on the inner piston
edge that slowed down the oxidization process within the crevice
region.
(5) EGR implemented at the mid- and high-load conditions was
effective in mitigating end-gas autoignition and NOx formation.
In addition, a leaner operation further promoted engine work and
lowered NOx emission because of the reduced combustion tem­
Fig. 23. Predicted (a) traces of pressure and MFB and (b) energy balance and perature and more advanced ST.
NOx emission with various CRs from low- to high-load conditions.
Declaration of Competing Interest
3.5. Overall performance evaluation at CR of 13 and 14
The authors declare that they have no known competing financial
The overall performance with CR of 13 and 14 at three engine loads interests or personal relationships that could have appeared to influence
was evaluated. In each case, the corresponding ST has been swept to the work reported in this paper.
achieve optimal performance. Table 3 summarizes details of the major
control parameters in simulations. With the growth in engine load, a Data availability
higher EGR rate was employed to inhibit NOx emission and end-gas
autoignition. Note that to reduce the use of auxiliary fuel supply, the Data will be made available on request.
passive PCC mode was utilized at the mid- and high-load conditions.
Fig. 23 shows the predicted traces of pressure and MFB, energy
balance, and NOx emission from low- to high-load conditions at two CRs. Acknowledgments
From the low to high load, a higher ITE was generally obtained owing to
a smaller wall HT loss during the transition. Note that there was a sig­ This paper is based on work supported by Saudi Aramco Research
nificant decline in wall HT loss when transiting from the low to mid load. and Development Center FUELCOM program under Master Research
Because no auxiliary fuel was supplied in the PC at mid load, less intense Agreement Number 6600024505/01. FUELCOM (Fuel Combustion for
reacting jets were generated, considerably attenuating the jet flame- Advanced Engines) is a collaborative research undertaking between
piston interaction. As expected, compared to the cases with CR = 13, Saudi Aramco and KAUST intended to address the fundamental aspects
the relatively lower incomplete loss was yielded at CR = 14, primarily of hydrocarbon fuel combustion in engines, and develop fuel/engine
due to the narrower squish height that promoted the combustion in the design tools suitable for advanced combustion modes. The computa­
crevice. Furthermore, because of the higher tendency to lead to end-gas tional simulations utilized the clusters of the KAUST Supercomputing
autoignition at CR =14, the ST was further postponed compared with Laboratory. The authors thank Convergent Science Inc. for providing the
CR = 13. Therefore, slightly smaller ITE and NOx emissions were CONVERGE license.
generated at the high load. On the other hand, higher ITEs were ob­
tained at low and mid loads with CR = 14, because of the increased Supplementary materials
expansion ratio as well as the higher pressure build-up in the PC that
resulted in the faster combustion process in the MC. Supplementary material associated with this article can be found, in
the online version, at doi:10.1016/j.jaecs.2023.100192.
4. Conclusions
References
This work conducted a computational investigation on methanol
PCC for heavy-duty engine applications. An optical engine combustion [1] Reitz RD, Ogawa H, Payri R, Fansler T, Kokjohn S, Moriyoshi Y, Agarwal A, et al.
IJER editorial: the future of the internal combustion engine. Int J Engine Res 2020;
experiment was performed to visualize the jet flame details. Extensive 21:3–10.
simulations were performed to optimize the engine performance from [2] Ramirez A, Sarathy SM, Gascon J. CO2 derived E-Fuels: research trends,
low- to high-load operations. The effects of overall λ, piston and PC misconceptions, and future directions. Trends Chem 2020;2:785–95.
[3] Xu L, Bai XS. Numerical investigation of engine performance and emission
geometries, PCFR values, EGR, and CR on the engine combustion
characteristics of an ammonia/hydrogen/n-heptane engine under RCCI operating
characteristics and emissions were evaluated. The major conclusions are conditions. Flow Turbul Combust 2023. https://doi.org/10.1007/s10494-023-
00453-y.

12
X. Liu et al. Applications in Energy and Combustion Science 15 (2023) 100192

[4] Tountas AA, Ozin GA, Sain MM. Solar methanol energy storage. Nature Catalysis [29] Liu X, Marquez MEcheverri, Sanal S, Silva M, AlRamadan AS, Cenker E, Sharma P,
2021;4:934–42. et al. Computational assessment of the effects of pre-chamber and piston
[5] Zang G, Sun P, Elgowainy A, Wang M. Technoeconomic and Life Cycle Analysis of geometries on the combustion characteristics of an optical pre-chamber engine.
Synthetic Methanol Production from Hydrogen and Industrial Byproduct CO2. Fuel 2023;341:127659.
Environ Sci Technol 2021;55:5248–57. [30] Attard WP, Fraser N, Parsons P, Toulson E. A turbulent jet ignition pre-chamber
[6] Matamis A, Lonn S, Luise L, Vaglieco BM, Tuner M, Andersson O, Alden M, et al. combustion system for large fuel economy improvements in a modern vehicle
Optical characterization of methanol compression-ignition combustion in a heavy- powertrain. SAE Int J Engines 2010;3:20–37.
duty engine. Proc Combust Inst 2021;38:5509–17. [31] Hua J, Song Y, Zhou L, Liu F, Wei H. Operation strategy optimization of lean
[7] Li Y, Bai XS, Tunér M, Im HG, Johansson B. Investigation on a high-stratified direct combustion using turbulent jet ignition at different engine loads. Appl Energy
injection spark ignition (DISI) engine fueled with methanol under a high 2021;302:117586.
compression ratio. Appl Therm Eng 2019;148:352–62. [32] Shah A, Tunestal P, Johansson B. Effect of relative mixture strength on
[8] Gong C, Yi L, Zhang Z, Sun J, Liu F. Assessment of ultra-lean burn characteristics performance of divided chamber ‘avalanche activated combustion’ ignition
for a stratified-charge direct-injection spark-ignition methanol engine under technique in a heavy duty natural gas engine. SAE Tech Pap 2014. https://doi.org/
different high compression ratios. Appl Energy 2020;261:114478. 10.4271/2014-01-1327.
[9] Yao C, Hu J, Geng P, Shi J, Zhang D, Ju Y. Effects of injection pressure on ignition [33] Shah A, Tunestal P, Johansson B. Effect of pre-chamber volume and nozzle
and combustion characteristics of diesel in a premixed methanol/air mixture diameter on pre-chamber ignition in heavy duty natural gas engines. SAE Tech Pap
atmosphere in a constant volume combustion chamber. Fuel 2017;206:593–602. 2015. https://doi.org/10.4271/2015-01-0867.
[10] Li Y, Jia M, Liu Y, Xie M. Numerical study on the combustion and emission [34] Hlaing P, Echeverri Marquez M, Singh E, Almatrafi F, Ben Houidi M, Johansson B,
characteristics of a methanol/diesel reactivity controlled compression ignition Cenker E. Effect of pre-chamber enrichment on lean burn pre-chamber spark
(RCCI) engine. Appl Energy 2013;106:184–97. ignition combustion concept with a narrow-throat geometry. SAE Tech Pap 2020.
[11] Aljabri H, Liu X, Al-lehaibi M, Cabezas KM, AlRamadan AS, Badra J, Im HG. Fuel https://doi.org/10.4271/2020-01-0825.
flexibility potential for isobaric combustion in a compression ignition engine: a [35] Wang B, Xie F, Hong W, Du J, Chen H, Su Y. The effect of structural parameters of
computational study. Fuel 2022;316:123281. pre-chamber with turbulent jet ignition system on combustion characteristics of
[12] Liu X, Wang H, Zheng Z, Liu J, Reitz RD, Yao M. Development of a combined methanol-air pre-mixture. Energy Convers Manag 2022;274:116473.
reduced primary reference fuel-alcohols (methanol/ethanol/propanols/butanols/ [36] Harrington A, Hall J, Bassett M, Cooper A. Effect of jet ignition on lean methanol
n-pentanol) mechanism for engine applications. Energy 2016;114:542–58. combustion using high compression ratio. SAE Tech Pap 2023. https://doi.org/
[13] Xu L, Treacy M, Zhang Y, Aziz A, Tuner M, Bai XS. Comparison of efficiency and 10.4271/2023-01-0319.
emission characteristics in a direct-injection compression ignition engine fuelled [37] Almatrafi F, Hlaing P, Echeverri Marquez M, Ben Houidi M, Johansson B. Narrow-
with iso-octane and methanol under low temperature combustion conditions. Appl throat pre-chamber combustion with ethanol, a comparison with methane. SAE
Energy 2022;312:118714. Tech Pap 2020. https://doi.org/10.4271/2020-01-2041.
[14] Liu H, Zhang X, Zhang Z, Wu Y, Wang C, Chang W, Zheng Z, et al. Effects of 2- [38] Hlaing P, Echeverri Marquez M, Burgos P, Cenker E, Ben Houidi M, Johansson B.
ethylhexyl nitrate (EHN) on combustion and emissions on a compression ignition Analysis of fuel properties on combustion characteristics in a narrow-throat pre-
engine fueling high-pressure direct-injection pure methanol fuel. Fuel 2023;341: chamber engine. SAE Tech Pap 2021. https://doi.org/10.4271/2021-01-0474.
127684. [39] Hlaing P, Said A, Cenker E, Im HG, Turner J. Comparing unburned fuel emission
[15] Hua J, Zhou L, Gao Q, Feng Z, Wei H. Influence of pre-chamber structure and from a pre-chamber engine operating on alcohol fuels using FID and FTIR
injection parameters on engine performance and combustion characteristics in a analyzers. SAE Tech Pap 2022. https://doi.org/10.4271/2022-01-1094.
turbulent jet ignition (TJI) engine. Fuel 2021;283:119236. [40] K. Richards, P. Senecal, E. Pomraning, Converge (v3.0), Madison (WI): Convergent
[16] Rajasegar R, Niki Y, García-Oliver JM, Li Z, Musculus MPB. Fundamental insights Science.
on ignition and combustion of natural gas in an active fueled pre-chamber spark- [41] Redlich O, Kwong JN. On the thermodynamics of solutions. V. an equation of state.
ignition system. Combust Flame 2021;232:111561. Fugacities of gaseous solutions. Chem Rev 1949;44:233–44.
[17] Tang Q, Sampath R, Marquez ME, Sharma P, Hlaing P, Houidi MB, Cenker E, et al. [42] Han Z, Reitz RD. Turbulence modeling of internal combustion engines using RNG
Optical diagnostics on the pre-chamber jet and main chamber ignition in the active κ-ε models. Combust Sci Technol 1995;106:267–95.
pre-chamber combustion (PCC). Combust Flame 2021;228:218–35. [43] O’Rourke PJ, Amsden AA. A spray/wall interaction submodel for the KIVA-3 wall
[18] Yu X, Zhang A, Baur A, Engineer N. The impact of pre-chamber design on part load film model. SAE Tech Pap 2000. https://doi.org/10.4271/2000-01-0271.
efficiency and emissions of a miller cycle light duty gasoline engine. SAE Tech Pap [44] Senecal PK, Pomraning E, Richards KJ, Briggs TE, Choi CY, McDavid RM,
2021. https://doi.org/10.4271/2021-01-0479. Patterson MA. Multi-dimensional modeling of direct-injection diesel spray liquid
[19] Silva MR, Ben Houidi M, Hlaing P, Sanal S, Cenker E, AlRamadan A, Chang J, et al. length and flame lift-off length using CFD and parallel detailed chemistry. SAE
The effects of piston shape in a narrow-throat pre-chamber engine. SAE Tech Pap Tech Pap 2003. https://doi.org/10.4271/2003-01-1043.
2022. https://doi.org/10.4271/2022-01-1059. [45] Liu X, Wang H, Zheng Z, Yao M. Development of a reduced primary reference fuel-
[20] Liu X, Aljabri H, Silva M, AlRamadan AS, Ben Houidi M, Cenker E, Im HG. PODE3-methanol-ethanol-n-butanol mechanism for dual-fuel engine simulations.
Hydrogen pre-chamber combustion at lean-burn conditions on a heavy-duty diesel Energy 2021;235:121439.
engine: a computational study. Fuel 2023;335:127042. [46] Lavoie GA, Heywood JB, Keck JC. Experimental and theoretical study of nitric
[21] Gussak L. High chemical activity of incomplete combustion products and a method oxide formation in internal combustion engines. Combust Sci Technol 1970;1:
of prechamber torch ignition for avalanche activation of combustion in internal 313–26.
combustion engines. SAE Techn Pap 1975:750890. [47] Zdanowicz A, Mohr J, Tryner J, Gustafson K, Windom B, Olsen DB, Hampson G,
[22] Glasson N, Lumsden G, Dingli R, Watson H. Development of the HAJI system for a et al. End-gas autoignition fraction and flame propagation rate in laser-ignited
multi-cylinder spark ignition engine. SAE Tech Pap 1996:961104. https://doi.org/ primary reference fuel mixtures at elevated temperature and pressure. Combust
10.4271/961104. Flame 2021;234:111661.
[23] Sampath R, Tang Q, Echeverri Marquez M, Sharma P, Hlaing P, Ben Houidi M, [48] Gholamisheeri M, Thelen B, Gentz G, Toulson E. CFD modeling of an auxiliary
Cenker E, et al. Study on the pre-chamber fueling ratio effect on the main chamber fueled turbulent jet ignition system in a rapid compression machine. SAE Tech Pap
combustion using simultaneous PLIF and OH* chemiluminescence imaging. SAE 2016. https://doi.org/10.4271/2016-01-0599.
Int J Adv Curr Pract Mobil 2020;3:137–49. [49] Gholamisheeri M, Wichman IS, Toulson E. A study of the turbulent jet flow field in
[24] Allison PM, de Oliveira M, Giusti A, Mastorakos E. Pre-chamber ignition a methane fueled turbulent jet ignition (TJI) system. Combust Flame 2017;183:
mechanism: experiments and simulations on turbulent jet flame structure. Fuel 194–206.
2018;230:274–81. [50] Pomraning E, Richards K, Senecal P. Modeling turbulent combustion using a RANS
[25] Benekos S, Frouzakis CE, Giannakopoulos GK, Bolla M, Wright YM, Boulouchos K. model, detailed chemistry, and adaptive mesh refinement. SAE Tech Pap 2014.
Prechamber ignition: an exploratory 2-D DNS study of the effects of initial https://doi.org/10.4271/2014-01-1116.
temperature and main chamber composition. Combust Flame 2020;215:10–27. [51] Silva M, Mohan B, Badra J, Zhang A, Hlaing P, Cenker E, AlRamadan AS, et al.
[26] Silva M, Sanal S, Hlaing P, Cenker E, Johansson B, Im HG. Effects of geometry on DoE-ML guided optimization of an active pre-chamber geometry using CFD. Int J
passive pre-chamber combustion characteristics. SAE Tech Pap 2020. https://doi. Engine Res 2022;0. 14680874221135278.
org/10.4271/2020-01-0821. [52] Liu X, Silva M, Mohan B, AlRamadan AS, Cenker E, Im HG. Computational
[27] Echeverri Marquez M, Hlaing P, Ben Houidi M, Magnotti G, Johansson B, Cenker E. Optimization of the Performance of a Heavy-Duty Natural Gas Pre-Chamber. Fuel
Optical diagnostics of pre-chamber combustion with flat and bowl-in piston 2023. https://doi.org/10.2139/ssrn.4441826.
combustion chamber. SAE Tech Pap 2021. https://doi.org/10.4271/2021-01- [53] Amsden AA. KIVA3V, Rel.2, Improvements to KIVA3V. LA report, LA-UR-99-915.
0528. Los Alamos National Lab; 1999.
[28] Kim J, Scarcelli R, Som S, Shah A, Biruduganti MS, Longman DE. Assessment of [54] Silva M, Liu X, Hlaing P, Sanal S, Cenker E, Chang J, Johansson B, et al.
turbulent combustion models for simulating prechamber ignition in a natural gas Computational assessment of effects of throat diameter on combustion and
engine. J Eng Gas Turbines Power 2021;143. turbulence characteristics in a pre-chamber engine. Appl Therm Eng 2022;212:
118595.

13

You might also like