Jatropha-Soybean Biodiesel Blends: A Signature Design with

advertisement
December 2013, Volume 4, No.6
International Journal of Chemical and Environmental Engineering
Jatropha- Soybean Biodiesel Blends: A Signature
Design with Optimum Low Temperature
Properties and Oxidation Stability
Kuhanesapathy Thavaras Pathy*, Suzana Yusup, Umer Rashid
Chemical Engineering Department, Universiti Teknologi PETRONA, Tronoh Perak, Malaysia
Corresponding Author E-mail:
kuhanesapathy.tp@petronas.com.my
*
Abstract:
With growing concerns brought by excessive usage and depletion of fossil fuels, the race for finding alternative energy as their
substitutes is getting heated up. In the midst of this, biodiesel has gained popularity as a renewable, environmental friendly alternative.
Exploration of non-edible seed oils as biodiesel raw materials seem feasible without significantly affecting the global food economy.
In the present study, blends of Jatropha and Soybean biodiesel have been explored to determine the optimum mix, achieving better
low temperature properties. A two-step process consisting of pre-esterification and transesterification was developed to produce
biodiesel from crude Jatropha curcas L. oil and Soybean. The process was carried out at optimized set of conditions: methanol/oil
molar ratio (6:1), sodium methoxide catalyst concentration (1.00 wt%), temperature (65oC) and mixing intensity (600 rpm). The
physicochemical properties showed that the methyl esters contained low moisture level (<150 ppm) and acid value (<0.2 mg KOH/
mg-oil) respectively. Characterization of the fatty acid methyl esters (FAMEs) was accomplished by gas-chromatography. The fuel
properties of the biodiesel blends produced were found to be within the standards specifications of ASTM D 6751 and EN 14214.
Keywords: Biodiesel Blends, Transesterification, Low Temperature Properties, Oxidation Stability
1. Introduction
The world is confronted with the twin crises of fossil fuel
exhaustion and environmental dilapidation. The
indiscriminate extraction and usage of fossil fuels have
led to a cutback in petroleum reserves [1]. Known crude
oil reserves could be dwindling in less than 50 years at the
current rate of usage. Thus, increased environmental
concerns, tougher clean air act standards require the
search for feasible alternative fuels, which are
environmental friendly [2]. Even though vegetative oils
can be fuel for diesel engines, but their high viscosities,
low volatilities and poor cold flow properties have
directed to the study of its myriad of derivatives [1]. With
the escalating need for fuel, biodiesel has gained
popularity and attractiveness as an alternative,
environmentally friendly energy resource [3]. The
attractive attributes of biodiesel fuel are: (i) it is plantderived, not petroleum-derived, and as such its
combustion does not increase current net atmospheric
levels of CO2, a “greenhouse” gas; (ii) it can be locally
produced, offering the option of reducing petroleum
imports; (iii) it is biodegradable; (iv) relative to
conventional diesel fuel, its combustion products have
reduced levels of particulates, carbon monoxide, sulphur
oxides, hydrocarbons, soot and under some conditions,
nitrogen oxides [4].
Biodiesel is an alternative fuel produced from renewable
vegetable oils, animal fats or recycled cooking oils by
transesterification reaction. Vegetable oils can be edible
such as cottonseed, groundnut, corn, rapeseed, soybean,
palm oil, sunflower, peanut, coconut, etc. and non-edible
such as Jatropha, Pongamia, neem, rubber seed, mahua,
silk cotton tree, jojoba, and castor oil [5]. However,
continuous and large scale production of biodiesel from
edible oil without appropriate planning may lead to
negative impact to the world, such as depletion of food
supply resulting in economic imbalance. In other words,
biodiesel is competing limited land availability with food
industry for plantation of oil crop [6]. On the other hand,
the major economic aspect to think about for input costs
of biodiesel production is the feedstock, which is about
80% of the total operating cost [7]. To overcome this
problem, non-edible oils and their methyl esters are
produced for blended biodiesel. This would not only
provide low-cost feedstock, but at the same time would
also not significantly affect the global food economy. As
such, methyl esters derived from edible oil and non-edible
oil are blended to get an optimum mix. The research on
biodiesel oxidation stability and low temperature
properties have been put on top priority with soybean as
the edible oil and Jatropha as the non-edible oil.
Soybeans flower produce 60-80 pods, each holding three
pea sized beans. In processing, soybeans are cleaned,
cracked, dehulled and rolled into flakes. This breaks the
Jatropha- Soybean Biodiesel Blends: A Signature Design with Optimum Low Temperature Properties and Oxidation Stability
oil cells for efficient extraction to separate the oil and
meal component. After producing the soybean oil, the
remaining flakes can then be processed into many types
of edible soy protein products, or used to make soybean
meal. [8]. On the other hand, Jatropha is a nut belonging
to the Euphorbiaceae family. It is grown in central and
South America, South East Asia, India and Africa [9]. In
Malaysia, wild Jatropha tree is also known as Jarak
Pagar particularly in Peninsular Malaysia area. Jatropha
can be cultivated well under adverse climatic because of
its low moisture demands, fertility requirements and
tolerance to high temperatures [10]. Each fruit contains 23 oblong black seeds which can produce oil. Jatropha
seeds are obtained, where the ripe seeds are collected and
the damaged seeds are removed. The seeds are cleaned,
de-shelled and dried in an oven. The seeds are ground to
powder using a grinder before extracting the oil. The
extraction of oil is then carried out by soxhlet extraction
using hexane [11]. The major toxic compounds in
Jatropha plants are curcin and purgative which are
dominantly found in seeds, fruits and sap [12]. This
makes it a non-edible vegetable oil. Biodiesel consists of
mono-alkyl esters of long chain fatty acid produced by
transesterification of vegetable oil or animal fat with
alcohol.
In the transesterification of vegetable oils, a triglyceride
reacts with an alcohol in the presence of a strong acid or
base, yielding a mixture of fatty acids alkyl esters and
glycerol as the by-product. The entire process is a
sequence of three successive and reversible reactions, in
which di- and monoglycerides are formed as
intermediates [14]. The stoichiometric reaction requires 1
mol of a triglyceride and 3 mol of alcohol. However, an
excess of alcohol is used to increase the yields of the alkyl
esters and to permit effective phase separation from the
glycerol formed. Figure 1 below details the
transesterification of vegetable oils [13].
The effect of acidity of the oil is vital in the studies of
biodiesel production. It was noted that the amount of
catalyst used depended very much on the acidity or
percentage of free fatty acids (FFA) in the oil itself. The
basic catalysts become ineffective partly because they
were consumed by neutralization [15]. A high FFA also
deactivates alkaline catalysts, and the addition of excess
amount of alkaline as compensation leads to the
formation of emulsion and gels, which in return increases
viscosity and causes loss to ester yield [16]. The FFA will
react with alkaline catalyst to produce soap that prevents
an effective separation of ester and glycerol. A two-step
transesterification is reported by several researchers as the
best method in the biodiesel production from non-edible
oil, which usually contains high FFA. At the preceding
step, the FFA content of oil is reduced by acid-catalyzed
esterification process; meanwhile, at the following step,
an alkaline-catalyzed transesterification is used to transform
oil and methanol to methyl esters and glycerol [6].
Low temperature properties, another vital attribute of
biodiesel are defined by the changes of biodiesel
properties such as crystallization, gelling or viscosity
increase due to temperature fluctuations that might
unfavourably impact the operability of the vehicles. These
properties are expressed by the values of cloud point and
pour point. These properties will be the vital factor to
determine if the biodiesel produced can be put into
operations in cold climate countries. This is significant
because at present, the largest demand of biodiesel is in
the European countries. The low temperature properties
of biodiesel are influenced by the types of fatty acids
present in the feedstock oil [6].
The effect of acidity of the oil is vital in the studies of
biodiesel production. It was noted that the amount of
catalyst used depended very much on the acidity or
percentage of the free fatty acids (FFA) in the oil itself.
Figure 1: Transesterification of vegetable oils.
The basic catalysts become ineffective partly because
they were consumed by neutralization [15]. A high FFA
also deactivates alkaline catalysts, and the addition of
excess amount of alkaline as compensation leads to the
formation of emulsion and gels, which in return increases
viscosity and causes loss to ester yield [16]. The FFA will
react with alkaline catalyst to produce soap that prevents
an effective separation of ester and glycerol. A two-step
transesterification is reported by several researchers as the
best method in the biodiesel production from non-edible
oil, which usually contains high FFA. At the preceding
step, the FFA content of oil is reduced by acid-catalyzed
esterification process; meanwhile, at the following step,
an alkaline-catalyzed transesterification is used to transform oil
and methanol to methyl esters and glycerol [6].
Low temperature properties, another vital attribute of
biodiesel are defined by the changes of biodiesel
properties such as crystallization, gelling or viscosity
increase due to temperature fluctuations that might
unfavourably impact the operability of the vehicles. These
properties are expressed by the values of cloud point and
pour point. These properties will be the vital factor to
determine if the biodiesel produced can be put into
operations in cold climate countries. This is significant
because at present, the largest demand of biodiesel is in
the European countries. The low temperature properties
of biodiesel are influenced by the types of fatty acids
present in the feedstock oil [6].
Eventually, with the execution of biodiesel as a substitute
fuel for petroleum-derived diesel oil, this may cause
depletion of edible-oil supply worldwide. As such, if the
price of palm oil continues to rise higher, consumers who
use palm oil as a source of edible oil will object [6]. On
the other hand, the major economic aspect to think about
354
Jatropha- Soybean Biodiesel Blends: A Signature Design with Optimum Low Temperature Properties and Oxidation Stability
for input costs of biodiesel production is the feedstock,
which is about 80% of the total operating cost [7]. To
overcome this problem, non-edible oils and their methyl
esters are produced for blended biodiesel. This would not
only provide low-cost feedstock, but at the same time
would also not significantly affect the global food
economy. As such, methyl esters derived from edible oil
and non-edible oil are blended to get an optimum mix of
them. The research on biodiesel oxidation stability and
low temperature properties have been put on top priority
with soybean as the edible oil and Jatropha as the nonedible oil.
The objective of the current study was to evaluate pure
biodiesel, soybean oil methyl ester (SOME) and Jatropha
oil methyl ester (JOME) along with its blends, at
compositions, (SOME:JOME,20:80), (SOME:JOME,
40:60), (SOME:JOME,60:40), (SOME:JOME,80:20) as
potential biodiesel fuel. Using standard methods low
temperature properties, oxidation stability, water content
and density were determined. It was further compared to
ASTM D6751 and EN 14214 standards to produce a
signature mix.
2. Methodology
Materials: Refined soybean oil (SBO) and crude Jatropha
oil (CJO) were purchased from Socma Trading (M) Sdn.
Bhd. and Bionas, Malaysia respectively. Methanol,
sodium hydroxide, potassium hydroxide, sodium
methoxide, anhydrous sodium sulfate, isopropanol,
toluene and phenolphthalein were purchased from Merck
(Darmstadt, Germany). All the chemicals used were
analytical reagent grade.
Determination of Acid Value: The acid value (AV) has an
impact on fuel aging. Besides, the alkaline-catalyzed
reaction is very sensitive to the content of FFA, which
should not exceed a certain limit recommended to avoid
deactivation of catalyst, formation of soaps and emulsion
[5]. Acid values (AV) of vegetable oil were determined
according to American Oil Chemists’ Society (AOCS)
Method Cd 3d-63.
Pre-treatment and Transesterification
I. Pre-esterification procedure for non-edible oil
Acid-catalyzed pre-treatment of CJO with an initial acid
value (AV) of 24.02 mg KOH g-oil-1 was accomplished in
a 500 mL three-necked round bottom flask connected to a
reflux condenser and a mechanical stirrer set to 1200 rpm,
and heated with a water bath to control the reaction
temperature. Initially, CJO (~188.00 g, 250 mL) and
methanol (88 mL, 35 vol%) were prepared in a flask,
followed by drop-wise addition of sulphuric acid (conc.,
1.0 vol%) [19]. The contents were heated at reflux for 4
hours. Upon cooling to room temperature, the phases
were separated. The esterification products were
separated in a separating funnel to obtain the upper oil
layer, which was then washed with ionized water several
times until the pH of washing water was close to 7.0. The
resultant pre-esterified oil was dried by anhydrous sodium
sulphate before subsequent transesterification
II. Transesterification procedure for edible and nonedible oils
200g of pre-treated Jatropha oil was taken in a threenecked round-bottomed flask. A water-cooled condenser
and a thermometer with cork are connected to the top and
side of the round- bottomed flask, respectively.
Meanwhile, the oil was warmed by placing the roundbottomed flask in the water bath maintained at a
temperature of 60oC. The required amount of methanol
(6:1 mol ratio) and catalyst, sodium methoxide, NaOCH 3
(1.0 wt%) were added into the oil for vigorous mixing by
means of a mechanical stirrer fixed into the flask. The
required temperature is maintained throughout the
reaction time (90 min) and the reacted mixture is then,
poured into a separating funnel. The mixture is allowed to
separate and settle overnight by gravity settling into a
clear, golden liquid biodiesel on the top with the light
brown glycerol at the bottom. The next day, the glycerol
is drained off from the separating funnel, leaving the
biodiesel/ester at the top [20]. Of the two separated
phases; the upper phase consist of methyl esters with
small amount of impurities such as residual alcohol,
glycerol and partial glycerides, while the lower is the
glycerol. The upper methyl esters layer collected is
further purified by distilling residual methanol at 800C
(external bath temperature). Some traces of impurities
such as remaining catalyst, residual methanol and
glycerol are removed by consecutive rinses with distilled
water. Residual water is then removed by drying esters
with anhydrous sodium sulphate, Na2SO4, followed by
filtration [21]. These procedures were repeated for
soybean oil.
Product analysis
Analysis of biodiesel (fatty acid methyl esters, FAME) for
fatty acid profile determination was performed with a
7890A gas chromatograph (Agilent Technology Inc.,
USA), equipped with a BP-X5 (SGE) capillary column
(30 m x 0.25 mm x 0.25 mm) and a flame ionization
detector. The column temperature was programmed at:
initial temperature 100oC, kept 3 mins, and then at a rate
of 20oC/min up to 170oC, kept 10 mins, followed up to
260oC. Helium was used as the carrier gas. The injector
temperature was programmed at 280oC. Approximately
0.1 mL of the sample was diluted in 1 mL of solvent
(hexane) and a sample volume of 0.1 µL was injected into
the column. The identification of the peaks was archived
by retention times by means of comparing them with
authentic standards analyzed under the same conditions.
Soybean oil methyl ester (SOME) and Jatropha oil methyl
ester (JOME) were examined, respectively.
Product characterization and properties of biodiesel
I. Density
The density of the methyl ester was calculated using a
gay-lussac pycnometer (25 mL) at an ambient
temperature, 20oC.
355
Jatropha- Soybean Biodiesel Blends: A Signature Design with Optimum Low Temperature Properties and Oxidation Stability
Table 1- Fatty acid profiles of Jatropha and soybean methyl esters.
II. Moisture Content
The moisture content of the methyl ester was estimated
following ASTM D95 using Mettler Toledo DL 39 (Karl
Fischer Coulometer). SOME and JOME were tested,
respectively.
Fatty Acid
Jatropha
Soybean
Myristic (C14/0)
-
0.06
0.04
Palmitic (C16/0)
16.31
16.42
8.43
Palmitoleic (C16/1)
0.89
1.20
0.06
-
0.12
0.06
Stearic (C18/0)
9.89
9.30
3.31
Oleic (C18/1)
37.33
38.46
31.70
Linoleic (C18/2)
35.58
34.09
54.94
Linolenic (C18/3)
-
-
0.67
Arachidic (C20/0)
-
0.26
0.24
Behenic (C22/0)
-
0.10
0.26
Lignoceric (C24/0)
-
-
0.08
3. Results and Discussion
Gondoic
-
-
0.21
3.1 Quality of Biodiesel Produced
Saturated
26.20
26.16
12.42
Table 1 depicts the fatty acid (FA) compositions of
SOME and JOME. Linoleic, oleic, palmitic and stearic
acids were the main components of SOME with
contribution present at 54.94, 31.70, 8.43 and 3.31%,
respectively. Minor constituents included linolenic
(0.67%), behenic (0.26%), arachidic (0.24%), gondoic
(0.21%), lignoceric (0.08%), palmitoleic (0.06%),
margaric (0.06%) and myristic acid (0.04%) also
identified. Soybean oil methyl ester was characterized by
a high percentage of unsaturated fatty acids (87.58%),
followed by saturated fatty acids (12.42%) constituting
the remaining. The fatty acid profiles of SOME in the
present study were in close agreement with that of other
profiles of SOME [19].
The major fatty acids in JOME were oleic, linoleic,
palmitic and stearic fatty acid. Oleic acid showed the
highest percentage of composition of 37.33% followed by
linoleic acid with 35.58%. Palmitic and stearic
contributed 16.31% and 9.89%, respectively. Other fatty
acid composition included palmitoleic (16.31%). The
content of unsaturated fatty acids was 73.8%, with
saturated fatty acids (26.2%) constituting the remaining.
These results are in accordance with previous report on
the FA profile of JOME [22]. In general, JOME contained
lesser unsaturated fatty acids (73.8%) compared to SOME
(87.58%).
Other properties that have been evaluated include the acid
value of oil. It has been reported [23] that
transesterification would not occur effectively if FFA
content in the oil was above 3%. Hence, the pre-treatment
of Jatropha oil has yielded an AV of 0.2525 mg KOH /goil. Besides, soybean oil recorded an AV of 0.0505 mg
KOH /g-oil. , on the other hand were reported as 870.8
and 875.9 kg/m3 for JOME and SOME, respectively.
Monounsaturated
38.22
39.75
31.97
Polyunsaturated
35.58
34.09
55.60
Margaric (C17/0)
III. Pour and Cloud Point
The cloud point and pour point were measured as per
ASTM D97 and D2500 test methods, respectively.
IV. Oxidation Stability
The oxidation stability was quantified by the induction
period (IP) of neat methyl esters and its blends. The IP
was evaluated according to the Rancimat method EN
14112 for biodiesel.
According to ASTM D6751 test methods, maximum
allowable limit for water content is 500 ppm. This study
obtained 148.1 ppm for JOME and 139.4 for SOME,
respectively.
3.2 Low Temperature Operability
At low temperatures, biodiesel is deemed less appropriate
for usage as compared to petro-diesel. Static tests are
carried out to identify first wax and non-flow
temperatures for the fuel. The low temperature properties
were evaluated in terms of cloud point (CP) and pour
point (PP) [21]. As seen in Figure 2, SOME provided CP
and PP values of 1oC and -3oC, respectively. The values
obtained for JOME were 4oC (CP) and 3oC (PP),
respectively. It is notable that the results obtained for
SOME in previous reports [19, 24] are almost comparable
with the current report. For comparison, a previous study
reported the low temperature properties for Jatropha, the
CP as 4oC [25] and PP as 2oC [21] closely conforming to
the current study. The blends of biodiesel were examined
to study the effect of both edible and non-edible oil
methyl esters on each other. As per results shown in
Figure 2, the blending of 20% soybean (edible oil) has
significantly reduced the cloud point of JOME to 3.6 oC
and pour point to 0.0oC. The increasing fraction of edible
oil methyl ester (SOME) in the blends has gradually
reduced the overall low temperature properties. It is
proven that soybean oil methyl esters exhibit better low
temperature properties as compared to Jatropha. With the
blending of soybean biodiesel in Jatropha biodiesel, we
could exhibit improved cloud and pour point properties.
Though blending of soybean biodiesel with Jatropha
biodiesel worsen the low temperature properties of
soybean biodiesel, the results are of least concern in the
356
Jatropha- Soybean Biodiesel Blends: A Signature Design with Optimum Low Temperature Properties and Oxidation Stability
temperate climate of Asia. Based on Malaysian Standard
for Diesel Fuel (MS123:1993), the recorded maximum
CP and PP are 18oC and 15oC, respectively [26]. Thus,
results achieved in this report conform to the standard.
The enhanced low temperature properties of SOME
versus JOME may be attributed to its lower content of
saturated fatty acids, 12.42% and 26.20%, respectively.
Based on previous reports, it has been determined that
small levels of saturated FAME have an inconsistent
influence on low temperature properties of biodiesel as a
result of their higher melting points [27, 28, 29]. Higher
content of polyunsaturated fatty acids also contributes
considerably to the increasing number of double bond
content. It should be noted that melting point decreases
significantly in the presence of double bond, as indicated
quality assurance, neat biodiesel has to be stored well to
maintain its properties and stability. The Rancimat
method (EN 14112) is listed as the oxidative stability
specification in ASTM D6751 and EN 14214. A
minimum induction period (IP) of 3 h is required for
ASTM D6751; whereas a more severe limit of 6 h or
greater is denoted in EN 14214 [19].
It was observed from Figure 3, that as the proportion of
Jatropha increases, oxidation stability also increases. For
neat biodiesel, soybean and Jatropha recorded an IP of
4.55 h and 7.63 h, respectively. It should be noted that
soybean yielded an IP of 5.00 h [19] and 3.80 h [19] in
previous studies. So, the IP obtained in the present study
is within the range. As for Jatropha, the IP obtained in the
current test is close to the results obtained in other study
[3] with an IP of 8 h. The blends were tested at
proportions of (SOME:JOME,20:80), (SOME:JOME,
40:60), (SOME:JOME::60:40), (SOME:JOME::80:20),
each yielding an IP of 7.05 h, 5.83 h, 5.41 h and 5.01 h,
respectively. Meeting the more stringent specification,
ASTM D6751, only blends of 80% of Jatropha and 20%
of soybean conforms.
The cause for good stability is associated to the resistance
to auto-oxidation, in the existence of high fraction of
saturated fatty acids. Unsaturated and polyunsaturated
fatty acids are significantly more reactive to oxidation
than saturated compounds. The reason for this is that the
unsaturated fatty acid chains contain the most reactive
sites which are particularly vulnerable to the free radical
attack [15]. The rate of auto-oxidation of fatty acid
methyl esters (FAME) is influenced by the presence of
double bonds that are separated by allylic methylene
positions (AMP) with bis-allylic methylene positions
(BAMP) being even more prone to oxidation. The reason
polyunsaturated fatty acids are particularly vulnerable to
auto-oxidation is due to the presence of BAMP [19]. In
this report, JOME displays (Refer Table 1) a lower
content (35.58%) of polyunsaturated compounds as
compared to SOME (55.60%). So, it is evident that JOME
would exhibit a better oxidative stability as compared to
SOME, and blends with higher
Figure 2: Low temperature properties for pure and blends of
biodiesel.
Figure 3: Comparison with blended biodiesel and commercial B5
by the melting point of methyl esters of stearic (C18:0;
37.7oC), oleic (C18:1; -20.2oC), linoleic (C18:2; -43.1oC),
and linolenic (C18:3; -57oC) acids [30]. Further proven by,
the fact that SOME contains higher amount of
polyunsaturated FAME (55.60%) as compared to JOME
(35.58%).
3.3 Oxidation Stability
Oxidation stability is a matter of interest as it cannot be
kept beyond a period. For best usage, standardization and
Figure 4: Oxidation stability for neat and blends of biodiesel
357
Jatropha- Soybean Biodiesel Blends: A Signature Design with Optimum Low Temperature Properties and Oxidation Stability
[6]
Gui, M. M., Lee, K., & Bhatia, S. (2008). Feasibility of edible oil
vs. non-edible oil vs. waste edible oil as biodiesel feedstock.
Energy , 33, 1646-1653.
[7]
Balat, M., & Balat, H. (2008). A critical review of bio-diesel as a
vehicular fuel. Energy Conversion and Management , 49, 2727-2741.
[8]
The American Soybean Association. (n.d.). Soy Stats 2011.
(United
Soybean
Board)
Retrieved
2011,
from
http://www.soystats.com/2011/Default-frames.htm
[9]
G. M., G., M., M., & M., T. (1996). Exploitation of the tropical oil
seed plant Jatropha Curcas L. International of Biosource
Technology , 58, 77-82.
[10] N., K., K., K., S., K., N., K., & S., R. (2007). Genetic variability
and divergence studies in seed traits and oil content of Jatropha
(Jatropha curcas L.) accessions. Biomass & Bioenergy , 31, 497-502.
[11] Salimon, J., & Abdullah, R. (2008). Physicochemical Properties of
Malaysian Jatropha Curcas Seed Oil. Sains Malaysiana , 37 (4), 379-382.
Figure 5: Comparison with blended biodiesel and Commercial B5
proportion of JOME would improve the fuel properties.
4. Conclusion
Soybean biodiesel, when blended with Jatropha methyl
ester produces a composition having efficient and
improved low temperature property. Food versus fuel, the
current heated argument could be overcome with this
optimum mixture, thus, reducing dependability over
edible oil. Soybean biodiesel exhibits good low
temperature properties, where as, Jatropha biodiesel has
poor low temperature properties. With the blend of these
esters, an additive effect could be achieved.
The contents of fatty acids in both biodiesel, in view of
saturated and unsaturated have extensively contributed to
the difference in two vital and critical properties of
FAME. A signature design of 80% Jatropha (non-edible
oil) and 20% soybean biodiesel (edible oil), conforming
to standards set by ASTM D6751, EN 14214 and
MS123:1993 can be an optimum mix with with enhanced
oxidation stability and low temperature property.
ACKNOWLEDGMENT
The author would like to express his utmost gratitude to
Universiti Teknologi PETRONAS for providing
necessary facilities in the completion of this study and
approval to undertake this research.
REFERENCES
[1]
1] Ma, F. R., & Hanna, M. A. (1999). Biodiesel production: A
review. Biosource Technology , 70, 1-15.
[2]
Devanesan, M. G., Viruthagiri, T., & Sugumar, N. (2007).
Transesterification of Jatropha oil using immmobilized
Pseudomonas fluorescens. African Journal of Biotechnology , 6
(21), 2497-2501.
[3]
Wang, R., Hanna, M. A., Zhou, W.-W., Bhadury, P. S., Chen, Q.,
& Song, B.-A. (2011). Production and selected fuel properties of
biodiesel from promising non-edible oils: Euphorbia lathyris L.,
Sapium sebiferum L. and Jatropha curcas L. Bioresource
Technology , 102, 1194-1199.
[4]
Antolin, G., & Tinaut, F. (2002). Optimization of biodiesel
production of sunflower oil transesterification. Bioresource
Technology , 83, 111-114.
[5]
Qiu, F., Li, Y., Yang, D., Li, X., & Sun, P. (2011). Biodiesel
production from mixed soybean oil and rapeseed oil. Applied
Energy , 88, 2050-2055.
[12] Jatropha Curcas L.: IPCSINTOX 2007. (n.d.). Retrieved 2011,
from
IPCSINTOX
Databank:
http://www.intox.org/databank/documents/plant/jatropha/jcurc.htm
[13] Schuchardt U., S. R. (1998). Transesterification of Vegetable Oils:
a Review. Journal Brazil Chemical Society , 9 (1), 199-210. B., F.,
R.O., B., & E. H., P. (1986). J. Am. Oil Chem. Soc. , 63, 1375.
[14] Karavalakis, G., Hilari, D., Givalou, L., Karonis, D., & Stournas, S.
(2011). Storage stability and ageing effect of biodiesel blends
treated with different antioxidants. Energy , 36, 369-374.
[15] Crabbe, E., & Nolasco-Hipolito, C. (2001, March). Biodiesel
Production from Crude Palm Oil and Evaluation of Butanol
Extraction and Fuel Properties. Process Biochemistry , 65-71.
[16] American Soybean Association. (2008). Myths and Realities
Behind Rising Food Prices. BASF The Chemical Company.
[17] Auto Focus Asia. (2007). Biodiesel 2020: the emerging market.
Industry Reports, Houston, TX.
[18] Moser, B. R., & Vaughn, S. F. (2010). Coriander seed oil methyl
esters as biodiesel fuel: Unique fatty acid composition and
excellent oxidative stability. Biomass and Bioenergy , 34, 550-558.
[19] Chitra, P., Venkatachalam, P., & Sampathrajan, A. (2005).
Optimisation of experimental conditions for biodiesel production
from alkali-catalysed transesterification of Jatropha curcus oil.
Energy for Sustainable Development , IX (3).
[20] Rashid, U., Anwar, F., Jamil, A., & Bhatti, H. N. (2010). Jatropha
Curcas Seed Oil as a Viable Source for Biodiesel. Pak. J. Bot , 42
(1), 575-582.
[21] M. C., C., & C. C., H. (2000). Surface Active Properties of Palm
Oil with respect to the Processing of Palm Oil. Journal of Oil Palm
Research , 12 (1), 107-116.
[22] Canakci M, Van Gerpen J. Biodiesel production via acid catalysis.
Am Soc Agric Eng 1999;42:1203–10.
[23] Ramadhas, A., Jayaraj, S., & Muraleedharam, C. (2005). Biodiesel
production from high FFA rubber seed oil. Fuel , 84, 35-40.
[24] Sarin, R., Sharma, M., Sinharay, S., & Malhotra, R. (2007).
Jatropha-Palm biodiesel blends: An optimum mix for Asia. Fuel ,
86, 1365-1371.
[25] Foon, C. S., May, C. Y., Liang, Y. C., Ngan, M. A., & Basiron, Y.
Palm Biodiesel: Gearing Towards Malaysian Biodiesel Standards.
Kuala Lumpur: Malaysian Palm Oil Board.
[26] Moser, B. (2008). Influence of blending canola, palm, soybean,
and sunflower oil methyl esters on fuel properties of biodiesel.
Energy Fuels 22, 4301-6.
[27] Dunn, R. (2008). Crystallization behavior of fatty acid methy
esters. J Am Oil Chem Soc , 85, 961-72.
[28] Suzana Yusup, Modhar Khan,”Basic properties of crude rubber
seed oil and crude palm oil blend as a potential feedstock for
biodiesel production with enhanced cold flow characteristics”
Biomass and Bioenergy, 34 (2010) 1523-1526.
[29] Dunn, R., & Knothe, G. (2009). A comprehensive evaluation of
the melting points of fatty acids and esters determined by
differential scanning calorimetry. J Am Oil Chem Soc , 86, 843-56.
358
Download