You are on page 1of 13

Fuel 259 (2020) 116290

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Lean combustion analysis using a corona discharge igniter in an optical T


engine fueled with methane and a hydrogen-methane blend

Valentino Cruccolinia, , Gabriele Discepolia, Alessandro Cimarelloa, Michele Battistonia,
Francesco Mariania, Carlo Nazareno Grimaldia, Massimo Dal Reb
a
Università degli Studi di Perugia, Department of Engineering, Italy
b
Federal-Mogul Powertrain, Tenneco Group, Italy

A R T I C LE I N FO A B S T R A C T

Keywords: The robustness of combustion initiation is one of the main issues of actual spark-ignition engines, especially for
Corona ignition highly-diluted or lean mixtures. In this work, the effects on combustion stabilization obtained by the usage of a
Lean combustion radio-frequency corona igniter were evaluated on a single-cylinder optical engine. The comparison with a
Methane conventional spark igniter was carried out using pure methane fuel and a blend of hydrogen and methane. For
Hydrogen
each combination of fuel and igniter, the combustion stability was explored at different air–fuel ratios, from
Optical engine
stoichiometric conditions to the lean stable limit (up to λ = 2.0 with the corona igniter and the hydrogen-
methane mixture). The combustion analysis was carried out by using the synchronized indicating and imaging
data. The latter is essential to estimate the contribution of the corona igniter, which was found to be considerable
only before the 5% of mass fraction burned. The corona effect igniter, with respect to a conventional spark
igniter, was able to extend the lean stable limit of about 0.15 λ units with methane fuel, and about 0.10 λ units
with the hydrogen-methane blend in the tested engine point. Early flame analysis confirmed the capability of
corona igniter to improve combustion onset speed and to obtain a more stable and repeatable flame kernel. The
findings of this study can help for a better implementation of corona ignition with gaseous low-carbon fuels, and
in particular to achieve a higher lean limit extension without the drawback of a performance decay given by a
substantial hydrogen enrichment.

1. Introduction system is able to deliver to the mixture. They found that only 2.5% is
deposited into the medium with respect to the primary energy input.
1.1. Background Many research groups studied how to increase discharge efficiency and
combustion stability, extending the conventional spark igniter con-
Nowadays, Spark Ignition (SI) engines have to face different issues. cepts. Multiple spark discharge [19,20], controlled electronic ignition
The increasingly restrictive pollutant emission regulations regarding [21], continuous discharge [22] and high energy ignition systems
nitrogen oxides (NOx) reduction require high Exhaust Gas Recirculation [23,24] show a small extension of the lean limit, as they maintain the
(EGR) dilution in the combustion chamber [1–5] and/or very lean issues related to the use of conventional spark plugs, like small thermal
air–fuel mixtures [6–9]. Moreover, in recent years, naturally aspirated plasma volume and heat losses through electrodes.
SI engines have been progressively replaced by boosted ones [10,11]. Recent theoretical and experimental studies involve a new genera-
These features are in general adopted in response to the demand for a tion of plasma-assisted igniters, mainly based on the production of
higher fuel economy [12–15] and a consequent carbon dioxide emis- transient non-thermal plasma [25–28]. Microwave assisted discharge
sion reduction. The recognized issue related to high EGR, lean com- [29–31] couples microwave radiation (frequency of 1–300 GHz) to a
bustion and boosted charge operation is the difficulty in ensuring a standard spark discharge typical of automotive engines. Nanosecond
stable and repeatable ignition and combustion process [16,17]. pulse discharge [32–35] is characterized by a high-frequency train of
A large body of literature involves studying conventional igniters, in high-voltage pulses. The reduced electric field that can be achieved is
order to determine optimization margins. Abidin and Chadwell [18] particularly high if compared to other plasma-assisted devices, due to
evaluated in a vessel the energy that a conventional spark ignition the proximity of the two electrodes. At the same time, the nanosecond


Corresponding author at: Università degli Studi di Perugia, Department of Engineering, via G. Duranti 93, 06125 Perugia, Italy.
E-mail address: valentino.cruccolini@studenti.unipg.it (V. Cruccolini).

https://doi.org/10.1016/j.fuel.2019.116290
Received 4 April 2019; Received in revised form 5 September 2019; Accepted 24 September 2019
0016-2361/ © 2019 Elsevier Ltd. All rights reserved.
V. Cruccolini, et al. Fuel 259 (2020) 116290

duration of the pulses avoids the breakdown phase. 1.2. Present contribution
Corona effect discharge [36–39] is the generation of non-equili-
brium plasma by a strong electric field, magnified by wires or needles, In this work, a comparison between the ignition performance of a
resulting in an ionization and an excitation of the surrounding medium corona igniter and a conventional spark igniter with pure methane and
without evolution in electric arc. This transient plasma can be con- a hydrogen-methane blend on an optical engine was carried out.
sidered the promoter of ignition thanks to the combination of three The objective is the assessment of the lean limit extension that
different effects [40]. First, the chemical kinetic enhancement, which corona can provide with respect to the conventional spark igniter with
consists in the production of active radicals and excited species via the same fuel, in particular with the hydrogen-methane blend.
electron impact. Second, the thermal effect, which is common to the In scientific literature, the lean limit extension characterization
equilibrium and near-equilibrium plasma: when plasma temperature given by hydrogen enrichment was carried out only with a conventional
rises, fuel oxidation becomes faster according to Arrhenius law [40,41]. spark igniter and on metal engines [62,63] without focusing on per-
Third, the corona wind, which is an ionic transport effect: it is relevant formance decay [62]; other works are with hydrogen-blended mixtures
in the ignition of a quiescent mixture, but negligible inside the com- on optical engines with spark igniters and do not focus on lean limit
bustion chamber of an internal combustion engine, due to the macro- extension [55]. Regarding corona ignition, studies were carried out on
scale charge motion [25,41,42]. Concerning the energy released by more conventional fuels such as methane and gasoline, with optical
thermal effect, our research group [43] evaluated it by means of corona [50,51] or metal engines [46]. The novelty of this paper is that, at the
discharge on quiescent air inside a constant volume chamber. The re- best of authors’ knowledge, it demonstrates for the first time that
lationship between this energy and the driving voltage at different corona ignition can be an effective way in achieving a higher lean stable
pressure levels was identified, and the amount of energy was compared limit instead of increasing H2 concentration in pure methane or in the
to the one released by conventional spark and multiple spark igniters. blend and without affecting engine performance.
Yu et al. [44] demonstrated the ability of a corona system to initiate Combustion was characterized from ignition and early flame de-
multiple ignition filaments on an air-propane mixture inside a constant velopment (image analysis of natural luminosity frames recorded by a
volume chamber, and found a non-negligible lean limit extension by high-speed camera) to combustion tail (indicating analysis). All the
increasing the discharge duration. Wang et al. [45], using a corona combinations between igniters (conventional spark igniter and RF-
system as well, showed a significant reduction in pressure rise time on corona igniter) and fuels (pure methane, hydrogen-methane blend)
quiescent air-hydrocarbon mixtures on a constant volume chamber and were tested. The spark-methane series will be considered as the base-
a pulse-detonation engine. Corona effect igniters were also tested on line, since it is obtained with a conventional igniter and a conventional
engines, showing an effective combustion improvement if compared to fuel. First, the combustion data from indicating analysis will be pre-
conventional systems, especially in lean or diluted mixtures. Compared sented: IMEP and its CoV, combustion duration data and heat release
to a conventional spark igniter, Pineda et al. [46] found a 10–15% and rate for all the tested points, as well as raw nitrogen oxides at the ex-
16–25% increase in EGR tolerance at boosted and naturally aspirated haust. Then, imaging data will be shown first in a qualitative way, and
conditions, respectively. Schenk et al. [47] compared the combustions then analyzed focusing up to λ 1.6 (which, for the baseline, is unstable,
of an 8% EGR diluted mixture triggered by a corona igniter, and a non- but close to lean limit): flame radii and flame speed evolution, as well as
diluted mixture triggered by a state-of-art transistor coil igniter. They flame temporal and spatial repeatability.
found a lower cycle-to-cycle variability with corona system. Bresler
et al. [48] compared corona to a conventional spark, a 3-plug igniter 2. Experimental apparatus and procedures
and a continuous discharge system. The best improvements regarding
both EGR tolerance and fuel consumption were found with the corona. 2.1. Engine and test bench
Optical access engines have been used to study corona widely.
Idicheria and Najt [49], using a gasoline-fueled engine, performed a The experimental campaign was carried out on a single-cylinder, 4-
morphological analysis of the flame front. Marko et al. [50] evaluated stroke research engine with optical access. The main features are
the projected flame area on a natural gas fueled engine. They found summarized in Table 1.
improvements in EGR tolerance using corona instead of conventional The pent-roof head has four valves and a centrally located igniter
spark. Two works of our research group [51,52] were focused on (Fig. 1a). The compression ratio (CR) value of the engine is 1–2 points
corona igniter on a gasoline-fueled optical engine. In the first, a lean lower than actual commercial engines, but it belongs to a well-studied
limit extension of 0.25 λ units in the tested point with respect to a range in which there is a trade-off between thermal efficiency and
conventional spark igniter was found. In the second, the effects on the power output [65,66]. The engine is characterized by an elongated
combustion behavior given by a variation in the electrical control cylinder and a Bowditch-type piston, provided with a quartz crown
parameters were evaluated at different air–fuel ratios. (30 mm radius). The latter replaces part of the original piston crown
Many of the papers reviewed above referred to natural gas fuel. and guarantees the optical access to the combustion chamber from the
Nowadays, natural gas is spreading in light-duty SI engine markets, as a bottom, by means of a 45-degree mirror (Fig. 1b). This optical access
valid alternative to gasoline-fueled engines. Its gaseous nature favors an configuration requires that the lubrication between cylinder liner and
efficient mixing inside combustion chamber, especially in direct injec- rings cannot be obtained by means of mineral oil, as occurs in the
tion applications. Methane, which is the major component of natural conventional engines. For this reason, the rings are made of graphite to
gas, is the lightest hydrocarbon with the smallest greenhouse effect ensure a sufficient lubrication. For all the other moving parts of the
upon combustion, among carbon-based fuels. For these reasons, a large engine, instead, a conventional mineral lubricant was used: its tem-
amount of research focused on methane-fueled engines, with both perature, together with coolant one, was set at 343 ± 0.2 K. This
conventional and innovative igniters [50,53–56]. Among many at- choice allows longer durability of the engine, with piston thermal ex-
tempts for further improving methane combustion stability, one possi- pansion close to the tolerance limit and piston crown temperature in the
bility is the addition of hydrogen in the fuel mixture. Indeed, it is well expected range of SI applications, even if coolant is about 20–25 K
known that H2 is able to considerably speed up the flame propagation lower than commercial power units [67].
[57–61] and to achieve combustion improvements, in terms of lean The engine can be controlled in terms of speed, ignition timing, air
stable limit or EGR dilution extension. and fuel flow rate. Speed control was obtained via an AVL 5700 dy-
namic brake, which is coupled to the optical engine. Airflow rate was
regulated by means of the throttle valve upstream of the intake mani-
fold. A research ECU was used to trigger the ignition and to control the

2
V. Cruccolini, et al. Fuel 259 (2020) 116290

Table 1
Engine specifications.
Displacement [cm3] 500
Bore [mm] 85
Stroke [mm] 88
Connecting Rod Length [mm] 139
Compression Ratio [−] 8.8:1
Number of Valves [−] 4
Exhaust Valve Open [CAD ATDCf] −553
Exhaust Valve Close [CAD ATDCf] −335
Intake Valve Open [CAD ATDCf] −380
Intake Valve Close [CAD ATDCf] −156
Internal residual mass fraction at the end of gas exchange process at 1000 rpm, IMEP 4.5 bar, λ = 1 (estimated by CFD-3D simulations) [−] 9%
Injection type [−] Port Fuel Injection

injector energizing time: the latter is the only parameter – at fixed Table 2
airflow – needed to control the air–fuel ratio. The gaseous fuel is in- Fuel properties [57,64,68]. H2 data, shown in Italic, are reported only as a
jected at 7.5 bar in the intake manifold by means of a Weber IAW003 comparison (pure hydrogen was not investigated in this paper).
injector for both methane and hydrogen-methane fuels. The hydrogen- CH4 H2-CH4 H2
methane blend has a 35% H2 composition by volume, equal to 6.3% by
mass; the remaining part is pure methane. This value is compatible with H2 concentration [%v/v] 0% 35% 100%
C/H ratio [−] 0.250 0.197 0
the idea of exploiting the existing CNG infrastructure to refuel vehicles
Density (at 293 K and 105 Pa) [kg/m3] 0.668 0.464 0.084
[57,63,64]: higher concentrations would require strong modifications Lower heating value [MJ/kg] 50.0 54.3 120.0
in both engines and infrastructures. Moreover, a higher hydrogen [MJ/m3] 33.4 25.2 10.1
concentration would imply an excessive volumetric heating value de- Stoichiometric air–fuel ratio [kgair/kgfuel] 17.1 18.2 34.2
crease [64]. In Table 2, properties of the tested fuels are shown. As a Energy in stoichiometric mixture [MJ/kgmix] 2.75 2.82 3.41
[MJ/m3mix] 3.21 3.18 3.01
comparison, also pure hydrogen values are reported.
A piezoelectric sensor (Kistler 6061B), flush-installed in the com-
bustion chamber, measured the in-cylinder pressure with an accuracy were synchronized with crankshaft angular position, detected by an
of ± 1%. A similar accuracy has been obtained for the intake port AVL 365C optical encoder, with 0.1 CAD resolution. Both air–fuel
pressure measurement, used as reference pressure for pegging and ac- equivalence ratio and nitrogen oxides concentration were acquired with
quired by a piezoresistive transducer (Kistler 4075 A5). Pressure data an accuracy of ± 2.5% by a fast Horiba MEXA-720 λ-NOx probe

Fig. 1. (a) pent-roof head (bore = 85 mm) with the 4 valves (exhaust on the left, intake on the right) and the central corona igniter. (b) the optical components:
bowditch-type piston, quartz crown (30 mm radius) and 45-degree mirror. (c) Schematic representation of the experimental setup (the same of a previous work of our
research group [52]).

3
V. Cruccolini, et al. Fuel 259 (2020) 116290

Table 3
Imaging specifications.
Image resolution pixel 512 × 512
Sampling rate [kHz] 20
Exposure time [μs] 49
Bit depth [−] 12
Spatial resolution [μm2/pixel] 123 × 123
Temporal resolution (@1000 rpm) [CAD] 0.3
Number of recorded consecutive combustion events [−] 63

installed in the exhaust pipe. For each operating point, 100 consecutive
combustion cycles were recorded and post-processed using a Kistler
KiBox indicating analyzer.

2.2. Imaging system and analysis

A Vision Research Phantom V710 high-speed CMOS camera coupled


with a Nikon 55 mm f/2.8 lens was used to record the natural lumin-
osity of the flames. The optical setting is shown in Table 3. For each
tested point, 63 consecutive combustions were recorded. The synchro-
nization with indicating data (Fig. 1c) allowed to match in-cylinder
pressure trace and flame evolution of the same cycle.
With the described apparatus, the early flame development can be
properly analyzed. The high temporal resolution allowed to record the
series of combustion onset images for each cycle: depending on the
flame speed, each combustion event was characterized by a number of
frames between 15 (fast combustion) and 120 (slow combustion). Due
to flame wrinkling, distortion and convection, the flame average radius
[51] that can be detected without reaching the optical boundary is
20 mm, even if the quartz piston crown radius is 30 mm. In the tested
engine the equivalent volume at 20 mm flame radius corresponds to 5%
of Mass Fraction Burned (MFB) detected by indicating system. The
high-speed imaging provides valuable geometrical information during
the early flame development, which complements the data acquired by
the indicating system.
The imaging post-processing was carried out by means of an in-
house MATLAB code, developed to quantify the combustion evolution.
The algorithm, reported in previous works [51,52], performs the op-
erations of imaging filtering, ignition detection and frame binarization.
More in detail, filtering is carried out by means of a 3x3 pixel spatial
median filter (Fig. 2d) in order to reduce salt-and-pepper noise. Ignition Fig. 2. Image post-processing main steps: (a) first ignition frame: the red square
detection is based on a frame-by-frame max gray level detection on a is the sub-area for ignition detection; (b) ignition detection principle; (c) raw
centrally-located subarea. A statistically-relevant number of frames combustion frame (it is the last in which corona discharge is still active and
before ignition, e.g. 50, is first chosen, and the average of the max gray streamers are visible); (d) combustion frame after median filter; (e) binarized
level over the statistic window is computed, together with the max combustion frame; (f) raw frame with flame boundaries and computed
equivalent flame radius. (For interpretation of the references to colour in this
absolute deviation with respect to the average. The threshold to re-
figure legend, the reader is referred to the web version of this article.)
cognize ignition is set to be higher than the max gray level detected in
the statistic window (Fig. 2a) and lower than the max gray level ex-
pected when ignition occurs (Fig. 2b). Particular attention must be paid ignition, a radio-frequency discharge (1.04 MHz) occurs between the 4-
to binarization, namely the conversion from grayscale frames to black- tip electrode and the combustion chamber, which acts as the counter-
and-white ones: white pixels correspond to flame or burned area, while electrode. Characterization of the equivalent circuit (shown in Fig. 3),
black ones to unburned mixture (Fig. 2e). The binarization threshold is as well as electronic behavior, were investigated by Mixell et al.
not fixed, but it is evaluated for each frame with a semi-automatic al- [70,71].
gorithm, proposed by Shawal et al. [69]. With this method, a great The driving voltage across the primary circuit and the discharge
accuracy in flame detection and separation from unburned mixture is duration can be set to best fit the operating conditions. If the driving
achieved, in all the combustion phases, from the ignition to the instant voltage – proportional to peak-to-peak voltage between firing end and
when the optical limit is reached. Once each frame has been converted combustion chamber – is too high, a conventional discharge, i.e. arcing,
to a black-and-white image, the algorithm evaluates the equivalent flame occurs, with a consequent loss of corona effect. Furthermore, the sta-
radius (Fig. 2f) and the horizontal component of the flame growth bilized arc condition is not desired because it can severely damage the
speed. igniter assembly. For this reason, an “arc prevention” function was
enabled so that switching from streamer to arc was limited to sporadic
2.3. Igniters events.
The corona igniter was compared with a conventional spark ignition
The corona effect igniter, named ACIS (Advanced Corona Ignition system. The latter was composed of a plug-top coil and a programmable
System), is controlled by a dedicated electronic system, which provides interface that allowed to set the number of commanded consecutive
energy to the igniter assembly. When the engine ECU commands sparks (from 1 to 17) for each ignition event. The coil, characterized by

4
V. Cruccolini, et al. Fuel 259 (2020) 116290

Fig. 3. (a) ACIS equivalent circuit; (b) igniter assembly [71].

conventional spark igniter, an Indicated Mean Effective Pressure


(IMEP) of 4.5 bar was targeted. This tested point was chosen because
the optical engine max speed is very limited (if compared to current
automotive production engines) and to maintain the same tested point
used in previous works of the same research group [51,52]. In any case,
it can be considered a valid choice, since it is well known that corona
ignition benefits are considerable especially at low-speed and low-load
[46].
At fixed throttle position and starting from stoichiometric condi-
tions, air–fuel ratio was increased by injecting less fuel, up to the first
lean unstable condition. For each tested point (combination of igniter,
fuel type and λ), the Maximum Brake Torque (MBT) ignition timing
was chosen. At fixed λ and fixed fuel type, corona requires a reduced
ignition advance with respect to the conventional spark to maximize
the brake torque, as it can be seen in the next section.
Preliminary tests were conducted for corona igniter in order to set
the driving voltage to the maximum level that avoids arc formation. As
explained in [51], since MBT requires higher spark advance angles with
leaner mixtures, the in-cylinder pressure at the time of ignition is lower
Fig. 4. Maximum allowable driving voltage for each tested point with the
and arcing events occur at lower driving voltage. So, it is necessary to
corona igniter.
strongly reduce the driving voltage approaching the lean limit. More-
over, gas type has a strong influence too: the presence of hydrogen
a nominal charge time of 3 ms, was operated in single spark mode. allows to increase the driving voltage at the same air–fuel ratio (Fig. 4),
Measurements of the released thermal energy during single spark as expected due to the different in-cylinder pressure and mixture
event and during corona discharge had been previously carried out by composition. The other corona parameter which can be set is the dis-
the same authors: in the pressure range typical of the optical engine charge duration. All the tests were performed with a fixed value of
tested points when ignition is required, the energy delivered to the 300 μs that is effective for the whole λ range, according to the previous
medium was in the range 2 ÷ 5 mJ for the tested spark, while between experience [52].
10 and 20 mJ for the corona igniter [43]. This fact must be taken into
account: beside the radicalization/ionization effects, typical of corona
3. Results and discussion
discharge and absent in spark ignition, the corona igniter is also able to
provide more energy to the medium with respect to the conventional
A first analysis is related to indicating data and in particular to
spark igniter. Nonetheless, it must be remarked that this higher thermal
combustion stability. In Fig. 5a, IMEP values can be found for each
energy, alone, is not sufficient to explain the increased effectiveness of
tested point. Since load control was performed by decreasing the
ACIS, because of the extremely reduced time from ignition to flame
amount of fuel at fixed air flow, IMEP decreases as λ increases. Using
presence [51]: the radicalization and ionization effects play a funda-
the same fuel type, no substantial differences were found between spark
mental role in this device [49,50].
and corona igniter (it must be remarked that for each tested point the
MBT ignition timing was chosen). Instead, the presence of hydrogen
2.4. Test campaign results in a lower IMEP, especially for stoichiometric and near-stoi-
chiometric mixtures, due to the low energy density of hydrogen if
The lean limit extensions provided by the corona system with pure compared to methane.
methane and hydrogen-methane were evaluated. The parameter used to evaluate combustion stability is the
Engine speed was set to 1000 rpm and throttle valve position was Coefficient of Variation (CoV) of the IMEP, namely the ratio between
fixed, in order to ensure constant airflow and tumble motion at the IMEP standard deviation and IMEP mean value. Among the different
same engine speed. In stoichiometric conditions with pure methane and limits proposed in literature, a stability limit band of 3–4% was chosen:

5
V. Cruccolini, et al. Fuel 259 (2020) 116290

Using pure methane, the corona igniter guarantees an extension of


about 0.15 λ units with respect to the conventional spark. Hydrogen-
addition effect, instead, is more intense: using the conventional spark, it
ensures a lean limit extension of 0.4 λ units. In the methane-spark,
methane-corona and hydrogen-methane-spark cases, a sharp transition
between stable and unstable conditions was found, and the last stable
point can be easily detected. On the contrary, using hydrogen-methane
with corona, the transition to unstable conditions is very smooth with
this configuration, differently from the sharp threshold behavior ob-
served in the other cases. A maximum λ value of 2.05 was reached with
a CoV value still below 5%; however, it is out of the stability band, and
the limit that was considered for this configuration was λ = 2.0. The
smooth trend at very high λ values is similar to what was found by Ma
et al. [62] and Poggiani et al. [63], which investigated the combustion
behavior of a spark-ignited engine with different hydrogen-methane
mixtures. The smoother trend is a result of the fact that the near-limit
points show the same CoV values (below 4%, still acceptable), so that
the higher the lean limit, the lower the slope of the λ-CoV curve [62].
These results suggest that corona ignition has a small, but not negli-
gible, effect on combustion stabilization with the tested hydrogen-me-
thane blend. It is interesting to quantify this stabilization with respect
to spark by comparing it with other fuels (Fig. 5c). With methane, as
said, an extension of about 0.15 λ units was found, while with the
blended mixture it was about 0.10 λ units. As a comparison, in the same
engine point (1000 rpm, 4.5 bar IMEP) with gasoline it was about 0.25
λ units, as reported in a previous work of our research group [51]. This
different improvement is in accordance with Ju and Sun [40] and
Starikovskiy and Aleksandrov [41]: the larger the fuel molecules, the
higher the dissociation and ionization pathways and the higher the
corona igniter effectiveness.
So, even if the lean stable limit found with the conventional spark
igniter on hydrogen blend was high (λ = 1.9), corona ignition was
Fig. 5. (a) IMEP against λ. (b) IMEP CoV against λ. A threshold between 3 and found to be able to further extend it. The near-limit lean region,
4% allows to distinguish between stable points (filled markers linked with solid moreover, was found to be wider than in the other fuel-igniter tested
lines) and unstable ones (empty markers linked with dashed lines). (c) Lean combinations, so that the transition between stable and unstable air–-
limit for spark ignited combustions (black bars) and corona ones (gray bars) fuel mixtures can be controlled easier, avoiding an abrupt switching
with different fuels. Gasoline tests are referred to the same experimental ap-
from regular to irregular engine operations.
paratus and the same engine point (1000 rpm, IMEP 4.5 bar) and are related to
In Fig. 6a, the ignition timing versus λ, depending on igniter and
a previous work [51].
fuel type, is shown. In the baseline case (methane-spark), ignition
timing increases with λ, since the combustion slows down. As expected,
combustion events with an IMEP CoV below 3% were considered stable, this trend is also obtained for all the other combinations of igniters and
and above 4% were considered unstable. In Fig. 5b it is possible to fuels. Hydrogen presence allows a 10–15 CAD spark advance reduction,
observe the trend of the IMEP CoV against λ for the different fuel-ig- strongly dependent on the air–fuel ratio. Corona effect on ignition ad-
niter combinations. vance reduction is more intense, especially for the richest mixtures. It is

Fig. 6. (a) Ignition timing vs λ; (b) ignition to 5% of mass fraction burned (CA0-5); (c) main combustion duration, namely 5 to 90% of mass fraction burned (CA5-
90).

6
V. Cruccolini, et al. Fuel 259 (2020) 116290

worth noticing that at λ = 1.6 the corona effect on methane ignition


timing has the same magnitude of hydrogen addition with the con-
ventional spark.
The effect on the first stage of combustion is shown in Fig. 6b. CA0-5
represents the angular interval from ignition timing to the 5% of mass
fraction burned. The trend is globally similar to ignition timing, shown
in Fig. 6a. In each tested point, corona is characterized by a strong CA0-
5 reduction with respect to spark (on average, 16 CAD with pure me-
thane and 9 CAD with hydrogen-methane blend). It is worth noticing
that with spark, at fixed λ, hydrogen addition effect on CA0-5 is lower
than corona ignition effect on pure methane. This denotes the large
impact of corona in the early combustion phase.
Main combustion duration (CA5-90, i.e. the angular interval from
5% to 90% of MFB, Fig. 6c) shows a completely different trend. At fixed
fuel type, in the stoichiometric and low-lean tested points, spark and
corona are characterized by the same values. When the lean limit is
approaching, the spark-ignited combustion duration starts to rise more
than the corona-ignited one. This behavior occurs for both methane and
hydrogen-methane blend and it is analogous to what found by Quader
[72], Ma et al. [62] and Suess et al. [73]: main combustion duration is
Fig. 8. Nitrogen oxides raw emission.
nearly the same when the engine approaches its lean limit, in-
dependently on igniter type. Although combustion duration is pro-
longed with engine leaning-out, there is an upper limit beyond which ignition timing. This confirms the higher burning rate given by the
the engine becomes unstable. A combination of fuel and igniter will corona igniter during the combustion onset.
have a larger lean operation ability if it provides shorter main com- Nitrogen oxides emissions trend (Fig. 8) is the same for all the tested
bustion duration at a given lambda. Fixed the fuel, if the main com- combinations. As expected, for stoichiometric or low-lean mixtures,
bustion duration is almost the same with the two igniters, the spark NOx emission is high, for λ values between 1.2 and 1.4 there is a rapid
case, which has a reduced lean limit, should rise before than the corona drop, while beyond λ = 1.6 the concentration falls below 200 ppm and
one, as shown in Fig. 6c. close to the sensor minimum detectable level (beyond λ = 1.8 the
Finally, hydrogen effect is plain, since it reduces the main com- emission is under the instrument sensitivity threshold). This occurs for
bustion duration of about 5 CAD on average with respect to pure me- all the measured data, independently of igniter or fuel. This behavior
thane. This is in agreement with literature data [57,63]. witnesses the need for reaching the combustion stability with very lean
Combustion can be also described by analyzing the rate of heat mixtures (λ ≥ 1.8). Fixed the igniter and the air–fuel ratio, the NOx
release. In Fig. 7 the HRR of the average cycle over crank angle AIT is emission with blended fuel is higher than the one with pure methane.
shown up to λ = 1.6, which is the highest air–fuel ratio that was tested This is in agreement other works, e.g. [58,74], and it is mainly due to
for all the fuel-igniter combinations. Fixed the fuel, the slope of corona the faster burning rate given by hydrogen addition. Finally, fixed the
curve is clearly higher than the spark one, and the maximum is closer to fuel and the air–fuel ratio, the nitrogen oxides emission was found to be

Fig. 7. Rate of heat release over crank angle AIT. Each series is referred to the average combustion cycle. The λ = 1.6 methane-spark series is represented in dashed
line and empty markers since the condition is unstable.

7
V. Cruccolini, et al. Fuel 259 (2020) 116290

Fig. 9. Flame fronts evolution; λ = 1.4. Images are cropped in correspondence Fig. 10. Flame fronts evolution; λ = 1.6. Images are cropped in correspondence
to the optical limit (30 mm radius). to the optical limit (30 mm radius).

lower with corona ignition, especially for stoichiometric and near-


stoichiometric cases (reduction in a range of 15 ÷ 30%). It must be
remarked that corona effect on nitrogen oxides emission is dependent
on many contrasting factors: combustion speed, heat release rate, later
MBT ignition timing (closer to TDC) and the lower contribution of NOx
emission during the corona discharge in pure air with respect to the
spark [75]. The first two, on their own, would affect NOx emission in a
detrimental way, while the last two would be beneficial. The resulting
effect, as shown, is on average an overall reduction.
Since the influence of Corona igniter is maximum in the early stage
of combustion, optical data can be particularly useful to better under-
stand flame kernel development. A first qualitative result can be ob-
tained by observing the flame development as a function of crank an-
gles. Figs. 9 and 10 show this evolution at λ = 1.4 and 1.6, respectively.
In each column the photograms are instantaneous images belonging to
the combustion sequence (among all the 63 events) which is closest to
the average behavior. The quickness of the deflagration front initiated
by the corona igniter is evident. By the time flames originated with a
spark become visible, corona ignited combustions have already reached
almost half of the optical window. Moreover, we can appreciate the
flame extent difference moving from λ = 1.4 to 1.6: methane series – in
proximity to the lean limit - are characterized by a non-negligible re-
duction if compared to λ = 1.4. In particular, the spark-ignited me-
thane-fueled flame 25 CAD after ignition event is still small and con-
fined in close proximity to spark electrodes.
In Fig. 11, the flame equivalent radii evolution (obtained by aver-
aging the computed flame equivalent radii of all the consecutive re-
corded combustions) is shown at different λ. The optical analysis is
focused on λ span from 1 to 1.6: in this interval it was possible to record
combustions for all the combination of igniters and fuels. λ = 1.6 is the
last tested point for spark-methane, already unstable; it is also close to Fig. 11. Average equivalent radius evolution over crank angle after ignition
corona-methane lean limit. All the curves show an S-shape and the timing for different air–fuel ratios.
flattening in the final part is due to the fact that flames have already

8
V. Cruccolini, et al. Fuel 259 (2020) 116290

reached the optical limit, while the combustion is still ongoing [54,67]. discharge is over there are 4 different flame fronts that first develop
For both methane and hydrogen-methane blend, at the same λ, independently and very quickly; this stage corresponds to the flame
corona radii evolution is much faster, reaching the optical limits several speed rise towards the maximum. Then, the 4 flame fronts coalesce and
degrees before the spark-ignited cases (as found from a qualitative point the growth continues in a less intense way, so that the flame speed
of view in Figs. 9 and 10). For both the fuels and at a fixed air–fuel decreases and converge to the plateau value. The only corona case in
ratio, the time to reach 20 mm of average equivalent flame radius from which this trend is not easily recognizable is the λ = 1.6 with pure
ignition timing is reduced of about 45–60% by using corona with re- methane: this case is stable but close to the lean limit, and the initial
spect to spark ignition. This reduction is more intense (close to 60%) at flame speed is low even if ignited via corona effect, so the 4 in-
stoichiometric ad near stoichiometric conditions, and less intense (close dependent flame fronts grow very slowly due to the poor combustion
to 45%) with leaner mixtures. This is in agreement with the indicated conditions. It must be noticed that the leaner the mixture and the higher
results previously discussed: qualitatively with the different slope of the noise in the flame growth speed (that is a derivative from a math-
heat release rate curves (Fig. 7) and quantitatively with the time to 5% ematical point of view), due to the increased cyclic variability during
of MFB (Fig. 6b). As discussed, hydrogen accelerates the flame propa- combustion onset, as clearly found in the λ = 1.6 corona-methane
gation. At λ = 1.6, for example, the flame extent of the spark-ignited series. Finally, the max speed values for both corona and spark cases are
blend, near the optical limits, becomes very close to the corona-ignited strongly dependent on the air–fuel ratio: max with stoichiometric
methane. This happens because corona flame is characterized by a fast- mixtures and decreasing with leaner conditions.
initial development followed by a slowdown, while hydrogen addition All these observations can be confirmed and refined by plotting the
on spark ignition is responsible for a lower, but more constant, flame flame growth speed versus the average equivalent flame radius
growth during the whole combustion process. (Fig. 12b), as proposed by Aleiferis and coworkers [76]. With this vi-
It is interesting to analyze also the flame growth speed, calculated as sualization, related to a spatial evolution of flame speed, instead of a
the time derivative of the average equivalent radius. In Fig. 12a it is temporal one, it is possible to better compare fast and slow-burning
plotted versus crank angle degree after ignition timing. The very first cases. The x-axis is cut when the average flame radius reaches 20 mm:
part is represented by means of dotted lines, in order to distinguish the in this point all the flame growth speed series reach the plateau value.
frames in which the ignition process (spark or streamers) is active and Moreover, lines are drawn starting from the first frame after the end of
would perturb the chamber luminosity, not due to combustion. The ignition phase. In particular, the spark event lasts about 50 μs (1 frame),
shape of the flame speed curves is dependent on the igniter: spark series while corona discharge lasts 300 μs (6 frames). Consider that after 50 μs
are characterized by a slow growth until reaching the max value, as (about 0.3 CAD at 1000 rpm) from spark ignition, flame kernel is on
expected for a “spherical” flame front, then speed tends to zero, due to average not already developed, and however it would not be visible: the
the reaching of the optical limit. Corona series, instead, are character- ground electrode would cover the flame kernel in the very first frames.
ized by a quick rise towards the max value, then the speed decreases Instead, after 300 μs (about 1.8 CAD at 1000 rpm) from ignition timing
and reaches a plateau, and finally tends to zero because of the optical the corona flame kernels, for the stoichiometric and near-stoichiometric
limit just like the spark case. This different behavior is due to the corona series, are well established, (see Fig. 2f as an example). For these rea-
features: ignition is volumetric and plasma-assisted, and when the sons, the average equivalent flame radius for corona ignition at the end

Fig. 12. Average flame growth speed versus crank angle after ignition timing (a) and average equivalent radii (b) for different air–fuel ratios. The λ = 1.6 methane-
spark series is represented in dashed lines and empty markers since the condition is unstable. The very first points for figure (a) are represented in dotted lines to
distinguish when ignition process is active.

9
V. Cruccolini, et al. Fuel 259 (2020) 116290

average flame radius of 9 mm (1% of MFB for the tested engine,


Fig. 14a) or 20 mm (5% of MFB, Fig. 14b) versus λ are shown. The
latter is in agreement with the CA0-5 distribution found with the in-
dicated analysis (Fig. 6b). Vertical bars refer to crank angle standard
deviation of the recorded 63 combustion events when the average
reaches the stated radius. With this kind of optical analysis, it is possible
to evaluate with more accuracy the passage from low to high flame
evolution variability, even comparing data series with an IMEP CoV
below 3% (so, stable enough according to indicated analysis). With the
same fuel, the effectiveness of corona (with respect to spark) in terms of
rapidity to reach a stated average flame radius and cycle variability
reduction is well evident (the crank angle standard deviation at the
same λ is lower). The rapid slope increase of methane curves when λ is
higher than 1.2–1.4 is also evident. Hydrogen-methane blend, instead,
is characterized by a less-rapidly rising slope over all the λ range.
Finally, in Fig. 15, the flame probability presence at 9 mm (about
1% of MFB) of average equivalent flame radius is shown for λ = 1.4
and 1.6. This kind of analysis was developed by Aleiferis and coworkers
[54]; it was used by Marko and coworkers [50] and by our research
Fig. 13. (a) Flame growth speed when the average equivalent radius reaches group as well [51,52]. Dark red pixels mean 100% of probability (over
20 mm against lambda for the tested fuels. The different igniter type was found the 63 recorded combustion events) to have a burned area there, while
not to affect this value. (b) Difference between blend v20 and methane v20 black pixels mean no flame in none of the recorded events (0% prob-
against lambda.
ability). Fixed the fuel and the air–fuel ratio, with the corona igniter
dark red areas are wider and the transition from dark red to black is
of the discharge can be way bigger than the spark counterpart, and the sharper with respect to the spark igniter. As a consequence, for a fixed
first plotted point of methane-corona (and hydrogen-methane-corona) average equivalent flame radius, the overall non-black area for the
is some millimeters far from the x-axis origin. Hydrogen-methane blend corona cases is smaller than the spark ones, denoting a higher repeat-
series, moreover, are characterized by an initial growth faster than ability of the entire combustion onset over different consecutive events.
methane, so the first plotted points for the corona-blend series are Fixed fuel and igniter, an air–fuel ratio variation effect is expected:
farther from the origin than the corona-methane counterparts. From the lower the lambda, the more repeatable the flame presence (wider
this graph it is evident that, despite influencing the overall flame dark red areas, smoother red to black transition, smaller overall non-
growth trend, igniter type was found not to affect the plateau value, black area, more regular and symmetric contour shape).
which is dependent only on air–fuel ratio and fuel type. Concerning the shape of the flame probability presence contours,
The “plateau” speed (Fig. 13) was analyzed as well: the average spark ignited cases show an asymmetric distribution, far from a circle
flame growth speed value when req avg is 20 mm (v20) is plotted versus ad not centered, with a preferable and biased motion towards the
the relative air–fuel ratio. For each fuel, v20 is almost linear with bottom part of the chamber (from high-speed camera point of view).
lambda in the tested range. This speed is higher with the blended fuel This fact is particularly evident in the pure methane cases, while in the
than with pure methane: at λ = 1 the first is 11.0 m/s and the latter spark-ignited blended cases the shape is more similar to a circle, de-
9.5 m/s, while at λ = 1.6 the difference is lower (7.0 m/s and 6.2 m/s noting a more stable combustion given by hydrogen addition. The
respectively). This reduced gap between methane and hydrogen-me- corona cases distribution, instead, seems to be on average more cen-
thane at higher λ (Δv20, Fig. 13b) can be explained with the lower tered and symmetrical, considering the different shape (4-leaf instead of
hydrogen amount inside the combustion chamber with leaner mixtures circular).
at fixed air mass flow rate. The λ = 1.6 corona – methane case is interesting since the 4-leaf
In Fig. 14 the crank angle intervals from ignition timing to an shape (common for all the other corona results) is difficult to recognize:

Fig. 14. Crank angle from ignition timing to 9 mm (a) or 20 mm (b) of average equivalent radius. Vertical bars represent the crank angle standard deviation.

10
V. Cruccolini, et al. Fuel 259 (2020) 116290

Fig. 15. Flame probability presence at 9 mm average equivalent flame radius (1% MFB). Top row: λ = 1.4. Bottom row: λ = 1.6. The white circles correspond to the
optical crown limit (30 mm radius).

other optical and indicating results.


This analysis is a confirmation of what previously shown from a
different point of view: it is possible to find in which conditions and in
which points flames are expected. Also with this optical method it can
be stated that the corona igniter is able to provide a more repeatable
combustion onset, resulting in a more stable combustion with respect to
the spark igniter.

4. Conclusions

Combustion analysis on an optical access engine was carried out


with two different igniters (a conventional spark plug and a radio-fre-
quency corona ignition system) and two different gaseous fuels (pure
methane and a blend of hydrogen and methane, with 35% of H2 in
volume) at 1000 rpm, low load, varying the air–fuel ratio until reaching
unstable conditions. The findings of this work are the following:

• As far as hydrogen addition is concerned, as expected its effect


considerably stabilizes combustion, with respect to the spark-me-
Fig. 16. Occurrence of corona-ignited combustion with only 3 flame kernels thane baseline, extending the lean limit of about 0.4 λ units.
(zoomed and intensified during the postprocess to better recognize the flame Hydrogen speeds up the whole combustion process (the CA5-90 is
areas). Frame acquired 3.6 CAD AIT. Tested case: corona igniter, methane fuel, about 5 crank angle degrees faster than baseline). The optical ana-
λ = 1.6. lysis confirmed the reduced cycle-to-cycle variability and the early
flame speed increase. As known, the drawbacks of hydrogen addi-
tion lay in a lower IMEP and, in the near-stoichiometric mixtures, a
actually, only 3 areas of high probability flame presence (orange-co-
lored areas) can be found. Beside the fact that this point is near the lean higher NOx production. On the other hand, beyond λ = 1.6, NOx
emission rapidly falls to values close to the lower scale of the in-
stable limit, so that the flame repeatability is much lower than other
corona cases, this can be explained by analyzing the variability of the strument.
flame kernels in the proximity of the igniter electrode: among the 63 • Corona ignition on methane-air mixture, with respect to baseline,
consecutive combustions recorded for this tested point, there are oc- was found to have a significant effect in the early phase of com-
bustion. The reduction of the CA0-5 with respect to baseline is
currences in which only 3 (out of 4) flame kernels are generated, in
particular the top left one is missing in some occurrences (Fig. 16). This stronger than with hydrogen addition, at least in the near-stoichio-
metric mixtures. The lean stable limit was increased of about 0.15 λ
is be due to either a lack of streaming activity from one of the 4 tips, or
the fact that the streamer generated from one of the tips is much weaker units with respect to baseline with no worsening on IMEP at the
same λ. Heat release rate is characterized by a sensibly higher slope
and shorter than the other 3, so that kernel formation is not ensured in
all the cases. This condition had been already found in other experi- in the first part of the combustion process. The optical analysis
highlights the strong improvement in early flame speed and in the
mental test cases with the same corona igniter, as reported in [52]. With
the blended fuel instead of methane, for both spark and corona igniters, reduction of combustion variability, also from a spatial point of view
(flame probability presence), that the indicated analysis is not able
the λ = 1.6 case is not critical in terms of flame repeatability (abundant
red areas, smooth red-to-blue transition), and this is coherent with the to evaluate.

11
V. Cruccolini, et al. Fuel 259 (2020) 116290

• Corona ignition with hydrogen-methane blend combines the bene- 0778.


fits of both radio-frequency igniter and hydrogen, resulting in a [19] Jung D, Iida N. An investigation of multiple spark discharge using multi-coil igni-
tion system for improving thermal efficiency of lean SI engine operation. Appl
higher lean limit, a shorter combustion duration and an even lower Energy 2018;212:322–32. https://doi.org/10.1016/j.apenergy.2017.12.032.
cyclic variability. The lean limit extension that corona guarantees is [20] Poggiani C, Cimarello A, Battistoni M, Grimaldi CN, Dal Re M, De Cesare M. Optical
about 0.10 λ units with respect to the spark-ignited hydrogen-me- investigations on a multiple spark ignition system for lean engine operation. SAE
Tech Pap 2016. https://doi.org/10.4271/2016-01-0711.
thane blend, reaching λ = 2.0 with stable conditions in the tested [21] Mazacioglu A, Gross M, Kern J, Sick V. Infrared borescopic evaluation of high-
engine point. energy and long-duration ignition systems for lean/dilute combustion in heavy-duty
natural-gas engines. SAE Tech Pap 2018;2018:1–14. https://doi.org/10.4271/
2018-01-1149.
To summarize, the RF corona igniter improves the lean limit toler- [22] Alger T, Gingrich J, Roberts C, Mangold B, Sellnau M. A High-Energy Continuous
ance with both the tested gaseous fuels, thanks to the quicker com- Discharge Ignition System for Dilute Engine Applications, 2013. doi: 10.4271/
bustion onset and the more repeatable flame kernel which is able to 2013-01-1628.
[23] Dale J. Application of high energy ignition systems to engines. Prog Energy
guarantee with respect to a conventional spark igniter. The lean ex-
Combust Sci 1997;23:379–98. https://doi.org/10.1016/S0360-1285(97)00011-7.
tension is hydrogen-dependent: the effect on the blended fuel is lower [24] Chen L, Wei H, Zhang R, Pan J, Zhou L, Feng D. Effects of spark plug type and
than with pure methane, but significant. ignition energy on combustion performance in an optical SI engine fueled with
Future work will be related to the sensitivity of corona igniter methane. Appl Therm Eng 2019;148:188–95. https://doi.org/10.1016/j.
applthermaleng.2018.11.052.
performance (in terms of lean limit extension and early flame evolu- [25] Ju Y, Sun W. Plasma assisted combustion: progress, challenges, and opportunities.
tion) on different hydrogen concentrations in the blend and at different Combust Flame 2015;162:529–32. https://doi.org/10.1016/j.combustflame.2015.
engine map points; the raw-gas analysis will be improved by detecting 01.017.
[26] Scarcelli R, Zhang A, Wallner T, Breden D, Karpatne A, Raja L, et al. Multi-di-
CO and HC as well by means of a dedicated analyzer. mensional modeling of non-equilibrium plasma for automotive applications. SAE
Tech Pap 2018;2018:1–10. https://doi.org/10.4271/2018-01-0198.
References [27] Merola SS, Marchitto L, Tornatore C, Valentino G, Irimescu A. Optical character-
ization of combustion processes in a DISI engine equipped with plasma-assisted
ignition system. Appl Therm Eng 2014;69:177–87. https://doi.org/10.1016/j.
[1] Wei H, Zhu T, Shu G, Tan L, Wang Y. Gasoline engine exhaust gas recirculation – A applthermaleng.2014.04.046.
review. Appl Energy 2012;99:534–44. https://doi.org/10.1016/j.apenergy.2012. [28] Kim GT, Seo BH, Lee WJ, Park J, Kim MK, Lee SM. Effects of applying non-thermal
05.011. plasma on combustion stability and emissions of NOx and CO in a model gas turbine
[2] Francqueville L, Michel J-B. On the effects of EGR on spark-ignited gasoline com- combustor. Fuel 2017;194:321–8. https://doi.org/10.1016/j.fuel.2017.01.033.
bustion at high load. SAE Int J Engines 2014;7. https://doi.org/10.4271/2014-01- [29] Hwang J, Kim W, Bae C, Choe W, Cha J, Woo S. Application of a novel microwave-
2628. assisted plasma ignition system in a direct injection gasoline engine. Appl Energy
[3] Catapan RC, Cancino LR, Oliveira AAM, Schwarz CO, Nitschke H, Frank T. Potential 2017;205:562–76. https://doi.org/10.1016/j.apenergy.2017.07.129.
for onboard hydrogen production in an direct injection ethanol fueled spark ignition [30] DeFilippo A, Saxena S, Rapp V, Dibble R, Chen J-Y, Nishiyama A, et al. Extending
engine with EGR. Fuel 2018;234:441–6. https://doi.org/10.1016/j.fuel.2018.07. the Lean Stability Limits of Gasoline Using a Microwave-Assisted Spark Plug, 2011.
023. doi: 10.4271/2011-01-0663.
[4] Lattimore T, Herreros JM, Xu H, Shuai S. Investigation of compression ratio and fuel [31] Ikeda Y, Padala S, Makita M, Nishiyama A. Development of Innovative Microwave
effect on combustion and PM emissions in a DISI engine. Fuel 2016;169:68–78. Plasma Ignition System with Compact Microwave Discharge Igniter, 2015. doi: 10.
https://doi.org/10.1016/j.fuel.2015.10.044. 4271/2015-24-2434.
[5] Briggs T, Alger T, Mangold B. Advanced ignition systems evaluations for high-di- [32] Singleton D, Pendleton SJ, Gundersen MA. The role of non-thermal transient plasma
lution SI engines. SAE Int J Engines 2014;7. https://doi.org/10.4271/2014-01- for enhanced flame ignition in C 2 H 4 –air. J Phys D Appl Phys
2625. 2011;44:022001https://doi.org/10.1088/0022-3727/44/2/022001.
[6] Xie F, Li X, Su Y, Hong W, Jiang B, Han L. Influence of air and EGR dilutions on [33] Pancheshnyi SV, Lacoste DA, Bourdon A, Laux CO. Ignition of propane-air mixtures
improving performance of a high compression ratio spark-ignition engine fueled by a repetitively pulsed nanosecond discharge. IEEE Trans Plasma Sci
with methanol at light load. Appl Therm Eng 2016;94:559–67. https://doi.org/10. 2006;34:2478–87. https://doi.org/10.1109/TPS.2006.876421.
1016/j.applthermaleng.2015.10.046. [34] Domingues E, Burey M, Lecordier B, Vervisch P. Ignition in an SI Engine using
[7] Alvarez CEC, Couto GE, Roso VR, Thiriet AB, Valle RM. A review of prechamber Nanosecond Discharges generated by a Spark Gap Plasma Igniter (SGPI), 2008. doi:
ignition systems as lean combustion technology for SI engines. Appl Therm Eng 10.4271/2008-01-1628.
2018;128:107–20. https://doi.org/10.1016/j.applthermaleng.2017.08.118. [35] Yu J, He L, Ding W, Zhao Z, Zhang H. Research on the impacts of air temperature on
[8] Jiang Y, Bao X, Sahu A, Ma X, Xu H, Thong A. Flame Kernel Growth and the evolution of nanosecond pulse discharge products. Appl Therm Eng
Propagation in an Optical Direct Injection Engine Using Laser Ignition. SAE Tech. 2016;98:265–70. https://doi.org/10.1016/j.applthermaleng.2015.12.013.
Pap., vol. 2017- October, 2017. doi: 10.4271/2017-01-2243. [36] Sher E, Ben-Yaish J, Pokryvailo A, Spector Y. A Corona Spark Plug System for Spark-
[9] Ran Z, Hariharan D, Lawler B, Mamalis S. Experimental study of lean spark ignition Ignition Engines, 1992. doi: 10.4271/920810.
combustion using gasoline, ethanol, natural gas, and syngas. Fuel 2019;235:530–7. [37] Mariani A, Foucher F. Radio frequency spark plug: an ignition system for modern
https://doi.org/10.1016/j.fuel.2018.08.054. internal combustion engines. Appl Energy 2014;122:151–61. https://doi.org/10.
[10] Galloni E, Fontana G, Palmaccio R. Effects of exhaust gas recycle in a downsized 1016/j.apenergy.2014.02.009.
gasoline engine. Appl Energy 2013;105:99–107. https://doi.org/10.1016/j. [38] Mariani A, Foucher F, Moreau B. The Effects of a Radio Frequency Ignition System
apenergy.2012.12.046. on the Efficiency and the Exhaust Emissions of a Spark-Ignition Engine. SAE Tech
[11] Teodosio L, De Bellis V, Bozza F. Combined effects of valve strategies, compression Pap 2013. https://doi.org/10.4271/2013-24-0053.
ratio, water injection and cooled EGR on the fuel consumption of a small turbo- [39] Starikovskaia SM. Plasma assisted ignition and combustion. J Phys D Appl Phys
charged VVA spark-ignition engine. SAE Tech Pap 2018:1–13. https://doi.org/10. 2006;39:R265–99. https://doi.org/10.1088/0022-3727/39/16/R01.
4271/2018-01-0854. [40] Ju Y, Sun W. Plasma assisted combustion: dynamics and chemistry. Prog Energy
[12] Jung D, Lee S. An investigation on the potential of dedicated exhaust gas re- Combust Sci 2015;48:21–83. https://doi.org/10.1016/j.pecs.2014.12.002.
circulation for improving thermal efficiency of stoichiometric and lean spark igni- [41] Starikovskiy A, Aleksandrov N. Plasma-assisted ignition and combustion. Prog
tion engine operation. Appl Energy 2018;228:1754–66. https://doi.org/10.1016/j. Energy Combust Sci 2013;39:61–110. https://doi.org/10.1016/j.pecs.2012.05.003.
apenergy.2018.07.066. [42] Kalman H, Sher E. Enhancement of heat transfer by means of a corona wind created
[13] Turner JWG, Popplewell A, Patel R, Johnson TR, Darnton NJ, Richardson S, et al. by a wire electrode and confined wings assembly. Appl Therm Eng 2001;21:265–82.
Ultra boost for economy: extending the limits of extreme engine downsizing. SAE https://doi.org/10.1016/S1359-4311(00)00038-7.
Int J Engines 2014;7. https://doi.org/10.4271/2014-01-1185. [43] Discepoli G, Cruccolini V, Dal Re M, Zembi J, Battistoni M, Mariani F, et al.
[14] Takahashi E, Kojima H, Furutani H. Advanced ignition technology for the Experimental assessment of spark and corona igniters energy release. Energy
achievement of high thermal efficiency of internal combustion engine. Synth Procedia 2018;148:1262–9. https://doi.org/10.1016/j.egypro.2018.08.001.
English Ed 2015;8:187–95. https://doi.org/10.5571/syntheng.8.4_187. [44] Yu S, Wang M, Zheng M. Distributed Electrical Discharge to Improve the Ignition of
[15] Hayashi N, Sugiura A, Abe Y, Suzuki K. Development of ignition technology for Premixed Quiescent and Turbulent Mixtures, 2016. doi: 10.4271/2016-01-0706.
dilute combustion engines. SAE Int J Engines 2017;10. https://doi.org/10.4271/ [45] Wang F, Liu JB, Sinibaldi J, Brophy C, Kuthi A, Jiang C, et al. Transient plasma
2017-01-0676. ignition of quiescent and flowing air/fuel mixtures. IEEE Trans Plasma Sci
[16] Reyes M, Tinaut FV, Giménez B, Pérez A. Characterization of cycle-to-cycle varia- 2005;33:844–9. https://doi.org/10.1109/TPS.2005.845251.
tions in a natural gas spark ignition engine. Fuel 2015;140:752–61. https://doi.org/ [46] Pineda DI, Wolk B, Chen J-Y, Dibble RW. Application of corona discharge ignition
10.1016/j.fuel.2014.09.121. in a boosted direct-injection single cylinder gasoline engine: effects on combustion
[17] Sen AK, Ash SK, Huang B, Huang Z. Effect of exhaust gas recirculation on the cycle- phasing, fuel consumption, and emissions. SAE Int J Engines 2016;9. https://doi.
to-cycle variations in a natural gas spark ignition engine. Appl Therm Eng org/10.4271/2016-01-9045.
2011;31:2247–53. https://doi.org/10.1016/j.applthermaleng.2011.03.018. [47] Schenk M, Schauer FX, Sauer C, Weber G, Hahn J, Schwarz C. Challenges to the
[18] Abidin Z, Chadwell C. Parametric study and secondary circuit model calibration ignition system of future gasoline engines – An application oriented systems com-
using spark calorimeter testing. SAE Pap 2015. https://doi.org/10.4271/2015-01- parison. Ignition Syst. Gasol. Engines Cham: Springer International Publishing;

12
V. Cruccolini, et al. Fuel 259 (2020) 116290

2017. p. 3–25. https://doi.org/10.1007/978-3-319-45504-4_1. [67] Irimescu A, Merola SS, Di Iorio S, Vaglieco BM. Investigation on the effects of bu-
[48] Bresler M, Attard W, Reese R. Investigation of alternative ignition system impact on tanol and ethanol fueling on combustion and PM emissions in an optically acces-
external EGR dilution tolerance in a turbocharged homogeneous direct injected sible DISI engine. Fuel 2018;216:121–41. https://doi.org/10.1016/j.fuel.2017.11.
spark ignited engine. SAE Int J Engines 2015;8. https://doi.org/10.4271/2015-01- 116.
9043. [68] White C, Steeper R, Lutz A. The hydrogen-fueled internal combustion engine: a
[49] Idicheria CA, Najt PM. Potential of Advanced Corona Ignition System (ACIS) for technical review. Int J Hydrogen Energy 2006;31:1292–305. https://doi.org/10.
Future Engine Applications. Ignition Syst. Gasol. Engines Cham: Springer 1016/j.ijhydene.2005.12.001.
International Publishing; 2017. p. 315–31. https://doi.org/10.1007/978-3-319- [69] Shawal S, Goschutz M, Schild M, Kaiser S, Neurohr M, Pfeil J, et al. High-speed
45504-4_19. imaging of early flame growth in spark-ignited engines using different imaging
[50] Marko F, König G, Schöffler T, Bohne S, Dinkelacker F. Comparative optical and systems via endoscopic and full optical access. SAE Int J Engines 2016;9. https://
thermodynamic investigations of high frequency corona- and spark-ignition on a CV doi.org/10.4271/2016-01-0644.
natural gas research engine operated with charge dilution by exhaust gas re- [70] Burrows J, Mixell K, Reinicke P-B, Riess M, Sens M. Corona Ignition – Assessment of
circulation. Ignition Syst. Gasol. Engines Cham: Springer International Publishing; Physical Effects by Pressure Chamber, Rapid Compression Machine, and Single
2017. p. 293–314. https://doi.org/10.1007/978-3-319-45504-4_18. Cylinder Engine Testing. In: Ignition Syst. Gasol. Engines - 2nd Int. Conf., 2014, p.
[51] Cimarello A, Grimaldi CN, Mariani F, Battistoni M, Dal Re M. Analysis of RF corona 87–107.
ignition in lean operating conditions using an optical access engine. SAE Tech Pap [71] Burrows J, Mixell K. Analytical and experimental optimization of the advanced
2017. https://doi.org/10.4271/2017-01-0673. corona ignition system. Ignition Syst. Gasol. Engines Cham: Springer International
[52] Cimarello A, Cruccolini V, Discepoli G, Battistoni M, Mariani F, Grimaldi C, et al. Publishing; 2016. p. 267–92. https://doi.org/10.1007/978-3-319-45504-4_17.
Combustion behavior of an RF corona ignition system with different control stra- [72] Quader AA. Lean Combustion and the Misfire Limit in Spark Ignition Engines, 1974.
tegies. SAE Tech Pap 2018;2018:1–19. https://doi.org/10.4271/2018-01-1132. doi: 10.4271/741055.
[53] Huang Z, Wang J, Liu B, Zeng K, Yu J, Jiang D. Combustion characteristics of a [73] Suess M, Guenthner M, Schenk M, Rottengruber H. Investigation of the potential of
direct-injection engine fueled with natural gas−hydrogen mixtures. Energy Fuels corona ignition to control gasoline homogeneous charge compression ignition
2006;20:540–6. https://doi.org/10.1021/ef0502453. combustion. Proc Inst Mech Eng Part D J Automob Eng 2012;226:275–86. https://
[54] Aleiferis PG, Serras-Pereira J, Richardson D. Characterisation of flame development doi.org/10.1177/0954407011416905.
with ethanol, butanol, iso-octane, gasoline and methane in a direct-injection spark- [74] Wallner T, Ng HK, Peters RW. The effects of blending hydrogen with methane on
ignition engine. Fuel 2013;109:256–78. https://doi.org/10.1016/j.fuel.2012.12. engine operation, efficiency, and emissions. Engineering 2007. https://doi.org/10.
088. 4271/2007-01-0474.
[55] Di Iorio S, Sementa P, Vaglieco BM, Catapano F. An experimental investigation on [75] Rehbein N, Cooray V. NO production in spark and corona discharges. J Electrostat
combustion and engine performance and emissions of a methane-gasoline dual-fuel 2001;51–52:333–9. https://doi.org/10.1016/S0304-3886(01)00115-2.
optical engine. SAE Int 2014. https://doi.org/10.4271/2014-01-1329. [76] Aleiferis PG, Taylor AMKP, Ishii K, Urata Y. The nature of early flame development
[56] Srivastava DK, Agarwal AK. Comparative experimental evaluation of performance, in a lean-burn stratified-charge spark-ignition engine. Combust Flame
combustion and emissions of laser ignition with conventional spark plug in a 2004;136:283–302. https://doi.org/10.1016/j.combustflame.2003.08.011.
compressed natural gas fuelled single cylinder engine. Fuel 2014;123:113–22.
https://doi.org/10.1016/j.fuel.2014.01.046.
[57] Verhelst S, Wallner T. Hydrogen-Fueled Internal Combustion Engines. Elsevier Ltd; Glossary
2009. doi: 10.1016/j.pecs.2009.08.001.
[58] Di Iorio S, Sementa P, Vaglieco BM. Analysis of combustion of methane and hy-
drogen–methane blends in small DI SI (direct injection spark ignition) engine using
Abbreviations
advanced diagnostics. Energy 2016;108:99–107. https://doi.org/10.1016/j.energy.
2015.09.012. ACIS: Advanced Corona Ignition System
[59] Salvi BL, Subramanian KA. Experimental investigation on effects of exhaust gas AIT: After Ignition Timing
recirculation on flame kernel growth rate in a hydrogen fuelled spark ignition en- ATDCf: After Top Dead Center firing
gine. Appl Therm Eng 2016;107:48–54. https://doi.org/10.1016/j.applthermaleng. CAD: Crank Angle Degree
2016.06.125. CCV: Cycle-to-Cycle Variability
[60] Hora TS, Shukla PC, Agarwal AK. Particulate emissions from hydrogen enriched CMOS: Complementary Metal-Oxide Semiconductor
compressed natural gas engine. Fuel 2016;166:574–80. https://doi.org/10.1016/j. CNG: Compressed Natural Gas
fuel.2015.11.035. CoV: Coefficient of Variation
[61] Gong C, Li Z, Chen Y, Liu J, Liu F, Han Y. Influence of ignition timing on combustion CR: Compression Ratio
and emissions of a spark-ignition methanol engine with added hydrogen under lean- ECU: Engine Control Unit
burn conditions. Fuel 2019;235:227–38. https://doi.org/10.1016/j.fuel.2018.07. EGR: Exhaust Gas Recirculation
097. HRR: Heat Release Rate
[62] Ma F, Wang Y, Liu H, Li Y, Wang J, Zhao S. Experimental study on thermal effi- IMEP: Indicated Mean Effective Pressure
ciency and emission characteristics of a lean burn hydrogen enriched natural gas MBT: Maximum Brake Torque
engine. Int J Hydrogen Energy 2007;32:5067–75. https://doi.org/10.1016/j. MFB: Mass Fraction Burned
ijhydene.2007.07.048. SI: Spark Ignition
[63] Battistoni M, Poggiani C, Grimaldi CN. Experimental investigation of a port fuel TDC: Top Dead Center
injected spark ignition engine fuelled with variable mixtures of hydrogen and
methane. SAE Tech Pap 2013. https://doi.org/10.4271/2013-01-0226.
[64] Serrano D, Laget O, Soleri D, Richard S, Douailler B, Ravet F, et al. Effects of me- Nomenclature
thane/hydrogen blends on engine operation: experimental and numerical in-
vestigation of different combustion modes. SAE Int J Engines 2010;3. https://doi.
CA0-5: Crank angle interval between ignition timing and MFB 5%
org/10.4271/2010-01-2165.
CA5-90: Crank angle interval between MFB 5% and 90%
[65] Bhatti SS, Verma S, Tyagi SK. Energy and exergy based performance evaluation of
req avg: average equivalent flame radius
variable compression ratio spark ignition engine based on experimental work.
v20: Average flame growth speed when req avg is 20 mm
Therm Sci Eng Prog 2019;9:332–9. https://doi.org/10.1016/j.tsep.2018.12.006.
λ: Air-fuel equivalence ratio
[66] Parlak A, Erbas Y, Yasar H, Soyhan H, Deniz C. First and second law analysis of a
Δθ(r): Crank angle interval between ignition timing and a stated value (9 or 20 mm) of
gasoline engine for various compression ratios. Int J Veh Des 2009;49:111. https://
average equivalent flame radius
doi.org/10.1504/IJVD.2009.024243.

13

You might also like