You are on page 1of 17

Applied Energy 191 (2017) 311–327

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Impact of spark plug gap on flame kernel propagation and engine


performance
Tawfik Badawy a, XiuChao Bao b,a, Hongming Xu a,c,⇑
a
Vehicle and Engine Technology Research Centre, University of Birmingham, Birmingham B15 2TT, UK
b
School of Automobile and Transportation, Xihua University, Chengdu, China
c
State Key Laboratory of Automotive Safety and Energy, Tsinghua University, Beijing, China

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Effects of the spark plug gap on flame


kernel growth area and radius were
examined.
 Flame front development on the
horizontal swirl plane were
investigated.
 Effects of the spark plug gap on
engine performance and emissions
were investigated.

a r t i c l e i n f o a b s t r a c t

Article history: Experimental optical and thermal tests were carried out in a constant-volume combustion chamber and a
Received 27 October 2016 single cylinder gasoline direct injection (GDI) engine to obtain a comprehensive understanding of the
Received in revised form 24 January 2017 effects of spark plug electrode gap on flame kernel development, engine performance, and emissions.
Accepted 26 January 2017
High-speed Schlieren visualization was utilized to study the flame kernel growth at different equivalence
ratios. Planar Laser Induced Fluorescence (PLIF) was employed to investigate the combustion zone and
the flame front development on the horizontal swirl plane after spark ignition. High-speed imaging tech-
Keywords:
nique was carried out to study turbulent flame propagation. Combustion analysis, using in-cylinder pres-
GDI engine
Spark plug gap
sure data and Mass Fraction Burned (MFB) was employed, along with exhaust emissions measurement to
Flame front obtain a better understanding of the spark plug gap effects on engine performance and emissions. It is
Combustion found that the flame kernel growth area increases as the spark plug gap increases. PLIF imaging for
Emissions the combustion process inside the GDI engine demonstrate a larger flame kernel associated with the lar-
ger gap. The maximum in-cylinder pressure, turbulent flame speed, heat release rate, and the mass frac-
tion burned increases with the spark plug gap. The engine output increases slightly and the combustion
process becomes more stable due to the reduction in cyclic variations as the spark plug gap increases.
With the maximum spark plug gap, the engine produces minimum hydrocarbon emissions and

⇑ Corresponding author at: University of Birmingham, Birmingham, UK.


E-mail addresses: tsb308@bham.ac.uk (T. Badawy), h.m.xu@bham.ac.uk (H. Xu).

http://dx.doi.org/10.1016/j.apenergy.2017.01.059
0306-2619/Ó 2017 Published by Elsevier Ltd.
312 T. Badawy et al. / Applied Energy 191 (2017) 311–327

Nomenclature

Dt time interval between the two images PLIF Planar Laser Induced Fluorescence
ATDC after top dead centre V the average flame speed
BTDC before top dead centre PN particulate number
CAD crank angle degree DS augmentation of the flame area
COV coefficient of variation MIE minimum ignition energy
ECU engine control unit AIT after ignition timing
EMOP exhaust maximum opening position u fuel-air equivalence ratio
SI spark-ignition TDC top dead centre
IMOP intake maximum opening position EGR exhaust gas recirculation
GDI gasoline direct injection ULG unleaded gasoline
HC hydrocarbon IMEP indicated mean effective pressure
HRR heat release rate
RFSI Radio Frequency sustained Plasma Ignition
MFB mass fraction burned

particulate number concentration. NOx emissions are increased as the spark plug gap becomes wider due
to the higher temperature accompanied with the increase in flame speed and in-cylinder pressure.
Ó 2017 Published by Elsevier Ltd.

1. Introduction hydrogen–air mixture. Their study sheds more lights on the roles
of the various physical phenomena in spark kernel formation and
Over the past several decades, a number of studies have been ignition, in particular the important effects of viscosity, pressure
conducted to investigate the effects of spark plug design on gradients, electrode geometry, and hot gas confinement.
spark-ignition (SI) engine performance. The spark plug firing end Recently, ignition systems utilized the plasma sustained igni-
design features such as gap projection, gap size, electrode size, tion systems as a promising alternative to conventional spark
and tip configuration demonstrated influences on the engine per- plugs. Mariani and Foucher [12] compared the impacts of a Radio
formance [1–4]. Cycle-to-cycle variations in Indicated Mean Effec- Frequency sustained Plasma Ignition System (RFSI) with a conven-
tive Pressure (IMEP) and engine emission levels were considered to tional spark plug on the performance of a spark ignition engine.
be important indicators to evaluate the effects of various spark They concluded that in all test conditions the RFSI improved engine
plug designs. Improved spark plugs must be employed to ignite efficiency, extended the lean limit of combustion and reduced the
leaner mixtures and to endure under harsh operating conditions cycle-by-cycle variability, in comparison with the conventional
for the high performance requirement [5]. Moreover, the spark spark plug. Furthermore, the adoption of the RFSI also had a posi-
plug electrode geometry design can be utilized in order to reduce tive impact on carbon monoxide and unburned hydrocarbon emis-
the heat losses from the flame kernel to electrode and speed up sions, whereas nitrogen oxide emissions increased due to higher
the kernel growth [6]. Likewise, the electrode material is of great temperatures in the combustion chamber.
practical importance to improving service life, ignitability, pre- Spark plug gap is considered one of the key factors that must be
ignition protection and fouling resistance of electrodes [7]. set properly before the plug is installed inside the engine for three
Lee and Boehler [8] investigated the effects of spark plug elec- reasons: (1) If the gap is too wide, the electrical voltage may not be
trode configurations on flame kernel formation and development, high enough to arc across, which would result in a misfire. (2) If the
and on engine performance. Three types of spark plugs were gap is too narrow, the spark may not ignite a ‘‘lean” air/fuel mix-
selected for evaluation. They were standard J-gap spark plug with ture, which would also result in a misfire. (3) The voltage require-
2.5 mm center electrode, J-gap spark plug with 0.6 mm center elec- ment of a spark plug is directly proportional to the size of the gap.
trode, and surface gap spark plug with 0.4 mm center electrode. Furthermore, the electrode gap influenced the early formation of a
Their study revealed that the spark plugs with 0.6 mm diameter flame (kernel), which played a dominant role in determining the
center electrode, referred to as fine wire spark plugs, were able subsequent behaviour of that flame, and thereby influenced the
to ignite the leanest mixtures and were the only spark plug that engine performance [13,14]. The kernel growth has been described
demonstrated the lowest predicted 0–2% mass fraction burned as a two-stage process: in the early short stage, the mass and
(MFB) times for both 0% and 20% exhaust gas recirculation (EGR). energy transfer processes are mainly controlled by the pressure
Also, the results displayed that fine wire spark plugs had signifi- wave and expanding plasma kernel; in the next much longer stage,
cant advantages in engine performance in terms of engine stability diffusion and thermal conduction dominate mass and energy
and fuel consumption rate. However, the difference between spark transfer when the flame is self-sustained [15]. In addition, Bhaskar
plug types becomes small as the average equivalence ratio [16] investigated the effect of different spark plug gaps of 0.4, 0.5,
approached 1.0. Han et al. simulated ignition in methane–air [9] 0.6 and 0.64 mm on the coefficient of variation (COV) of IMEP at
and hydrogen–air [10] mixtures and investigated the effects of different ignition timings and different engine loads. He concluded
temperature, electrode gap distance, electrode size (cylindrical that the minimum COV of IMEP was noticed for the spark gap of
electrodes), and spark duration on the predicted minimum ignition 0.6 mm and ignition timing of 18 CAD BTDC. Therefore, an incor-
energy (MIE). These studies used detailed chemistry and provided rect electrode gap may affect engine performance as the spark
new insight on the effect of different spark parameters on ignition. magnitude may be insufficient to ensure complete combustion of
Also, Bane et al. [11] examined the fluid mechanics following a the air-fuel mixture. Herweg and Ziegler [17] found that reducing
spark discharge and the effect on the ignition process in a the contact areas between the flame kernel and the spark plug can
T. Badawy et al. / Applied Energy 191 (2017) 311–327 313

be achieved either by reducing the electrode diameter and/or 2. Experimental setup


increasing the gap leads to a faster flame kernel development. Also,
the flow pattern near spark gaps is an important factor that gov- 2.1. Schlieren optical method
erns the flame kernel structure; the flow pattern is affected by
the spark electrode diameter, gap width, and spark duration [18]. Fig. 1 shows a detailed schematic of the Schlieren set up used in
Furthermore, increased gap spacing and gap projections are bene- the present study. A constant volume combustion vessel with two
ficial in improving the brake specific fuel consumption of the circular quartz observation windows (100 mm diameter) was uti-
engine, and the ability to ignite lean fuel/air ratios [2]. lized, alongside eight heating elements installed at each corner.
There are several ways in which a spark plug becomes inopera- Temperature modulation of the vessel was implemented via closed
tive, and these are outlined as follows: (1) Fouling which is pro- loop control to monitor the heating elements described and allow
duced by fuel wetting of the spark plugs or by deposition of continuous observation of the fuel-air mixture condition within
combustion products (soot and water) on the spark plug electrodes the chamber. The fuel was injected through a multi-hole GDI injec-
and insulator causing short circuit [19,20]. (2) Fouling with oil tor, which was mounted in the top cover of the chamber, and was
deposits from excessive passage of engine oil into combustion controlled by an Electronic control unit (ECU). For ignition process
chamber due to piston ring or valve guide seal leakage causing the spark plug was positioned in the centre of the top cover of the
open circuit [21]. (3) Breaking of the insulator. (4) Pre-ignition. vessel, and for safety purposes a pressure release valve was
(5) Conduction through the insulator. (6) Electrical puncture of mounted at the side of the vessel, and operating at 7 bar.
the insulator. The temperature of the spark plug’s firing end must 500 W xenon lamp combined with a lens array prior to an
be kept low enough to prevent pre-ignition, but high enough to adjustable aperture were employed to create a point light source.
prevent fouling and insure adequate electrode gap life [2]. This is Then concave mirror was utilized to receive this light and produce
called ‘‘Thermal Performance”, and is determined by the heat range a parallel beam, which passed through the circular quartz windows
selected. of the vessel, illuminating the region of interest for this test. Fol-
The current study presents a comprehensive examination of the lowing this, a second concave mirror on the opposite side of the
effect of the electrode spark plug gap on flame kernel development, vessel was used to collect the parallel beam, which integrated
engine performance, and emissions. High-speed Schlieren visual- the light prior to its intersection by a knife edge, to achieve the
ization was utilized to study the flame kernel growth at different desired Schlieren effect (two dimensional imaging). To document
equivalence ratio. Planar Laser Induced Fluorescence (PLIF) was the combustion events, a Phantom research V710 high-speed cam-
employed to investigate the combustion zone and the flame front era was utilized (synchronised with spark timing; no recorded
development on the horizontal swirl plane after spark ignition. delay), with a capture rate of 10 kHz (10,000 frames per second)
High-speed imaging technique was carried out to study turbulent and resolution of 800  800 pixels.
flame propagation. Combustion analysis, using in-cylinder pres- The burned gases scavenged through the exhaust by using com-
sure data and Mass Fraction Burned (MFB) was employed, along pressed air. After flushing and before each test, the vessel chamber
with exhaust emissions measurement to obtain a better under- was opened to the ambient air until the air temperature inside the
standing of the spark plug gap effects on engine performance and vessel stabilized at the test point. Once the temperature stabilized,
emissions. the valves to the chamber were closed and the fuel was injected

Fig. 1. Schematic diagram for Schlieren set up.


314 T. Badawy et al. / Applied Energy 191 (2017) 311–327

and to generate a homogenous fuel-air mixture, the mixture was Table 1


sustained for a minimum of five minutes to guarantee homogene- Key engine specifications.

ity and a relative state of inactivity. Following this, LAB VIEW pro- Combustion chamber design Pent-roof with a centrally mounted
gram was employed for triggering both the spark plug and the spark plug
camera simultaneously. After the ignition of the air-fuel mixture Displaced volume 562 cc
occurred, the burned products were extracted from the vessel Bore  stroke 89 mm  90.3 mm
chamber, enabling the experiment to be restarted. To ensure con- Compression ratio 11:1
Maximum intake valve lift/duration 10.5 mm/250 CAD
fidence in procedure, each test was repeated a minimum of three Maximum exhaust valve 9.3 mm/250 CAD
times, with the process being pursued at initial temperatures of lift/duration
90 °C and a variety of u from 0.8 to 1.2. Intake Maximum Opening Position 109 CAD BTDC
(IMOP)
Exhaust Maximum Opening Position 271 CAD ATDC
2.2. Single-cylinder optical engine
(EMOP)
Injector type Solenoid-actuated multi-hole (laser
A single cylinder 562 cc optical engine with extended Bowditch drilled)
piston arrangement was used in this work (see Fig. 2). The engine
utilized a pent roof combustion chamber with a centrally mounted
spark plug and four valves (two inlets, two exhausts). The piston thermodynamic properties to typical gasoline fuels, and 3-
crown included a 65 mm flat-topped quartz window, providing pentanone was chosen as a seeding tracer. The main physical and
optical access to the combustion chamber through a 45° stationary photophysical properties required of a fluorescent tracer for in-
mirror. A triangular quartz window was fitted on one side of the cylinder LIF studies include: boiling point and transport properties
combustion chamber to provide additional optical access. The closely match to the carrier fuel, absorption spectrum suitable for
engine specifications are summarized in Table 1. available laser wavelengths, satisfactory fluorescence quantum
The air/fuel equivalence ratio (k) was controlled by means of a yield, and insensitivity to oxygen quenching [22]. Previous studies
wide band lambda sensor (ETAS LA4-E). The lambda sensor had the of the photophysical behaviour of 3-pentanone indicated its
capability of supporting any fuel specified by H:C, O:C and N:C advantages over other common tracers for our work [23–25]. A
ratios. For the injection system, an electronic control unit (ECU) laser sheet with a thickness of approximately 1 mm and width of
was employed to adjust the amount of injection using the LabVIEW 49 mm was formed using wavelength matched mirrors and laser
program to change the air/fuel equivalence ratio. The lambda sen- sheet optic. This beam was directed into the combustion chamber
sor provided a visual display of relative air/fuel ratio (k), which through a triangular quartz window. The fluorescence signal was
enabled a range of lambda values from 0.4 to 10 to be logged via imaged at a right angle to the excitation plane via a fixed mirror
a data acquisition system. at 45° and was collected through a 105 mm f/4.5 UV-Nikkor lens
onto a gated image intensifier (Hamamatsu C10880-03F) linked
2.3. Planar Laser Induced Fluorescence (PLIF) system to the Phantom V710 CCD camera. The Camera gate width was
1 ls, and the intensifier was gated for a short duration of 280 ns
Fig. 3 presents a schematic diagram of the PLIF setup. A pulsed to reduce background light and combustion luminosity. A Schott
Nd: YAG laser with an output of 87 mJ/pulse at 266 nm was used to WG360 long-pass filter was placed in front of the camera to block
excite the molecules of the added tracer to the fuel to fluorescence, unwanted wavelengths, passing the LIF signal and reject elastic
at a specified crank angle. For the PLIF measurements, isooctane scattering from the walls. The repetition rate of the PLIF imaging
was used as a surrogate fuel due to its similar physical and system was limited to 10 Hz due to the repetition rate of the Nd:

Exhaust valves
Injector
Spark plug

Intake valves

(a) (b)
Fig. 2. Schematic of (a) the optical setup for high speed imaging experiments, (b) cylinder head configuration.
T. Badawy et al. / Applied Energy 191 (2017) 311–327 315

Quartz

(a) (b)
Fig. 3. The experimental system (a) Optical accesses of the single-cylinder engine, (b) schematic diagram of the PLIF setup.

YAG laser. Thus the optical engine was run at 1200 rpm. A shaft flame kernel formation. These experiments were performed in a
encoder was used to generate the trigger pulse at a specified crank constant-volume combustion chamber over a wide range of equiv-
angle of the engine. The Nd: YAG Laser and imaging system (cam- alence ratios in a carefully controlled environment and under lam-
era and intensifier) were controlled by a commercial time box and inar conditions. The flame kernel area was defined as the outer
were synchronised (with one crank angle degree resolution) with envelope of the flame, whilst the flame kernel radius was defined
the engine via a reference signal from a LabVIEW engine control as the vertical distance from the end of the center electrode to
unit. the boundary of the flame, see Fig. 4.
An in-house developed MATLAB program was used for image
processing for both the combustion vessel and engine results. To
2.4. Spark plug gaps, flame kernel (area & radius) definition and define the boundary of the flame, the raw image data were cor-
turbulent flame speed calculation rected by background subtraction and noise signal removal
through median filtering. Then, the background corrected image
For the current study, spark plugs with three different gaps of was converted to binary image by thresholding. Finally, the bound-
1 mm, 1.2 mm and 1.4 mm were used to investigate their impacts ary of the flame shape was identified by the software as shown in
on flame kernel growth, engine performance and emissions. The Figs. 4 and 5. For the engine raw flame images, a circular mask with
ignition system was a single coil-on-plug (COP) type, powered by a diameter slightly smaller than the piston window was laid over
a 12 V 9 A supply. The dwell time (which is defined as the period the original image to remove reflection from the metal housing
of time that the coil is turned on) was maintained at 6 ms provid- of the optical crown. Consequently, the computer program will
ing ignition energy of 35–40 mJ and spark duration of approxi- be able to distinguish between the location of the window area
mately 2 ms. A spark plug with a thin laser welded iridium tip and the flame boundary.
central electrode and J-type ground electrode (NGK-ILKAR6C10) Regarding the calculation of turbulent flame speed inside GDI
was used for all fired engine testing. The spark energy was fixed engine, the average flame speed in two adjoining images was
through the whole test and consequently the voltage was held con- defined as:
stant for all the three gaps.
Initially, a series of experiments were conducted to investigate DS
V¼ ð1Þ
the effects of different spark plug gaps on the early stages of the L Dt

(a) (b)
Fig. 4. (a) Spark plug, (b) flame kernel area and radius definitions.
316 T. Badawy et al. / Applied Energy 191 (2017) 311–327

Fig. 5. Flame speed calculation procedure.

Table 2
Summaries experimental tests, techniques and fuels used in this work.

Experiment Flame kernel area and Flame front detection Combustion flame imaging Combustion pressure analysis and
radius measurements emissions measurements
Technique Schlieren high-speed PLIF measurements High-speed imaging of flame Horiba MEXA-7100DEGR gas
imaging of flame analyser and DMS 500
Test environment Constant volume chamber Single- cylinder optical engine Single-cylinder optical engine Single-cylinder thermal engine
90 °C initial temperature Speed:1200 rpm, spark Speed:1200 rpm, spark Speed:1200 rpm, spark
and 1 bar initial pressure timing:25CAD BTDC, injection timing:25CAD BTDC, injection timing:25CAD BTDC, injection
timing:280CAD BTDC timing:280CAD BTDC timing:280CAD BTDC
Fuel Gasoline (ULG95) with 3.3% iso-octane doped with Gasoline (ULG95) with 3.3% ethanol Gasoline (ULG95) with 3.3% ethanol
ethanol 3-pentanone (3% by mass)

In which DS is the augmentation of the flame area; Dt is the time 1.4 mm is approximately 672.8 and 903 mm2 respectively com-
interval between the two images; and L is the length of the flame pared to 286.8 mm2 for 1 mm gap.
boundary. Fig. 5 is an illustration of the calculation process. This big difference in the average flame kernel area can be due
For the current study all the experimental tests, techniques and to the extension of the spark plug gap from 1 to 1.2 and 1.4 mm,
fuels are summarized in Table 2. leading to higher ignition energy with a larger plasma volume
and more contact with unburned gas. As a result, a faster flame
kernel is developed, which accelerates the mass fraction burnt,
3. Results and discussion resulting in higher heat release rate [26]. In addition, spark elec-
trode configuration plays a dominant role in flame kernel develop-
3.1. Flame kernel propagation for different spark plug gaps ment due to the amount of heat loss to the spark electrode.
Therefore, for narrow gaps a small core flame will be produced
Fig. 6 shows the flame kernel growth at equivalence ratio of and consequently the energy lost will be increased due to the heat
u = 1 for three different spark plug gaps of 1, 1.2 and 1.4 mm. transfer from the flame kernel to the electrodes. Furthermore, for
The fuel used was gasoline at initial temperature and pressure of narrow gaps, a portion of the spark energy in the electrical fall
90 °C and 1 bar, respectively. At the beginning of the flame kernel regions of the cathode and anode may be lost to the electrodes
initiation and up to 1 ms from the start of spark, the difference in because of the fall regions being in close proximity to the electrode
the flame kernel size between the three gaps is relatively small surface [27–29]. Consequently, the flame propagation rate will be
especially for 1.4 mm and 1.2 mm compared to 1 mm gap. As the limited [18]. Ishii et al. [18] revealed that as the spark plug gap
time after the start of flame kernel initiation is further increased, increased, the amount of heat losses reduced. By contrast, for small
spark plug gap of 1.4 mm and 1.2 mm gap generate a flame kernel plug gap a large amount of hot gas was captured in the spark gap
area significantly larger compared to that of spark plug gap of and then the recirculation flow moved the captured gas toward the
1 mm. At 2 ms the average calculated flame kernel area of gaps spark electrode tip and contributed to the increment of heat loss.
1.2 and 1.4 mm is approximately 21.3 and 35.6 mm2 respectively The wider the gap, the greater is the volume of air-fuel mixture
compared to 8.6 mm2 for 1 mm gap. Further increasing of the time exposed to the spark which assists in ignition lean mixtures [1].
up to 8 ms, the average calculated flame kernel area of gaps 1.2 and As the electrode gaps increases, lower quenching losses occur,
T. Badawy et al. / Applied Energy 191 (2017) 311–327 317

Fig. 6. Comparison of typical flame growth for three different spark plug gaps of 1 mm, 1.2 mm and 1.4 mm at equivalence ratio of u = 1.

Fig. 7. Comparison of typical flame growth for three different spark plug gaps of 1 mm, 1.2 mm and 1.4 mm at different equivalence ratio of u = 0.9, u = 1, u = 1.1 and u = 1.2
at 6 ms after start of flame kernel initiation.

and this leads to improved conditions for ignition. In contrast, lar- not affected by the spark plug gap especially for gaps of 1.2 mm
ger gaps demand higher ignition voltages. and 1.4 mm. The reasons for a faster flame speed at slightly rich
Fig. 7 shows the flame kernel growth for three different spark mixture setting with an equivalence ratio near to u = 1.2, that more
plug gaps of 1, 1.2 and 1.4 mm at different equivalence ratios of fuel molecules are presented in the chamber during combustion
u = 0.9, u = 1, u = 1.1 and u = 1.2 at 6 ms after start of flame kernel and hence more radicals are formed ahead of the flame front,
initiation. For lean and stoichiometric conditions, the difference in and significantly higher flame temperatures are achieved [30].
flame kernel growth is high between the three spark plug gaps. However, at very high rich conditions the flame temperature will
Spark plug gap of 1.2 mm and 1.4 mm has a significant larger flame start to decrease again due to the incomplete combustion process,
kernel compared to gap 1 mm. The effect of the spark plug gap and hence less thermal energy will be released. Likewise, for lean
reduces once it proceeds to richest conditions of u = 1.1 and mixtures, the flame temperature decreases due to the lower fuel
u = 1.2. For rich condition of u = 1.2 the flame kernel growth is mass flow rate available which releases less thermal energy [31].
318 T. Badawy et al. / Applied Energy 191 (2017) 311–327

Fig. 8. Flame kernel area development as a function of time with spark plug gap of 1, 1.2 and 1.4 mm for different equivalence ratio of u = 0.9, u = 1, u = 1.1 and u = 1.2.

3.1.1. Flame area particularly between gaps of 1.2 mm and 1.4 mm. At 6 ms, the
Fig. 8 shows a quantitative comparison of the average flame average calculated areas are summarized in Table 3.
area as a function of time after spark initiation with spark plug The ignition energy and the heat losses are the key parameters
gaps of 1, 1.2 and 1.4 mm at different equivalence ratios. To ensure which play a dominant role on the flame kernel propagation. When
high confidence in data reliability, each test was repeated a mini- the gap is widened, the breakdown energy increases almost in pro-
mum of three times with a maximum uncertainty of ±22.5 mm2 portion to the gap width [32]. These enhancements in ignition
in flame area calculation. For all the equivalence ratios, the energy of the wider spark gap can be attributed to the extended sur-
1.4 mm gap consistently has a larger flame area compared to the face of the plasma, which requires more energy to sustain itself,
other gaps. At 1 ms all the average flame area trends are closely particularly in conditions where the energy density is held con-
matched for all the spark plug gaps. As the time after ignition stant, resulting in a constant temperature gradient along the
increases, the effect of the electrode gap becomes significant espe- plasma surface [33,34]. This increase in the ignition energy is
cially for lean and stoichiometric conditions. As the mixture derived from the larger initial kernel volume. The enlarged kernel
becomes richer, the effect of the electrode gap starts to diminish stimulates the flame growth which burns faster because it has a lar-
ger flame surface area making contact with the unburned gas. As a
result, when the electrode gap is extended from 1 mm to 1.2 mm
Table 3 and 1.4 mm, the flame area after 6 ms and u = 1 enlarges by about
The average flame area in mm2 at 6 ms after ignition. 3.5 and 5 times, respectively compared to that of 1 mm gap.
1 mm gap 1.2 mm gap 1.4 mm gap
u = 0.9 26.5 108.6 147 3.2. Flame tomography imaging using PLIF technique
u=1 88 301.4 456.5
u = 1.1 388.7 547.8 626.6 PLIF imaging of isooctane & 3-pentanone was utilized to obtain
u = 1.2 717.8 795.2 820.3
planar details of the combustion zone and the flame front develop-
T. Badawy et al. / Applied Energy 191 (2017) 311–327 319

ment on the horizontal swirl plain after spark ignition. PLIF was averaged. Afterwards, the instantaneous flame which closely
employed for accurate identification of the flame front in a certain matched the average flame image was chosen for comparison as
plain just below the spark plug that could not be identified by shown in Fig. 9.
chemiluminescence imaging due to the projected line-of-sight nat- Fig. 9 shows that at 10 CAD after ignition timing (AIT), the flame
ure of the latter technique [35,36]. Furthermore, PLIF system was kernel appears on the left side of the spark plug. Furthermore, the
designed to meet certain criteria, including finding all edges with- spark plug gaps of 1.2 and 1.4 mm demonstrate a larger flame area
out any false detections and to give higher fidelity. The PLIF system compared to that of 1 mm gap. During the flame expansion, the
was synchronised with the engine crank angle to capture images at number of sizes and frequencies of turbulent length scales increase
25 CAD BTDC which was the ignition timing. PLIF tests were all as the flame kernel grows [37,38]. Those scales that are larger than
performed under stoichiometric conditions at 1200 rpm, with the the size of the kernel will primarily ‘convect’ it without distortion
engine coolant and oil temperature fixed at 90 °C, injection timing (e.g. away from the spark-plug electrodes), whilst those smaller
at 280 CAD BTDC and injection pressure of 150 bar. Due to the lim- than the kernel will distort and/or wrinkle it depending on the
ited optical access for laser excitation, the laser beam was adjusted degree of turbulence. The PLIF images reveal the presence of fresh
just below the spark plug ground electrode to cover the area at gases trapped in the burned gas area, which may be explained by
vicinity of the spark plug where crucial early flame developments the complex wrinkling of the flame front and an isotropic flame
took place. The flame front detection was determined based on the development [36,39]. The calculated curvature of the flame front
disappearance of 3-pentanone concentration in the flame front. is an important factor in understanding turbulence and chemistry
Additionally, 3-pentanone also did not induce any fluorescence interaction phenomena, and can be used to perform quantitative
through the flame zone. The fluorescence signal intensity analysis of flame fronts. Fig. 9 illustrates at 15 CAD AIT the flame
decreased between the unburned and burn gas. After the ignition, front contours boundary between burned and unburned gases
the fluorescence images were recorded to provide a qualitative regions. The spatial location variation and the wrinkling of the
measurement of the combustion process. For each test point, two flame front characterize the influence of cycle-to-cycle dispersion
sets of 200 backgrounds without any flame and 200 flame and turbulence motion on the flame development. The spark plug
images were captured. The background images were gap of 1.4 mm consistently has a larger flame area compared to the
averaged and then subtracted from the raw flame images to other gaps. At 15 CAD AIT the flame area of the spark plug
eliminate any background noise. The raw flame images were then gap1.4 mm is not totally covered due to the limited area of the

Injector

Fig. 9. Instantaneous PLIF images of the three spark plug gaps for u = 1 at different crank angle.
320 T. Badawy et al. / Applied Energy 191 (2017) 311–327

Fig. 10. Effect of the spark plug gap on the engine load (IMEP) for different Fig. 11. Effect of the spark plug gap on the COV of IMEP for different equivalence
equivalence ratio. ratio.

laser sheet. At 20 CAD AIT for all the spark gaps the flame area is
variation increases as the mixture becomes leaner with excess air
not totally covered by the laser sheet.
or more dilute with a higher burned gas fraction from residual
gases.
3.3. Impact of the spark plug gap on the load and COV of IMEP The smallest gap of 1 mm consistently results in higher COV of
IMEP for the entire range of equivalence ratio tested. The COV of
Load variations versus equivalence ratios for the three different IMEP of 1.2 mm and 1.4 mm gaps is significantly lower compared
spark plug gaps investigated are presented in Fig. 10. It can be to that of 1 mm gap for stoichiometric and rich conditions. Whilst
noticed that as the equivalence ratio increases, the load increases for lean conditions the difference in COV of IMEP between the gaps
until it reaches the maximum and then starts to decrease again is reduced. Le Coz [41] demonstrated that the interactions by the
for very rich conditions. The lower mean effective pressure associ- flow field are categorized as the convection by the mean large-
ate with the lean and very rich conditions is fundamentally linked scale velocity field (low frequency velocity) and the wrinkling by
to the flame speed. The maximum flame speed occurs when the the small-scale turbulence (high frequency fluctuation intensity).
mixture strength for hydrocarbon fuels is about 10% rich. There- He revealed that the cyclic variation in flame initiation duration is
fore, for both lean and very rich mixtures, the flame speed predominantly affected by flame kernel convection which is deter-
decreases. Lean mixtures release less thermal energy, resulting in mined by the large scale fluid motion when the mixture is lean and
lower flame temperatures and hence lead to lower flame speeds. homogeneous. Also, he stated that there would be no correlation
Whilst, very rich mixtures experience incomplete combustion, between turbulence and early combustion if the engine is operated
hence also release less thermal energy resulting in lower flame at air-fuel ratio leaner than u = 0.625. The small scale turbulence
speeds. The reduction of the flame speed leads to an increase in affected the initiation stability mostly for the equivalence ratio
the burning time losses, and consequently reduces both the indi- higher than u = 0.8. It is generally accepted that the more homoge-
cated mean effective pressure and engine power [40]. The smallest neous the mixture is, the lower the cyclic variability will be [42].
gap of 1 mm consistently results in lower loads for the entire range The higher COV of IMEP of smaller spark plug gaps can be linked
of equivalence ratios tested. For equivalence ratio smaller than 1, to the cyclic variations in heat loss due to the cyclic variations in
the larger gap of 1.4 mm results in higher load, but for the equiva- the contact area between flame and electrodes. These variations
lence ratios higher than 1, the 1.2 and 1.4 mm gaps result in similar in the heat losses cause significant cyclic variations in flame devel-
loads. This behaviour can be justified by the fact that lean mixtures opment [43,18]. The contact area between flame and electrodes is
are more susceptible to the initial kernel size, whereas there is a largely controlled by the local flow field in the spark plug vicinity:
limit to the effect of initial kernel size on the combustion beha- the cycles in which the flame is convected away from the elec-
viour of rich mixtures. These results coincides with the aforemen- trodes have a smaller contact area than the cycles in which the
tioned observations regarding the average flame kernel area, which flame remains centred in the spark gap. The heat transfer from
demonstrates that the spark plug gap has a significant effect on the the hot gas kernel to the electrodes is mainly controlled by conduc-
flame kernel size at lean mixtures, whilst this effect nearly dimin- tion and convection. The heat loss from the flame kernel to elec-
ishes in rich mixtures. trode affects the flame growth through two mechanisms: (1) it
Fig. 11 demonstrates the effect of the three different spark plug decreases the kernel temperature, leading to a relative contraction
gaps investigated on the coefficient of variation (COV) of IMEP for of the kernel, and (2) it takes heat out of the flame front, and
different equivalence ratios. Cyclic variations in the combustion thereby decreases the burning velocity. Consequently, the lower
process are caused mainly by (1) the variations in mixture motion burning velocity will lead to slower flame development and the
within the cylinder at the time of spark from cycle-to-cycle. (2) cycle-by-cycle variations in the phasing of the main combustion
Variation in the amounts of air and fuel feed to the cylinder each event become larger.
cycle. (3) Variation in the mixing of fresh mixture and residual Moreover, a magnificent difference in the COV of IMEP between
gases within the cylinder each cycle, especially in the vicinity of the gaps is noticed especially for stoichiometric and rich condi-
the spark plug. The magnitude of cycle-by-cycle combustion tions. This can be linked to the shrouding effect of closer gaps,
T. Badawy et al. / Applied Energy 191 (2017) 311–327 321

which reduces the benefits of mixture turbulence within the gap, Fig. 14 shows the effect of variations of spark plug gaps on the
and consequently leads to large contact area between flame and in-cylinder pressure at different equivalence ratios. In-cylinder
electrodes, resulting in more heat losses from the flame kernel to pressure measurements in the engine were made using a Kistler
electrode. In addition, stoichiometric and slightly rich conditions 6051A piezoelectric pressure transducer along with a Kistler
have higher flame temperature, which increases the heat losses 5011B charge amplifier, which was used to amplify the signal from
by conduction to the electrode due to the less turbulent flow near the transducer. Both the pressure transducer and the charge ampli-
the closer gap [43]. fier had a linearity error of less than 0.5% of the full scale. In order
At stoichiometric condition u = 1, the reduction in COV of IMEP to acquire detailed in-cylinder pressure data with a good resolu-
of the gap 1.2 mm and gap 1.4 mm is 22.3% and 24.7% respectively, tion, the in-cylinder pressure data was recorded at high speed with
compared to that for the gap of 1 mm. whilst for the rich condition a resolution of 0.1 CAD for 300 consecutive cycles, and the data
of u = 1.1 the reduction in COV of IMEP of the gap 1.4 mm and gap recorded by LabVIEW program. Experimental uncertainty was
1.2 mm is 27.5% and 20.5% respectively, compared to gap 1 mm. evaluated at 95% confidence level for the in-cylinder pressure data
For the rich condition of U = 1.2, the reduction in COV of IMEP of to ensure confidence that the difference between the gaps peak
the gap 1.4 mm and gap 1.2 mm is 31.96% and 9.59% respectively, pressure was not due to the measurements error. It was found that
compared to gap 1 mm. the maximum uncertainty in the peak pressure measurements was
approximately 0.78%.
3.4. Impact of the spark plug gap on flame speed, ROHR, MFB and in- At each testing condition, the in-cylinder pressure results were
cylinder pressure averaged from 300 consecutive cycles. The maximum in-cylinder
peak pressure is demonstrated at stoichiometric condition of
The single cycle combustion images shown in Fig. 12 represent u = 1. For both lean and rich conditions, the associated flame speed
the typical cycle for each spark plug gap at stoichiometric condi- will be lower, and consequently the peak pressure will be reduced.
tion of U = 1, 25 CAD BTDC spark timing, 280 CAD BTDC injection For lean mixture conditions, the mass flow rate of fuel is low which
timing and the IMEP vary approximately around 4.5 bar due to produces lower exothermic energy resulting in lower flame tem-
the change of the spark plug gap. In all the images, the flames tend perature and hence lower flame speed. Whilst for rich mixtures
to propagate towards the exhaust valves due to the higher local the in-cylinder temperatures are reduced due to the incomplete
temperature and the swirl, as shown in previous studies [44,45]. combustion, and consequently also produce less thermal energy
Furthermore, this can be attributed to the spark plug’s position clo- resulting in lower flame speed. This reduces the mean effective
ser to the exhaust valves and the clockwise tumble motion in the pressure and consequently the engine power reduces. For all
combustion chamber. In addition, the installation of the injector equivalence ratios, the maximum in-cylinder pressure occurs for
near to the exhaust valves along with the high possibility of spray spark gaps of 1.4 and 1.2 mm. Furthermore, due to the faster burn-
flash boiling causes the fuel to be relatively richer around this ing of the spark plug gaps 1.4 and 1.2 mm the peak pressure
region of the combustion chamber, resulting in higher flame prop- slightly shifts towards TDC compared to that of spark plug gap
agation speed adjacent to the exhaust valve side [36]. For this con- 1 mm.
dition, spark plug gaps of 1.4 and 1.2 mm demonstrate consistently Figs. 15 and 16 show the effect of variations of spark plug gaps
higher flame speed compared to that of 1 mm gap. This may be on the heat release rate and mass fraction burned characteristics
occurs when the spark plug gap is extended from 1 mm to calculated from the averaged in-cylinder pressure data. The MFB
1.2 mm and1.4 mm, where the ignition energy is increased, result- was defined as the accumulated heat released in successive crank
ing in large plasma volume which contacts more unburned gas. As angle ranging from the start to the end of combustion divided by
a result, a faster flame kernel will be developed, which accelerate the total released heat in the whole combustion process [46].

the mass fraction burnt which derives higher heat release rate. The heat release rate dQ was calculated using the following
dh
Fig. 13 shows the average calculated flame speed of gasoline equation [31];
fuel for different spark plug gaps at equivalence ratios of U = 0.8,
   
1 and 1.2. The flame speed calculation is obtained from the dQ c dV 1 dP
¼ P þ V  ð2Þ
MATLAB according to Fig. 5. In order to ensure that the program dh c1 dh c1 dh
only process the flame data contained within the respective fields
of view, i.e. piston crown window, a mask has been applied to each where the heat capacity ratio (c) is the ratio of specific heats
image individually before the filtering takes place. Furthermore, (Cp/Cv); h is the crank angle; Q is the released heat; P and V are
Flame speed is shown to increase up to crank angles near to 30 the pressure and cylinder volume.
CAD AIT, as the flame begins to expand into the unburned charge The highest rate of increase in the HRR curve is obtained from
of the combustion chamber. Beyond this crank angle the flame 1.4 mm gap, followed by 1.2 and 1 mm gap. Once more, the leanest
speed decreases as the flame reaches the end of the optical field conditions have a significantly lower peak HRR compared to that of
of view and begins to interact with the cylinder walls. Therefore, the richest condition. The mixture strength influences the rate of
for the current study, the flame speed is investigated mainly up combustion and amount of heat generated. The maximum flame
to 30 CAD AIT. Fig. 13 shows that for the all equivalence ratios speed for all hydrocarbon fuels occurs at nearly 10% rich mixture.
the flame speed of spark plug gap of 1.4 mm and 1.2 mm is higher Flame speed is reduced for lean as well as very rich mixture. Lean
compared to the spark plug gap of 1 mm. At equivalence ratio U = 1 mixture releases less heat resulting in lower flame temperature
the spark plug gap of 1.4 mm has the highest flame speed of about and lower flame speed. Very rich mixture results incomplete com-
6.8 m/s at 362 CAD, nearly 1.5% faster than the peaks of the spark bustion (C and CO instead of CO2) and also results in production of
plug gap of 1.2 mm and 7% for gap of 1 mm. The flame speed of less heat and flame speed remains low.
spark plug gap of 1.4 mm at 350 CAD is about 3.52 m/s, 9.33% The results demonstrate that as the spark plug gap increases,
higher than that of 1.2 mm gap and 11.5% to 1 mm gap. This means the ignition energy is increased, which derives higher heat release
shortly after the ignition, the flame speed of 1.4 mm gap becomes rate and accelerates the mass fraction burnt. The heat release rate
significantly faster than the other two gaps. Additionally, spark is seen to have a profile similar to an isosceles triangle, with a max-
plug gap of 1.4 mm has the shortest ignition delay among the three imum of 23 J/deg at about 13 CAD ATDC for u = 1 for gap of
gaps. These two reasons result in the very fast flame propagation in 1.4 mm, corresponding to the 50% value on the mass fraction
1.4 mm gap combustion, as shown in previous figures. burned curve. Table 4 also summarize the mass burnt fraction at
322 T. Badawy et al. / Applied Energy 191 (2017) 311–327

CAD AIT Spark plug gap 1 mm Spark plug gap 1.2 mm Spark plug gap 1.4 mm

Exhaust

10

Intake
Intake

15

20

25

30

Fig. 12. Instantaneous flame images of gasoline Fuel at stoichiometric condition of U = 1 under different spark plug gaps.

different equivalence ratio. For u = 0.8, as the gap is increased from rich conditions, the initial flame growth becomes faster and exhibit
1 mm to 1.4 mm, the initial flame growth becomes faster by about more heat release rate.
1.5 CAD. These results come from the increase of ignition energy The mass fraction burned curve has an exponential profile. In
and the expansion of the plasma volume which contacts more the initial phase of the combustion process, the mass burning rate
unburned gas [26]. As a result of faster kernel development, the will be dependent on the laminar flame speed, which is in turn a
maximum heat release rate is increased. For stoichiometric and function of mixture strength. Therefore, the fastest mass burning
T. Badawy et al. / Applied Energy 191 (2017) 311–327 323

Fig. 13. Flame speed development at various equivalence ratio (a) u = 0.8, (b) u = 1, (c) u = 1.2 under different spark plug gaps.

Fig. 14. Effect of the spark plug gap on the in-cylinder pressure for different equivalence ratio.

Fig. 15. Effect of the spark plug gap on heat release rate for different equivalence ratio.

is displayed at stoichiometric conditions compared to that of both minimum for spark gap of 1.4 mm with 27.32 CAD while for
lean and rich conditions. The spark plug gap of 1.4 mm is expected 1.2 mm is 27.8 CAD and for 1 mm is 28.15 CAD. The decrease in
to have a faster mass burning due to its large plasma volume, the ignition delay for spark gap of 1.4 mm is due to the increase
which contacts more unburned gas compared to that of the other occurs in the ignition energy. For 50% mass fraction burned, the
gaps. Consequently, a faster flame kernel will be produced, which spark gap of 1.4 is faster with 37.5 CAD compared with spark
accelerates the mass fraction burnt, and derives higher heat release gap of 1.2 mm with 38 CAD and spark gap of 1 mm with 39.14
rate. For stoichiometric condition of u = 1 the ignition delay is CAD. At the lower flame temperatures corresponding to leaner
324 T. Badawy et al. / Applied Energy 191 (2017) 311–327

Fig. 16. Effect of the spark plug gap on mass burned fraction for different equivalence ratio.

Table 4
Crank angle position ASOC for different mass burned fraction.

Gap u = 0.8 u=1 u = 1.2


t0–10 (ignition delay) t0–50 t0–90 t0–10 (ignition delay) t0–50 t0–90 t0–10 (ignition delay) t0–50 t0–90
1 33 46 65 28.15 39.14 56.5 28.5 39.5 57.5
1.2 32 45 63 27.8 38 55 27.5 39 56
1.4 31.5 44.8 62.5 27.32 37.5 54 27 38.5 54.5

mixtures the rate of thermal and molecular diffusion is decreased, decreased. That is due to the interaction between the effects of
both which restrict flame propagation [47]. Therefore, the slowest increasing temperature greatly at slightly rich conditions, and Oxy-
mass burned fraction is noticed for U = 0.8. gen availability and gasoline high heating value on volume basis
for lean conditions. In order to interpret the NOx emissions beha-
viour, the trend of NOx emissions should be linked to the combus-
3.5. Impact of the spark plug gap on emissions
tion phasing and in-cylinder temperature distribution, which are
considered as main parameters of in-cylinder NOx formation in
3.5.1. Hydrocarbon and NOx emissions
the previous studies [31]. It is noticed that the spark plug gaps of
HC and NOx emissions for the (ULG95) at 150 bar injection pres-
1.2 and 1.4 mm consistently are associated with higher NOx emis-
sure were measured using a Horiba MEXA-7100DEGR gas analyser
sions for all equivalence ratios compared to that of 1 mm spark
with a resolution of 1 ppm. Fig. 17 shows the influence of the
plug gap. This can be demonstrated based on the in-cylinder pres-
engine operating equivalence ratio on the indicated specific NOx
sure data shown in Fig. 15. This data has clarified that the larger
emissions (g/kW h) for three different values of the spark plug
spark gaps consistently have higher peak pressure, and conse-
gaps. NOx emissions are maximum at lean condition of u = 0.9
quently higher temperature inside the cylinder resulting in higher
and decrease rapidly as the equivalence ratio is increased or
NOx emissions [48]. The spark plug gaps of 1.4 mm and 1.2 mm

Fig. 17. Effect of the spark plug gap on NOx emissions for different equivalence
ratio. Fig. 18. Effect of the spark plug gap on HC emissions for different equivalence ratio.
T. Badawy et al. / Applied Energy 191 (2017) 311–327 325

Fig. 19. Particulate number concentration of PN emission with NGK spark plug for equivalence ratio (a)-U = 0.8, (b)-U = 1, (c)-U = 1.

show a higher NOx emission of 10.04% and 9.42%, respectively mixture. With fuel rich mixture condition the level of HC increase
compared to 1 mm gap at lean condition of u = 0.8. At stoichiomet- due to incomplete combustion and crevice volume. For all equiva-
ric conditions the spark plug gaps of 1.4 mm and 1.2 mm display lence ratios the spark plug gap of 1.4 mm shows the minimum
an increase in NOx emission with 6.3% and 4.8%, respectively com- hydrocarbon emissions compared to spark plug gaps of 1.2 mm
pared to 1 mm gap. On the other hand, stoichiometric and lean and 1 mm. The maximum decrease in hydrocarbon emissions is
mixtures exhibit higher NOx emission variability. As it is revealed noticed at u = 0.8 for spark plug gap of 1.4 mm at with 12.4% com-
in Fig. 11, stoichiometric and lean mixtures lead to higher COV of pared to that of 1 mm gap. followed by 1.2 mm gap which dis-
IMEP, translated into more intense combustion variability, which played 8.7% decrease in hydrocarbon emissions compared to
leads to higher NOx emission variability [49]. 1 mm gap.
Fig. 18 shows the influence of the engine operating equivalence Based on the aforementioned results, as the spark gap is
ratios on the indicated specific HC emissions (g/kW h) for three dif- widened, a reduction in the HC emission is achieved. These results
ferent values of the spark plug gaps. The hydrocarbon emissions are consistent with the findings of Burgett et al. [1], who concluded
are minimum for lean condition of u = 0.9 and increase for both that wider gaps were responsible for decreases of HC emissions.
very lean conditions and richest conditions. If air-fuel ratio is too The wider the gap, the greater is the volume of air-fuel mixture
lean, poorer combustion occurs and the temperature is too low exposed to the spark which assists in the ignition of lean mixtures.
for hydrocarbon to burn late in the expansion stroke, this result Small gaps can cause the engine to misfire intermittently because
in large amounts of HC emissions, the extreme is total misfire at the spark plug electrodes quench the flame kernel due to the pos-
leaner air-fuel ratios. One misfire out of 1000 cycles gives exhaust sible heat transfer from the flame to electrode. Furthermore, closer
emissions of 1 g/kg of fuel used [50]. Furthermore, the low flame gap also provides a shrouding effect, and hence the benefits of mix-
speed at low u means that the flame may not even reach all the ture turbulence within the gap will be lost. Consequently, this will
326 T. Badawy et al. / Applied Energy 191 (2017) 311–327

lead to improper mixing around the spark plug, resulting in less from 1 mm to 1.2 mm and 1.4 mm the total particulate number
efficient combustion process and more HC emissions. concentration reduces with 20.7% and 33.2%. for rich conditions
u = 1.2, as the spark plug gap increases from 1 mm to 1.2 mm
3.5.2. Impact of the spark plug gap on the PN emissions and 1.4 mm the total particulate number concentration reduce
Particulate number (PN) concentration was measured in a range with 5.3% and 5.6%.
of 5–1000 nm using Cambustion DMS500 Fast Particulate Spec-
trometer. The device combines electrical mobility measurements 4. Summary and conclusions
of particles with sensitive electrometer detectors to give outputs
of particle size, number and mass in real time. The DMS500 was Experimental optical and thermal tests were carried out in a
connected directly to engine exhaust flow via its heated sampling constant-volume combustion chamber and a single cylinder gaso-
line close to the exhaust port. The sampling system incorporates line direct injection (GDI) engine, to obtain a comprehensive under-
two stages of dilution. The first diluter uses compressed dry air standing of the effects of spark plug electrode gap on flame kernel
metered by the DMS500 for low ratio dilution up to 5:1. Sample development, engine performance and emissions. High-speed Sch-
flow rate was 7.6 l/m and the heated line temperature was set at lieren visualization was utilized to study the flame kernel growth at
150 °C to avoid nucleation of hydrocarbons. The second diluter different equivalence ratios. Planar Laser Induced Fluorescence
uses a rotating disc to provide high dilution ratios which can be (PLIF) was employed to investigate the combustion zone and the
varied to maintain good signal to noise ratio. This dilution ratio flame front development on the horizontal swirl plane after spark
was adjusted automatically between 1 and 20 in orders to keep ignition. High-speed imaging technique was carried out to study
particulate number concentration in the range of 0.5–5%. turbulent flame propagation. Combustion analysis and examination
The device has 10% uncertainty in particle number measure- of engine emissions were carried out over different spark plug gaps.
ment over the full spectrum. Fig. 19 presents total PN concentra- The conclusions drawn from the work are as follows:
tion over different equivalence ratio for the three spark plug
gaps. For this discussion, particles with mobility diameters below 1. At the beginning of the flame kernel initiation and up to 1 ms
40 nm will be referred to as nucleation mode particles. Particles from the start of spark the difference in the flame kernel size
in this size range are due to incipient soot, small agglomerates, between different spark plug gaps is relatively small especially
and oxidized particles [51]. Above 40 nm in mobility diameter, for larger gaps such as 1.2 and 1.4 mm. As the time after igni-
particles are likely to be larger agglomerates and will be referred tion progresses, the larger spark plug gaps generate a signifi-
to as agglomeration mode particles (sometimes referred to as accu- cantly larger flame kernel areas compared to that of the spark
mulation mode particles). plug gap of 1 mm.
The data show the sensitivity of the engine out particle size to 2. The effect of the spark plug gap is dominant at lean and stoi-
global equivalence ratio. It was noticed that for the leanest cases, chiometric conditions and diminishes for rich conditions.
they are dominated by the smallest particle sizes, whereas the 3. The smallest spark gap of 1 mm generates lower engine load
richest cases show higher concentrations of large particles. Also output for the entire range of equivalence ratio tested. For
as u increases, the distribution consistently shifts toward agglom- equivalence ratio smaller than 1, the spark gap of 1.4 mm
eration mode particles. For richest conditions the amount of fuel results in higher load, but for equivalence ratios higher than
injected increases and consequently the injection duration 1, the 1.2 and 1.4 mm gaps result in similar loads. The smallest
increases and thus increases the likelihood of spray impingement gap of 1 mm consistently results in higher COV of IMEP for the
on in-cylinder surfaces and liquid fuel films. This wetting increases entire range of equivalence ratio tested. The COV of IMEP of
the probability of locally rich mixtures, causing more soot to be 1.2 mm and 1.4 mm gaps is significantly lower compared to
formed under partially premixed or diffusive combustion condi- that of 1 mm gap for stoichiometric and rich conditions. Whilst
tions rather than premixed conditions [52]. This can contribute for lean conditions the difference in COV of IMEP between the
to the larger number of agglomeration mode particles seen in the gaps is reduced.
study under rich equivalence ratios. Furthermore, due to the higher 4. The in-cylinder pressure, flame speed, heat release rate and the
in-cylinder temperatures, the temperature late in the expansion mass fraction burned increase with in spark plug gap. The max-
stroke and during the exhaust stroke will remain higher. This imum spark plug gap gives minimum hydrocarbon emissions
higher temperature can promote greater growth of particles under and particulate number concentration, but with higher NOx
rich conditions late in the expansion stroke and potentially even emissions due to the high temperature combustion tempera-
into the exhaust stroke. This higher temperature, combined with ture as well as the increase in flame speed and in-cylinder
lower oxygen concentrations and higher hydrocarbon concentra- pressure.
tions, can aid in soot growth.
For all the equivalence ratios tested the results of 1.2 mm and
1.4 mm gaps consistently have lower PN concentration compared References
to the 1 mm gap. This reduction in PN concentration emissions is
mainly due to the associated lower HC emissions generated from [1] Burgett R, Leptich J, Sangwan K. Measuring the effect of spark plug and ignition
system design on engine performance. SAE technical paper 720007; 1972.
the wider gaps as shown in Fig. 18. Furthermore, small gap rela- http://dx.doi.org/10.4271/720007.
tively experiences les turbulent intensity within the gap, and con- [2] Craver R, Podiak R, Miller R. Spark plug design factors and their effect on
sequently tend to degrade the fuel-air mixing around the spark engine performance. SAE technical paper 700081; 1970. http://dx.doi.org/10.
4271/700081.
plug, which increases the HC emissions and PN concentration [3] Nishio K, Oshima T, Ogura H. A study on spark plug electrode shape. Int J Veh
emissions [53]. Moreover, based on the aforementioned results, Des 1994;15:119–30.
the wider gaps produce lower COV of IMEP and higher in- [4] Yamaguchi T, Nakamura S, Oshima T. Spark plug and its electrode
configuration. Google Patents; 1987.
cylinder pressure, turbulent flame speed and heat release rate.
[5] Shimanokami Y, Matsubara Y, Suzuki T, Matsutani W. Development of high
These parameters will enhance the combustion efficiency and ignitability with small size spark plug. SAE technical paper 2004-01-0987;
reduce the PN concentration emissions [54]. At u = 0.8, as the spark 2004. http://dx.doi.org/10.4271/2004-01-0987.
plug gap increases from 1 mm to 1.2 mm and 1.4 mm the total par- [6] Hood S. The V-grooved electrode spark plug. SAE technical paper 901535;
1990. http://dx.doi.org/10.4271/901535.
ticulate number concentration reduces with 9.8% and 10.6%. Whilst [7] Lenk M, Podiak R. Copper cored ground electrode spark plug design. SAE
for stoichiometric conditions u = 1, as the spark plug gap increases technical paper 881777; 1988. http://dx.doi.org/10.4271/881777.
T. Badawy et al. / Applied Energy 191 (2017) 311–327 327

[8] Lee Y, Boehler J. Flame kernel development and its effects on engine [31] Heywood JB. Internal combustion engine fundamentals. New York: McGraw-
performance with various spark plug electrode configurations. SAE technical Hill; 1989.
paper 2005-01-1133; 2005. http://dx.doi.org/10.4271/2005-01-1133. [32] Song J, Sunwoo M. Analysis of flame kernel development with Schlieren and
[9] Han J, Yamashita H, Hayashi N. Numerical study on the spark ignition laser deflection in a constant volume combustion chamber. Proc Inst Mech Eng
characteristics of a methane–air mixture using detailed chemical kinetics: Part D: J Automobile Eng 2002;216:581–90.
effect of equivalence ratio, electrode gap distance, and electrode radius on MIE, [33] Arcoumanis C, Bae C. Correlation between spark ignition characteristics and
quenching distance, and ignition delay. Combust Flame 2010;157:1414–21. flame development in a constant-volume combustion chamber. SAE technical
http://dx.doi.org/10.1016/J.COMBUSTFLAME.2010.02.021. paper 920413; 1992. http://dx.doi.org/10.4271/920413.
[10] Han J, Yamashita H, Hayashi N. Numerical study on the spark ignition [34] Arcoumanis C, Bae C. Visualization of flow/flame interaction in a constant-
characteristics of hydrogen–air mixture using detailed chemical kinetics. Int J volume combustion chamber. SAE technical paper 930868; 1993. http://
Hydrogen Energy 2011;36:9286–97. http://dx.doi.org/10.1016/j. dx.doi.org/10.4271/930868.
ijhydene.2011.04.190. [35] Schießl R, Dreizler A, Maas U. Comparison of different ways for image post-
[11] Bane SP, Ziegler JL, Shepherd JE. Investigation of the effect of electrode processing: detection of flame fronts. SAE technical paper 1999-01-3651;
geometry on spark ignition. Combust Flame 2015;162:462–9. http://dx.doi. 1999. http://dx.doi.org/10.4271/1999-01-3651.
org/10.1016/j.combustflame.2014.07.017. [36] Aleiferis PG, Behringer MK. Flame front analysis of ethanol, butanol, iso-octane
[12] Mariani A, Foucher F. Radio frequency spark plug: an ignition system for and gasoline in a spark-ignition engine using laser tomography and integral
modern internal combustion engines. Appl Energy 2014;122:151–61. http:// length scale measurements. Combust Flame 2015;162:4533–52. http://dx.doi.
dx.doi.org/10.1016/j.apenergy.2014.02.009. org/10.1016/j.combustflame.2015.09.008.
[13] Hill P, Kapil A. The relationship between cyclic variations in spark-ignition [37] Abdel-Gayed R, Bradley D, Lawes M. Turbulent burning velocities: a general
engines and the small structure of turbulence. Combust Flame correlation in terms of straining rates. In: Proceedings of the royal society of
1989;78:237–47. http://dx.doi.org/10.1016/0010-2180(89)90128-4. London a: mathematical, physical and engineering sciences; 1987. p. 389–413.
[14] Bradley D, Lung F-K. Spark ignition and the early stages of turbulent flame [38] Bradley D. How fast can we burn? In: Symposium (international) on
propagation. Combust Flame 1987:71–93. http://dx.doi.org/10.1016/0010- combustion; 1992. p. 247–62.
2180(87)90022-. [39] Sacadura JC, Robin L, Dionnet F, Gervais D, Gastaldi P, Ahmed A. Experimental
[15] Kravchik T, Sher E. Numerical modeling of spark ignition and flame initiation investigation of an optical direct injection SI engine using fuel-air ratio laser
in a quiescent methane-air mixture. Combust Flame 1994;99:635–43. http:// induced fluorescence. SAE technical paper 2000-01-1794; 2000. http://dx.doi.
dx.doi.org/10.1016/0010-2180(94)90057-4. org/10.4271/2000-01-1794.
[16] Bhaskar H. Effect of spark plug gap on cycle-by-cycle fluctuations in four [40] Gupta HN. Fundamentals of internal combustion engines. PHI Learning Pvt.
stroke spark ignition engine. Int J Innovative Res Dev 2016;5. ISSN 2278–0211. Ltd; 2012.
[17] Herweg R, Ziegler G. Flame kernel formation in a spark-ignition engine. In: [41] Le Coz J. Cycle-to-cycle correlations between flow field and combustion
International symposium COMODIA. p. 173–8. initiation in an SI engine. SAE technical paper 920517; 1992. http://dx.doi.org/
[18] Ishii K, Tsukamoto T, Ujiie Y, Kono M. Analysis of ignition mechanism of 10.4271/920517.
combustible mixtures by composite sparks. Combust Flame 1992;91:153–64. [42] Pundir B, Zvonow V, Gupta C. Effect of charge non-homogeneity on cycle-by-
http://dx.doi.org/10.1016/0010-2180(92)90097-9. cycle variations in combustion in SI engines. SAE technical paper 810774;
[19] Quader AA, Dasch CJ. Spark plug fouling: a quick engine test. SAE technical 1981. http://dx.doi.org/10.4271/810774.
paper 920006; 1992. http://dx.doi.org/10.4271/920006. [43] Pischinger S, Heywood J. How heat losses to the spark plug electrodes affect
[20] Collings N, Dinsdale S, Hands T. Plug fouling investigations on a running flame kernel development in an SI-engine. SAE technical paper 900021; 1990.
engine-an application of a novel multi-purpose diagnostic system based on the http://dx.doi.org/10.4271/900021.
spark plug. SAE technical paper 912318; 1991. http://dx.doi.org/10.4271/ [44] Ma X, Jiang C, Xu H, Richardson S. In-cylinder optical study on combustion of
912318. DMF and DMF fuel blends. SAE technical paper 2012-01-1235; 2012.
[21] Hall C, Beaubier R, Marckwardt E, Courtney R. Spark plug fouling-a survey-test [45] Ma X, Xu H, Jiang C, Shuai S. Ultra-high speed imaging and OH-LIF study of
procedures-fuel factors. SAE technical paper 570254; 1957. http://dx.doi.org/ DMF and MF combustion in a DISI optical engine. Appl Energy
10.4271/570254. 2014;122:247–60.
[22] Baritaud T, Heinze T. Gasoline distribution measurements with PLIF in a SI [46] Andersson J, Collier A, Garrett M, Wedekind B. Particle and sulphur species as
engine. SAE technical paper 922355; 1992. http://dx.doi.org/10.4271/922355. key issues in gasoline direct injection exhaust. Nippon Kikai Gakkai
[23] Ghandhi J, Felton P. On the fluorescence behavior of ketones at high 1999;15:449–52.
temperatures. Exp Fluids 1996;21(2):143–4. http://dx.doi.org/10.1007/ [47] Sakai S, Hageman M, Rothamer D. Effect of equivalence ratio on the particulate
BF00193918. emissions from a spark-ignited, direct-injected gasoline engine. SAE technical
[24] Grossmann F, Monkhouse P, Ridder M, Sick V, Wolfrum J. Temperature and paper 2013-01-1560; 2013. http://dx.doi.org/10.4271/2013-01-1560.
pressure dependences of the laser-induced fluorescence of gas-phase acetone [48] Alasfour F. NOx emission from a spark ignition engine using 30% iso-butanol–
and 3-pentanone. Appl Phys B 1996;62(3):249–53. http://dx.doi.org/10.1007/ gasoline blend: Part 2—Ignition timing. Appl Therm Eng 1998;18:609–18.
BF01080952. [49] Karvountzis-Kontakiotis A, Ntziachristos L, Samaras Z, Dimaratos A, et al.
[25] Braeuer A, Beyrau F, Leipertz A. Laser induced fluorescence of ketones at Experimental investigation of cyclic variability on combustion and emissions
elevated temperatures for pressures up to 20 bars by using a 248 nm of a high-speed SI engine. SAE technical paper 2015-01-0742; 2015. http://
excitation laser wavelength: experiments and model improvements. Appl Opt dx.doi.org/10.4271/2015-01-0742.
2006;45(20):4982–9. [50] Ganesan V. Internal combustion engines. McGraw Hill Education (India) Pvt
[26] Song Jeonghoon, Seo Youngho, Sunwoo Myoungho. Effects of ignition energy Ltd; 2012.
and system on combustion characteristics in a constant volume combustion [51] Maricq MM, Harris SJ, Szente JJ. Soot size distributions in rich premixed
chamber. SAE, 2000-05-0016; 2000. ethylene flames. Combust Flame 2003;132:328–42. http://dx.doi.org/10.1016/
[27] Ko Y, Anderson RW. Electrode heat transfer during spark ignition. SAE S0010-2180(02)00502-3.
technical paper 0148-7191; 1989. http://dx.doi.org/10.4271/892083. [52] Maricq MM. Soot formation in ethanol/gasoline fuel blend diffusion flames.
[28] Ziegler GF, Wagner EP, Maly RR. Ignition of lean methane-air mixtures by high Combust Flame 2012;159:170–80. http://dx.doi.org/10.1016/
pressure glow and arc discharges. In: Symposium (international) on j.combustflame.2011.07.010.
combustion. p. 1817–24. [53] Zhao F, Lai M-C, Harrington D. Automotive spark-ignited direct-injection
[29] Swett Jr CC. Spark ignition of flowing gases; 1956. gasoline engines. Elsevier; 2000.
[30] Brown AG. Measurement and modelling of combustion in a spark ignition [54] Eastwood P. Particulate emissions from vehicles, vol. 20. John Wiley & Sons;
engine. Brunel University School of Engineering and Design PhD Theses; 1991. 2008.

You might also like