Sie sind auf Seite 1von 12

Energy 52 (2013) 143e154

Contents lists available at SciVerse ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

PFI (port fuel injection) of n-butanol and direct injection of biodiesel


to attain LTC (low-temperature combustion) for low-emissions
idling in a compression engine
Valentin Soloiu*, Marvin Duggan, Spencer Harp, Brian Vlcek, David Williams
Department of Mechanical Engineering, Georgia Southern University, Statesboro, 30460 GA, United States

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 30 June 2012
Received in revised form
10 January 2013
Accepted 11 January 2013
Available online 19 February 2013

In this study, n-butanol (port fuel injection) PFI was investigated in a direct injection compression
ignition engine while at idling speeds, and loads, 1e3 bar IMEP (indicated mean effective pressure) in
order to determine the effects on combustion, efciency, emissions, and specically, a modied tradeoff
of soot and nitrogen oxides. As a result, the engine entered into (low-temperature combustion) LTC
regions, for selected loads and speeds. Compared with the baseline taken with ultra-low sulfur diesel no.
2, the heat release with n-butanol in (premixed charge compression ignition) PCCI mode, has resulted in
a 75% reduction from the maximum values, while a secondary peak appeared where the diffusion
combustion typically occurs in the power stroke. At 3 bar IMEP an early, (bottom dead center) BTDC lowtemperature heat release was found that began 6 earlier than for the diesel reference cycle, and corresponding to 1200 K. Soot emissions showed a massive decrease of about 98%, concurrently with a 74%
reduction of nitrogen oxides at 3 IMEP by controlling the combustion phases and by modifying the
classical NOxesoot tradeoff. The results of this work prove that biodiesel combined with n-butanol PFI in
PCCI and LTC are very effective in simultaneously reducing soot and NOx at idling speeds.
2013 Elsevier Ltd. All rights reserved.

Keywords:
n-Butanol combustion
LTC (low-temperature combustion)
PCCI (premixed charge compression
ignition)
PFI (port fuel injection)

1. Introduction
Over their lifetime, the transport vehicles such as trucks, locomotives, buses, accumulate millions of hours of idling [1]. The engine idling is necessary for air conditioning, keeping the cabin
warm or cool, and loading and unloading the vehicle. Supplementary, many US states mandate how long a vehicle can idle, in an
effort to decrease the hazardous emissions associated with the
process. The EPA (U.S. Environmental Protection Agency) also
suggests idling emissions reduction technologies and tax exemptions for using them [1]. The long hours associated with idling can
have a signicant impact on environment and the purpose of this
research is to the use of biofuels (n-butanol and biodiesel) in the
LTC (low-temperature combustion) regime that was presumed to
have a signicant impact in reducing the soot and NOx emissions.
The (Energy Independence and Security Act) EISA 2007 mandates
that 36 billion gallons of renewable fuel is to be blended into US
transportation fuels by 2022 [2], nevertheless the performance of
biofuels in low-emissions, low-temperature combustion engines
remains under much investigation. In the United States fuel

* Corresponding author.
E-mail address: vsoloiu@georgiasouthern.edu (V. Soloiu).
0360-5442/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.energy.2013.01.023

consumption is presently dominated by gasoline, and the


(Department of Energy) DOE, (U.S. Energy Information Association)
EIA projects that in response to stringent (Corporate Average Fuel
Economy) CAFE standards, diesel and diesel-compatible biofuel
consumption will increase through the year 2035. To increase
marketability of clean diesel engines, substantial work has been
directed at developing technologies that reduce the costs related to
their emissions treatment systems (catalytic converters, retro tting, etc.). A promising way to achieve that is to mitigate the
emissions by controlling their formation during combustion in the
engine cylinder. One of the most capable of the new technologies is
the LTC in which the combustion is controlled to occur in predened relative airefuel ratioetemperature zones that simultaneously limit the formation of NOx, particulate matter, and soot, as
seen in Fig. 1. In general, such research has been focused on fossil
fuels and their particular vaporization and kinetic properties [3e8].
Effective means of achieving LTC usually includes EGR (exhaust gas
recirculation), but at idling speeds and low loads, port fuel injection
with no EGR can be an effective ways of achieving LTC. The biodiesel used in this study was obtained from the (US National Peanuts Research Laboratory) NPRL and produced from waste peanuts
of insufcient quality to enter the food market, making them a nonfood source. Bioethanol was considered originally for this study as
an ethanolediesel mixture due to widespread availability as an

144

V. Soloiu et al. / Energy 52 (2013) 143e154

Fig. 1. PCCI soot and NOx.

alternative fuel. However, the use of bioethanol presented limitations [9,10] since it has low boiling point, causing increased evaporation even at moderate temperatures, it has a lower energy
density compared to n-butanol, low miscibility with petroleum
which may cause phase segregation, and it is highly hygroscopic,
which leads to lower lubricity and early wear combined with corrosion of the injection system [19]. A comparison of ethanol/nbutanol fuel properties can be seen in Table 1.
n-Butanol is a high-energy content alternative fuel [11], similar
to ethanol in molecular structure [12], which can be produced from
agricultural waste (non-food sources). n-Butanol (bio-butanol
when produced biologically), has a 4 carbon atoms chain in the
structure and the molecular formula C4H9OH [19]. There is abundant plant biomass available as low-value agricultural commodities
or processing wastes requiring proper disposal to avoid pollution
problems [13] that can otherwise be used to produce n-butanol. Its
LHV (lower heating values) is higher than that of ethanol, has lower
water absorption, has better blending ability, is less corrosive to
pipelines, and can be used in conventional combustion engines
without modication [14]. Butanol/diesel blends as an alternative
for petroleum diesel have been previously studied [15e17], and the
results indicate that it may be a better alternative than ethanol. The
results of this work also indicate that butanol can be an effective
alternative biofuel.
2. Present state of knowledge and literature review
An Omnivorous engine (multifuel engine) employing new combustion modes e PCCI (premixed charge compression ignition) and
LTC e has been instrumented at Georgia Southern University. The
engine works by PFI of alcohols (n-butanol) while being fueled with
diesel/biodiesel by direct injection. The focus of PCCI and LTC with
biofuels is on local mixture preparation and combustion temperature control by balancing the physical timescale of spray development and local mixture formation with the ignition chemistry
Table 1
n-Butanol and ethanol fuel properties.
Property

n-Butanol

Ethanol

CN
Density
Viscosity at 40  C
LHV
Mw
Saturation degree
Flash Pt.
Cloud Pt.

24.80
0.79 g/cm3
1.8 mm2/s
32 MJ/kg
74.1 g
e
37  C
e

54
0.0066 g/cm3
27 MJ/kg
46.07 g
e
13  C
e

timescale [18]. Diesel with a high cetane number has chainbranching reactions produces a large number of (hydroxyls) OH
and other active radicals, triggering a cool-ame reaction at the end
of the compression stroke [19]. n-Butanol has the potential to form
homogenous mixtures leading to PCCI since it has a much lower
boiling point and distillation temperature, compared to biodiesel. In
the zones with richer fueleair mixtures, the exothermic reaction and
the large amounts of (hydrogen peroxide) H2O2 that accumulate may
lead to early ignition. A secondary fuel e alcohol (n-butanol) e has
been used to change the reaction pathways in order to increase the
mixtures ignition delay and favorably change the combustion
phasing by reducing the OH formation rates and increasing the
OH consumption rates. The premixed combustion and hot
ignition would be delayed due to the retardation of the cool-ame
period. As a reaction suppressor, the alcohol reduces the rate of
low-temperature oxidation (BTDC (before top dead centre)) and,
consequently, delays the onset of high-temperature oxidation, and it
may eliminate the local non-homogenous rich zones resulting in
reduction of the maximum combustion temperature and, subsequently, the NOx and PM (particulate matter) emissions. This
research suggests that PM and NOx formation can be drastically
reduced by controlling in-cylinder combustion temperature at idle
and low loads, while maintaining the maximum combustion temperature between 1500 K and 1800 K, to ensure reduced formation of
PM and NOx, and adequate oxidation of (carbon monoxide) CO and
unburned (hydrocarbons) HC while maintaining engine efciency.
This technically challenging investigation has not often been
attempted, and much remains unknown [20]. It has the potential to
advance knowledge within and across elds by investigating biofuels
for LTC in a diesel engine. Most of the previous work has focused on
low-temperature kinetics that controls the start of ignition in (homogeneous charge compression ignition) HCCI in diesel engines and
knock in spark-ignition engines. The intermediate temperature
ignition responsible for these phenomena begins at about 1000 K,
when H2O2 decomposes into two OH radicals, followed rapidly by
a transition to high-temperature chain-branching [21]. Nevertheless,
the kinetic processes distinctive to the low-temperature regime,
from 650 K to 800 K, release sufcient heat to shorten the time at
which the mixture reaches the ignition temperature of about 1000 K.
Theoretical, kinetic modeling and experimental studies were conducted [22e25] that placed these processes on a solid theoretical
basis. In red engine research [26,27], studies have revealed how
these low-temperature reactions would accelerate ignition in diesel
and HCCI engines. Much remains to be investigated about the impact
of higher pressures and binary mixtures, from biofuels, on many of
these processes.
Low-temperature combustion is expected to benet overall engine efciency, primarily because of reduced cylinder heat loss and
the potential of molecular properties of the expanding combustion
gases from dilute combustion to allow more of the energy to be
extracted in the expansion stroke. McCormick et al. (2002) developed a study to reduce soot in a diesel engine. It was determined
that blending very high percentages of FishereTropsch diesel can
produce a NOx neutral fuel [28]. Van Gerpen et al. (2001) developed
a study on biodiesel oxidation on engine performance and emissions characteristics. The results of the study demonstrated that
there was no major difference in thermal efciency between diesel
and biodiesel [29]. Wallner et al. (2009) compared two fuel blends,
10% by volume ethanol and 10% by volume 1-butanol. The study
suggested that 10% by volume 1-butanol can be used as an effective
oxygenate, with an improvement in fuel economy. The study also
showed the 1-butanol had the lowest nitrogen oxide emissions [30].
Lujaji et al. (2010) developed a study evaluating the effects of blends
containing biodiesel, butanol, and diesel on engine performance,
combustion, and emission characteristics. The results of the study

V. Soloiu et al. / Energy 52 (2013) 143e154

show that fuel properties of vegetable oils are improved by the


blending of vegetable oil, butanol, and diesel fuel [31]. Rakopoulos
et al. (2010) developed a study in which the effects on combustion of
various volumes of ethanol or n-butanol, blended with diesel were
evaluated. In terms of heat release the study showed that with the
use of biofuels, there was reduction in smoke opacity and nitrogen
oxides [32]. Kim et al. (2004) studied the effect of premixed gasoline
fuel on the combustion characteristics of compression ignition engine. The results showed that a partial HCCI engine with a premixed
gasoline fuel shows single-stage ignition and the increase in premixed ratio of gasoline fuel results in the advance of ignition timing
and the remarkable reduction of nitrogen oxide emissions [33]. Lu
et al. (2011) studied the effect of multiple injection strategies in
diesel PCCI combustion. The study showed that at low engine loads,
the combination of PCCI and EGR can simultaneously reduce soot,
CO, HC, and NOx emissions [34]. Park et al. (2011) investigated the
effects of EGR and pilot injection on PCCI combustion in a diesel
engine. The results of his work showed an optimum injection timing
of 25 BTDC (before top dead center) for NOx [35]. Neely et al. (2004)
investigated PCCI-DI (direct injection) combustion on emissions in
a light-duty diesel engine. The study suggested that early single
pilot injections resulted in moderate NOx reductions with increases
in CO and fuel consumption, while multiple injections were effective for a simultaneous reduction in CO, BSFC (break specic fuel
consumption), and NOx emissions [36]. Karra et al. (2008) studied
the characteristics of engine emissions using biodiesel blends in LTC
regimes. The results showed that biodiesel blends generally produced lower soot emissions and higher NOx emissions compared to
regular diesel fuel [37]. Altun et al. (2011) studied the effect of using
n-butanol in conventional diesel fuelebiodiesel blends on engine
performance and exhaust emissions. For exhaust emissions, carbon
monoxide and hydrocarbons emissions decreased, while NOx
remained unchanged at low engine loads and decreased at high
engine loads [38].
3. Thermo-physical properties of selected fuels
3.1. Fuel characteristics
Ultra-low sulfur diesel no. 2 was used for the baseline cycles for
combustion and emissions with the biodiesel/n-butanol in PCCI
regimes. The biodiesel (B100) used in this experiment was derived
from peanuts and provided by the NPRL; biodiesel is of tremendous
interest because it is simple to produce and there is abundant
biomass waste for production [39]. The EIA reported that by the
year 2035 the need for energy imports must be offset by the
increased use of domestically produced biofuels, and demand reductions of said imports with new vehicle fuel economy standards,
and growing energy prices [2]. The n-butanol has been chosen
because of its high-energy content, low evaporative pressure and
low (cetane number) CN. Also n-butanols high heat of vaporization
may be benecial in achieving LTC because more heat is expected
be absorbed during mixing, which would delay the ignition [40],
and result in cool ames at low temperatures. The n-butanol was
introduced through PFI in the intake manifold for precise control of
the amount injected per cycle and to avoid seizing the plunger in
the barrel of the injection pump because of reduced lubrication. The
characteristics of the fuels used in the experiment can be seen in
Table 2.
3.2. Cetane number
The CN for the peanut fatty acid methyl ester can be determined
with good accuracy by using a regression model based on fatty acid
composition percentage, proposed by Bamgboye and Hansen [41]

145

Table 2
Properties of selected fuels.
Property

USLD#2

Peanut FAME

n-Butanol

CN
Density
Viscosity at 40  C
LHV
Mw
Saturation degree
Flash Pt.
Cloud Pt.

47
0.85 g/cm3
2.32 mm2/s
42.6 MJ/kg
233 g
NA
100  C min
16.1  C

53
0.87 g/cm3
5.2 mm2/s
36 MJ/kg
w292 g
93
176  C
17  C

24.80
0.79 g/cm3
1.8 mm2/s
32 MJ/kg
74.1 g
e
37  C
e

and calibrated in Ref. [19]. From these values, the cetane numbers
for different blends were calculated based on the percentage of
each fuel type based on Table 3. The regression equation used
appeared as follows:

CN 61:1x1 0:088x2 0:133x3 0:152x4  0:101x5


 0:039x6  0:243x7  0:395x8

(1)

The variables x1, x2, x3, x4, x5, x6, x7, and x8 represent the percentage
of fatty acid methyl esters, and the coefcients for the saturated
FAMEs (fatty acid methyl ester) are positive and increase with an
increase in the carbon number [19] The coefcients of the unsaturated FAMEs are due to the reduction in the overall cetane
number associated with unsaturation [42]. Based on this model, the
cetane numbers for 100% peanut FAME obtained were approximately 54. Fig. 2 displays the results of the cetane numbers
obtained from the model above [49,50].
3.3. Dynamic viscosity
Biodiesels viscosity affects spray atomization, ignition delay,
combustion phasing, emissions, and also engine component wear.
At lower ambient temperatures the increase in viscosity leads to
poor spray atomization and high emissions.
From previous studies [43,44], it was determined that the viscosity of biodiesel is higher than that of diesel but of less magnitude
than the original crude bio-oil used for production. For all the binary ULSD#2-FAME mixtures developed, (0e100% peanut FAME),
the viscosity was investigated at various temperatures to determine its changes with respect to temperature and percentage of
FAME. Test comparisons were done with a Brookeld Viscometer
DV II Pro type, tted with the Small Sample Adapter attachment,
and the viscosity tests were operated from 26 to 60  C at a spindle
speed of 200 rpm. The results indicated that the viscosity of the test
fuels and ULSD blends is higher than that of pure diesel and is
heavily dependent upon the biodiesel content. The data collected,
displayed in Fig. 3, showed that at 40  C, the viscosity of ULSD#2 up
to B100 ranged from 2.3 to 5.2 cP.

Table 3
Peanut biodiesel fatty acid composition.
Fatty acid

Carbon atoms

(%)

Palmitic
Stearic
Oleic
Linoleic
Linolenic
Arachidic
Eicosenoic
Behenic
Nervonic
Total

16
18
18
18
18
20
20
22
24
174

8.78
2.95
58.74
21.84
0.23
1.4
1.82
3.22
1.02
100

146

V. Soloiu et al. / Energy 52 (2013) 143e154

70

60
40

80

55
50
45

20
60
0
40

DTA uV/mg

TGA [ % mass]

60

Cetane Number [CN]

Diesel
B100
Butanol

100

Peanut FAME

65

-20
20
-40

40

0
35

100

30
0

20

40

60

80

200

300

400

500

-60
600

Temperature [C]

100

Fig. 4. Diesel, biodiesel, n-butanol TGADTA.

Percentage of FAME in ULSD[%]


Fig. 2. Cetane number peanut FAME.

The results show that the viscosity increases with higher content of FAME in the mixtures, but as the temperature increases, the
viscosity decreases sharply. These experiments determined that the
Peanuts-FAME/ULSD#2 mixtures were within the ASTM viscosity
standard when using up to 50% FAME.

n-butanol has a strong endothermic reaction at 90  C, where it


vaporizes, and begins oxidization almost immediately at 115  C. The
vaporization speeds are presented in Fig. 5, with n-butanol and
biodiesel having twice the vaporization speed as diesel.

4. Experimental method
4.1. Experimental engine setup

3.4. Thermogravimetric and differential thermal analysis


Thermal analysis studies were performed in a controlled environment using an air purged atmosphere from room temperature
up to 600  C. From results in Fig. 4, it can be seen that ULSD#2 will
vaporize about 90% of its original mass by 200  C, with an elevated
exothermic reaction around 250  C as seen in Fig. 4. The biodiesel
was stable up to 175  C, proving that it required a higher temperature to begin the vaporization. From 275 to 320  C, the oxidation
took place, which reaches a peak at 300  C, seen in Fig. 4. The

Studies were performed with an experimental 4 stroke, direct


injection, single cylinder, compressioneignition engine with an
omega combustion chamber in the piston. The engine was liquid
cooled and naturally aspirated, with two valves per cylinder. The
experiments were performed without the use of any EGR. The engine parameters, given in Table 4, were used in the thermodynamic
calculations presented throughout this paper and the experimental

0.2

B20
B35
B50
B100
Diesel

Dynamic Viscosity (cP)

Rate of Vaporization [mg/sec]

10

Diesel
B100
n-Butanol

0.15

0.1

0.05

0
25

30

35

40
45
50
Temperature [C]

55

Fig. 3. Dynamic viscosity of diesel, biodiesel, and biodiesel blends.

60

100

200

300

400

500

Temperature [C]
Fig. 5. Diesel, biodiesel, and n-butanol rate of vaporization.

600

V. Soloiu et al. / Energy 52 (2013) 143e154


Table 4
Engine parameters.
Maximum power
Maximum torque
Bore  stroke
Displacement
Injection nozzle
Injection pressure
Injection timing
Compression ratio

16.9 kW
77.47 N m
112 mm  115 mm
1132 cc
4 orices  0.300 mm
200 bar
20BTDC
16:1

setup is presented in Fig. 6. The engine was converted to a PCCI


system by introducing an injector, for n-butanol, into the intake
manifold for which ow rate was controlled using an Electronic
Control Unit with the mass ow rate measured using the AVL KMA
4000 ex fuel consumption measurement system. n-Butanol was
injected at two xed ow rates named: (B100B) and (B100C),
respectively. LTC was achieved through PCCI by advancing fuel injection timing in the intake manifold, which led to longer premixing
time resulting in a lean mixture and lower ame temperatures.
Particulate emissions were measured from the exhaust line using an
AVL 483 Micro Soot Sensor at a dilution ratio of 1:9. Gaseous
emissions were measured using an AVL SESAM FTIR V4 (30 species).
Air ow into the engine was measured using a computerized Meriam Laminar Flow Meter, in-cylinder pressure was measured using
a Kistler 6053 cc uncooled piezoelectric pressure sensor, and the
injection pressure was measured using a Kistler 6229 in-line
injector pressure sensor with Kistler 6533A11 clamp adapter. An
Omron optical encoder was used to measure the (crank angle degree) CAD for data acquisition at 50 kHz, that is at every 0.1 CAD at
800 rpm representing 0.02 ms (millisecond), by using an Yokogawa
750 DL high speed data acquisition system. Experiments were carried out at the engine speed of 800 rpm, representing low speeds
and low to medium loads for the experimental engine. The engine
was allowed to run until exhaust gas temperature, cooling water

147

temperature, and lubrication oil temperature became steady. Gaseous emissions were measured continuously at a rate of 1 Hz for
each respective load and fuel combination. The fuels properties used
during experimentation can be seen in Table 1.
4.2. Combustion investigations of biodiesel with PFI n-butanol in LTC
The engine investigations were performed at loads of 1e3 bar
IMEP that represent partial to mid load of the engine at 800 rpm;
higher loads are typically not applicable at idling speeds. In
Figs. 7 and 8, the inuence of the n-butanol PCCI on the maximum
cylinder pressure is shown for 1 bar and 3 bar IMEP. It can be seen
that the PFI of n-butanol caused the cycles maximum pressure to
decrease by 20%, at and the peak pressure was delayed by 2 CAD
(crank angle degree) at 1 IMEP, and as much as 7 CAD for 3 IMEP
compared with ULSD#2. Switching to B100 did not have any impact
on injection timing; the injection timing at which the injector needle
was open is presented in Fig. 9. It can be seen that for both the
ULSD#2 and B100, injector was open at 20 CAD BTDC and closed at
12 CAD, BTDC for 3 bar IMEP.
4.3. Heat release
Using the rst law of thermodynamics and the ideal gas law
with equivalent substitutions and assumptions for a closed system,
the rate of heat release was determined using equation (2). The
following three assumptions were used in this determination: the
mixture inside the combustion chamber behaves as an ideal gas,
the gaseous contents within the combustion chamber are homogeneous, and the system shall be considered a closed system with
constant mass throughout the cycle (neglecting varying mass due
to crevice ow, blow-by, and fuel injection).

g
dQ
1
dP
dv

V
P
g  1 dq g  1 dq
dq

Fig. 6. Experimental setup.

(2)

148

V. Soloiu et al. / Energy 52 (2013) 143e154

80

80
Diesel
B100
B100 B
B100C
Motored

B100

Cylinder Pressure [bar]

60

Diesel

70

50
40
30

60

1.5

Motored

50
40

30
20

0.5

Diesel

20

Needle Lift [mm]

Cylinder Pressure [bar]

70

B100

10

10
0
340

0
330

345 350 355 360 365 370


Crank Angle [CAD]

375

380

340

350

360

0
380

370

Crank Angle [CAD]


Fig. 9. Cylinder pressure at 800 rpm 3 bar IMEP.

Fig. 7. Cylinder pressure at 800 rpm 1 bar IMEP.

The apparent heat release for each testing point is shown in


Figs. 10 and 11 with ULSD#2, B100, B100B, and B100C at 1 and
3 bar IMEP. The net heat-release proles for the n-butanol PCCI
show a combustion mode comprising of a diminished premixed
combustion phase, and a delayed combustion phase similar to
the diffusion ame period, especially at the lower loads. The
low-temperature heat-release occurred at 16 CAD BTDC, directly
after the start of injection of biodiesel at 20 CAD BTDC, seen in
Fig. 11. The ignition delay at 3 IMEP was 9 CAD for ULSD#2, 7 CAD
for B100, 6 CAD for B100B, and 3 CAD for B100C. The ignition delay
timing ranged from 1.9 ms for ULSD#2 to 0.6 ms for B100C that
contained high volumes of n-butanol. The net heat-release proles
for diesel and B100 show the conventional diesel combustion
consisting of a premixed combustion phase and a diffusion

controlled phase. The ignition delay and rate of vaporization


determined the premixed burning phase, while the rate of air/fuel
mixing determines the diffusion combustion phase. In the hightemperature combustion phase, the ames tend to initialize
and propagate to regions with air/fuel ratios that are nearstoichiometric, resulting in the diesel combustion typical soot and
NOx tradeoff. However, the net heat-release rate associated with
the n-butanol in PCCI, indicated signicant differences as seen in
Figs. 10 and 11. The B100C at 3 bar IMEP showed a peak in heat
release of 21 J/deg at 10 BTDC as a result of low-temperature
oxidation reactions at 1200 K and a second peak during the conventional diffusion ame phase, of 40 J/deg late in the power stroke
at 5 ATDC (after top dead center) resulting from the high-temperature heat release.

100
80
Diesel
B100
B100 B
B100 C
Motored

60

Diesel
B100
B100 B
B100C

80

Heat Release Rate [J/deg.]

Cylinder Pressure [bar]

70

50
40
30
20

60
40
20
0

10
0
340

345

350

355

360

365

370

Crank Angle [CAD]


Fig. 8. Cylinder pressure at 800 rpm 3 bar IMEP.

375

380

-20
340

345 350

355

360

365 370

Crank Angle [CAD]


Fig. 10. Heat-release rate at 800 rpm 1 bar IMEP.

375

380

V. Soloiu et al. / Energy 52 (2013) 143e154

100

100
C90

Diesel
B100
B100 B
B100C

80

Mass Burned [%]

80

Heat Release Rate [J/deg.]

149

60
40
20

60
C50
40
Diesel
B100
B100 B
B100 C

20

C10
-20
340

345 350

355

360

365 370

375

0
340

380

Crank Angle [CAD]

350 360

370 380 390 400


Crank Angle [CAD]

410

420

Fig. 13. Mass burned 800 rpm 3 bar IMEP.

Fig. 11. Heat-release rate 800 rpm 3 bar IMEP.

4.4. Mass burned

4.5. Instantaneous volume-averaged cylinder temperature

The mass percent of the ULSD#2, B100, B100B, and B100C for 1
and 3 bar IMEP is shown in Figs. 12 and 13, and it was determined
by integrating the gross heat release in the cylinder during combustion. The diesel and B100 show a trend in which the mass of fuel
burned increases quickly 15 BTDC in the compression stroke, and
after TDC (top dead center), during the diffusion ame phase, the
rate of mass burned increases slowly until this phase is over; that
holds true for each load (1e3 bar IMEP). The B100B and B100C
show a relatively constant rate of mass burned throughout the
entire combustion phase especially at lower loads; however as the
load is increased the B100B (lower amounts of n-butanol per cycle)
began to follow a similar trend to the ULSD#2 and biodiesel as
displayed Fig. 13.

At low loads the maximum cylinder temperature decreased


from 1500 K with conventional combustion, as low as 1250 K, with
the PCCI/LTC, shown in Fig. 14. The decrease in temperature is
a result of the alcohol injection; a signicant amount of the heat
was absorbed in the process of vaporizing n-butanol within the
cylinder, while the main combustion phase takes place later in the
power stroke. There was no signicant difference in maximum
cylinder temperature between the fuels at 3 bar IMEP, as a result of
the increase in load, Fig. 15.
4.6. Heat uxes
In order to obtain the heat transfer within the cylinder during
combustion, the heat uxes had to be calculated, and several

100

2000

C90
Cylinder Gas Temperature [K]

Mass Burned [%]

80

60
C50
40
Diesel
B100
B100 B
B100 C

20
C10
0
340

350 360

370

380

390 400

Crank Angle [CAD]


Fig. 12. Mass burned 800 rpm 1 bar IMEP.

410

420

1800

Diesel
B100
B100 B
B100C

1600
1400
1200
1000
800
340

345 350 355 360 365 370


Crank Angle [CAD]

375

Fig. 14. Cylinder gas temperature at 800 rpm 1 bar IMEP.

380

150

V. Soloiu et al. / Energy 52 (2013) 143e154

2
Diesel
B100
B100 B
B100 C

1800

1.5

Heat Flux [MW/m]

Cylinder Gas Temperature [K]

2000

1600
1400
Diesel
B100
B100 B
B100C

1200

Total Heat Flux


1

0.5

Convection

1000
800
340

345 350

355

360

365 370

375

0
320

380

Crank Angle [CAD]


Fig. 15. Cylinder gas temperature at 800 rpm 3 bar IMEP.

S,N,D
30,ma

(3)

In which the air viscosity is calculated with the following


equation:



1273:15 110:4
TA a 1:5
 105
,
TA a 110:4
1273:5

ma 4:94,

Radiation
360
380
400
Crank Angle [CAD]

term displayed diminishing values by 50e70% for the fuels with


alcohol PFI (B100B at and B100C-highest rate of n-butanol), seen in
Fig. 16, which correlates very well with the very low measured soot
emissions, as seen in Figs. 19 and 22.
4.7. Heat transfer
Based upon the heat uxes, the heat transfer throughout the
cycle was determined, taking into account the area of the combustion chamber exposed to the gases (piston omega combustion
chamber piston crown cylinder head the lateral (variable)
area of the cylinder) and the results are presented in Fig. 17 for B100
and Fig. 18 for B100C at 3 bar IMEP. The innermost curves for each
fuel represent the (apparent (net/apparent) heat-release rate)

(4)

The heat uxes for each fuel and fuel mixture were obtained
using the Annand model [45] further developed by Soloiu [46], and
were calculated using the instantaneous volume-averaged gas
properties every 0.02 ms (millisecond) throughout the cycle, as
presented in equation (5), where the rst term denotes the convection heat ux, and the second term denotes the radiation heat
ux during combustion. It was found in Ref. [47] that the optimum
results are obtained with an average piston speed compared with
instantaneous piston speed.

_ a A
q



4
Rea0:7 TA a  TW s, TA4 a  Tw

lA a
D

(5)

The air thermal conductivity coefcient is displayed in the following formula:

lA a 1:2775,108 ,TA a 7:66696,105 ,TA a


0:00444888

420

Fig. 16. Heat ux 800 rpm 1 bar IMEP.

parameters were necessary to determine these, including the


instantaneous volume-averaged in-cylinder Reynolds number
which has been calculated as presented in equation (3), the gas
viscosity by equation (4), gas density and convection heat transfer
coefcient, obtained with equation (5). The model for obtaining the
Re number gives relatively accurate results and displays a very
distinct difference between Re number at TDC in exhaust/intake
stroke of about 10,000 versus 100,000 at TDC in combustion. On the
other hand, the calculated Reynolds number showed essentially the
same values and trends for all fuels throughout the cycle.

Rea ra

340

(6)

The maximum convection and radiation uxes for the ULSD#2


and B100 were practically equal for each load, but the radiation

Fig. 17. Heat transfer at 800 rpm 3 bar IMEP, B100.

V. Soloiu et al. / Energy 52 (2013) 143e154

151

20

NOx [g/kWh]

15
Diesel
B100
B100 B
B100 C

10

1.5

2.5

IMEP [bar]

Fig. 18. Heat transfer at 800 rpm 3 bar IMEP, B100C.

Fig. 20. NOx emissions at 800 rpm.

AHRR. The area between the inner curve and the middle curve (blue
area in web version) represents the loss in energy due to the convective heat transfer and is displayed at each CAD. The area between the outermost curve and the middle curve (red area)
represents the radiated heat transfer while the outermost curve
(envelope) represents the gross heat released by the fuels combustion. While the biodiesel (B100) displayed almost no heat losses
for the fast heat release in the premixed charge combustion phase
between 350 and 355 CAD, the heat losses in convection (blue area
in web version) and radiation (red area) are visible in the power
stroke, Fig. 17. Conversely, the n-butanol mixture (B100C) displays
much lower radiation, and that correlates exceptionally well with
the experimental measurement results of almost no soot formation
presented in Fig. 19 (soot emissions), Fig. 21 (sooteNOx strategy for

all fuels at one load-2 bar IMEP), and Fig. 22 (sooteNOx tradeoff at
various loads).
5. Emissions characteristics of n-butanol and biodiesel in
PCCI & LTC
5.1. Nitrogen oxides and soot
Soot and NOx are produced during the premixed phase of
combustion at the boundary of the fuel spray combined with
high-temperature lean regions. In conventional diesel combustion,
in general, there exists a tradeoff in which a solution in decreasing
one causes an increase in the other. In order to change this tradeoff
favorably and reduce both emissions concomitantly, lean premixed

20

0.1

B100
15

NOx [g/kWh]

0.08

Soot Emissions [g/kWh]

Diesel

Diesel
B100
B100 B
B100 C

0.06

0.04

2 bar IMEP

10

B100 B

5
0.02

B100 C
0

0
1

1.5

2
IMEP [bar]

Fig. 19. Soot emissions at 800 rpm.

2.5

0.01

0.02
0.03
Soot [g/kWh]

0.04

Fig. 21. SooteNOx tradeoff at 2 bar IMEP.

0.05

152

V. Soloiu et al. / Energy 52 (2013) 143e154


Table 5
Unburned hydrocarbons emissions for each load in g/kWh.

20

NOx [g/kWh]

15

10

Diesel
B100
B100 B
B100 C

0
0

0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08


Soot [g/kWh]
Fig. 22. SooteNOx tradeoff at various loads.

(low-temperature combustion) LTC regions have been developed at


limited speed and load operations presented in this study.
At lower temperatures (less than 1600 K), soot formation is
reduced as temperatures are too low for fuel pyrolysis. Soot emissions showed signicant decreases as much as 98% at 3 bar IMEP
compared with ULSD#2, as seen in Fig. 19. The reduction in radiation results from the low soot production associated with the
n-butanol PCCI, and the biofuels oxygenated properties also contribute to this reduction. Khan et al. (1973) suggest that in order to
reduce net soot release, one must decrease its rate of formation by
decreasing the proportion of the diffusion combustion phase, and
a direct relationship exists between soot and combustion duration
resulting in the conventional sooteNOx tradeoff [48]. Lyns paper

10

Diesel
B100
B100 B
B100 C

CO Emissions [g/kWh]

10

Load [IMEP]

Diesel

B100

B100B

B100C

1
2
3

15.12
5.82
6.09

6.29
5.87
4.77

162.62
27.20
12.48

274.43
127.85
25.19

on diesel combustion [47], suggests that a shorter ignition delay


yields a higher amount of diffusion burning because there is less
time applicable for mixing prior to ignition. Since diffusion combustion phase has a slower burning rate than the premixed combustion phase, combustion duration increases as the diffusion
combustion phase increases; this was also found in this research.
Therefore, the soot should increase when the inuence of diffusion
burn to the total heat release would increase soot [19]. In this study
the soot has been reduced by 98% while the NOx by 75% with B100C
at 3 bar IMEP as seen in Figs. 19 and 20. At the same time the diffusion combustion was present only at very low loads and has
transformed in a second stage of high-temperature heat release by
adding the n-butanol, as seen in Figs. 18 and 19 and therefore
altering the conventional sooteNOx tradeoff favorably.
The impact of the fueling strategy on the sooteNOx tradeoff is
presented in Fig. 22. The study started with diesel, moved to biodiesel and introduced gradually a higher content of n-butanol. It
can be easily seen the difference from ULSD#2 to B100 is marginal
but the impact of n-butanol in the fuels B100B and B100C is very
signicant with substantial reduction of both pollutants.
If all the loads are taken into account as seen in Fig. 22, the results of the study are conrmed; for the fuels B100B and B100C the
classical sooteNOx tradeoff has completely changed and the reductions in emissions are very important. Further improvement
could be made by incorporating the EGR technology or using variable injection timings.
5.2. Carbon monoxide and unburned hydrocarbons
The disadvantages of n-butanol PCCI and LTC operation found in
this research are the higher CO, Fig. 23 and HC, Table 5 emissions
primarily due to incomplete combustion during the premixed burn
phase from over-lean areas in the combustion chamber and
showing an increase of 10e20 times dependent upon the speed and
load. Other factors that may explain this high increase are the lack
of intake manifold heating that probably is producing a fuel pooling
in the intake, and a crude manifold injection strategy that has not
been correlated yet with the valves timing, thus allowing the
passage of some butanol directly from intake into the exhaust
manifold.

10

5.3. Formaldehyde

1000

The use of alcohol and biodiesel in LTC regimes caused increases


in formaldehyde due to the oxygenated characteristics associated
with both fuels, and the incomplete combustion caused by LTC. This
increase is more prominent at lower loads and decrease as the load
is increased, Table 6. The low power output for idling also

100
10

Table 6
Formaldehyde emissions for each load in g/kWh.

1
1

1.5

2
IMEP [bar]

Fig. 23. CO emissions at 800 rpm.

2.5

Load [IMEP]

Diesel

B100

B100B

B100C

1
2
3

0.69
0.31
0.23

0.92
0.45
0.26

30.70
5.33
2.37

19,909
10.82
4.19

V. Soloiu et al. / Energy 52 (2013) 143e154

7. Conclusions

200

Specific Energy Consumption [MJ/kWh]

153

Diesel
B100
B100 B
B100 C

150

100

50

0
1

1.5

2.5

IMEP [bar]
Fig. 24. Specic energy consumption at 800 rpm and various loads.

contributed to increases in formaldehyde emissions as opposed to


diesel and biodiesel that exhibit a more complete combustion.
6. Engine efciency
Dual fuels were supplied into the engine simultaneously,
requiring the BSFC to be represented by the specic energy consumption per kilowatt-hour. Fig. 24 suggests that the efciency
converges at higher loads towards the ULSD#2 and B100 values, but
the research could not proceed further due to cycle irregularity
since the engine has a limited load range available at idling speed,
with a maximum load of 3 bar IMEP. The overall efciency constantly increased with the increase in load, for both fuels B100B and
B100C; B100B almost achieved the same efciency as ULSD#2 and
B100 3 bar IMEP as seen in Fig. 25.

The low-temperature combustion of the n-butanolebiodiesel


mixtures was achieved by a combination of single-shot direct
biodiesel injection and low pressure sequential n-butanol port injection. LTC was achieved for low to medium loads when combined
with PCCI at idling speeds.
The premixed combustion phase was drastically diminished
with the injection of n-butanol and for high contents of alcohol and
medium loads an early low-temperature heat release was detected.
At 3 bar IMEP, the early low-temperature heat release of nbutanol based fuel starts 6 (1.25 ms) earlier than the diesel reference heat release, with a peak at 10 BTDC corresponding to
1200 K and maintaining the LTC regime.
The diffusion combustion was present only at very low loads and
was transformed into a second stage of high-temperature heat
release by adding the n-butanol, resulting into a modication of
the sooteNOx tradeoff. NOx was signicantly reduced by 74% at
3 bar IMEP.
Heat losses from radiation of burned gas in the combustion
chamber decreased by 10e50%, in good correlation with the major
decrease in soot emissions of about 98% for 3 bar IMEP.
High CO and HC have been detected in the exhaust gas, primarily due to partially combusted fuel escaping the premixed
combustion process and showing an increase of 10e20 times
depending of speed and load.
The specic fuel energy consumption increased when the engine was fueled with n-butanol, however, for 3 bar IMEP, the value
for all fuels was in the same range.
The research was able to prove that a clean idling technology
can be developed based on n-butanol PFI combined with PCCI and
LTC in a biodiesel fueled engine. Reductions in soot and NOx were
very substantial and their tradeoff was modied favorably.
Future research should mitigate the shortcomings found by
heating the intake manifold to eliminate fuel pooling in the intake
with an improved manifold injection strategy correlated with the
valves timing to avoid the passage of butanol directly from intake
into the exhaust manifold, and an EGR strategy to wider further LTC
range with sooteNOx reduction.
Acknowledgments

30
25

Overall Efficiency [%]

National Science Foundation for their support, Dr. Chris Butts


from the U.S. National Peanuts laboratory, Dr. Brian Koehler from
Georgia Southern University for their contributions.

Diesel
B100
B100 B
B100 C

References

20
15
10
5
0
1

1.5

2
IMEP [bar]

Fig. 25. Overall efciency at 800 rpm.

2.5

[1] U.S. Environmental Protection Agency. Veried idling reduction technologies.


U.S. EPA. Web, www.epa.gov; Sept. 2011, http://www.epa.Gov/smartway/
technology/idling.htm; Sept. 2011.
[2] United States Government. Energy Independence and Security Act of 2007;
2007. p. 1e310. [Print].
[3] Lapuerta Margin, Octavio Armas, Jose Rodrigeuz-Fernandez. Effect of biodiesel
fuels on diesel engine emissions. Progress in Energy and Combustion Science
2008;34(2):198e223.
[4] Najt P, Foster D. Compression-ignited homogeneous charge combustion. SAE
Technical Paper 830264, http://dx.doi.org/10.4271/830264; 1983.
[5] Thring R. Homogeneous-charge compression-ignition (HCCI) engines. SAE
Technical Paper 892068, 1989; http://dx.doi.org/10.4271/892068.
[6] Ryan T, Callahan T. Homogeneous charge compression ignition of diesel fuel.
SAE Technical Paper 961160, http://dx.doi.org/10.4271/961160; 1996.
[7] Gray III AW, Ryan III TW. Homogeneous charge compression ignition (HCCI) of
diesel fuel. SAE paper 971676; 1997.
[8] Stanglmaier R, Roberts C. Homogeneous charge compression ignition (HCCI):
benets, compromises, and future engine applications. SAE Technical Paper
1999-01-3682, http://dx.doi.org/10.4271/1999-01-3682; 1999.
[9] Saisirirat P, Foucher F, Chanchaona S, Mounaim-Rouselle C. Effects of ethanol,
nbutanolen-heptane blended on low temperature heat release and HRR

154

[10]

[11]
[12]

[13]

[14]
[15]

[16]

[17]

[18]

[19]

[20]

[21]
[22]

[23]

[24]

[25]

[26]

[27]

[28]

[29]
[30]

[31]

V. Soloiu et al. / Energy 52 (2013) 143e154


phasing in diesel-HCCI. SAE Technical Paper 2009-24-0094. http://dx.doi.org/
10.4271/2009-24-0094.
Wang Xiangang, Cheung CS, Di Yage, Huang Zuohua. Diesel engine gaseous
and particle emissions fueled with dieseleoxygenate blends. Fuel April 2012;
94:317e23.
Peter D. Biobutanol: an attractive biofuel. Biotechnology Journal 2007;2(12):
1525e34.
Lebedevas Sergejus, Lebedeva Galina, Sendzikiene Egle, Makarevicience Violeta.
Investigation of the performance and emission characteristics of biodiesel fuel
containing butanol under the conditions of diesel engine operation. Energy &
Fuels 2010;2010(24):4503e9 [Print].
Ramey David, Yang Shang-Tian. Production of butyric acid and butanol from
biomass. United States Department of Energy. Web, www.afdc.energy.gov;
2004, http://www.afdc.energy.gov/afdc/pdfs/843183.pdf; 2004.
Pucher H, Sperling E. N-butanol-diesel mixture as alternative fuel for diesel
motors. Erdol & Kohle Erdgas Petrochemie 1986;39(8):353e6. 15.
Miers SA, Carlson RW, McConnell SS, Ng HK, Wallner T, Esper JL. Drive cycle
analysis of butanol/diesel blends in a light-duty vehicle. SAE Technical Paper
2008-01-2381, 2008; http://dx.doi.org/10.4271/2008-01-2381.
Weiskirch C, Kaack M, Blei I, Eilts P. Alternative fuels for alternative and
conventional diesel combustion systems. SAE Technical Paper 2008-01-2507,
2008; http://dx.doi.org/10.4271/2008-01-2507.
Zldy M, Hollo A, Thernesz A. Butanol as a diesel extender option for internal
combustion engines. SAE Technical Paper 2010-01-0481, 2010; http://
dx.doi.org/10.4271/2010-01-0481.
Yao M, Zheng Z, Haifeng L. Progress and recent trends in homogeneous charge
compression ignition (HCCI) engines. Progress in Energy and Combustion
Science 2009;35:398e437.
Soloiu V, Duggan M, Ochieng H, Williams D, Molina G, Vlcek B. Investigation of
low temperature combustion regimes of biodiesel with n-butanol in the
intake manifold of a compression ignition engine. ASME Internal Combustion
Engine Division; 2012. ICEF2012e92053.
Bunting B, Eaton S, Crawford R, Xu Y, Wolf LR. Performance of biodiesel blends
of different FAME distributions in HCCI combustion. SAE paper 2009-01-1342,
2009; http://dx.doi/org/10.4271/2009-01-1342.
Westbrook CK. Chemical kinetics of hydrocarbon ignition in practical combustion systems. Proceedings of the Combustion Institute 2000;28:1563e77.
Leppard W. A comparison of olen and parafn autoignition chemistries:
a motored-engine study. SAE Technical Paper 892081, 1989; http://dx.doi.org/
10.4271/892081.
Grifths JF, Halford-Maw PA, Rose DJ. Fundamental features of hydrocarbon
autoignition in a rapid compression machine. Combustion and Flame 1993;
95:291e306.
Grifths JF, Halford-Maw PA, Mohamed C. Spontaneous ignition delays as
a diagnostic of the propensity of alkanes to cause engine knock. Combustion
and Flame 1997;111:327e37.
Westbrook C, Pitz W, Leppard W. The autoignition chemistry of parafnic
fuels and pro-knock and anti-knock additives: a detailed chemical kinetic
study. SAE Technical Paper 912314, 1991; http://dx.doi.org/10.4271/912314.
Sjberg M, Dec JE. Comparing late-cycle autoignition stability for single- and
two-stage ignition fuels in HCCI engines. Proceedings of the Combustion
Institute 2007;31:2895e902.
Sjberg M, Dec J, Hwang W. Thermodynamic and chemical effects of EGR and
its constituents on HCCI autoignition. SAE Technical Paper 2007-01-0207,
2007; http://dx.doi.org/10.4271/2007-01-0207.
McCormick R, Alvarez J, Graboski M, Tyson K, Vertin K. Fuel additive and
blending approaches to reducing NOx emissions from biodiesel. SAE Technical
Paper 2002-01-1658, 2002; http://dx.doi.org/10.4271/2002-01-1658.
Monyem A, Van Gerpen J. The effect of biodiesel oxidation on engine performance and emissions. Biomass & Bioenergy 2001;20:317e25.
Wallner T, Miers S, McConnel S. A comparison of ethanol and butanol as
oxygenates using a direct-injection, spark-ignition engine. Journal of Engineering for Gas Turbines and Power 2009;131.
Lujaji Frank, Kristf Lukcs, Bereczky Akos, Mbarawa Makame. Experimental investigation of fuel properties, engine performance, combustion

[32]

[33]

[34]

[35]

[36]

[37]

[38]

[39]

[40]
[41]

[42]

[43]

[44]

[45]

[46]

[47]

[48]

[49]

[50]

and emissions of blends containing croton oil, butanol, and diesel on a CI


engine. FUEL The Science and Technology of Fuel and Energy 2010;90:505e
10 [Print].
Rakopoulos D, Rakopoulos C, Papagiannakis R, Kyritsis D. Combustion heat
release analysis of ethanol or n-butanol diesel fuel blends in heavy-duty DI
diesel engine. Fuel The Science and Technology of Fuel and Energy May 2011;
90(5):1855e67.
Kim D, Kim M, Lee C. Effect of premixed gasoline fuel on the combustion
characteristics of compression ignition engine. Energy & Fuels 2004;18:
1213e9.
Lu Y, Yu W, Su W. Using multiple injection strategies in diesel PCCI combustion: potential to extend engine load, improve trade-off of emissions and
efciency. SAE Technical Paper 2011-01-1396, 2011; http://dx.doi.org/
10.4271/2011-01-1396.
Park Y, Bae C. Inuence of EGR and pilot injection on PCCI combustion in
a single-cylinder diesel engine. SAE Technical Paper 2011-01-1823, 2011;
http://dx.doi.org/10.4271/2011-01-1823.
Neely G, Sasaki S, Leet J. Experimental investigation of PCCI-DI combustion on
emissions in a light-duty diesel engine. SAE Technical Paper 2004-01-0121,
2004; http://dx.doi.org/10.4271/2004-01-0121.
Karra P, Veltman M, Kong S. Characteristics of engine emissions using biodiesel blends in low-temperature combustion regimes. Energy & Fuels 2008;
2008:3763e70 [Print].
Altun S, Oner C, Yasar F, Adin H. Effect of n-butanol blending with a blend of
diesel and biodiesel on performance and exhaust emissions of a diesel engine.
Industrial & Engineering Chemistry Research 2011;50:9425e30.
Kawamoto K, Araki T, Shinzawa M, Kimura S, Koide S, Shibuya M. Combination of combustion concept and fuel property for ultra-clean DI diesel. SAE
Technical Paper 2004-01-1868, 2004; http://dx.doi/org/10.4271/2004-011868.
Benajes J, Molina S, Novella R, Amorim R. Study on low temperature combustion for light-duty diesel engines. Energy & Fuels 2010;24:355e64.
Bamgboye A, Hansen A. Prediction of cetane number of biodiesel fuel from the
fatty acid methyl ester (FAME) composition. International Agrophysics 2008;
22:21e9.
Wadumesthrige K, Ng K, Salley S. Properties of butanol-biodiesel-ULSD ternary mixtures. SAE International Journal of Fuels and Lubricants 2010;3(2):
660e70. http://dx.doi.org/10.4271/2010-01-2133.
Soloiu V, Lewis J, Covington A, Vlcek B, Schmidt N. The inuence of peanut
fatty acid methyl ester blends on combustion in an indirect injection diesel
engine. In: Proceedings of the ASME 2011 internal combustion engine division
fall technical conference ICEF2011; October 2e5, 2011, Morgantown, West
Virginia, USA.
Steidley Kevin R, Knothe Gehard. Kinematic viscosity of biodiesel components
(fatty acid alkyl esters) and related compounds at low temperatures. Fuel
2007;86(16):2560e7 [Print].
Annand WJD, Ma TH. Instantaneous heat transfer rates to the cylinder head
surface of a small compression-ignition engine. Proceedings of the Institution
of Mechanical Engineers 1971;185(16):976e87.
Soloiu V, Nelson D, Covington A, Lewis J. Investigations of a fatty acid methyl
ester from poultry fat in a triple vortex separate combustion chamber diesel
engine stage one-combustion investigations. SAE Technical Paper 2011-011188, 2011; http://dx.doi.org/10.4271/2011-01-1188.
Lyn W. Study of burning rate and nature of combustion in diesel engines.
Proceedings of the Combustion Institute. Ninth Symposium (International) on
Combustion. NY: Academic Press; 1963. p. 1069e82.
Khan I, Greeves G, Wang C. Factors affecting smoke and gaseous emissions
from direct injection engines and a method of calculation. SAE Technical
Paper 730169, 1973; http://dx.doi.org/10.4271/730169.
Pishinger G,M, Falcon A,M, Siekmann R,W, Fernandes F,R. Methyl esters of
plant oils as diesel fuels, either straight or in blends vegetable oil fuels; 1982.
SAE Paper No. 4-82.C.
Marchetti JM, Miguel VU, Errazu AF. Possible methods for biodiesel production. Renewable and Sustainable Energy Reviews August 2007;11(6):1300e
11. http://dx.doi.org/10.1016/j.rser.2005.08.006.

Das könnte Ihnen auch gefallen