Sunteți pe pagina 1din 9

Combustion and Flame 159 (2012) 1100–1108

Contents lists available at SciVerse ScienceDirect

Combustion and Flame


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

Variations in the chemical composition and morphology of soot induced


by the unsaturation degree of biodiesel and a biodiesel blend
Maurin Salamanca a, Fanor Mondragón a, Jhon Ramiro Agudelo b, Pedro Benjumea c,
Alexander Santamaría a,⇑
a
Institute of Chemistry, University of Antioquia, A.A. 1226 Medellín, Colombia
b
Engineering Faculty, University of Antioquia, A.A. 1226 Medellín, Colombia
c
Faculty of Mines, National University of Colombia, A.A. 1026 Medellín, Colombia

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

Article history: This work is about the influence of the molecular structure of the fatty acid esters present in two neat
Received 23 February 2011 biodiesel fuels and their blend (50% by volume) on particulate matter emission. Experiments were per-
Received in revised form 19 July 2011 formed in a four-cylinder direct injection automotive diesel engine under carefully controlled operating
Accepted 11 October 2011
conditions, so that the difference in performance and emissions were affected only by biodiesel fuels
Available online 7 November 2011
composition and properties. The results indicated that the composition and degree of unsaturation of
the methyl ester present in biodiesel plays an important role in the chemical composition of particulate
Keywords:
matter (PM) emitted. It was observed that linseed biodiesel (BL100) produces more PM and hydrocarbons
Biodiesel
Soot
(HC) than Palm biodiesel (BP100) as a consequence of more unsaturated compounds in its composition,
Unsaturation degree which favor the soot precursor’s formation in the combustion zone. Thermogravimetric analysis (TGA)
Aliphatic component showed that the amount of volatile material in the soot from biodiesel fuels was slightly lower than that
FT-IR of diesel fuel, but not significant differences were observed among biodiesels. Similarly, the chemical
characteristics of the hydrocarbons of volatile material present in the particulate matter (referred in
the literature as SOF-soluble organic fraction), showed an increase in the aliphatic component as the
unsaturation degree of the fatty acid methyl ester increased. Additionally, it is concluded that there
are not significant nano-structural differences in the soot obtained from pure biodiesel fuels, even if they
have very different degrees of unsaturation.
Ó 2011 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction However, in some cases nitrogen oxides (NOx) can be emitted in lar-
ger quantities than those produced by diesel fuels [3–5].
The use of biodiesel is rapidly expanding around the world, Particular attention has been paid to the reduction of PM be-
making imperative to understand the impacts of biodiesel on the cause of environmental and health concerns, since it has been
diesel engine combustion process and pollutant formation. Accord- found that these particles are mutagenic and potentially carcino-
ing to the World Customs Organization (WCO) biodiesel is a genic [6]. The solid particles emitted from a diesel engine consist
renewable, biodegradable, environmentally friendly and energy of carbonaceous material accompanied of various organic and inor-
efficient fuel, which can fulfill energy security needs without sacri- ganic compounds, which are generated when the fuel is burned in
ficing the engine’s operational performance. Biodiesel is a mixture an environment where the temperature is high and oxygen con-
of mono alkyl esters of long-chain [C16AC18] fatty acids, usually centration is low. So far, soot formation reactions have not yet been
methyl or ethyl esters, obtained by transesterification of the tri- fully explained, and it is thought that soot composition depends on
glycerides contained in vegetable oils and animals fats [1,2]. many parameters such as combustion conditions, engine operating
Some advantages of using biodiesel or its blends on engine mode (engine speed and torque), and the quality of fuel and the
performance, apart from reducing dependency on imported petro- lubricating oil [4,7].
leum, are: the high flash point, the inherent lubricity and the reduc- Most studies have found sharp PM reductions when biodiesel or
tion of most regulated exhaust emissions (particulate matter (PM), diesel–biodiesel blends are used as compared to diesel fuel. These
hydrocarbons (HC), carbon monoxide (CO)) compared to diesel fuel. variations are mainly caused by reduction in the formation of soot
precursors as well as by the enhanced soot oxidation. Larger oxy-
gen content combined with the absence of aromatic and sulfur
⇑ Corresponding author. Fax: +57 42196565. compounds in biodiesel have been pointed out as the main reasons
E-mail address: alsp04@matematicas.udea.edu.co (A. Santamaría). for this behavior [3,4,8].

0010-2180/$ - see front matter Ó 2011 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.combustflame.2011.10.011
M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108 1101

Frijters and Baert [9] evaluated various biodiesel fuels among NOx emissions, while PM emissions remained unchanged. In addi-
other oxygenates and found a good correlation between PM emis- tion, fully saturated fatty acid chains (C12, C16 and C18), showed a
sions and the oxygen content in the fuel. Schmidt and Van Gerpen NOx emissions increased with decreasing chain length, while PM
[10] observed that the need for additional oxygen to obtain a emissions did not show significant changes.
certain reduction in PM emissions was lower when using oxygen- Lapuerta et al. [14] carried out experiments in order to identify
ates instead of oxygen-enriched air as combustion reagent. This the effect of the degree of unsaturation of biodiesel fuels on the per-
experiment proved the advantage of fuel-bonded oxygen as a conse- formance and emissions of diesel engines. In their experiments the
quence of the higher accessibility of the oxygen atom during iodine number ranged from 90 to 125 and the results showed that
combustion. the degree of unsaturation had significant effects on PM emissions.
Kohse et al. describe in their review that the chemical decompo- As the biodiesel fuel became more unsaturated, PM emissions de-
sition mechanism of biodiesel during combustion processes is creased by 20%. Regarding particle size distributions, unsaturated
much more complex than its counterpart (diesel), not only because biodiesel fuels showed a smaller mean particle diameter.
of the size of the fuel molecules, but also as a consequence of the Sarathy et al. [17] carried out a study using diffusion and pre-
additional reactions of oxygen-containing species. For example, mixed flames in order to investigate the effect of chemical structure
methyl esters can decompose into two separate reactive oxygen of saturated and unsaturated fatty acid methyl esters (FAME) on the
carriers which could contribute to soot-precursor reduction, or composition of combustion products. They found that the unsatu-
the OACAO structure in the molecule may form CO2. Also, It has rated methyl ester fuel (methyl crotonoate) led to the generation
been found that depending of the type of alkyl ester (methyl or of higher levels of oxygenated species, as well as much higher levels
ethyl esters), the decomposition path can vary significantly. For of C2H2, 1-C3H4, 1-C4H8, 1,3-C4H6 and benzene than the saturated
the methyl esters, the unimolecular decomposition will break a methyl ester (methyl butanoate), all of them considered soot pre-
CAO bond to form CH3 or CH3O radicals, whereas ethyl esters cursors. This suggests that unsaturated FAME would have a greater
can decompose through a unimolecular elimination reaction pro- tendency to soot formation than saturated FAME.
ducing C2H4 via a six-membered pericyclic transition state [11]. Puhan et al. [18] tested biodiesel from linseed, jatropha and coco-
Also, it has been found that the soot particle structure coming nut oils, in order to study the effect of molecular structure on the
from biodiesel differs from that produced by conventional diesel combustion characteristics and emissions in a single cylinder direct
fuels. By using TGA under inert atmosphere followed by oxidation injection engine. They found that emissions of NOx, smoke, Total
of devolatilized PM coming from diesel and a blend of diesel with hydrocarbons (THCs) and CO increased with the degree of unsatura-
20 wt% biodiesel, Boehman et al. [12] found that samples from the tion. Schönborn et al. [15] reported similar results. In this case they
B20 blend had a lower oxidation temperature, due to the presence carried out a series of experiments on a single-cylinder engine to
of oxygenated complexes at the surface while amorphous carbon investigate the influence of the molecular structure of fatty acid al-
was found in the inner core of soot particles. Song et al. [13] ex- kyl ester molecules under diesel conditions. It was observed that the
tended this work to pure soybean-oil biodiesel, and confirmed fas- emission of particulate matter showed correlation with the number
ter oxidation of the soot emitted by the neat biodiesel. They found of double bonds in the fuel molecules for the case of large accumu-
that the oxidation mechanism of the biodiesel soot leads to a cap- lation mode particles, and with the boiling point of the fuel samples
sule-type oxidation and eventual formation of graphene ribbon for the case of the smaller, nucleation mode particles.
structures. This mechanism starts with the development of micro- Benjumea et al. [19] conducted an experimental investigation
porosity during the course of the early stages of oxidation through on a high-speed direct-injection automotive diesel engine fueled
the partial removal of the amorphous aliphatic carbon-rich sec- with three mixtures of fatty acid methyl esters. The fuel matrix
tions present at the outermost shell. Once sufficient micropores was designed such that the effect of the degree of unsaturation
have been developed to penetrate to the particle core, the soot is of the tested biodiesel fuels was isolated. Results showed that THCs
likely to become hollow inside through internal burning of the emissions, smoke opacity, and NOx emissions increased with de-
more reactive internal carbon, forming ‘‘capsules’’. However to ex- gree of unsaturation. They also found that a higher degree of unsat-
plain this transition, the authors suggest the existence of an inter- uration of biodiesel fuels led to a longer ignition delay-ID (lower
mediate layer (that is less graphitic than the outermost shell, but cetane number) and, consequently, a more retarded start of com-
still more graphitic than the inner core) that survives this process. bustion (SOC), if this is true, it is expected an increment in the
After the removal of the inner core, physical factors such as the in- PM emission, especially for those biodiesel fuel with cetane num-
crease in layer mobility arising from reduced cross-linking and ber below 45, as it is indicated in Ref. [4].
minimization of strain energy arising from the hollowing out of This work is an extension of the already published work of
the soot particle can lead to layer rearrangement and coalescence Benjumea et al. who showed the effect of unsaturation degree of
where the wavy layers become much flatter and longer, with the biodiesel in engine performance [19]. In the present research, the
appearance of a ribbon-like structure [13]. main purpose was to evaluate the effect of biodiesel fueling on
The above reported results show that the fuel-bonded oxygen in the chemical composition and morphology of particulate matter
biodiesel plays a key role in PM reduction as compared to diesel emitted by a high speed direct injection (HSDI) diesel engine. In or-
fuels. Other investigations have been carried out aiming to explain der to give insights about the relationship between the chemical
the effect of biodiesel composition, especially unsaturation degree, structure of fatty acid methyl esters and PM emissions characteris-
on PM emissions and their chemical composition and morphology, tics, two types of biodiesel fuels representing extreme degrees of
which means that the study on this topic is not conclusive unsaturation and their blend at 50% by volume, were tested. The
[4,14,15]. effect of FAME chemical structure on engine emissions is becoming
McCormick et al. [16] tested 21 biodiesel fuels derived from a key criterion to choose raw materials for biodiesel production.
several feedstock and several pure fatty acid methyl esters in a die-
sel engine using the US heavy-duty federal test procedure. The ef- 2. Methodology
fects of fatty acid chain length and the number of double bonds on
NOx and PM emissions were investigated. They found a direct cor- 2.1. Fuels characteristics
relation between the number of double bonds (quantified as iodine
number, which is defined as a measure of the amount of iodine ab- The evaluated fuels were neat palm oil biodiesel, neat linseed oil
sorbed in a given time by a chemically unsaturated material) and biodiesel and their blend at 50% by volume. These fuels are described
1102 M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108

as BP100, BL100 and BPL50, respectively. Tables 1 and 2 show the developed by authors in order to translate this transient cycle into
chemical composition and the main physicochemical properties of steady operation modes (torque–speed) [20].
the tested fuels. The engine was coupled to an eddy current dynamometer
The methyl ester content of the neat biodiesel fuel was above (Schenck W230). The air consumption was measured with a hot-
97% by weight. The chemical composition of the test fuels quanti- wire sensor (Magnetrol TA2), and fuel consumption with an elec-
fied by gas chromatography indicated that palm oil biodiesel is tronic balance mass flow sensor. For recording the instantaneous
characterized by having a balance between saturated and mono- in-cylinder pressure, a Kistler 6056A piezoelectric pressure trans-
saturated methyl esters. On the other hand, linseed oil biodiesel ducer installed in the glow plug and a Kistler 5011B charge ampli-
is predominantly unsaturated (90.8%), having significant contents fier were used.
of di-unsaturated and especially tri-unsaturated methyl esters. A One hundred pressure curves were registered in each engine
commercial grade no. 2 diesel was used as reference fuel with an operation mode in order to guarantee confidence in the combustion
elemental mass composition of 87.2% carbon, 12.8% hydrogen diagnosis. That number of pressure curves led to a coefficient of var-
and 0.0225% sulfur, and an aromatic content of 29.3% (13% mono- iation (COV) of the indicated mean effective pressure under 3% for
aromatics, 13.3% diaromatics and 3% polyaromatics). all tests performed. The instantaneous piston position was deter-
mined using an angular encoder with a resolution of 1024 pulses/
2.2. Engine parameters and sampling procedure revolution (Heidennhain ROD 426) coupled to the crankshaft at
the opposite extreme of the fly-wheel. High speed data were ac-
Experiments were carried out in an instrumented automotive quired using TonéÒ LabView based software and National Instru-
diesel engine (Table 3). The engine was operated at 2420 rpm ments™ data acquisition system (Model PCI-MIO-16E-4 board).
and 95 Nm. These conditions were chosen because it was the point From a software called Caribe, which was developed to run the
of minimum air–fuel ratio and maximum smoke opacity according thermodynamic combustion diagnosis of direct injection diesel en-
to the Federal test procedure (FTP-75) homologation cycle. The gines based on in-cylinder pressure trace. A single-zone thermody-
dynamic characteristics of the vehicle Chevrolet Luv 2.5 Turbo, namic model described in Ref. [21] is used to determine the heat
equipped with an Isuzu 4JA1 diesel engine used in this study, release and the bulk in-cylinder temperature’’. Smoke opacity was
and the FTP-75 speed profile were programmed in a software measured with an AVL Dicom 4000 optical sensor.
Before starting experiments with a new fuel, the lines were
Table 1 drained, then the fuel was added and the engine operated at least
Chemical composition of neat biodiesel fuels. for 1 h in order to purge any fuel remaining from previous experi-
BP100 BPL50 BL100 ments. Measurements were carried out during 5 h of continuous
Lauric (12:0) 0.31 0.17 0.00
stationary operation in order to guarantee enough particle matter
Miristic (14:0) 1.03 0.60 0.00 to be collected (30–50 mg). Each fuel was randomly tested three
Palmitic (16:0) 43.3 25.9 5.30 times on different days and experiments were carried out at
Stearic (18:0) 4.20 4.1 3.90 1500 m above sea level (patm = 640 mm Hg) without any modifica-
Palmitoleic (16:1) 0.15 0.12 0.07
tion of the engine or the fuel injection system.
Oleic (18:1) 41.8 32.1 20.5
Linoleic (18:2) 9.10 12.8 17.1 Soot particles were taken from the exhaust pipe using a 50 cm-
Linolenic (18:3) 0.15 24.3 53.1 length stainless steel probe connected to a vacuum system in line
Total saturated (%) 48.8 30.8 9.2 with a Teflon filter of 0.25 lm pore diameter. This filter was used
Total unsaturated (%) 51.1 69.2 90.8
to collect the PM and low molecular weight compounds. A cold
trap was used to capture water and light compounds that passed
through the filter. The sampling collection time was carried out
Table 2
every 15 min during a whole cycle in order to obtain 15–20 filters
Physicochemical properties of evaluated fuels.
by each fuel tested. Then, 10 filters randomly selected were
Properties Diesel BP100 BL100 BPL50 weighted and the average weight value, including the standar devi-
no. 2
ation of them were reported in Table 4.
Chemical formulac C14.7H28,8 C18.1H34.9O2 C18.9H33.5O2 C18.4H34.2O2
Molecular weight 205.2 284.2 292.6 287.6
Density at 15 °C (kg/ 859.3 871.6 893.5 885.2
2.3. Particulate matter chemical characterization
m3)
Lower heating value 42.2 37.1 37.1 37.1
(MJ/kg)a 2.3.1. Thermogravimetric analysis-TGA
Oxygen (wt%)b 0 11.3 10.9 11.1 TGA of bulk PM samples collected were subjected to a thermal
Cetane numberd 46 63.7 36.5 51.3
treatment at 800 °C in nitrogen atmosphere to remove water and
Sulfur content (ppm) 225 0 0 0
the volatile fraction of particles. Once this temperature was
a
Calculated from ultimate composition and measured gross heating value. reached, the sample was iso-thermally held during 12 min in nitro-
b
Calculated from composition. gen and then oxidized in air to induce the combustion of the fixed
c
Calculated from fatty acid methyl esters profile.
d
Calculated as mass-weighted average.
carbon content, which is sometimes associated with the elemental

Table 4
Table 3
Engine performance with different fuels.
Automotive diesel engine characteristics.
Fuel Break power Equivalence Max. in-cylinder PM (mg)a Opacity
Reference ISUZU 4JA1
(kW) ratio Temp. (°C) (%)
Type Turbocharged, direct injection, rotating pump
Swept volume 2499 cm3 BP100 23.9 ± 0.1 0.37 ± 0.01 1710 ± 4.2 1.02 ± 0.23 0.6 ± 0.1
Configuration 4 in-line cylinders BPL50 23.9 ± 0.1 0.37 ± 0.01 1731 ± 5.2 1.93 ± 0.29 0.9 ± 0.1
Diameter  stroke 93 mm  92 mm BL100 24.4 ± 0.1 0.37 ± 0.01 1780 ± 3.1 2.75 ± 0.36 1.3 ± 0.2
Compression ratio 18.4 Diesel 24.0 ± 0.4 0.37 ± 0.01 1674 ± 5.5 2.96 ± 0.38 1.9 ± 0.2
Rated power 59 kW (80 hp) at 4100 rpm no. 2
Maximum torque 170 Nm at 2300 rpm a
Mass of PM collected in Teflon filters.
M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108 1103

carbon of soot samples. All experiments were performed in a TA Data in Table 2 shows that the average chain length and oxygen
2950 unit and the total flow rate of gases was adjusted at content of biodiesel remained almost constant indicating that the
100 mL/min during each test. The uncertainty was estimated by main structural difference among them was their degree of unsat-
the change in mass profiles for three soot samples run under the uration. On the other hand, properties such as density, cetane num-
same conditions. ber, in-cylinder temperature and smoke opacity were found to be
sensitive to the unsaturation degree of biodiesel. For instance, it
2.3.2. Infrared spectroscopy characterization (FT-IR) was found that the smoke opacity, a property related with the par-
For qualitative FT-IR analysis a KBr pellet was prepared at ticulate matter emission, was higher for diesel compared to biodie-
1 wt%. of PM. Each spectrum was the result of 300 scan accumula- sel fuels. However, this property increased among the biodiesel
tion, which provided the best signal/noise ratio. A Nicolet Magna fuels according to the degree of unsaturation, result that also agrees
560 spectrometer was used with a MCT/A detector operated in a with the mass of particulate matter collected in the filters (see Table
wavenumber range of 600–4000 cm 1. Three measurements of 4). Sarathy et al. confirmed this tendency and suggested that an
each sample were taken in order to estimate the method reproduc- unsaturated FAME would have a greater tendency to soot formation
ibility. The uncertainty of the IR measurements was less than 5%. than a saturated FAME [17]. This is consistent with the EPA engine
data [24] which determined that the more unsaturated soybean
2.3.3. Raman spectroscopy analysis methyl esters formed more soot than the more saturated beef tal-
Raman spectra of the PM samples were recorded with a LabR- low methyl esters. However, some engine data did not follow this
aman HR Horiba microscope system using a 632.8 nm He/Ne laser trend.
as an excitation source. Seven spectra were taken for each sample On the other hand, density and cetane number have shown some
in different spots. Spectra of the samples were in the range of relation to particulate matter emission [4]. For instance, when the
100–3600 cm 1 with a 50X magnification objective and 20 s expo- biodiesel density was higher than 895 kg/m3 or the cetane number
sition time, a source power of 0.17 mW (in order to avoid altering lower than 45 the PM emissions increased considerably. So that it is
or burning the sample). Curve fitting for the determination of first expected that PM emission from BL100 will be similar to diesel, but
order spectral parameters was made with the LabPlot software. higher than BP100 and BPL50 respectively (see Table 4). Also, it was
Following the recommendations by Sadezky et al. [22], two found that as the degree of unsaturation of the test fuels increased,
Lorentzian functions (for the graphite band G, located around the peak in-cylinder bulk-gas-averaged temperature obtained at
1580 cm 1, and the defect band D1 located around 1360 cm 1) the high load condition also increased due to an increase in the pre-
and one Gaussian function (for the defect band D3, located around mixed fraction of combustion, resulting in more energy being
1500 cm 1) were used. Seven different spots were analyzed and released over a short time scale with a nearly constant combustion
averaged in each sample in order to improve the statistical signif- chamber volume. This result was previously reported by authors in
icance. For each Raman spectra, the fitting procedure was repeated Ref. [19] and it is introduced here to support the interpretation of
three times in order to ensure the reproducibility. the possible chemical changes observed for PM when it is exposed
to a high temperature regimen.
2.3.4. Energy dispersive X-ray analysis
Energy dispersive X-ray (EDX) spectroscopy coupled with scan- 3.2. Chemical characterization of particulate matter obtained from
ning electron microscopy (SEM) has been widely used to detect neat biodiesel fuels
elements in the surface of single airborne particles [23]. Identifica-
tion and semi-quantitative analysis of elements present on parti- Table 5 summarizes the dry basis and dry ash-free basis proxi-
cles collected on filters was performed in an EDX Oxford mate analysis of particulate matter samples produced when neat
Instruments Inca Penta Fet x3 unit operated at an excitation volt- palm and linseed oil biodiesels and the 50% mixture of them were
age of 50 kV during a collection time of 30 s. The compositional re- used in the engine, for comparison, results from diesel PM samples
sult for each sample was reported as a nine points average. run under the same conditions were introduced. In general, the
proximate analysis of the four samples was very similar. However,
it can be observed that the amount of volatile material produced
2.4. Particulate matter morphological analysis
under inert atmosphere between 120 °C and 800 °C is slightly less
in the PM coming from neat biodiesel than for Diesel no. 2.
2.4.1. Transmission Electron Microscopy (TEM)
Similarly, this behavior was also reflected in the small increase
Information of size and morphology of PM was obtained by
of the fixed carbon content, especially among the samples coming
TEM. A small amount of soot collected in Teflon filters was depos-
from neat biodiesel fuels, whose values were slight above of the
ited on the Formvar grid and then placed in a JEOL JEM 1200EX
reference fuel (44.1% DAF base). This trend is opposite to what
microscope operated at a voltage range between 20 and 120 kV.
has been reported in literature, for instance, Boehman et al. [12]
Measurement of the particle size was done by hand using the free
and Williams [25] found an increase in the soluble organic fraction
ImageJ software, taken a 100 nm scale as reference.
followed by a reduction in the elemental carbon of PM taken at the

3. Results and discussion


Table 5
3.1. Engine results Proximate analysis of PM determined by TGA. In all cases the uncertainty was
estimated to be less than 0.1%.

Table 4 shows the engine performance results comparing biodie- Fuel Volatile matter %(w/ Fixed carbon %(w/ Residue %(w/w)
sel fuels with diesel no. 2 fuel. Each result is the average of five w) 120–800 °C w) Comb. 800 °C

measurements (one per hour of test). Break power was almost the Dry DAF Dry DAF Dry
same for all fuels, which guarantees that measurements were car- BL100 48.9 55.1 39.8 44.9 11.3
ried out under the same energy output. Equivalence ratio was almost BP100 48.3 53.7 41.6 46.3 10.1
the same for all fuels, so that, the differences in performance and BPL50 47.9 52.9 42.7 47.1 9.4
Diesel no. 2 50.9 55.9 40.2 44.1 8.3
emissions were not affected by this parameter, but by differences
in biodiesel chemical composition and properties. Note: dry = dry basis and DAF = dry ash-free basis.
1104 M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108

exhaust of a Cummins ISB 5.9 L turbodiesel engine by using 20% fuel, which promotes the carbon oxidation and therefore the rela-
soy biodiesel blended with ultralow sulfur diesel (ULSD) and tive ash percentage in the soot particles increases. The explanation
100% soy biodiesel, feature that gives to the primary particles an of the presence of the ash in biodiesel soot is given later based on
amorphous and disordered nanostructure, which facilitates the the EDX analysis.
oxidative attack. However, we believe that the differences of the Figure 1 shows a typical FT-IR spectrum of PM from a diesel
aforementioned work with our results are related to the sulfur con- engine operated with diesel. The PM samples of the tested biodiesel
tent of diesel, since in our case is around 225 ppm. In fact, in a pre- fuels showed a similar peak pattern, which indicates that the
vious paper, Boehman et al. [12] assessed diesel–biodiesel blends, chemical functional groups present in those particles are similar.
using low sulfur diesel (LSD 325 ppm sulfur) and ultralow sulfur The main characteristic signals observed in the spectra correspond
diesel (ULSD 15 ppm sulfur) and it is implied that the SOF content to AOH stretching mode of phenolic and acid groups as well as water
in the particulate matter generated in a Cummins ISB 5.9 L turbo- adsorbed on KBr powder, CAH stretching of aliphatic groups
diesel engine, decreases as the sulfur content in the fuel increases. (2975 cm 1, 2925 cm 1 and 2850 cm 1), C@O stretching of car-
So that, this fact along with others characteristics such as biodiesel bonyl-carboxylic groups (1720 cm 1), C@C stretching of aromatics
type and biodiesel esters composition, fuel oxygen content and en- (1600 cm 1), and SAO stretching mode of sulfate groups
gine operating conditions, can favor the oxidation process leading (1200 cm 1 and 1100 cm 1).
to a more carbonized material, as the one obtained here. In previous publications, the signal corresponding to the SAO
Similarly, comparison of the neat biodiesel series indicates that stretching mode was not evident in soot coming from hydrocar-
volatile matter content (or SOF) of particulate matter increases as bons flames (ethylene and benzene), since these fuels have not sul-
the unsaturation degree of methyl esters present in biodiesel in- fur in their composition [27]. Whereas particulate matter samples
creases according to the following order: BL100 > BPL50 > BP100. obtained in this work come from the combustion of diesel (CxHyS),
This result agrees with the theory, since the linseed biodiesel has which contains 225 ppm of sulfur, therefore, the presence of this
in its composition a high content of linoleic acid and linolenic acid, heteroatom in fuel and lubricating oil, promotes the increased in
with two and three unsaturations, respectively. Therefore, it is be- the intensity of the signals between 1200 cm 1 and 1100 cm 1.
lieved that the chemical decomposition of these compounds con- The same analysis can be done for PM samples of biodiesel, but
tributes in some extent to the formation of soot precursor in this case, the SAO signal comes only from the lubricating oil,
species. For instance, Kohse et al. [11] illustrated in their review since biodiesel does not contain sulfur in its composition.
the chemical decomposition path for ethyl propanoate and a surro- The presence of the CAH stretching in aliphatic groups comes
gate biodiesel containing methyl decanoate and methyl decenoate. mainly from methyl, methylene, and methine groups bonded to
The authors indicated that the main decomposition products con- aromatic rings in PAHs or methylene bridges (fluorine-type) main-
sist of oxygenated compounds (ketones, aldehydes, peroxides, cyc- taining the interconnection of PAHs within the network [27]. Some
lic ethers etc.) and hydrocarbon compounds (CH4, C2H4 and others other authors have proposed that the presence of an aliphatic sig-
C2AC6 hydrocarbons), fact that was also confirmed by using n- nal may also come from the unburned fuel condensed on the par-
hexadecane as surrogate rapeseed biodiesel fuel [26]. The presence ticle surface at low temperature [15,28,29]. Studies done in flames
of these hydrocarbons in the combustion zone is considered to be have also reported that the presence of aliphatic moieties in soot
the key for the polycyclic aromatic hydrocarbons (PAHs) forma- and soot precursors, especially at the soot inception point contrib-
tion, which subsequently are transformed into soot through isom- utes to the increases of mass growth process of soot just after the
erization reactions. However, it is important to highlight that the first particle has been formed [30,31].
mole fraction of these hydrocarbons is still lower than those found It is important to highlight that the lack of resolution or the
in comparable fuel-rich hydrocarbon flames, confirming a ten- overlapping of the CAH aromatic signal around 3050 cm 1 caused
dency of ester fuels to reduce soot precursors. by the broadening of the AOH stretching mode and the low
Other important feature observed in Table 5 is related to the ash intensity of the CAH out-of-plane vibrations for aromatics, usually
content of PM samples. The residue varied between 9% and 12%, observed for carbonaceous material between 700 and 900 cm 1,
being slightly larger for samples of neat biodiesel, a behavior that suggest that soot particles and their precursors (mainly poly-
can be explained by the increase of the oxygen content into the cyclic aromatic hydrocarbons-PAHs) are highly carbonized and

Fig. 1. Typical FTIR spectrum of PM from a diesel engine operated with diesel no. 2.
M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108 1105

composed of poly condensed aromatic units with low hydrogen increase in the unsaturation degree of the fuels. This result was
content at the periphery that are not visible by IR through the unexpected because of the well known decrease in soot formation
CAH stretching mode. However, the high intensity of C@C stretch- as in-cylinder temperature increases [35]. A possible explanation
ing mode of aromatic species observed at 1600 cm 1 supports this for this was presented by Gaïl et al. [36] and Sarathy et al. [17],
reasoning [27,32]. Others studies done using carbon nanotubes who showed through experiments and kinetic modeling of an
show that the infrared analysis of these materials corroborates that opposed flow diffusion flame and a jet stirred reactor, that con-
the larger the aromatic cluster, the lower the aromatic CAH out-of- centrations of soot precursors were much higher with unsaturated
plane vibration intensity becomes [33,34]. (MC) than with saturated (MB) pure methyl esters: pC3H4 (propyne),
Additionally, the prominent broad and intense band observed in aC3H4 (allene), 1,3-butadiene, and benzene. The use of model
the spectrum of Fig. 1 between 3600 cm 1 and 3000 cm 1 is as- compounds let them to identify the effect of the double bond as
signed to contributions from a variety of AOH stretching modes, the key reaction pathway that led to the unsaturated ester forming
including coupled absorption functionalities between acid and higher levels of soot precursors and unsaturated hydrocarbons,
alcoholic groups of many different chemical and carbon environ- when compared to the saturated ester. This result was also
ments. corroborated by Garner et al. [37] who observed that acetylene
Figure 2 part a) shows the infrared region corresponding to the yield produced in the pyrolysis of methyl octanoate and methyl
signals of the aliphatic groups present in the particulate matter octenoate increased with the number of carbon–carbon double
produced in the combustion of BP100, BL100 and BPL50 fuels. bonds.
For comparison purposes the spectra were normalized based on Figure 2 part b displays the infrared region between 700 and
the total area indicating that the fraction of aliphatic species pres- 2000 cm 1 for the PM evaluated in this study. It is worth while to ob-
ent in the soot particles is favored in the BL100 not only by its serve that the signal corresponding to C@C stretching vibration
chemical composition, but also by the increase in the unsaturation mode of alkenes or aromatic groups do not change significantly
level (C16:1, C18:2, C18:3) of the methyl esters [3], which favors among samples. Nevertheless the low intensity of the CAH out-of-
the precursors species formation in the combustion zone. plane bending mode signal (700 cm 1) of these groups indicates that
For instance, Sarathy et al. [17] used premixed and non-premixed the deposits are comprised of condensed polyaromatic hydrocar-
combustion experiments to investigate the effect of the chemical bons [27].
structure of saturated (methyl butanoate; CH3CH2CH2C(@O)OCH3) Finally, most of the oxygen content of PM samples is involved in
and unsaturated (methyl crotonate; CH3CH@CHC(@O)OCH3) fatty the sulfate groups as well as other metal oxide formation. The
acid methyl esters on the combustion species. The experimental re- spectra presented here do not show the presence of carbonyl
sults indicated that methyl crotonate combustion produced much stretching vibration (1720 cm 1), suggesting that under the condi-
higher levels of C2H2, 1-C3H4, 1-C4H8, and 1,3-C4H6 than methyl but- tions used in this study most of these groups evolved as CO2 or CO.
anoate. In opposed flow diffusion flames, the methyl crotonate also The corresponding SAO stretching mode signal of sulfate groups is
produced benzene while for methyl butanoate it was not detected. evident in all the samples in spite the fact that the neat biodiesel
All these species are considered soot precursors. The results also ex- does not contain sulfur in its composition. The origin of this signal
plain why the methyl ester (mostly saturated compounds) present is attributed to the sulfur present in the lubricant oil [28]. This is in
in BP100 and BPL50 disfavors the routes that lead to soot precursors, agreement with results reported in the scientific literature, where
a fact that is in agreement with the tendency observed in the nom- the emissions of sulfur by a series of biodiesels or their blends are
inal amount of PM collected in filters as well as with the smoke significantly smaller, which in turns favors the reduction of PM,
opacity reported in Table 4. The above results follow the increasing since the nucleation process by the sulfur mechanism is reduced
order: BL100 > BPL50 > BP100, which is the same sequence of the [8,28]. Up to now there is no clear explanation of the sulfur

Fig. 2. Variation in the FT-IR signals intensities in: (a) aliphatic and (b) fingerprint regions of the spectra of PM according to the biodiesel fuel used.
1106 M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108

reduction observed, but it is suspect that the high lubricity of bio- that the ID/IG ratio of the samples coming from neat biodiesels is
diesel can imply less consumption of lubricant oil [2]. lower than that obtained for the reference fuel (diesel). This result
Table 6 summarizes the elemental analysis results of the PM indicates that the PM coming from neat biodiesel is more struc-
samples obtained from neat biodiesel. In fact the interpretation tured (less amorfous) compared to the diesel soot.
given here can be extended to the composition of the residue ob- The nano-structural differences between diesel and biodiesel
tained by TGA. For comparison purposes the elemental analysis PM can be attributable to the extent of oxidation, the facility of cre-
of the reference material (particulate matter coming from diesel) ating oxygenated complexes at the surface of particles and the pres-
was introduced. It is worthwhile to mention that the data obtained ence of metallic species during combustion, being higher for
by EDX corresponds to a surface analysis and not the bulk. There- biodiesel than for diesel, since neat biodiesel posses 10% oxygen
fore, all of the interpretations based on this technique are merely as well as metallic residues in its composition. So that, the oxidation
qualitative. Besides the common elements present in organic struc- increase caused by biodiesel combustion at the operation mode
tures, C, O and H (the last one no identified by EDX), the elemental used this study facilitates the removal of the cross-link moieties,
composition of the particulate matter evaluated in this study, permitting the crystallites to realign themselves to energetically
shows the presence of S and Zn in the diesel and Ca, K, Si, Na, Fe, more favorable orientations [13,40].
Mg and P observed in the neat biodiesel. Seong et al. evaluated the impact of intake oxygen enrichment on
Some authors have reported that these elements may arise from oxidative reactivity and properties of diesel soot produced in a four-
the unwanted combustion of lubricants and their additives (e.g. cylinder turbo-charged common rail diesel engine operated at low
Zn). The source of other elements (e.g. Si) may be from inorganic and high load conditions. They found that the effect of oxygen
components or organic complexes present in biomass, these ele- enrichment in the crystalline structure of soot is higher when there
ments can be converted into aerosols during the combustion pro- is a high load on the engine. Although the surface oxygen content of
cesses. Alkali (e.g. Na) and alkaline earth (Ca) elements can come soot does not show a consistent trend at low and high loads with
from the transesterification process. The tendency observed for oxygen enrichment, the soot generated from high load contains
the oxygen content is not very conclusive since its variability some metallic species, which are attributed to the oxidation of lubri-
was very high among samples. As mentioned before, the sulfur that cating oil by a synergistic effect of high oxygen concentration and
appears in the soot comes from the lubricant oil and/or fuel, as in high temperature regimen, then, those metallic oxides can promote
the diesel case [15,28,29]. soot particle oxidation, inducing changes in the crystalline structure
RAMAN spectroscopy was used to get further information of the [40].
carbon structure. This method provides detailed data which allows On the other hand, upon comparing the PM coming from the
distinction of order–disorder effects on carbonaceous materials. three biodiesel fuels evaluated in this study, it is observed that the
Figure 3 shows the RAMAN spectrum of particulate matter sample ID/IG ratio does not change significantly from a statistical standpoint,
originating from a diesel system. indicating that the carbonization or carbon rearrangement of PM de-
The Raman spectra of carbonaceous material can be divided pends on the synergistic effect of high temperature regimen and oxi-
into areas of first-order (<1800 cm 1) and second-order (2200– dation process only, besides, the fuel oxygen content of biodiesels
3400 cm 1) peaks, the last one no shown in this paper. The bands was similar and does not depends on the unsaturation degree of
that appear at the first region are the G (graphite) band at around methyl esters present in biodiesel.
1580 cm 1 which is attributed to the natural sp2 vibration mode of
the six-member ring planes in the graphite-like structure, as well 3.3. Particulate matter morphology obtained by TEM
as the D (amorphous carbon) band around 1380 cm 1 (Fig. 3a),
whose origin can be explained as a manifestation of the in-plane The particulate matter formation processes and its morphology
vibrational mode at the surface of the sp2 domains [32]. The high have been extensively investigated and a wealth of information has
signal intensity between the two maxima peaks can be attributed been accumulated in the last two decades. It is well known that
to another band at 1500 cm 1, which has been designated the D3 particulate matter emitted from a diesel engine is usually observed
band in earlier studies of Cuesta et al. [38] and Jawhari et al. [39] as chain-like aggregates (secondary particles) composed of several
who suggested that the D3 band originates from the amorphous tens to hundreds of primary ‘‘spherical’’ particles [41].
carbon fraction of soot (organic molecules, fragments or functional TEM images of PM (Fig. 4) also confirm that they consist of
groups). The G-peak position of the soot is higher than that as- aggregated particles that form condensed structures. All of the
signed to graphitic materials, 1580 cm 1, providing evidence of images are at the same scale, with a 100 nm reference marker.
the presence of a disordered carbon fraction within the soot sam- The PM produced by biodiesel presents very well defined chain-
ple. The presence of the D-peak in all the spectra indicates that the like aggregates composed of nearly primary spherical particles,
symmetry of the graphene sheets is broken by lattice discontinu- partially covered with tarry material, which is always associated
ities or defects [32]. with soluble organic fraction (SOF) that evolves during the pyroly-
In this study, the integrated ID/IG ratio was used to investigate sis process in the temperature range between 120 and 500 °C, as
the graphite-like structure of soot materials. As this ratio decreases observed by TGA.
the higher the graphitic-like structure becomes. Figure 3b shows Aggadi et al. [29] observed that soot particles composed mainly
of graphitic phases can coexist with amorphous states where PAHs
Table 6 and aliphatic units may be present. Under the conditions used in
Elemental composition of PM obtained at the exhaust of a diesel engine obtained by
this study, the primary particle size does not significantly change
EDX.
among the series of neat biodiesel (22.0 nm average, see Table 7),
Fuel Elements although its general tendency seems to indicate that biodiesel fuel
C O S Zn Ca K Si Na Fe Mg P can produce up to 11% particle size reduction (BL100 y BPL50)
Diesel 76.4 18.9 4.4 0.4 ND ND ND ND ND ND ND compared to diesel, since biodiesel can increase the reactivity to-
no. 2 ward oxidation, causing a shrinkage of particles. This tendency cor-
BP100 79.5 18.6 0.5 0.8 ND ND ND ND ND ND ND roborates what has been observed by RAMAN, since PM coming
BPL50 80.5 16.4 1.3 ND 0.6 0.2 0.3 0.5 ND ND ND from neat biodiesel has a more defined crystalline structure than
BL100 78.8 17.4 0.6 0.4 1.0 0.2 ND 0.7 0.4 0.6 0.6
diesel soot and the slightly particle shrinkage observed in samples
ND: not detected (0.2%). does not depends on the unsaturation degree of biodiesel fuel.
M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108 1107

Fig. 3. Typical Raman spectrum from PM diesel (a) and Id/IG ratio of samples (b).

Diesel BPL50

100 nm 100 nm

BP100 BL100

100 nm 100 nm

Fig. 4. TEM images of PM originating from diesel and biodiesel.

Some authors did not find any significant effect of biodiesel on the 4. Conclusions
particle sizes. Turrio-Baldassarri et al. [42] evaluated rapeseed oil
in 20% blends with diesel fuel in a 6-cylinder engine, and con- Biodiesel emissions were cleaner than regular diesel emissions
cluded from their TEM analysis that both particle size and mor- due to a substantial decrease in HC and nominal PM mass emission
phology remained within the same range for all fuels. Chen and caused by the increase of oxygen content in the fuel, low sulfur con-
Wu [23] tested a single-cylinder engine under three steady modes tent and the absence of aromatic species. However, it is believed that
of operation and found no significant differences in the mean the chemical composition of the methyl ester present in biodiesel
diameters obtained with SMPS between diesel and soybean-oil can play an important role in the chemical characteristics of HC
biodiesel fuels, although they did find decreases in the mass (24– and PM emitted upon comparing through a series of neat biodiesel.
42%) and number (40–49%) of emitted particles. From the experimental work carried out to investigate the effect of
the degree of unsaturation of biodiesel fuels on combustion and
Table 7 chemical characteristics of PM emission, the following conclusions
TEM results for the primary particles obtained from
can be drawn:
biodiesel.

Fuel Average (nm) (1) It was found that the smoke opacity, a property related with
Diesel no. 2 24.0 ± 6.0 the PM emission increased among the biodiesel fuels accord-
BP100 23.9 ± 5.8 ing to the degree of unsaturation, result that also agrees with
BPL50 21.3 ± 5.1 the mass of particulate matter collected in the filters. Also,
BL100 21.4 ± 6.1
properties such as density, in-cylinder temperature and
1108 M. Salamanca et al. / Combustion and Flame 159 (2012) 1100–1108

cetane number were found to be sensitive to the unsatura- [4] M. Lapuerta, O. Armas, J. Rodriguez-Fernandez, Prog. Energy Combust. Sci. 34
(2008) 198–223.
tion degree of biodiesel, leading to an increase in the first
[5] R. Awang, C.Y. May, Fuel Chem. Div. Prepr. 47 (2002) 618–621.
two proprieties and a reduction in the last one. [6] A.F. Sarofim, J.S. Lighty, E.G. Eddings, Fuel Chem. Div. Prepr. 47 (2002) 618–
(2) It was found based on filter collection that linseed biodiesel 621.
(BL100) produces 60% more PM than palm biodiesel (BP100), [7] E. Sendzikiene, V. Makareviciene, P. Janulis, Pol. J. Environ. Stud. 16 (2007)
259–265.
since linseed biodiesel has more unsaturated compounds [8] M. Lapuerta, J. Rodrıguez-Fernandez, J.R. Agudelo, Bioresour. Technol. 99
(17.1% C18:2 and 53% C18:3) in its composition than palm, (2008) 731–740.
which could favors the soot precursor’s formation in the [9] P.J.M. Frijters, R.S.G. Baert, Int. J. Vehicle Des. 41 (2006) 242–255.
[10] K. Schmidt, J.H. Van Gerpen, The Effect of Biodiesel Fuel Composition on Diesel
combustion chamber. Combustion and Emissions SAE Paper, 961086, 1996.
(3) It was seen that volatile matter content of particulate matter [11] K. Kohse, P. Obwald, T.A. Cool, T. Kasper, N. Hansen, F. Qi, C.K. Westbrook, P.R.
increases as the unsaturation degree of methyl esters pres- Westmoreland, Angew. Chem. Int. Edit. 49 (2010) 3572–3597.
[12] A.L. Boehman, J. Song, M. Alam, Energy Fuels 19 (2005) 1857–1864.
ent in biodiesel increases according to the following order: [13] J. Song, M. Alam, A.L. Boehman, U. Kim, Combust. Flame 146 (2006) 589–604.
BL100 > BPL50 > BP100, due to the high content of linoleic [14] M. Lapuerta, O. Armas, J. Rodríguez-Fernández, Effect of the Degree of
acid and linolenic acid in linseed, with two and three unsat- Unsaturation of Biodiesel Fuels on NOx and Particulate Emissions, SAE paper
No. 2008-01-1676, 2008.
urations, respectively. Therefore, the chemical decomposi- [15] A. Schönborn, N. Ladommatos, J. Williams, R. Allan, J. Rogerson, Combust.
tion of these compounds may contribute in some extent to Flame 156 (2009) 1396–1412.
the formation of soot precursor species. [16] R.L. McCormick, M.S. Graboski, T.L. Alleman, A.M. Herring, K.S. Tyson, Environ.
Sci. Technol. 35 (9) (2001) 1742–1747.
(4) It was observed that the aliphatic content of PM produced by
[17] S.M. Sarathy, S. Gail, S.A. Syed, M.J. Thomson, P. Dagaut, Proc. Combust. Inst. 31
BL100 was twice the one obtained from BP100 and BPL50, (2007) 1015–1022.
result that is in good agreement with the increase in the [18] S. Puhan, N. Saravanan, G. Nagarajan, N. Vedaraman, Biomass Bioenergy 34
degree of unsaturation of the methyl esters present in this (2010) 1079–1088.
[19] P. Benjumea, J.R. Agudelo, A.F. Agudelo, Energy Fuels 25 (1) (2011) 77–85.
biodiesel (C16:1, C18:2, C18:3), which favors the formation [20] J. Agudelo, R. Moreno, J. Pérez, Revista Facultad de Ingeniería 51 (2009) 79–87.
of precursors species of the C2, C3 and C4 type, as well as [21] P. Benjumea, J.R. Agudelo, A.F. Agudelo, Fuel 88 (2009) 725–731.
benzene. [22] A. Sadezky, H. Muckenhuber, H. Grothe, R. Niessner, U. Poschl, Carbon 43
(2005) 1731–1742.
(5) From a morphological point of view it was observed that [23] C.H. Luo, W.M. Lee, J.J. Liaw, J. Environ. Sci. 21 (2009) 452–457.
neat biodiesel does not affect significantly the size of the [24] U.S.E.P.A. Draft Technical Report EPA420-P-02-001, 2002.
PM emitted. However, soot particles produced by neat bio- [25] A. Williams, R.L. McCormick, R.R. Hayes, J. Ireland, H.L. Fang, Effect of Biodiesel
Blends on Diesel Particulate Filter Performance, NREL/CP-540-40015,
diesel are more carbonized than diesel particles, according Presented at the Powertrain and Fluid Systems, Conference and Exhibition
to ID/IG ratio, suggesting that possible differences between Toronto Canada, October 2006.
them are caused by an increase in the oxidation rate, instead [26] P. Dagauta, S. Gaıl, M. Sahasrabudhe, Proc. Combust. Inst. 31 (2007) 2955–
2961.
of the unsaturation degree of biodiesel, result that agrees [27] A. Santamaria, N. Yang, E.G. Eddings, F. Mondragon, Combust. Flame 157
with the TGA and Raman results. (2010) 33–42.
[28] R.L. Vander Wal, V.M. Bryg, M.D. Hays, Aerosol Sci. 41 (2010) 108–117.
[29] N. Aggadi, C. Arnas, F. Bénédic, C. Dominique, X. Duten, F. Silva, K. Hassouni,
D.M. Gruen, Diamond Relat. Mater. 15 (2006) 908–912.
Acknowledgments [30] A. Ciajolo, R. Barbella, A. Tregrossi, L. Bonfanti, Combust. Sci. Technol. 153
(2000) 19–32.
Authors wish to thank the Colombian Ministry of Agriculture [31] B. Öktem, M.P. Tolocka, B. Zhao, H. Wang, M.V. Johnston, Combust. Flame 142
(2005) 364–373.
and Rural Development and the Environmental authority from [32] K. Al-Qurashi, A.L. Boehman, Combust. Flame 155 (2008) 675–695.
Medellín (Área Metropolitana del Valle de aburrá) for the financial [33] U.J. Kim, C.A. Furtado, X. Liu, G. Chen, P.C. Eklund, J. Am. Chem. Soc. 127 (2005)
support of project No. 003 2007D3608-67-07 (Biodiesel project). 15437–15445.
[34] Y.C. Hsieh, Y.C. Chou, C.P. Lin, T.F. Hsieh, C.M. Shu, Aerosol Air Qual. Res. 10
The authors also would like to thank the Sostenibilidad program (2010) 212–218.
2011–2012 of the University of Antioquia for financial support. [35] J.B. Heywood, Internal Combustion Engine Fundamentals, McGraw-Hill, 1988.
M.S. thanks COLCIENCIAS and the University of Antioquia for the [36] S. Gaïl, S.M. Sarathy, M.J. Thomson, P. Diévart, P. Dagaut, Combust. Flame 155
(2008) 635–650.
PhD scholarship. [37] S. Garner, R. Sivaramakrishnan, K. Brezinsky, Proc. Combust. Inst. 32 (1) (2009)
461–467.
References [38] A. Cuesta, P. Dhamelincourt, J. Laureyns, A. Martinez-Alonso, J.M.D. Tascon,
Carbon 32 (1994) 1523–1532.
[39] T. Jawhari, A. Roid, J. Casado, Carbon 33 (1995) 1561–1565.
[1] WCO, Draft Report of the 36th Session of the Harmonized System Review Sub-
[40] H.J. Seong, A.L. Boehman, Energy Fuels 25 (2011) 602–616.
committee. World Customs Organization, Brussels, Belgium. NR0722E1A (RSC/
[41] T. Ishiguro, Y. Takatori, K. Akihama, Combust. Flame 108 (1997) 231–234.
36/nov. 2007), 2007.
[42] L. Turrio-Baldassarri, C.L. Battistelli, L. Conti, R. Crebelli, B. DeBerardis, A.L.
[2] A.A. Refaat, Int. J. Environ. Sci. Technol. 4 (2009) 677–694.
Iamiceli, Sci. Total Environ. 327 (2004) 147–162.
[3] G. Knothe, C.A. Sharp, T.W. Ryan, Energy Fuels 20 (2006) 403–408.

S-ar putea să vă placă și