You are on page 1of 9

Fuel Processing Technology 91 (2010) 13641372

Contents lists available at ScienceDirect

Fuel Processing Technology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / f u p r o c

An investigation of the effects of spray angle and injection strategy on dimethyl ether
(DME) combustion and exhaust emission characteristics in a common-rail
diesel engine
Seung Hyun Yoon a, June Pyo Cha a, Chang Sik Lee b,
a
b

Graduate School of Hanyang University, Hanyang University, 17 Haengdang-dong, Sungdong-gu, Seoul, 133-791, Republic of Korea
Department of Mechanical Engineering, Hanyang University, 17 Haengdang-dong, Sungdong-gu, Seoul 133-791, Republic of Korea

a r t i c l e

i n f o

Article history:
Received 18 January 2010
Received in revised form 16 April 2010
Accepted 22 April 2010
Keywords:
Alternative fuel
DME (dimethyl ether)
Combustion characteristics
Exhaust emissions
Nano-particle

a b s t r a c t
An experimental investigation was performed on the effects of spray angle and injection strategies (single
and multiple) on the combustion characteristics, concentrations of exhaust emissions, and the particle size
distribution in a direct-injection (DI) compression ignition engine fueled with dimethyl ether (DME) fuel. In
this study, two types of narrow spray angle injectors (spray = 70 and 60) were examined and its results
were compared with the results of conventional spray angle (spray = 156). In addition, to investigate the
optimal operating conditions, early single-injection and multiple-injection strategies were employed to
reduce cylinder wall-wetting of the injected fuels and to promote the ignition of premixed charge. The
engine test was performed at 1400 rpm, and the injection timings were varied from TDC to BTDC 40 of the
crank angle.
The experimental results showed that the combustion pressure from single combustion for narrow-angle
injectors (spray = 70 and 60) is increased, as compared to the results of the wide-angle injector
(spray = 156) with advanced injection timing of BTDC 35. In addition, two peaks of the rate of heat release
(ROHR) are generated by the combustion of air-fuel premixed mixtures. DME combustion for all test
injectors indicated low levels of soot emissions at all injection timings. The NOx emissions for narrow-angle
injectors simultaneously increased in proportion to the advance in injection timing up to BTDC 25, whereas
BTDC 20 for the wide-angle injector. For multiple injections, the combustion pressure and ROHR of the rst
injection with narrow-angle injectors are combusted more actively, and the ignition delay of the second
injected fuel is shorter than with the wide-angle injector. However, the second combustion pressure and
ROHR were lower than during the rst injection, and combustion durations are prolonged, as compared to
the wide-angle injector. With advanced timing of the rst injection, narrow-angle injectors with multiple
injections could achieve low NOx levels and soot levels similar to single-injection cases.
Crown Copyright 2010 Published by Elsevier B.V. All rights reserved.

1. Introduction
Diesel engines are superior power sources among the automobile
engines because of their excellent performance, higher fuel efciency,
and lower exhaust emissions of, hydrocarbons (HCs), carbon
monoxide (CO), and carbon dioxide (CO2), as compared to gasoline
engines. On the other hand, the amounts of particulate matter (PM)
and nitrogen oxide (NOx) emitted are generally higher, as compared
with gasoline engines.
In recent years, with increasing population and industrialization,
fossil fuel prices have increased due to limited petroleum resources.
With regard to diesel engines, the regulations for PM and NOx
emissions have been strengthened to reduce negative health impacts.
Corresponding author. Tel.: + 82 2 2220 0427; fax: +82 2 2281 5286.
E-mail addresses: ysh3790@hanyang.ac.kr (S.H. Yoon), matia@hanyang.ac.kr
(J.P. Cha), cslee@hanyang.ac.kr (C.S. Lee).

Also important are reductions in greenhouse emissions in order to


mitigate climate change. Therefore, research on alternative fuels
which can be substituted for fossil fuels in diesel internal combustion
engines, has been conducted [13].
Dimethyl ether (DME, CH3OCH3) is a representative alternative
fuel which can be synthesized from natural gas, coal, crude oil, as well
as from non-fossil fuel feedstocks, such as biomass and waste
products. DME fuel is advantageous in compression ignition (CI)
engines, with an oxygenated molecular structure composed of an
oxygen atom (about 35% by volume) between two methyl radicals
(CH3). Due to the absence of direct carboncarbon (CC) bonding,
DME can drastically reduce or suppress the formation and development of soot during combustion while still providing conventional
diesel-like thermal efciency. In addition, DME has good ignition
capability in engines because it has a relatively high cetane number, as
compared with conventional diesel fuel and the high latent heat of
DME fuel leads to lower cylinder temperature of air-fuel mixtures

0378-3820/$ see front matter. Crown Copyright 2010 Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.fuproc.2010.04.017

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

the increase in unburned hydrocarbons (UHCs) and higher PM


concentration, are accompanied by fuel impingement on the cylinder
wall. The increased wall wetting upon earlier injection timing may be
caused by the PM, HC, and CO emissions during unstable and incomplete
combustion in the engine. Also, the wall impingement and wetting
effects in the conventional combustion chamber reduce the rate of fuel
evaporation, spray penetration, and fuel atomization [10].
Thus, more detailed investigation of the nano-particle emissions,
characteristics of reduced exhaust emissions, and combustion
performance of a DME fuel engine is necessary.
The purpose of this study, then, is to investigate the effects of a
narrow injection angle injector with a multiple-injection strategy on
these variables in a small common-rail DI (direct-injection) diesel
engine fueled with DME. Analysis results regarding the combustion
performance and emission characteristics upon split fuel injection and
partial premixed compression ignition with narrow injection angle
injectors are reported herein and compared with the results obtained
using conventional (wide) injection angle injectors.

Nomenclature
ATDC
BTDC
CA
CI
DI
DME
EGR
IMEP
FSN
LHV
PM
TDC
UHC

1365

after top dead center


before top dead center
crank angle
compression ignition
direct injection
dimethyl ether
exhaust gas recirculation
indicated mean effective pressure
ltered smoke number
lower heating value
particulate matter
top dead center
unburned hydrocarbon

early in the combustion phase. DME also has good atomization


properties due to its low boiling point. Due to these benets in diesel
engines, DME is the most promising alternative fuel for diesel engines
[4,5].
However, substituting with DME does not address the problem of
nitrogen oxide (NOx) emissions [6,7]. In order to reduce NOx emissions
in diesel engines, research has been performed using various injection
strategies to lower the combustion temperature. Kim et al. [8] reported
that multiple injection strategies affected the reduction in NOx emissions
when using CI engine-applied DME fuel. They carried out an experimental investigation with three injection mode cases: split injection,
pilot injection, and partial premixed charge compression ignition. The
results showed that a multiple injection strategy reduces NOx emissions
by limiting the increasing rate of the combustion temperature. When
investigating the effect of compression ratio and fuel injection angle on
HCCI (homogeneous charge compression ignition) combustion, Kim et
al. [9] reported that early fuel injection resulted in simultaneous
reduction in concentrations of NOx and soot emissions, as compared to
conventional injection timing. They showed the results of singleinjection, enhanced premixing, and low ame temperatures due to the
formation of overall lean mixtures. In conventional diesel fuel injection,
the early direct-injection and high fuel injection pressure, responsible for

2. Experimental setup and procedure


2.1. Experimental conguration
The engine used for this study is based on a single-cylinder,
naturally aspirated DI diesel engine with a double overhead cam
(DOHC) and a re-entrant combustion chamber. The engine has a bore
of 84.5 mm, a stroke of 75.0 mm, a displacement volume of 373.3 cm3,
and a compression ratio of 17.8. The experimental apparatus for the
DME engine consists of the test engine, DME fuel injection system,
dynamometer and control systems, combustion analyzer, exhaust
emissions analyzer, and nano-particle analyzer, as shown in Fig. 1.
The engine load and speed were controlled using a DC dynamometer (55 kW) system. The in-cylinder pressure was measured by a
piezoelectric pressure transducer (6052B1, Kistler), coupled to a
charge amplier (5011B, Kistler). For each test case, the combustion
pressure data were measured during 1000 cycles and acquired by a
DAQ board (PCI-MIO-16E-1, National Instrument) with a sampling
interval of 0.1 crank angle (CA) to ensure accurate ignition timing
and phasing of heat release. The in-cylinder pressure data, obtained as
a function of crank angle, was averaged in order to eliminate the effect

Fig. 1. Schematic of the experimental apparatus.

1366

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

of cycle-to-cycle variations and to calculate the rate of heat release


(ROHR) for analysis of combustion characteristics under each test
condition. An injection pressure controller (TDA-1100, TEMS) maintained the fuel pressure of the common-rail. An injector driver (TDA3300, TEMS) synchronized with a crank angle sensor is used to control
the injection timing and injection mass. The injection mass of DME
was measured using a fuel ow meter system (GP-30K, AND).
Exhaust emissions measurements without catalytic treatments
were made with a NOx analyzer (BCL-511, Yanaco), a soot analyzer
(AVL-407, AVL), and HCCO analyzer (MEXA-554JK, Horiba). The
measurements of the exhaust gas components were recorded after
the test engine had sufciently stabilized to steady state engine
(temperature) conditions. The temperatures of the intake air, exhaust
gas, coolant, and fuel were monitored; the ow rate of intake air was
controlled and measured using a ow control system (GFC67,
Aalborg). Nano-particle number and size distribution were measured
using a scanning mobility particle sizer (SMPS, Model 3936, TSI) and a
condensation particle counter (CPC). The measured mobility equivalent particle diameter ranged from 10.4 nm to 392.4 nm. The
sampled raw gas was heated and diluted using a raw gas diluter
(MD19-2E, Matter Engineering). Prior to each test, the combustion
pressure transducers and emission analyzers were calibrated to
ensure measurement consistency.
The geometry of the combustion chamber and test injectors is
shown in Fig. 2. In this study, three types of injectors are tested. Two
kinds of narrow spray angle injectors (spray = 60, 70) were
examined because Fig. 2 shows that, with early injection timing
(tinj = BTDC 40), the use of a conventional 156 degree injector caused
spray onto the cylinder liner or wall, which, therefore, increased PM,
HC, and CO emissions. For the early injection strategy, a narrow spray
angle injector is required to target the bowl. Advanced injection
timing, with a narrow spray angle, was adopted to obtain a partially
premixed compression ignition combustion pattern. This system
involves different mixture conditions, such as mean drop size,
penetration, and injection velocity, in the cylinder. These conditions
change during the injection, depending on the ambient conditions,
the injection pressure, and the spray angle [11]. Specic test injectors
are listed in Table 1.

2.2. Test fuels


DME (CH3OCH3) is one of the simplest ether compounds. The
physical and chemical fuel properties of DME are shown in Table 2. As
shown in this table, DME has only CH and CO bonds, without CC
bonds, and contains about 34.8% oxygen. Because of these properties,
DME combustion produces low PM emissions and can tolerate a
higher EGR rate to reduce NOx emissions to a greater extent than with
conventional diesel fuel. The lower heating value of DME, only 64.7%
of that of diesel fuel, requires a larger amount of injection fuel to
ensure the same engine power. The cetane number of DME is higher
than that of diesel fuel, which demonstrates good ignition capability.
The latent heat of evaporation of DME is much higher than that of
diesel fuel, which is benecial for reducing the mixture temperature.

Fig. 2. Schematic diagram of the combustion chamber and fuel spray (tinj = BTDC 40).

Table 1
Specications of test injectors.
Injector type

Common-rail (electronically controlled)

Injection pressure
Nozzle type
Number of nozzle holes
Nozzle hole diameter

60 MPa
Mini-sac
6
0.128 mm

Base injector (wide spray angle)


Spray angle (spray) Case 1

156

Modied injectors (narrow spray angle)


Spray angle (spray) Case 2
Spray angle (spray) Case 3

70
60

However, DME is in a gaseous state under standard atmospheric


conditions. Therefore, DME needs to be pressurized to over 0.5 MPa to
keep it in a liquid state under ambient temperature and pressure
conditions. The fuel delivery pressure should be increased to 1.7
2.0 MPa under engine operating conditions to prevent vapor lock in
the fuel injection system. The low viscosity of DME causes leakage in
the fuel supply system, which relies on small clearances for sealing. Its
lower lubricity characteristics result in intensied surface wear on the
moving parts within the fuel injection system. The compressibility of
DME is generally four to six times higher than that of diesel in a closed
system, and the compression work by the fuel delivery pump is
greater with DME than with diesel fuel [12,13].
In general, DME deteriorates the rubber seals, mainly due to its
corrosive nature. For that reason, all existing rubber seals in injection
systems should be replaced with non-corrosive materials. In this
study, to improve the durability of the injection system, 1000 ppm
lubricity additive (539 M, Lubrizol) was added to the DME fuel. In
addition, the pressure and temperature of returned DME fuel were
controlled and maintained by the pressure regulator and heat
exchanger to reduce changes in fuel properties, respectively.

2.3. Experimental conditions and procedure


All tests were conducted using a dynamometer system under a
constant engine speed of 1400 rpm. The coolant and oil temperatures
were maintained at 70 1 C to reduce variation in the test results.
Also, the injection pressure of DME fuel was xed at a constant
pressure of 60 MPa. With single injection, the fuel amount of 10 mg
per stroke was injected, and injection timings including the start of
injection (SOI) and the end of injection (EOI) were varied from BTDC
40 to TDC in steps of 5 with three types of injectors.

Table 2
Properties of DME fuel.
Properties

DME

Chemical formula
Mole weight
Carbon content
Hydrogen content
Oxygen content
Carbonhydrogen ratio
Relative density-gas
Relative density-liquid
Kinematic viscosity
Surface tension
Boiling point
Cetane number
Auto-ignition temperature
Stoichiometric A/F ratio
Lower heating value
Modulus of elasticity

CH3OCH3
46.07 (g/mol)
52.2 (mass%)
13.0 (mass%)
34.8 (mass%)
0.33
1.59 (15 C/1 bar)
0.66 (15 C/1 bar)
b0.1 (cST)
0.012 (N/m)
120130 (C)
5560
235 (C)
9.0
27.628.8 (MJ/kg)
6.37E8 (N/m2)

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

1367

In this study, the effect of multiple-injection strategies on


combustion and exhaust emissions was investigated. The strategy
was to divide the pulse signal for a single injection into two pulse
signals with dwell time. In the multiple-injection strategy, the total
amount injected matched the quantity of the single injection
(minj = 10 mg/stroke). First, injection occurred from BTDC 35 to
BTDC 15 in steps of 10. The rst injection mass of 2 mg was used to
form a premixed charge in the cylinder. The second injection occurred
at BTDC 5 and TDC, and the injection mass was 8 mg to control the
initiation of combustion.
For the nano-particle measurements, the sampled raw gas was
diluted 200 to 1, and then diluted gas was heated to 120 C to avoid
particle nucleation and condensation of the components of the
sampled raw gas. Detailed test conditions and the injection strategy
are shown in Table 3 and in Fig. 3, respectively.
Fig. 3. Injection current for single- and double-injection strategies (rst tinj = 35 BTDC,
and second tinj = 5 BTDC).

3. Experimental results and discussion


3.1. Effect of spray angle on the combustion and exhaust emissions
characteristics
The combustion pressures and rate of heat release (ROHR) of DME
fuel in single-injection mode with three types of injectors and
different injection timings at BTDC 5 and BTDC 35 are shown in
Fig. 4(a) and (b), respectively.
For the case of injection timing at BTDC 5, the trends of the heat
release curve for all cases indicated dominant premixed combustion
patterns, which generally appear in small DI diesel engines with high
injection pressure. The results of Case 1 (spray = 156) indicated a
relatively higher peak combustion pressure, ROHR value, and slightly
shorter ignition delay, as compared to Case 2 (spray = 70) and Case 3
(spray = 60). Moreover, the diffusive ame (combustion) phase of
Case 1 is considerably shorter than those of the narrow-angle
injectors (Case 2 and Case 3). These results can be explained by
observing that a conventional (wide) spray angle and combustion
chamber design are generally suitable for injection timings close to
the end of the compression stroke [14]. At that point, most injected
fuels are accepted into the bowl regions of the piston, and the injected
DME fuel is promptly auto-ignited due to the higher ambient pressure
and temperature. On the other hand, for Case 2 and Case 3, injected
fuel impinged directly along the surface of the conic extrusion area in
the middle of the piston bowl, as the distance from the nozzle tip to
the face of the piston bowl is considerably shorter due to the narrow
spray angles. For this reason, the formation of a homogenous air-fuel
mixture is hampered, and premixed spikes of heat release, lower

combustion pressures, and prolonged combustion phases are reduced, as compared to Case 1.
In Fig. 4(b), at the advanced injection timing of BTDC 35, most of
early-injected DME fuel for Case 1 is directed toward the piston head
and cylinder wall. In contrast, injected fuels in Cases 2 and 3 are
spread out along the piston cone wall and conned to the piston bowl,
as shown in Fig. 2, to congure the interaction between spray angle
and piston bowl. For this reason, the combustion pressures and heat
releases in Cases 2 and 3 are increased, as compared to that in Case 1.
In addition, the peak pressure and ROHR of Case 2 are slightly higher
than those of Case 3, with the same trends for the case of BTDC 5.
Injection timing of BTDC 35, in which the compression pressure and

Table 3
Experimental test conditions.
Test fuels

DME

Engine speed
Coolant temperature
Oil temperature
Intake-air temperature
Intake-air pressure

1400 (rev./min)
70 1 (C)
70 1 (C)
25 1 (C)
0.1 MPa (N/A)

Single-injection mode
Injection timing (tinj)
Mass of injection (minj)

BTDC 40 TDC
10 (mg/cycle)

Multiple (pilot)-injection mode


First injection timing (t1st)
Second injection timing (t2nd)
Mass of injection (minj)

BTDC 35, 25, 15


BTDC 5
2 + 8 (mg/cycle)

Measurement of nano-particle
Dilution rate
Dilution temperature

200 to 1
120 5 C

Fig. 4. Effect of spray angle and injection timings on the combustion characteristics in
single-injection mode (Pinj = 60 MPa, minj = 10 mg).

1368

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

in-cylinder charge temperature are very low, resulted in considerably


increased ignition delays. However, it can be suspected that earlyinjected DME fuel formed a well-premixed mixture during the
compression stroke, which resulted in homogenous charge combustion and a more active combustion process. Even when fuel injection
occurs earlier, a superior evaporation rate, higher cetane number, and
active low-temperature fuel oxidation reactions prevail. Consequently, the diffusive combustion phases disappear, and higher combustion
pressures and heat releases occur, as compared to the case with
retarded injection timing (tinj = BTDC 5). In particular, to consider
the patterns of heat release, two ROHR peaks are generated by the
combustion of air-fuel premixed mixtures. These combustion patterns
are similar to homogenous charge compression ignition (HCCI),
which consists of low-temperature combustion and high temperature
combustion phases [1517].
Especially, early direct injected DME fuel with narrow spray angle
injectors in this study gained similar results of the port injected DME
(premixed mixture) with in-cylinder diesel fuel HCCI combustion.
Ying et al. [18] carried out an investigation about the premixed DMEair charge and the conventional direct-injection compression ignition
combustion. They found that the shape of the heat release curve for
HCCI-DI case was not typical of DICI combustion (diffusion-controlled). A two-stage heat release process, low-temperature reaction
(cool ame) and high temperature reaction (hot ame) existed in
both HCCI-DI and HCCI combustion.
Fig. 5 shows the effect of injection timing on the indicated mean
effective pressure (IMEP), as a function of the start of injection (SOI),
for different spray angles with single-injection mode. The overall
decreasing trend of the IMEP with advanced injection timing for test
injectors was similar, and the values for Case 2 were much higher than
for other cases in the range from TDC to BTDC 25. As the injection
timing was advanced from TDC, the combustion process also
advanced. This advance occurred during the compression stroke,
linearly reducing combustion performance.
At further advanced injection timing (BTDC 30BTDC 40), the
values of IMEP for Case 1 (spray = 156) are rapidly decreased due to
the increased fuel wall-wetting phase and impinged fuel on the piston
head surface caused by longer spray tip penetration with immoderately early timing, despite the high fuel evaporating property of DME.
In general, the charge pressure and temperature, timed early during
the compression stroke, are much lower than those at TDC, causing
spray-wall interaction.
However, in the cases of narrow-angle injectors, Case 2 and 3
combustions showed that the values of IMEP were slightly decreased
and maintained at levels of BTDC 25, although injection timing
advanced to BTDC 40. Employing narrow spray angle injectors
avoided the interaction between fuel and the cylinder wall during

early injection timings. The spray-wall-wetting phenomenon was


subsequently reduced, and deterioration of combustion performance
was prevented. Furthermore, comparing Case 2 with Case 3, the IMEP
of Case 2 (spray = 70) is higher, with a constant difference of nearly
10%, as compared to Case 3 (spray = 60) in all test ranges. Examining
the impingement angle between the fuel spray and the piston cone
surface, smaller in Case 2 than Case 3, required the connement of
injected fuel to the piston bowl when the piston rises during the
compression stroke.
Fig. 6 shows the effect of spray angles and injection timings on the
concentrations of PM and indicated specic NOx (ISNOx) emissions in
single-injection mode. In most CI combustion processes, the formation of exhaust emissions are complicated and primarily inuenced by
chemical and physical fuel properties, fuel spray, as well as
atomization and combustion characteristics.
DME contains no sulfur, and its excellent evaporation characteristics can achieve better fuel atomization for reduced wall wetting.
DME combustion has been widely demonstrated to reduce unburned
emissions, such as PM, HC, and CO, as compared to petroleum diesel
combustion.
As shown in Fig. 6, DME combustion for wide (spray = 156) and
narrow injectors (spray = 70, 60) indicated low lter smoke number
(FSN, b0.006) at all injection timings because the high oxygen content
of the fuel, the absence of soot precursors (sulfur and aromatic
contents), and the short diffusion combustion process promote
smokeless combustion. Furthermore, the smoke numbers of narrow
injectors (Case 2 and 3) are considerably lower than those in Case 1. In
Case 2, nearly zero FSN was achieved at advanced timings of BTDC 20.
ISNOX emissions for Cases 2 and 3 increased in proportion to the
advance in injection timing up to BTDC 25 and, for Case 1, up to BTDC
20 due to the increase of combustion temperature and pressure. At
further advanced injection timings, the concentrations of ISNOx
emissions rapidly decreased. In particular, at test ranges of BTDC 35
and BTDC 40, Cases 2 and 3 had slightly lower ISNOx emissions, as
compared to Case 1. These results are explained by fuel injected early
promptly evaporating and then forming a homogeneous and fairly
lean premixed charge within a combustion chamber during the
ignition delay period. Also, due to the higher latent heat of
evaporation of DME, the charge temperature was relatively low
before the initiation of combustion. For these reasons, early-injected

Fig. 5. Effect of spray angle and injection timing on the combustion performance in
single-injection mode (Pinj = 60 MPa, minj = 10 mg).

Fig. 6. Effect of spray angles and injection timing on the soot and ISNOx emissions in
single-injection mode (Pinj = 60 MPa, minj = 10 mg).

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

DME may contribute to a reduction in NOx emissions, due to the


decrease in peak combustion temperature, as shown in the combustion characteristics in Fig. 4.
In general, DME fuel has relatively faster evaporation and
vaporization and better fuel atomization characteristics than standard
diesel. DME also has higher oxygen content, lower CH and CO ratios,
and a shorter diffusive combustion phase. These physical fuel
properties, spray characteristics, and combustion characteristics
result in a complete combustion process while avoiding unburned
exhaust emissions.
For ISHC and ISCO emissions, Cases 2 and 3 generally have lower
emissions at the same test conditions than Case 1, as shown in Fig. 7.
However, Case 1 also had low levels of ISHC and ISCO emissions at test
ranges from TDC to BTDC 25. At more advanced injection timings
(beyond BTDC 30), the ISHC and ISCO emissions rapidly increase in
Case 1. These results can be attributed to the portion of fuel spray
injected toward the piston head and bowl lip edge, then entering the
squish area of the combustion chamber, in the investigation of the
piston and crank angle geometry during the low-temperature and incylinder pressure condition. Consequently, fuel spray in the squish
region produced larger amounts of unburned emissions. These higher
CO and HC emissions results can be also found in the premixed diethyl
ether (DEE) HCCI-DI diesel engine caused by the premixed fuel charge
trapped in the cylinder crevices and lower combustion temperatures
[19].
The characteristics of particulate size distributions and particle
concentrations for spray angles under different injection timings of
BTDC 5 and BTDC 35 are shown in Fig. 8(a) and (b), respectively. The
dilution rate between exhaust gas and ambient air was maintained at
a 200 to 1 ratio, and the dilution gas temperature was controlled at
120 C to prevent condensation of the sampled raw gas. For improved
accuracy and reliability, experiments were conducted after the test
engine stabilized, and then measurements were taken.
In this gure, results from Cases 2 and 3 indicate similar patterns of
particle distribution without variation in particle concentrations at
both injection timings of BTDC 5 and BTDC 35. In comparison with
the ranges of particle diameter, the nucleation mode particles
composed of volatile condensates (below 40 nm) and larger size
particles (over 200 nm) for Cases 2 and 3 are at slightly higher levels
than in Case 1 at both injection timings. On the contrary, the number

1369

Fig. 8. Effect of spray angles and injection timings on the nano-particle distribution
characteristics in single-injection mode (Pinj = 60 MPa, minj = 10 mg).

of particles ranging from 40 to 200 nm for Case 1 is much higher than


for Cases 2 and 3. The differences in particle concentrations are
increased at BTDC 35. These results correlate with the combustion of
fuel in the squish regions.
3.2. Effect of spray angles and the multiple-injection strategy on
combustion and exhaust emissions characteristics

Fig. 7. Effect of spray angles and injection timing on the ISHC and ISCO emissions in
single-injection mode (Pinj = 60 MPa, minj = 10 mg).

In this study, the pulse signal of single injection is divided into two
pulse signals with short dwell time in order to reduce excessive spray
penetration and interaction with cylinder walls when DME was
injected at early timings. The total injected fuel mass was maintained
constant in single-injection mode (minj = 10 mg/stroke). First, injection mass was xed at 2 mg, and injection timing was varied from
BTDC 35 to BTDC 15 in step with 10 CA. Second, injection mass and
injection timing were held constant at 8 mg and BTDC 5, respectively.
To reduce interference between the rst and second injections, a
minimum dwell time of 10 CA was selected.
The effects of varying the spray angle and injection mode on the
combustion pressures and rate of heat release (t1st = BTDC 35,
t2nd = BTDC 5) are illustrated in Fig. 9. Combustion phases occur
twice due to double-injection, and the peak combustion pressures and
peak ROHR of the multiple-injection strategy for all test injectors are
slightly higher or showed equivalent levels compared with those of
the single combustion cases, as shown in Fig. 4. In comparison of
injection angles, combustion pressure and ROHR upon the rst
injection for Cases 2 and 3 are higher than those for Case 1, leading
to more thorough combustion due to the absence of the wall-wetting

1370

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

Fig. 9. Effect of spray angle and injection timings on the combustion characteristics in
multiple-injection mode (Pinj = 60 MPa, minj = 2 mg (BTDC 35) + 8 mg (BTDC 5)).

problem. Also, the ignition delay between rst and second injections
in Cases 2 and 3 was shorter than that in Case 1. However, upon the
second injection, combustion pressure and ROHR for Cases 2 and 3
were lower and the combustion durations prolonged, as compared to
Case 1. The reason is that the rst combustion actively consumed the
oxygen in the charge, resulting in by-products of combustion,
undiluted charge, and insufcient of oxygen in the combustion
chamber. These factors resulted in the deterioration of the second
combustion process and temperature.
Regarding the rst injection in Case 1, the fuel stream was angled
toward the cylinder wall, and the subsequent over-lean condition
inside the piston bowl led to incomplete combustion. For this reason,
the ignition delay preceding the second combustion was increased,
and combustion occurred as in the single injection with the rapid
premixed combustion phases.
Fig. 10 shows a comparison of IMEP values according to the
injection strategy (single and multiple-injections), the variation in
rst injection timings (BTDC 35, 25, 15), and spray angles. With the
single injection, the values of BTDC 5 for test injectors are chosen for
the maximum value at all test ranges. The line and dot-line indicate
the values of single injection in the results gure.
In this gure, the IMEP values of multiple injections for each test
injector are higher than those of single injections. These higher
combustion performances may be explained if the main combustion
process started at the end of the compression stroke and combustion
was maintained during the early expansion stroke according to the
conversion rate of heat to work. The results of Case 2 are higher than
those of the other cases and show the highest IMEP value during the
retarded rst injection timing of BTDC 15, which has a dwell time of

Fig. 10. Effect of spray angle and rst injection timings on the combustion performance
in multiple-injection mode (Pinj = 60 MPa, minj = 2 mg + 8 mg).

10 CA between the rst and second injections. The use of the narrowangle injectors and multiple injections for DME fuel to avoid spraywall wetting problems at the fairly advanced injection timings
revealed out secure for achieving homogeneous combustion while
maintaining the relatively high combustion performances [20,21].
Fig. 11 shows that the effects of spray angle and injection strategy
on FSN concentration and ISNOx emissions varied according to rst
injection timing. In this gure, it can be seen that ISNOx emissions of
multiple injection for all injectors were gradually reduced according
to the advance of rst injection timings. Soot emissions were similar
for all rst injection timings. In this study, the optimal rst injection
timing for simultaneous reduction in NOx and soot emissions
appeared to be BTDC 35 with a multiple-injection strategy. Multiple
injections cause a dilution effect and lower in-cylinder temperature
during the combustion process, resulting in lower NOx emissions.
Moreover, soot emissions can simultaneously be reduced due to lower
CH ratio and oxygen content in DME combustion.
The effects of spray angle and rst injection strategy on the
concentrations of ISHC and ISCO emissions are shown in Fig. 12. As
rst injection timing advances, the concentrations of ISHC and ISCO
emissions for Case 1 are linearly and simultaneously increased. These
emissions in Cases 2 and 3 are also increased with advances in the rst
injection, but the rates of increase are relatively low. In addition, ISHC
and ISCO emissions for Case 2 indicate the lowest levels at all test
ranges and also indicate relatively low levels in comparison to the
single-injection levels. These higher HC and CO emissions are mainly
associated with the fuel combustion during the compression stroke
after the rst injection, which reduced the oxygen in the charge and
deterioration during the complete combustion phase following
second injection, producing unburned exhaust gas emissions. However, at retarded injection timing, the ISHC and ISCO emissions for all
cases decreased to nearly the same emissions levels as with single
combustion since fuel injection at high ambient pressure and
temperature conditions increases fuel evaporation and enhances
combustion.
The distributions and concentrations of particle size for a multiple
injection strategy and spray angles with rst injection timing of BTDC
35 and second injection timing of BTDC 5 are shown in Fig. 13. In
general, combustion with multiple injection in a CI engine is an
effective method to reduce NOx emissions as the result of low
combustion temperature. Conversely, the formation of particulates
increases because fuel injected earlier occasionally forms diluted
charge regions in the combustion chamber and, thus, results in lower
combustion performance when fuel is injected later. Subsequently,
the injection strategy and DME oxygen content inuence particle
distribution and particle concentration.

Fig. 11. Effect of spray angles and rst injection timings on the soot and ISNOx emissions
in multiple-injection mode (Pinj = 60 MPa, minj = 2 mg + 8 mg).

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

Fig. 12. Effect of spray angles and rst injection timings on the ISHC and ISCO emissions
in multiple-injection mode (Pinj = 60 MPa, minj = 2 mg + 8 mg).

In this gure, in spite of multiple injections, the trends for particle


number concentration in Cases 2 and 3 are the same as those with
single injection; particle size (ranging 10 nm340 nm) is similar
between multiple and single injection, as well. However, particles in
the 30250 nm range for Case 1 are considerably larger, as compared
to Cases 2 and 3 and are also about twice the concentration of those
with single injection. Thus, inefcient combustion may inuence the
volume and weight of smoke emissions.
Fig. 14 shows the relationship between the total particle volume
and the total particle number for different spray angles with single
and multiple injection strategies. With single injection, the total
numbers of particles in Cases 2 and 3 are slightly increased, but the
total volumes of particles rapidly decrease due to the decrease in large
particles and the increase in small particles as injection timing
advances from BTDC 35 to BTDC 5. However, the total number and
total volume of particles for Case 2 increase more considerably with
advancing injection timing.
With multiple-injection, the number of particles larger than
100 nm, which inuences the total weight of particles, is signicantly
higher in Case 1 than in Cases 2 and 3.
4. Conclusions
To analyze the effects of spray angle and injection strategy on
combustion characteristics, performance, and reduction characteris-

1371

Fig. 14. Effect of spray angles on the total particle volume and particle number in
multiple-injection mode (Pinj = 60 MPa, minj = 2 mg + 8 mg).

tics of exhaust emissions, a small, common-rail DI diesel engine fueled


with DME was investigated. The following conclusions are drawn
from the results:
1. At advanced injection timing of BTDC 35, the combustion
characteristics of single combustion for Cases 2 and 3 are increased,
compared with the results of Case 1. In addition, the peak pressure
and ROHR of Case 2 are slightly higher than those of Case 3, with
the same trends for the case of BTDC 5. In particular, considering
the patterns of heat release, two peaks of ROHR are generated by
the combustion of air-fuel premixed mixtures.
2. DME combustion for all test injectors showed low levels of PM
emissions at all injection timings. The NOx emissions simultaneously increased in proportion to the advance in injection timing
up to BTDC 25 for Cases 2 and 3 and BTDC 20 for Case 1. ISHC and
ISCO emissions of Cases 2 and 3 are generally lower than those of
Case 1.
3. In multiple-injection mode, the combustion pressure and ROHR of
the rst injection for Cases 2 and 3 are higher. The ignition delay of
the second injection fuel is shorter in Cases 2 and 3 than in Case 1.
However, the second combustion pressure and ROHR are lower,
and combustion duration was prolonged, as compared to Case 1.
The IMEP values for each test injector are higher than those of
single injection.
4. At advanced rst injection timing, narrow-angle injectors and
multiple injections achieved low NOx emissions and similar levels
of PM emissions, as compared with single-injection cases. Retarded
rst injection timing was needed to reduce the levels of HC and CO
emissions.
5. The particle number of Cases 2 and 3 is similar between multipleand single injection. In addition, Case 2 showed a smaller total
particle volume than Case 2 in spite of higher total particle number.
However, Case 1 had higher total number and total volume of
particles due to the increase in particle size.

Acknowledgement

Fig. 13. Effect of spray angles on the nano-particles distribution characteristics in


multiple-injection mode (Pinj = 60 MPa, minj = 2 mg + 8 mg).

This work was supported, in part, by the Center for Environmentally Friendly Vehicle (CEFV) of the Eco-STAR project of the Ministry of
the Environment (MOE) in Seoul, Republic of Korea, and the Second
Brain Korea 21 Project. This work was also nancially supported by a
manpower development program for Energy & Resources and the
project for development of a clean alternative fuelled power-train
system (10033863-2009-11), supported by the Ministry of Knowledge and Economy (MKE).

1372

S.H. Yoon et al. / Fuel Processing Technology 91 (2010) 13641372

References
[1] S.C. Sorenson, Dimethyl ether in diesel engines: progress and perspectives,
Journal of Engineering for Gas Turbine and Power (2001) 652658.
[2] G. Denis, O. Herwig, S. Daniel, S. Eddie, A.W. Marc, The performance of a heavy
duty diesel engine with a production feasible DME injection system, SAE
Technical Paper Series, 2001, 2001-01-3629.
[3] K. Shuichi, O. Mitsuharu, M. Tomoya, DME fuel blends for low emission directinjection diesel engines, SAE Technical Paper Series, 2000, 2000-01-2004.
[4] H. Teng, J.C. McCandless, J.B. Schneyer, Compression ignition delay (physical
+ chemical) of dimethyl ether an alternative fuel for compressionignition
engines, SAE Technical Paper Series, 2003, 2003-01-0759.
[5] Y. Mingfa, Z. Zunqing, X. Sidu, F. Maoling, Experimental study on the combustion
process of dimethyl ether (DME), SAE Technical Paper Series, 2003, 2003-01-3194.
[6] R. Egnell, Comparison of heat release and NOx formation in a DI diesel engine
running on DME and diesel fuel, SAE Technical Paper Series, 2001, 2001-01-0651.
[7] T.C. Tow, D.A. Pierpont, R.D. Reitz, Reducing particulate and NOx emissions by
using multiple injections in a heavy duty DI diesel engine, SAE Technical Paper
Series, 1994, p. 940897.
[8] M.Y. Kim, S.H. Yoon, K.H. Park, C.S. Lee, Effect of multiple injection strategies on
the emission characteristics of dimethyl ether (DME)-fueled compression ignition
engine, Energy & Fuels 21 (5) (2007) 26732681.
[9] M.Y. Kim, J.W. Kim, J.H. Lee, C.S. Lee, Effect of compression ratio and spray
injection angle on HCCI combustion in a small DI diesel engine, Energy & Fuels 20
(1) (2004) 6976.
[10] S. Naoki, T. Tadashi, N. Terukazu, Dual Mode Combustion Concept with Premixed
Diesel Combustion by Direct Injection Near Top Dead Center, SAE Technical Paper
Series, 2003, 2003-01-0742.
[11] S. Wanhua, L. Tiejian, P. Yiqiang, A Compound Technology for HCCI Combustion in
a Diesel Engine Based on the Multi-pulse Injection and the Bump Combustion
Chamber, SAE Technical Paper Series, 2003, 2003-01-0741.

[12] Z. Longbao, W. Hewu, J. Deming, H. Zuohua, Study of Performance and


Combustion Characteristics of a DME-Fueled Light-duty Direct-Injection Diesel
Engine, SAE Technical Paper Series, 1999, 1999-01-3669.
[13] S.C. Sorenson, M. Glensvig, D. Abata, Di-methyl Ether in Diesel Fuel Injection
Systems, SAE Technical Paper Series, 1998, p. 981159.
[14] B. Walter, B. Gatellier, Near zero NOx emissions and high fuel efciency diesel
engine: the NADI concept using dual mode combustion, oil & gas science and
technology-Rev, IFP 58 (1) (2003) 101114.
[15] D.S. Kim, M.Y. Kim, C.S. Lee, Reduction of nitiric oxides and soot by premixed fuel
in partial HCCI engine, Journal of Engineering for Gas Turbines and Power 128
(2006) 497505.
[16] D.S. Kim, M.Y. Kim, C.S. Lee, Combustion and emission characteristics of a partial
homogeneous charge compression ignition engine when using two-stage
injection, Combustion Science and Technology 179 (2007) 531551.
[17] G.D. Neely, S. Sasaki, J.A. Leet, Experimental Investigation of PCCI-DI Combustion
of Emissions in a Light-duty Diesel Engine, SAE Technical Paper Series, 2004,
2004-01-0121.
[18] W. Ying, H. Li, Z. Jie, Z. Longbao, Study of HCCI-DI combustion and emissions in a
DME engine, Fuel 88 (2009) 22552261.
[19] C. Cinar, O. Can, F. Sahin, H.S. Yucesu, Effects of premixed diethyl ether (DEE) on
combustion and exhaust emissions in a HCCI-DI diesel engine, Applied Thermal
Engineering 30 (2010) 360365.
[20] G.A. Lechner, T.J. Jacobs, C.A. Chryssakis, D.N. Assanis, R.M. Siewert, Evaluation of a
Narrow Spray Cone Angle, Advanced Injection Timing Strategy to Partially
Premixed Compression Ignition Combustion in a Diesel Engine, SAE Technical
Paper Series, 2005, 2005-01-0167.
[21] S. Kook, C. Bas, Combustion Control Using Two-stage Diesel Fuel Injection in a
Single-Cylinder PCCI engine, SAE Technical Paper Series, 2004, 2004-01-0938.

You might also like