Sunteți pe pagina 1din 184

Chapter 1

Introduction
1.1 General Introduction: Biodiesel
The world is presently confronted with the twin crises of fossil fuel depletion and
environmental degradation. Indiscriminate extraction and lavish consumption of fossil
fuels have led to reduction in reserve carbon resources. The search for alternative fuels,
which promises a harmonious correlation with sustainable development, energy
conservation, efficiency and environmental preservation, has become highly
pronounced in the present context. Bio-fuels can provide a feasible solution to this
worldwide petroleum crisis. Scientists around the world have explored several
alternative energy resources that have a potential to quench the ever-increasing energy
thirst of today's population. Various biofuel energy resources explored include biomass,
biogas, primary alcohols, vegetable oils, biodiesel, etc. These alternative energy
resources are largely environment friendly but they need to be evaluated on case-to-
case basis for their advantages, disadvantages and specific applications. Some of these
fuels can be used directly while others need to be formulated to bring the relevant
properties closer to conventional fuels. The present energy scenario has stimulates the
active research interest in non-petroleum, renewable and non-polluting fuels.

Time is ripe to strike a balance between energy security and energy usage in the
face of enormous growth of world population, increased technical development and
standard of living in the industrial nations. The prices of crude oil keep rising and
fluctuating on a daily basis. The variation in the energy prices over last decade has
necessitated development of commercial fossil fuel alternatives from bio-resources.

Several sources of energy, especially for driving the automotive are being
developed and tested. This may well be the main reason behind the growing awareness
and interest for non-conventional bioenergy sources and fuels in various developing
countries which are striving hard to offset the oil monopoly. This introduction presents

1
detailed information on biodiesel together with its emission benefits and the prospects
of biodiesel as an alternative source from non-edible oil.

Bio-diesel is a fast-developing alternative fuel in many developed and developing


countries of the world. A number of feedstock options for production of bio-fuel have
been considered in different countries (Bhasabatra & Sutiponpeibum 1982 a, b).

1.2. Global biofuel scenario


Biomass has been recognized as a major world renewable energy source to
supplement declining fossil fuel resources. Biomass appears to be an attractive
feedstock for three main reasons. First, it is a renewable resource that could be
sustainably developed in the future. Second, it appears to have formidably positive
environmental benefits with no net releases of carbon dioxide (CO2) and very low sulfur
content. Third, it appears to have significant economic potential provided that fossil fuel
prices increase in the future. Ligno-Cellulosic bio-methanol also have low emissions
because the carbon content of the alcohol is primarily, derived from carbon that was
sequestered in the growing of the bio-feedstock and is only being re-released into the
atmosphere.

Transports sector is a major consumer of petroleum. Fuels such as diesel,


gasoline, liquefied petroleum gas (LPG) and compressed natural gas (CNG).This sector is
likely to suffer badly because of following reasons: (a) Prices of petroleum in global
market have a raising trend; (b) Petroleum reserves are limited and it is a monopoly of
some oil-importing countries and rest of the world depends on them; (c) Number of
vehicles based on petroleum fuels is on increase worldwide. Many research programs
recently focused on the development of concepts such as renewable resources,
sustainable development, green energy, eco-friendly process, etc. in the transportation
sector. In developed countries there is a growing trend towards employing modern
technologies and efficient bio-energy conversion using a range of biofuels, which are
becoming cost-wise competitive with fossil fuels. Advantages of bio-fuels are the
following:(a) Bio-fuels are easily available from common biomass sources; (b) they

2
reduce CO2 footprints (c) bio-fuels are environmentally friendly; (d) they help in energy
security and improve economy; (e) they are biodegradable and contribute to
sustainability. Renewable resources are more evenly distributed than fossil and nuclear
resources, and energy flows from renewable resources are more than three orders of
magnitude higher than current global energy use. Today’s energy consumption is
unsustainable because of equity issues as well as environmental, economic and
geopolitical concerns that have implications far into the future

1.3. What is biodiesel?


Biodiesel is considered as clean fuel that contains no petroleum, no sulphur, and
no aromatics; therefore it can be blended at any level with petroleum diesel to create a
bio-diesel blend or can be used as neat. Just like petroleum, biodiesel operates in
compression ignition that essentially require very little or no engine modifications
because biodiesel has properties similar to petroleum diesel fuel. It can be stored just
like the petroleum diesel fuel and doesn’t require separate infrastructure. The use of
bio-diesel in conventional diesel engines results in substantial reduction of un-burnt
carbon monoxide, hydrocarbons and particulate matters. It has about 10% built in
oxygen that helps it to burn efficiently. Its higher cetane number improves the ignition
quality even when blended with petroleum diesel.

The best way to use non-edible oil as fuel is to convert it in to biodiesel.


Biodiesel is the name of a clean burning mono-alkyl ester-based oxygenated fuel made
from natural, renewable sources such as new/used vegetable oils and animal fats. The
resulting biodiesel is quite similar to conventional diesel in its main characteristics.
Biodiesel contains no petroleum products, but it is compatible with conventional diesel
and can be blended in any proportion with mineral diesel to create a stable biodiesel
blend. The level of blending with petroleum diesel is referred as Bxx, where xx indicates
the amount of biodiesel in the blend (i.e. B10 blend is10% biodiesel and 90% diesel. It
can be used in CI engine with no major modification in the engine hardware.

3
1.4 Biodiesel production
The general process of biodiesel production is shown in reaction above. A fat or
oil reacts with an alcohol in the presence of a catalyst to produce glycerin and methyl
esters or biodiesel. The methanol is charged in excess to assist in quick conversion and
recovered for reuse. The catalyst is usually sodium or potassium hydroxide shown in
reaction 1, Charts 1 & 2.

Reaction 1:- General chemical reaction process of biodiesel production


CH2-O-COR KOH, 6h CH2-OH
| → |
CH-O-COR + 3R’OH 3RCOOR’ + CH-OH
| |
CH2-O-CO-R CH2-OH

(1 kg) → (450 gm) (10 gm) (98 kg) (10gm)


Oil Alcohol Bio diesel Glycerin

Chart 1:- Basic scheme for Biodiesel production

4
Chart 2:- General process of biodiesel production

n Alcohol
Alcohol
Recover
y
Biodiesel

Reactor Settler Washing Purification Evaporation

Non- edible oil

Glycerin
Mineral Neutralization Settler Evaporation
Acid
Distillation
Catalyst
Fatty Acid

1.5 Characteristics of biodiesel

1.5.1 Physical properties


Table 1: Comparison of the physico-chemical parameters of the investigated sample
to ASTM and EN standards

Parameter ASTM ASTM EN-diesel EN- Crude oil Transesterified


diesel biodiesel biodiesel Biodiesel

Color Gold. Gold. Gold. Gold. Gold. Gold. yellow


yellow yellow yellow yellow yellow
Kinematic 2.4-4.1 1.9-6.0 2.0-4.5 3.5-5 27-11 4-8
viscosity
mm2/s

Specific - - 0.820- 86-0.90 0.92 0.87


gravity 0.845
Free fatty - - - - 3.1 0.25
acid (%)
Acid value - 0.80 - 0.50 6.3 0.49
(mg KOH/g)
Source: - Traore S. et al (2007)

5
Investigated properties and the range of color, kinematic viscosity, specific
gravity, free fatty acids and acid value in ASTM diesel and biodiesel, EN diesel, crude oil
and transesterified biodiesel are presented on Table 1.

1.5.2 Emission Characteristics


The methods, limits, unit of emission characteristics of pure biodiesel according
to the ASTM specification are given in Table 2.

Table 2:- ASTM Specification for B100 limit

Property ASTM Method Limits Units


Flash Point D93 130 min. ºC
Water & Sediment D2709 0.050 max. % Volume
Kinematic Viscosity (40degree C) D445 1.9-6.0 mm2/sec
Sulfated Ash D874 0.020 max. % mass
Sulphur D5453 0.05 max. % mass
Copper Strip Corrosion D130 No.3 max.
Cetane number D613 47 min.
Cloud Point D2500 Report ºC
Carbon Residue (100% Sample) D4530* 0.050 max. % mass
Acid Number D664 0.80 max. mg KOH/g
Free Glycerin D6584 0.020 max. % mass
Total Glycerin D6584 0.240 max. % mass
*The carbon residue shall run out in the 100% sample
Source: Canaki M. (2007)

Biodiesel is the only alternative fuel to have a complete evaluation of emission


results and potential health effects submitted to the U.S. EPA under the Clean Air Act
Section 211(b). These programs include the most stringent emissions testing protocols
ever required by EPA for certification of fuels in the U.S. Emission results for pure
biodiesel (B100) and mixed biodiesel (B20-20% biodiesel and 80% petro diesel)
compared to conventional diesel are given in Table 3.

6
Table 3:- Biodiesel emissions comparison to conventional diesel
Emissions B100 B20
Regulated Emissions
Total Unburned Hydrocarbons -93% -30%
Carbon Monoxide -50% -20%
Particulate Matter -30% -22%
NOx +13% +2%
Non-Regulated Emissions
Sulphates -100% -20%*
Polycyclic Aromatic Hydrocarbons -80% -13%
(PAH)**
NPAH (Nitrated PAHs)** -90% -50%***
Ozone Potential of Speciated HC -50% -10%
Life-Cycle Emissions
Carbon Dioxide (LCA) -80%
Sulphur Dioxide (LCA) -100%
*Estimated from B100 results. **Average reduction across all compounds measured.
***2-nitroflourine results were within test method variability.

Source:- www.epa.gov/otaq/models/analysis/biodsl/po2001.pdf, U.S. EPA (2001)

Graph:-1 Impact of biodiesel

Average emission impacts of biodiesel compared with fossil diesel


Source: EERE (2006)
Opinion regarding emissions of nitrogen dioxides varies from one study to
another study. Some fleet tests concluded Nox emissions to have increased with the use

7
of biodiesel as fuel while other studies proved that emissions of Nox can be controlled,
if not decreased, by adjustments like retarding the injection timing or by adding heavy
alkylate replacing 20% of the fuel of B20% blend biodiesel. Average emission impacts of
biodiesel for heavy-duty highway engines are as given in Graph1.

1.5.3 Lubricity of Biodiesel


Biodiesel blends offer superior lubricating properties which may reduce engine
wear and extend the life of fuel injection systems. Tests with two leading lubricity
measuring systems the BOCLE machine and the HFRR machine have shown that
biodiesel blends offer better lubricating properties then conventional petroleum diesel.
The result of a lubricity test done by Exxon with petro diesel and biodiesel blends is
given in Table 4.

Table 4:- Lubricity test performance with different blends of petro diesel and
biodiesel in machine
Fuel Type Scar Friction Film %
Conventional low sulphur diesel 492 0.24 32
Blend (80% petro diesel + 20% biodiesel) 193 0.13 93
Blend (70% petro diesel + 30% biodiesel) 206 0.13 93
Petro diesel + 1000 ppm lubricity additive 192 0.13 82
Petro diesel + 500 ppm lubricity additive 215 0.14 94
Petro diesel + 300 ppm lubricity additive 188 0.13 93
Source: Exxon & Interchem Environmental Inc. Lubricity Results (HFRR Machine)

1.6 Biodiesel Specifications


The key components that determine the quality of biodiesel are monoalkyl
esters, water & sediment, kinematic viscosity, ash copper strip corrosion, aromaticity
etc. Specification for B100 and a provisional specification for biodiesel and petro diesel
are also notified by the ASTM given in Table 5. Table 6 summarizes standards for
biodiesel in various countries and shows a comparison of selected properties of
biodiesel and petro diesel.

8
Table 5:- Fuel properties for petro diesel and biodiesel
Property ASTM Petro diesel fuels Biodiesel
Method

Flash Point D93 325 K min. 403 K min


Water & Sediment D2709 0.05 max. vol % 0.05 max. vol%
Kinematic Viscosity (40˚ C) D445 1.3-4.1 mm2/sec 1.9-6.0 mm2/sec
Sulfated Ash D874 - 0.02 mass wt.%
Ash D5453 0.01 max. wt. % -
Sulfur D130 0.05 max wt. %. -
Sulfur D613 - 0.05 max wt.%
Copper Strip Corrosion D2500 NO 3 max NO 3 max
Cetane number D4530* 40 min. 47 min
Aromaticity D664 35 min vol %. -
Carbon Residue (100% Sample) D6584 -. 0.05 max mass %
Carbon D6584 0.35 max mass % -
Distillation Temperature (90% D4951 555 K min-611 Kmax. -
Recovered)
*The carbon residue shall run out in the 100% sample.
Source: Demirbas A. (2007) Energy Policy

Table 6:- Biodiesel Standards of different countries


Specifications Units Australia France Germany Italy Sweden USA Draft EU
Standard/Specification ONC1191 - DINE UNI SS1 ASTMD EN14214
51606 10635 55436 6751
Introduction Date Jly 1997 Spt1997 Spt1997 Apr1997 Nov1996 Dec2001 2001
3
Density @15 C g/cm 0.85-0.89 0.87-.89 .875- 0.86- 0.87- - 0.86-
0.90 0.90 0.90 0.90
2
Viscosity@40 C mm /s 3.5-5.0 3.5-5.0 3.5-5.0 3.5-5.0 3.5-5.0 1.9-6.0 3.5-5.0
Flash Point ċ ≥100 ≥100 ≥110 ≥100 ≥100 ≥130 ≥130
CFPP ċ 0/-15 - 0-10/- - -5 - 0/-15
20
Pour Point ċ - -10 - 0/-15 - - -
Sulphur %max 0.02 0.02 0.01 0.01 0.01 0.05 0.01
CCR 100%max 0.05 - 0.05 - - 0.05 -
10%disti.residue %max - 0.3 - 0.5 - - 0.3
Sulphated Ash %max 0.02 - 0.03 - - 0.02 0.01
(Oxid) Ash, mx %mass - - - 0.01 0.01 - -
Water max. mg/kg - 200 300 700 300 ≤0.05 500
Total mg/kg - - 20 - 20 - -
Contaminants

9
Cu Corrosion 3h/50ċ - - 1 - - NO.3 1
Cetane No. ≥49 ≥49 ≥49 - ≥48 ≥47 ≤≥49
Neutral No. ≤0.8 ≤0.5 ≤0.5 ≤0.5 ≤0.6 ≤0.8 ≤0.02
Methanol %mass ≤0.20 ≤0.01 ≤0.3 ≤0.02 ≤0.02 - ≤0.02
Ester Content %mass - ≥96.5 - ≥98 ≥98 - ≥96.5
Monoglyceride %mass - ≤0.8 ≤0.8 ≤0.8 ≤0.8 - ≤0.8
Diglyceride %mass - ≤0.2 ≤0.4 ≤0.2 ≤0.1 - ≤0.20
Triglyceride %mass - ≤0.2 ≤0.4 ≤0.1 ≤0.1 - ≤0.03
Free Glycerol %mass ≤0.02 ≤0.02 ≤0.02 ≤0.05 ≤0.02 ≤0.02 0.25
Total Glycerol %mass 0.24 0.25 0.25 - - 0.24 ≤115
Iodine No. ≤120 ≤115 ≤115 - ≤125 - -
C18:3 & higher ≤15 - - - - - 10
acids
Phosphorous ppm ≤20 ≤10 ≤10 ≤10 ≤10 ≤10 10
Alkaline Matter (Na,K) - ≤5 ≤5 ≤10 ≤10 - ≤360
Distillation 95% ċ - ≤360 - - - ≤360 *
IBP min. ċ - - - - - *
Bound Glycerin - - - - - Max 0.8
Oxidation Hrs. - - - - - 6 min.
Stability
Sediment - - - - - ≤0.05
Cloud Point - - - - * -
Source: - National renewable energy laboratory report (2002-04)

1.7 Toxicity of Biodiesel


Impacts on human health represent significant criteria as to the suitability of the
fuel for commercial applications. Health effects can be measured in terms of fuel
toxicity to the human body as well as health impacts due to exhaust emissions. Tests
conducted by various laboratories showed the acute oral toxicity of pure biodiesel fuel
as well as B20 in a single dose study on rats, and that biodiesel is not toxic and there is
no hazards anticipated from ingestion incidental to industrial exposure. According to
NIOSH (National Institute for Occupational Safety & Human Health), a 96-hr. lethal
concentration of biodiesel for bluegills was greater than 1000 mg/l and this aquatic
toxicity is deemed as insignificant. It is less than the irritation produced by 4% soap and
water solution.

10
1.8 Sources / options for biofuel
India is one among largest petroleum consuming and importing countries. India
imports about 70% of its petroleum demand; petroleum imports is currently about
Rs.600 billion (about 30% of total import bill) compared to current trade deficit of about
Rs.500 billion. The currently yearly consumption of diesel in India is approximately 40
million tones forming 40% of the total petroleum products consumption. The ongoing
economic expansion would increase the demand for transportation fuel in short and
medium terms at high rates. According to international Energy Agency (IEA) scenario
developed for the USA and the EU indicate that near-term targets of up to 6%
displacement of petroleum fuels with biofuels appear feasible using conventional
biofuels. A 5% displacement of gasoline in the USA and 8% in EU requires about 5% of
available cropland to produce ethanol. A 5% displacement of diesel requires 13% of USA
cropland and 15 % in the EU, it is anticipated.

The dwindling fossil fuel sources and the increasing dependency on imported
crude oil have led to a major interest in expanding the use of bio-diesel. The recent
commitment by the USA government to increase bio-energy three fold in 10 years has
added impetus to the search for viable biofuels. The EU have also adopted a proposal
for a directive on the promotion of the use of bio-fuels with measures ensuring that bio-
fuels account for at least 2% of the market for gasoline and diesel sold as transport fuel
by the end of 2005.

Non-edible seeds like Jatropha, Pongamia and Neem etc. will be best options
as sources for feedstock for the oil and biofuel production in India. Detailed information
is required to use these sources for the biodiesel production and new technologies
regarding enhancement of use and production of biofuel need to be furnished

Justification of study
This study brings out the problems that lead and justify the feasibility of Jatropha
and non-edible oil to be utilized as fuel for the transport sector in India.

11
This study includes collection and screening of high oil yielding varieties of J.
curcas. Cycle of biodiesel produced from non-edible oil by transesterification, analysis
mixed feedstock as a biodiesel has been investigated.

Comparison of J. curcas seeds oil with Pongamia and Neem seed oils has been
made in order to provide an estimate of their potentials as biodiesel; and engine
performance when mixed oils are used. The study would provide useful leads on
application of Jatropha curcas feedstock for biodiesel.

Objectives of the study


 To screen germplasm of J. curcas for oil quantity and quality from
different provenances of India (Rajasthan and Uttaranchal).

 To standardize the protocols to maximize oil yield.

 To assess of physico-chemical properties of J. curcas and other non-


edible oils.

 To develop protocols for esterification and purification so that the crude


oil can be improved for use as diesel substitute for stationary motors.

 To evaluate application of biodiesel in stationary engines.

 To evaluate and compare Neem oil, Pongamia oil with that of J. curcas oil
for developing the biodiesel from mixed feedstock.

12
Chapter 2
Review of Literature

Feedstock for biodiesel


There is no single feedstock that can be used throughout the year and in all
regions of any country. Mixed feedstock is an option but demands change in process
and technology.

Many developed countries are using edible oil-seed crops such as soybean,
rapeseed, groundnut, sunflower for production of bio-diesel. However, developing
countries like India, who import huge quantities of edible oil to meet their
requirements, cannot afford to use edible oils for bio-diesel production. Many
alternatives plants have been identified as feedstock for bio-fuel; and Jatropha curcas is
one of them. The genus Jatropha has 476 species which are distributed throughout the
world. Among them, 12 species are recorded from India. Jatropha curcas Linn. (Physic
nut or Ratanjot) belongs to family Euphorbiaceae. According to Dehgan & Webster
(1979) and Schultze-Motel (1986) the genus name Jatropha derives from the Greek
word jatr´os (doctor) and troph´e (food), which implies its medicinal uses. This species is
native of tropical America, but is now found abundantly throughout the arid, semi-arid,
tropical and subtropical regions of the world (Makkar et al. 1997; Heller 1996). Jatropha
curcas Linn. is a deciduous shrub that grows up to a height of 3–5 m, and has a
productive life of 50 years. It bears fruits from the second year of its establishment and
the economic yield stabilizes from the fourth or fifth year onwards (Hikwa 1995;
Henning 1996; Makkar et al. 2001). In India, it is found in semi-wild conditions in the
vicinity of villages and is one of the most promising drought tolerant perennial plants
adapted to various soil conditions. It can tolerate drought conditions and animals do not
browse its leaves.

Major feedstock production areas region-wise:


Soybeans: Europe (Ukraine, Russia, Italy, France and Rest of Europe); North
America (USA, Canada); Latin America (Argentina, Brazil); Asia-Pacific (China, India,

13
Indonesia, Korea, Japan, Thailand and Rest of Asia-Pacific). Feedstock used for
biodiesel: Only Europe and North America.

Cotton Seeds: Europe (Greece, Spain); USA; Latin America (Brazil, Argentina);
Asia-Pacific (China, India, Pakistan, Australia and Rest of Asia-Pacific). Feedstock used
for biodiesel: Only Europe and North America.

Rapeseeds: Europe (Germany, France, United Kingdom, Poland, Czech Republic,


Denmark, Slovakia, Sweden, Austria and Rest of Europe); North America (USA, Canada);
Asia-Pacific (China, India, Australia, Pakistan and Bangladesh). Feedstock used for
biodiesel: Only Europe and North America

Groundnuts: USA; Asia-Pacific (China, India, Indonesia, Thailand and Pakistan);


Latin America (Brazil, Argentina).

Sunflower Seeds: Europe (France, Hungary, Spain, Italy, Slovakia, Russia and Rest
of Europe); North America (USA, Canada); Latin America; Asia-Pacific (China, India,
Pakistan and Australia).

Palm Kernel: Latin America (Brazil, Colombia); Asia-Pacific (Indonesia, Malaysia


and Thailand). Feedstock used for biodiesel: Malaysia, Indonesia
Copra: Asia-Pacific (Indonesia, India and Thailand) and Rest of World.
Castor Seeds: Brazil; Asia-Pacific (India, China and Thailand).

Jatropha Curcas: India, Africa, Malaysia, Indonesia. Feedstock used for biodiesel:
India, Pakistan, Africa.

In India Jatropha curcas has been accepted as a major feedstock other than
Pongamia pinnata and other non-edible vegetable oils. However, the west and other
countries are still dependent upon vegetable oil as a source of biodiesel. According to
Agarwal (2007) biodiesel can be blended in any proportion with mineral diesel to create
a biodiesel blend or can also be used as neat. According to him the main resources for
biodiesel production are non-edible oils obtained from Jatropha curcas (Ratanjyot),

14
Pongamia pinnata (Karanj), Calophyllum inophyllum (Nagchampa), and Hevea
brasiliensis (Rubber plant).

Two plant species, Sapindus mukorossi and Jatropha curcas were discussed as
newer sources of oil for biodiesel production by Chhetri et al (2008). Experimental
analysis conducted by them showed that both oils have great potential to be used as
feedstock for biodiesel production.

Jatropha is a fast growing and long lived plant, easy to propagate, found to be
growing in many parts of the country. It can grow and survive with minimum inputs in
marginal land and not browsed by animals and seeds not even eaten by away by birds.
The organic matter from shed leaves enhances earthworm activity in the soil around the
root zone of the plants, which improve the fertility of diesel has important implications
for meeting the demands of rural energy services and also exploring practical
substitutes for fossil fuels to counter greenhouse gas accumulation in the atmosphere
(Parida & Eganathan, 2007).

Studies for the synthesis and characterization of biodiesel from non-edible oils
like Jatropha curcas, Pongamia glabra (Karanj), Madhuca indica (Mahua) and Salvadora
oleoides (Pilu) have been carried out at our laboratory at the National Botanical
Research Institute. The seeds of Madhuca indica (Mahua) produce oil that can be
converted to biodiesel by transesterification. The cake left after extraction of oil can be
used as a fertilizer (Behl et al., 2007). Pongamia pinnata oil has multiple uses. It is a very
good source of vegetable oil that can be converted to biodiesel and meet diesel
requirement of the country. Biodiesel made from Karanj oil has been evaluated for its
efficacy by National Botanical Research Institute and IIP, Dehradun (Behl et al., 2007).

The seeds of Salvadora oleoides are rich in oil. The oil extracted from the seeds
can serve as a local (Rajasthan and Gujarat) resource that can be used as feedstock for
biodiesel in desert areas. The seeds are rich in oil and contain Lauric, myristic, and
palmitic acids (Behl et al. 2007).

15
Huber et al (2007) reported that the yield of straight chain alkanes increases
when sunflower oil is mixed with vegetable oil. They illustrated that dilution of heavy
vacuum oil (HVO) can improve the reaction chemistry. According to them mixing of
sunflower oil with heavy vacuum oil does not decrease the rate of desulfurization
indicating that sunflower oil does not inhibit the hydro treating of heavy vacuum oil.

Packages process for the cultivation, seed processing for oil extraction and
the production of the methyl esters from J. curcas oil was described by Foidl et al
(1996).

Agarwal (2007) reviewed production, characterization, current status of


vegetable oil and biodiesel from well-to-wheel including greenhouse gas emissions,
well-to-wheel efficiencies, fuel versatility, infrastructure, availability, economics, engine
performance emissions, effect on wear and lubricating oil etc. He reported that ethanol
is an attractive alternative fuel because it is a renewable bio-based resource and it is
oxygenated, thereby providing a potential to reduce particulate emissions in
compression-ignition engines. In this review, he also reported properties and
specifications of ethanol blended with diesel and gasoline fuel. Effect of the fuel on
engine performance and emissions (SI as well as compression ignition (CI) engines) and
material compatibility were also studied. According to his study biodiesel from Jatropha
curcas (Ratanjyot), Pongamia pinnata (Karanj), Calophyllum inophyllum (Nagchampa),
Hevea brasiliensis (Rubber) etc. oil can be blended in any proportion with mineral diesel
to create a biodiesel blend or can be used in its pure form. Biodiesel in compression
ignition (diesel) engine requires very little or no engine modifications.

Jatropha curcas
Jatropha curcas is a land race in India and occurs in several states and agro
climatic conditions. Screening for variability can help in genetic selection and
improvement of stocks. Several investigators have collected and evaluated Jatropha
curcas accessions and other non-edible oilseeds from India and abroad for use as
biodiesel feedstock.

16
Germplasm evaluation
Kaushik et al (2007) studied genetic variability in seed traits of 24 accessions of J.
curcas collected from Haryana. Seeds of collected accessions were accessed for oil
content and reported a range between 28 to 38 %. They reported a higher genotypic
coefficient of variation as compared to phenotypic, indicating the predominant role of
environment. They also found seed weight to have positive correlation with seed length,
breadth, thickness and oil content. On the basis of their research, it was suggested that
the crossing between accessions will result in wide spectrum in the form of cluster for
variability in subsequent generations.

Nambisan (2007) undertook genetic analysis of Jatropha species for yield


characteristics in order to identify quantitative trait loci (QTLs) and improving useful
characters of gene inherited in a multigenic fashion for the yield of oil. She standardized
and applied RAPD and AFLP to detect the variations at DNA level rather than the
phenotypes.

A systematic collection of 162 accessions of J. curcas was carried out from four
distinct eco-geographic zones of peninsular India in 2005 along with passport data,
documentation of important plant traits in-situ, eco-geographic parameters studies by
Sunil et al (2007). Assessment of variability among the collected accessions was also
undertaken by them.

Morphological characteristics of selected germplasm seeds


Morphological and physico-chemical studies of oil of non-edible oilseeds can be
helpful to screen out the best germplasm for the oil yield and oil components.

Parkia biglobbossa and Jatropha curcas seeds were analyzed by Akintayo (2004)
for their proximate composition like oil extraction, physic-chemical characteristics, fatty
acid composition, lipid classes and sterols of extracted oil. Proximate composition
analysis revealed that percentage of crude protein; crude fat and moisture in Parkia
biglobbossa were 32.4%, 26.52%, 10.18% while in J. curcas, it was 24.6%, 47.25%,
5.54%. Campesterol, stigmasterol, b-sitosterol, D5-avenasterol and D7-stigmasterol

17
were identified in P. biglobbossa seed oils, however b-sitosterol was found most
abundant, constituting 71.9% in J. curcas and 39.5% in P. biglobbossa. Fatty acid
composition was analyzed for P. biglobbossa and J. curcas oil. J. curcas oil had 72.7%
unsaturated fatty acids with oleic acid; and P. biglobbossa had 62% unsaturated fatty
acids with linoleic acid, being the most abundant. Lipid classification showed triglyceride
as the dominant lipid species in the seed oils. Physico-chemical analysis of the oils
showed that the oil extracted from P. biglobbossa and J. curcas are applicable for resin
and soap manufacture.

Visvanathan et al (1996) determined physical properties of Neem seed such as


dimensions, crushing strength, 1000 nut mass, relative mass of kernel and shell, angle of
repose, porosity, bulk density, particle density and coefficient of static friction. They
reported Neem seed moisture ranged from 7·6% to 21%, stem-end diameter ranged
from 12·87 to 16·20 mm. The crushing strength of the nut was measured along the
longitudinal axis and a diametric axis which decreased with increase in moisture
content; mass of 1000 nuts, percent content of the mass of the kernel and the angle of
repose of Neem nut was found to increase with increase in moisture content; the
porosity, bulk density, particle density decreased linearly as the moisture content
increased; and coefficient of static friction on various surfaces increased with increase in
moisture content.

Oil extraction
Use of appropriate oil extraction processes and extraction solvents is very
important factor to be followed in extraction of non-edible oil.

Many petroleum and non-petroleum solvents have been used to extract oil from
oil seeds. Hydrocarbon solvents (hexane, heptane and pentane) were used for the
standardization of solvent for high oil yield. Usually n-hexane was used as it is free from
the nitrogen, sulphur, unsaturated compounds and found sufficiently stable to be used
indefinitely. Adriaans (2005) observed that commercial heptane might be preferred for
the extraction of castor oil which is not freely miscible with hydrocarbons except at

18
elevated temperature. As a result only water has been found applicable as a solvent for
the extraction of oil from palm, olive and coconut. 1:2 ratio of hexane-water mixture
were found economical only in the processing of olive oil with high potential for future
use. Use of hexane for the extraction of oil from the oilseeds has been suggested and
found economically feasible.

Liauw et al (2008) studied Neem oil extraction using different solvents to


increase the oil yield. Maximum 44.29% oil yield was reported using n-hexane while
41.11% oil was extracted when ethanol at 50°C was used. After the extraction, the effect
of solvents in physico-chemical characteristics of oil was also studied by them.

Hexane and isopropanol were compared as solvents for use in ambient-


temperature equilibrium extraction of rice bran oil by Procter and Bowen (1996). 20 ml
of isopropanol solvent was found as effective to extract the oil from 2 g of bran as
compared to hexane. Free fatty acid level was found between 2-3% in both the solvents.
Large-scale production of oil was done by using 30 g of bran in 150 ml of solvent which
had similar free fatty acid content and a phosphorus level of approximately 500 ppm. It
was observed that the oil extracted with isopropanol was significantly more stable to
heat-induced oxidation than hexane.

Isopropanol was significantly more stable to heat-induced oxidation than hexane


and antioxidants that are more easily extracted by isopropanol than hexane may be
responsible for the increase in stability of oil as reported by Procter and Bowen (1996)
during their study on extraction of oil from rice bran oil.

Shah et al (2005) studied extraction of oil from Jatropha seeds by enzymatic


reaction. They treated the seeds using ultra sonication method and found that oil
extraction procedure was easier and took less time for extraction.

Mechanical press and solvent extraction methods for oil extraction were studied
by Adriaans (2005) for various oil seeds. It was observed that in a pre-press solvent
extraction the press was operated to give a pressed cake with 15 – 18% oil. The expeller
and solvent method for proper extraction of oil was modified to extract oil from the

19
cake with a solvent which offers a way to reduce the loss of oil content to less than 1%
in pressed cake.

Oil content of 162 accessions of J. curcas was estimated by Sunil et al (2007)


using Soxhlet extraction method. It ranged from 22% to 42%. They reported multi
location evaluation using an in-situ method developed to facilitate the selection of
promising accessions and for the identification of superior lines by assessing the
phenotypic traits of plants.

Protocol for the esterification of Karanj oil was developed by Raheman and
Phadatare (2004) which consist of heating of oil, addition of KOH and methyl alcohol,
stirring of mixture, separation of glycerol, washing with distilled water and heating for
removal of water.

Physicochemical properties of oil & biodiesel


Physico-chemical properties are crucial since these will govern quality and
process of trans esterification as well as performance of the blend. These parameters
can directly influence the quality of oil. Physico-chemical properties of fatty acid methyl
ester of non-edible seed oils like Azadirachta indica, Calophyllum inophyllum, Jatropha
curcas and Pongamia pinnata were found most suitable for biodiesel and match major
specification of biodiesel.

Physico-chemical properties of non-edible seed oils like Azadirachta indica,


Calophyllum inophyllum, Jatropha curcas and Pongamia pinnata were found most
suitable for biodiesel and match specification of biodiesel as per USA, Germany and
European standards (Azam et al, 2005). They concluded that these plants have great
potential for biodiesel production.

Soapnut oil was reported (Chhetri et al, 2008) to have an average of 9.1% free
FA, 84.43% triglycerides, 4.88% sterol and 1.59% others. Jatropha oil contains
approximately 14% free FA, approximately 5% higher than Sapindus mukorossi (soap
nut) oil. Soapnut oil biodiesel contains approximately 85% unsaturated FA while
Jatropha biodiesel was found to have approximately 80% unsaturated FA. Oleic acid was

20
found to be the dominant FA in both Soapnut and Jatropha biodiesel. Over 97%
conversion to FAME was achieved for both Soapnut and Jatropha oil.

Kaul et al (2007) studied the synthesis and characterization of biodiesel from


non-edible oils like Jatropha curcas, Pongamia glabra, Madhuca indica and Salvadora
oleoides.

The density and viscosity of the Polanga oil methyl ester formed after triple
stage trans esterification were found to be close to those of petroleum diesel oil. The
flash point of all the blends of Pongamia or polanga oil methyl ester was found higher
than that of diesel oil. Based on the exhaustive engine tests undertaken by Sahoo et al
(2007), it was concluded that polanga based biodiesel can be adopted as an alternative
fuel for the existing conventional diesel engines without any major modifications in the
engine system. Particularly, 100% biodiesel showed higher flash point than petroleum
diesel oil. All these tests for characterization of biodiesel demonstrated that almost all
the important properties of biodiesel were in very close agreement with the diesel oil
making it a potential fuel for the application in compression ignition engines for
complete replacement of diesel fuel.

Physico-chemical properties of oil and fatty acid methyl ester of non-edible seed
oil like Azadirachta indica, Calophyllum inophyllum, Jatropha curcas and Pongamia
pinnata were found most suitable for biodiesel which matches the major specification of
biodiesel with USA, Germany and European standards. Physico-chemical properties like
specific gravity, acid value, free fatty acids content, refractive index etc. of oil and
biodiesel have been determined and analyzed using Bureau of Indian standards BIS: 548
(1976).

Gerpen et al (2002, 2004) analyzed the basics of biodiesel production by basic


pilot plant as equipment and biodiesel plant logistics. They studied long term storage of
biodiesel and physico-chemical properties after extraction of oil of soybean. They
carried out studies on different methods of fuel property measurement, soap catalyst

21
measurement, fatty acids and total glycerol measurements and analyzed for the
physiochemical properties as per ASTM specification.

A central composite rotatable design was prepared to study the effect of


methanol quantity, acid concentration and reaction time on the reduction of free fatty
acids content of Mahua oil during its pretreatment for making biodiesel by Ghadge and
Raheman (2006). According to them, all the three variables significantly affected the
acid value of the product; methanol being the most effective followed the reaction time
and acid catalyst concentration.

Important fuel properties of methyl esters of Pongamia oil (Biodiesel) were


compared for the physico-chemical properties like (viscosity = 4.8 Cst @ 40°C and flash
point = 150°C) with ASTM and German biodiesel standards and were found good
compatibility as commercial diesel as reported by Karmee and Chadha (2005). Physico-
chemical properties of the Karanj methyl ester after esterification were found to be very
close to petroleum diesel oil by Srivastava and Verma (2007).

Study were undertaken by Sarathy et al (2007) on the detailed effect of the


FAME molecular structure on the saturated (i.e., methyl butanoate) and an unsaturated
(i.e., methyl crotonate) combustion chemistry. They studied that the C 4 FAME was
oxidized in an opposed flow diffusion flame and a jet stirred reactor. Consistent trends
were seen in both the experiments. Both fuels had similar reactivity. As a result it was
observed that methyl crotonate combustion produces much higher levels of C 2H2, 1-
C3H4, 1-C4H8 and 1,3-C4H6, benzene in the opposed flow diffusion flames and 2-propenal,
methanol and acetaldehyde than methyl butanoate. The methyl butanoate combustion
had higher levels of C2H4 while for methyl crotonate it was not detected.

According to the Liauw et al (2008) extraction of Neem oil using hexane and
ethanol as a solvent was effective for physico-chemical properties of the oil. They
studied kinetic reactions which indicated that extraction process was endodermic,
irreversible and spontaneous. As a result they observed that increase in temperature

22
during extraction increased oil yield, saponification value and peroxide value but
decreased the iodine value and the oil quality.

Nambisan (2007) reported Pongamia oil has higher quantity of unsaponifiable


matter than Jatropha oil while the acid value was similar for both. Presence of high
unsaponifiable matter in Pongamia inhibited its processing for biodiesel production.
Nambisan (2007) reported that the energy rating of J. curcas is comparatively low (40
MJ/kg) as compared to other species of Jatropha e.g. J. glandulifera (57.1MJ/kg), which
is a reason for low oil yield in J. curcas; however crossing these two species may result in
plants with higher oil and energy content.

Toxicity in non-edible oil


Biologically active substances such as phorbol esters (a family of compounds
known to cause a large number of biological effects such as tumor promotion and
inflammation) are responsible for the toxicity of J. curcas oil. These are responsible for
degradation of quality of oil non-edible oil seed cake. Reports related to toxicity in
detoxification are a major issue for use of cake as an animal feed.

Phorbol esters were isolated and their molluscidal, insecticidal, fungicidal


properties were analyzed in lab-scale experiments as well as in field trials by Gubitz et al
(1999). Biotechnological processes for exploitation of J. curcas have been developed by
them that include genetic improvement of the plant, biological pest control, enzyme-
supported oil extraction, anaerobic fermentation of the press cake, isolation of anti-
inflammatory substances and wound-healing enzymes.

Haas and Mittelbach (2000) studied the toxic agent as well as technical
restriction due to detoxification of seed oil of J. curcas plant. They reported that seed oil
of Jatropha curcas contains phorbol esters and it was necessary to find feasible routes
for detoxification of this oil. The same was done by treating the oil for refining process.
Several refining steps were optimized for detoxification. Almost no effect was observed
with degumming and deodorization whereas de-acidification and bleaching reduced the
phorbol esters content by 55%.

23
Shelf life of oil and biodiesel
Biodiesel as well as oils have a shelf life and cannot be stored for too long for
various reasons. Oils do have tocopherols that provide stability to some extent yet they
are prone to degradation. Various factors may effects the quality of oil and can degrade
the oil during storage. There have been several studies on shelf life of vegetable oils and
bio diesel.

Bouaid et al (2007) conducted 30-months study on high oleic sunflower oil


methyl ester (HOSME), high-erucic Brassica oil (HEBO) ME, low-erucic Brassica oil (LEBO)
ME and used frying oil (UFO) ME and found that all biodiesel samples were very stable
because they do not have rapid increase in peroxide value (PV), acid value (AV), viscosity
(ν), and insoluble impurities (II). However, there was a deterioration of the fuel after 12
months of storage. Significant differences were found in the value of the measured
parameters for all fuel type and storage conditions with the passage of time. For all
biodiesel samples, peroxide, acid value, viscosity, and insoluble impurities tended to
increase and iodine value (IV) decreased over time. Fuels exposed to daylight tended to
degrade at faster rate than did the others fuels, particularly as indicated by their
peroxide and acid values. The specification limit of the parameters studied exceeded in
biodiesel samples after a storage time of 12 months as reported by them.

Presence of fully converted monoalkyl esters is the major requirement in quality


biodiesel (Fernando, 2007). There is a high propensity of substandard biodiesel entering
the market and being used in compression ignition engines due to high associated costs
of testing and widespread production of biodiesel. It is important to understand how
low grade biodiesel with a lower methyl ester conversion affects the parameters of
quality standards, he reported, since this effects engine performance and durability.
Performance of fatty acid methyl esters with different proportions of unconverted
triglycerides has been evaluated by Fernando (2007). The study comprehensively
evaluated effect of unconverted triglycerides on flash point, water, sediment, kinematic
viscosity, sulfur content, sulfated ash, copper strip corrosion, cetane number, cloud

24
point, carbon residue, acid number, free glycerin, total glycerin, phosphorus content
and distillation temperature.

One of the major technical issues facing biodiesel is its susceptibility to oxidation
upon exposure to oxygen in ambient air. This susceptibility is due to its content of
unsaturated fatty acid chains, especially those with bis-allylic methylene moieties.
Oxidation of fatty acid chains was complex process that proceeds by a variety of
mechanisms. Besides the presence of air and various other factors influence the
oxidation process of biodiesel including presence of light, elevated temperature,
extraneous materials such as metals which may be even present in the container
material, peroxides, antioxidants as well as the size of the surface area between
biodiesel and air.

Gupta et al (2008) extracted oil from rice bran, Jatropha curcas and Karanj oil
and extracted oils were stored for one year to find out change in the quality during
storage. Physico-chemical properties like viscosity, free fatty acid content and density
were monitored. All the three parameters showed an increasing trend during the
storage period. This trend was observed in washed as well as unwashed bio-diesel of all
the three oils.

Addition of antioxidants or modification of the fatty ester profile is a common


approach to improve biodiesel oxidative stability. Knothe (2007) suggested factors that
influence biodiesel oxidative stability.

Leung et al (2006) investigated biodiesel degradation characteristics under


different storage conditions. Quality of twelve biodiesel samples, which were divided
into 3 groups, stored at different temperatures and environmental conditions were
monitored at regular interval over a period of 52 weeks. Experimental results
demonstrated that the biodiesel under test degraded less than 10% within 52 weeks for
those samples which were stored at 4 and 20 °C while nearly 40% degradation was
found for those samples stored at a higher temperature, i.e. 40°C. The results suggested

25
that high temperature together with air exposure, water content (due to hydrolysis)
greatly increases the biodiesel degradation rate.

Comparative study of crude oil, filtered oil and refined oil was reported by Mittal
et al (1964) who described the method of refining oil with Boume solution for
vegetative and industrial oil as per the FFA content following bailey’s method.

Conversion of vegetable oil to biodiesel


Trans-esterification is a process to convert oil to biodiesel. During conversion
various factors have to be identified that influence the yield and the properties of
biodiesel. Studies of relevant factors for biodiesel conversion have been undertaken
extensively.

Palm kernel oil has been identified as a renewable resource for biodiesel. The
effect of ethanol–PKO ratio on (PKO) biodiesel yield was studied by Alamu et al (2007)
with a view to obtain optimal feedstock ratio. Experiments were conducted for ethanol–
PKO ratios 0.1, 0.125, 0.15, 0.175, 0.2, 0.225 and 0.25 at trans esterification conditions
at 60°C temperature for 120 min reaction time in 1.0% KOH catalyst concentration.
29.5%, 54%, 75%, 89%, 96%, 93.5% and 87.2% average PKO biodiesel yields were
obtained for the respective feedstock ratios. This showed the increase in biodiesel yield
with ethanol–PKO ratio up to 0.2 in trans esterification reaction. Biodiesel as a fuel was
found within the biodiesel standard specifications.

Berchmans and Hirata (2007) developed a technique to produce biodiesel from


crude J. curcas oil having high free fatty acids (15% FFA) by trans esterification using acid
and alkali treatments. High FFA level of J. curcas oil was reduced to less than 1% by a
two-step pretreatment process. In the first step, reaction was carried out with 0.60 w/w
methanol-to-oil ratios in the presence of 1% w/w H2SO4 as an acid catalyst for 1h
reaction at 50°C. In the second step, trans esterification was done by using 0.24 w/w
methanol to oil and 1.4% w/w NaOH as alkaline catalyst to oil at 65°C. After the reaction
was over, the mixture was allowed to settle for 2 h and the methanol–water mixture
was separated and top layer of methyl ester was removed resulting in a 90% yield.

26
The kinetics of the esterification of free fatty acids (FFA) in sunflower oil with
methanol in the presence of sulphuric acid at concentrations of 5 and 10 wt.% relative
to free acids as catalyst and methanol/oleic acid mole ratios from 10:1 to 80:1 was
studied by Berrios et al (2007). Experimental results showed that a first-order kinetic law
was fit for the forward reaction and a second-order for the reverse reaction. The
influence of temperature on the kinetic constants was determined by fitting the results
to the Arrhenius equation. The energy of activation for the forward reaction decreased
with increasing catalyst concentration from 50.745 to 44.559 J/mol. Based on the
results, a methanol/oleic acid mole ratio of 60:1, a catalyst (sulphuric acid)
concentration of 5 wt. % and a temperature of 60°C provided a final acid value of the
biodiesel lower than 1 mg KOH/g oil within 120 min. and this was a widely endorsed
limit for efficient separation of glycerin and biodiesel during production.

Cvengros and Cvengrosova (2004) have used frying oils or fats (UFO) for the
production of methyl esters (ME) of higher fatty acids as alternative fuels for diesel
engines. They were targeting quality with an acidity number up to 3 mg KOH/g and
water content up to 0:1 wt.% after treatment. They reported that vacuum distillation
evaporator was an effective method for reducing free fatty acids which simultaneously
decreased the content of FFA and water in UFO. Final distillation of raw methyl ester in
an 8ml vacuum evaporator resulted in practically all parameters required by the
standard. Undesirable low-temperature properties of methyl ester derived from UFO
due to higher fraction of saturated acyls could be adjusted by the addition of
depressants.

Demirbas (2009) studied the trans esterification process of vegetable oils in


supercritical methanol without using any catalyst. He found that the most important
variables that may effects methyl ester yield during the trans esterification reaction
were molar ratio of alcohol and the reaction temperature to vegetable oil. Supercritical
methanol has a high potential for both transesterification of triglycerides and methyl
esterification of free fatty acids to methyl ester production for the diesel fuel substitute.

27
He et al (2007) developed a system for continuous transesterification of
vegetable oil using supercritical methanol by using a tube reactor. They observed
increase in the proportion of methanol; reaction pressure and reaction temperature can
enhance the production yield effectively. However, side reactions of unsaturated fatty
acid methyl esters (FAME) occurred when the reaction temperature was found over
300°C, which lead to much loss of material. There was also a critical value of residence
time at high reaction temperature and the production yield got decreased. The optimal
reaction condition under constant reaction temperature process was 40:1 of the molar
ratio of alcohol to oil, 25 min of residence time, 35 MPa and 310°C. Maximum yield of
77% was found in the optimal reaction conditions.

The kinetics in hydrolysis and subsequent methyl esterification was studied by


Minami and Saka (2006) to elucidate reaction mechanism. Fatty acid composition was
found to act as acid catalyst and simple mathematical models were proposed in which
regression curves were fit well with experimental results. Fatty acid was thus concluded
to play an important role in the two-step supercritical methanol process.

Biodiesel was synthesized enzymatically with Novozym-435 lipase in presence of


supercritical carbon dioxide by Rathore and Madras (2007). Effect of reaction variables
such as temperature, molar ratio, enzyme loading and kinetics of the reaction was
investigated for enzymatic synthesis in supercritical carbon dioxide at 200 to 400°C. As a
result very high conversions (>80%) were obtained within 10 minutes and nearly
complete conversions were obtained at within 40 min. However, conversions of only
60–70% were obtained in the enzymatic synthesis even after 8 h were observed.

The conversion process of oil to biodiesel by three types of reaction like trans
esterification, hydrolysis of triglycerides and methyl esterification studied by Susiana
and Saka (2004) for the conversion of water containing vegetable oil proceeded
simultaneously by free supercritical methanol as a catalyst to produce high yield and to
see the effect of water on the yield of methyl esters during conversion process. They
observed that the presence of free fatty acids and water always produced negative
effects which may cause soap formation, consumed catalyst and reduced catalyst
28
effectiveness, all of which resulted in a low conversion. They found that presence of
water at a certain amount could enhance the methyl esters formation. These results
were compared with those of methyl esters prepared by acid and alkaline catalyzed
methods which were found that supercritical methanol, crude vegetable oil as well as its
wastes could be readily used for biodiesel fuel production in a simple preparation.

Gerpen (2005) found conventional processing of oil to biodiesel involving an


alkali catalyzed process unsatisfactory for its cost; high free fatty acid feedstock and
soap formation. He showed that pretreatment processes using strong acid catalysts
provided good conversion yields and high-quality final products. These techniques have
been extended to allow biodiesel production from feedstocks like soap stock that are
often considered to be waste.

Optimum combinations for reducing the acid level of Mahua oil to less than 1%
after pretreatment was 0.32 v/v methanol-to-oil ratio, 1.24% v/v H2SO4 catalyst and 1.26
h reaction time at 60 °C as observed by Ghadge and Raheman (2006). After the
pretreatment of Mahua oil, trans esterification reaction was carried out with 0.25 v/v
methanol-to-oil ratio (6:1 molar ratio) and 0.7% w/v KOH as an alkaline catalyst to
produce biodiesel. Properties of biodiesel prepared from Mahua were analyzed and
matched to the requirements of both the American and European standards.

Five flow sheet options have been reported by Harding et al (2007) in their study
to investigate the alkali and enzyme catalyzed production routes from rapeseed oil, use
of methanol or ethanol for transesterification and the effect of efficiency of alcohol
recovery. They concluded that the enzymatic production route was environmentally
more favorable. Acidification, and photochemical oxidation were reduced by 5% with
considerable benefit on global warming. Certain toxicity levels have been reduced to
more than half. These results were achieved mainly due to lower steam requirements
for heating in the biological process.

He et al (2007) proposed a new technology of gradual heating that can


effectively reduce the loss caused by the side reactions of unsaturated FAME at high

29
reaction temperature with the new reaction technology; the methyl esters yield was
found to be more than 96%.

Cooking oil was mixed with canola oil at various ratios in order to make use of
used cooking oil for production of economical biodiesel by Issariyakul et al (2007).
Methyl and ethyl esters were prepared by KOH-catalyzed trans esterification from the
mixtures of both the oils. Water content, acid value and viscosity of esters after
conversion to biodiesel met ASTM standard except for ethyl esters prepared from used
cooking oil. Although ethanolysis was proved to be more challenging, ethyl esters
showed reduction in the crystallization temperature (−45.0 to −54.4°C) as compared to
methyl esters (−35.3 to −43.0 °C).

Karmee and Chadha (2005) prepared biodiesel from non-edible oil of Pongamia
pinnata by trans esterification of the crude oil using in methanol in the presence of KOH
as catalyst. As a result maximum conversion of 92% (oil to ester) was achieved using a
1:10 molar ratio of oil to methanol at 60°C. Tetrahydrofuran (THF) was used as a co-
solvent which increased the conversion to 95%. Solid acid catalysts viz. Hb-Zeolite,
montmorillonite K-10 and ZnO were also used for transesterification.

Morin et al (2007) studied heteropolyacids (HPA) with kegging structure and


evaluated homogeneous brinstead acid catalysts in the reaction of rapeseed oil trans
esterification with methanol and ethanol at 358 K atmospheric pressure. Rapeseed oil
trans esterification with ethanol over anhydrous keggin HPAs lead to higher conversion
level than H2SO4 compared at equivalent H+ concentration and H2O/H+ molar ratio. This
demonstrated the advantages of strong brinstead acids in vegetable oil
transesterification with ethanol in mild conditions. The proton solvation with water
molecule was shown to be a crucial parameter then Mo samples exhibited higher
activities due to their ability to lose crystallization water at lower temperatures
compared to samples. It was observed that higher trans esterification rates were
obtained with ethanol than methanol in presence of HPA.

30
Nabi et al (2006) converted non-edible Neem oil to biodiesel by trans
esterification and investigated the combustion, exhaust emissions of transesterified
biodiesel-diesel blends with neat diesel fuel.

Production of biodiesel from edible oils like palm and groundnut oil as well as
non-edible oils like Pongamia pinnata and Jatropha curcas was investigated by Rathore
and Madras (2007). They reported that variables affecting the conversion during
transesterification are molar ratio of alcohol to oil, temperature and time. They also
investigated use of supercritical methanol and ethanol without using any catalyst.

Sahoo et al (2007) have standardized method for the extraction of Polanga


(Calophyllum inophyllum L.) seed oil, conversion of oil to biodiesel along with the testing
of physico-chemical and mechanical properties. It was observed that the viscosity of
vegetable oil get reduced substantially after transesterification. The density and
viscosity of the Polanga oil methyl ester formed after triple stage transesterification
were found to be close to those of petroleum diesel oil. The flash point of all the blends
of Pongamia or Polanga oil methyl ester was found higher than that of diesel oil.

Based on the exhaustive engine tests undertaken by Sahoo et al (2007), it was


concluded that Polanga based biodiesel can be adopted as an alternative fuel for the
existing conventional diesel engines without any major modifications in the engine
system. Particularly, 100% biodiesel showed higher flash point than petroleum diesel oil.
All these tests for characterization of biodiesel demonstrated that almost all the
important properties of biodiesel were in very close agreement with the diesel oil
making it a potential fuel for the application in compression ignition engines for
complete replacement of diesel fuel.

Sunflower oil methanolysis was undertaken by Stamenkovic et al (2007) in a


stirred reactor at different agitation speeds. Measurements of drop size, drop size
distribution and the degree conversion demonstrated the effects of the agitation speed
in both non-reaction (methanol/sunflower oil) and reaction (methanol/KOH/sunflower
oil) systems.

31
Talens et al (2007) attempted to decrease the consumption of materials, energy
and promote the use of renewable resources. They suggested use of Energy Flow
Analysis (ExFA) as an environmental assessment tool and applied to the process of
biodiesel production to account wastes, emissions, exergetic efficiency, compare
substitutes and other types of energy sources. As a result they showed that the
production process had a low energy loss (492 MJ). The energy loss was reduced by
using potassium hydroxide/sulphuric acid as process catalysts and it was observed that
loss during biodiesel production can be further minimized by improving the quality of
the used cooking oil.

Factors effecting trans-esterification


Studies have been carried out by Marchetti et al (2007) using different oils as
raw material, different alcohol (methanol, ethanol, butanol) as well as different catalysts
such as sodium hydroxide, potassium hydroxide, sulfuric acid and supercritical fluids and
heterogeneous ones such as lipases enzymes. They evaluated the advantages,
disadvantages and kinetics of reaction.

Meher et al (2006) studied transesterification of Karanj oil in methanol for


production of biodiesel. The reaction parameters such as catalyst concentration,
alcohol/oil molar ratio, temperature and rate of mixing were optimized for production
of Karanj oil methyl ester. Fatty acid methyl esters content in the reaction mixture were
quantified by HPLC and 1H NMR. The yield of methyl esters from Karanj oil under the
optimal condition was 97–98%.

Ngamcharussrivichai et al (2007) studied the heterogeneously catalyzed trans


esterification of palm kernel oil with methanol over various modified dolomites at 60°C.
The modification of dolomite was performed via a conventional precipitation method
using various nitrate salt solutions of alkali earth metals and trivalent metals. Influences
of a variety of metals, calcination temperature of the parent dolomite, methanol/oil
molar ratios, reaction time, catalyst amount, and catalyst reuse were also investigated.
The results indicated that the calcination temperature of the parent dolomite was

32
crucial factor affecting the activity and the basicity of the resulting catalyst. The catalyst
modified from dolomite claimed at 600 and 700˚C, followed by the precipitation from Ca
(NO3)2 and the subsequent calcination at 800°C, exhibited the most active catalyst giving
the methyl ester content as high as 99.9% under the suitable reaction conditions, the
methanol/oil molar ratio of 15, amount of catalyst of 10 wt. % and reaction time of 3 h.

Sharma and Singh (2007) studied biodiesel production from extracted Karanj oil.
Molecular weight of the oil was 892.7(g). Both the acid as well as alkaline esterification
were found to be applicable to get biodiesel. They concluded that NaOH was a better
catalyst than KOH in terms of reaction yield. Maximum yield 89.5% was achieved at 8:1
molar ratio for acid esterification and 9:1 molar ratio for alkaline esterification in
presence of 0.5 wt. % catalysts (NaOH/KOH) with regular mechanical stirring.

Shah and Gupta (2007) evaluated use of lipase enzyme from Pseudomonas
cepacia for conversion of Jatropha oil to biodiesel. Commercial grade ethanol was found
compatible with enzyme-based process. The mono-ethyl esters of the long chain fatty
acids (biodiesel) were prepared by alcoholysis of Jatropha oil by enzyme lipase. The
optimization process consisted of screening of various commercial lipase preparations,
pH tuning, immobilization, varying water content in the reaction media. Varying amount
of enzyme were used and different temperature of the reaction. 98% yield (w/w) was
obtained by using P. cepacia lipase immobilized on celite at 50-58ºC in the presence of
4–5% (w/w) water in 8 h. They reported that this biocatalyst could be useful four times
without loss of any activity.

Tiwari et al (2007) worked on the response surface methodology (RSM) based on


central composite rotatable design (CCRD) was used to optimize the three important
reaction variables that was methanol quantity (M), acid concentration (C) and reaction
time (T) for reduction of free fatty acid (FFA) content of the oil to around 1% as
compared to methanol quantity (M0) and reaction time (T0) and for carrying out trans
esterification of the pretreated oil. Using (RSM), quadratic polynomial equations were
obtained for predicting acid value. The optimum combination for reducing the FFA of
Jatropha curcas oil from 14% to less than 1% was found in 1.43% v/v H2SO4 acid catalyst,
33
0.28 v/v methanol-to-oil ratio and in 88 min. reaction time at a reaction temperature of
60ºC as compared to 0.16 v/v methanol-to-pretreated oil ratio and 24 min. of reaction
time at a reaction temperature of 60ºC for producing biodiesel. This process gave an
average yield of biodiesel more than 99%. The fuel properties of J. curcas biodiesel
obtained were found comparable good to those of diesel and matches American and
European standards.

Vicente et al (2007) worked for the development and optimization of the


potassium hydroxide as a catalyst for the synthesis of fatty acid methyl esters (biodiesel)
from sunflower oil. Variables during the reaction were temperature, initial catalyst
concentration by weight of sunflower oil and the methanol: vegetable oil molar ratio,
with respect to the production of biodiesel purity and yield. It was observed that the
initial catalyst concentration was the most important factor, having a positive influence
on biodiesel purity, but a negative one on biodiesel yield. Temperature has a significant
positive effect on biodiesel purity and a significant negative influence on biodiesel yield.
The methanol: vegetable oil molar ratio was only significant for the biodiesel purity,
having a positive influence. The best conditions demonstrated were 25ºC, 1.3% wt. for
the catalyst concentration and a 6:1 methanol: sunflower oil molar ratio for the higher
yield as well purity of biodiesel.

The factorial design of experiments and a central composite design have been
used by Vicente et al (2007) to evaluate the influence of operating conditions on the
process of trans-esterification of sunflower oil with respect to the yield and the yield
losses due to triglyceride saponification and methyl ester dissolution in glycerol while
the variables studied were temperature, initial catalyst concentration and the methanol:
vegetable oil molar ratio. They observed that the yield increased and yield losses
decreased by decreasing catalyst concentration and temperature. However, the
methanol: sunflower oil molar ratio did not affect the material balance variables
significantly. Second-order models were obtained to predict the yield and both yield
losses.

34
Zhang et al (2003) studied the economic feasibilities of four continuous
processes to produce biodiesel including both alkali and acid-catalyzed processes using
waste cooking oil and the standard process using virgin vegetable oil as the raw
material. They reported that the alkali catalyzed process using virgin vegetable oil had
the lowest fixed capital cost and acid-catalyzed process using waste cooking oil was
found more economically feasible. On the basis of these economic calculations,
sensitivity analyses were done by Zhang et al (2003). Plant capacity and prices of
feedstock oils were found most significant factors affecting the economic viability of
biodiesel manufacture.

Use of biodiesel in engines & emissions


The diesel engine exhaust emissions cause a range of health problems. However,
Demirbas (2006) observed that biodiesel is an environmentally friendly fuel that will be
useful in any diesel engine without modification.

Ideal engine fuel should produce least pollution and at the same time have
higher engine life and optimal performance. How far biodiesel can meet these demands
is a question being investigated by several workers, both engineers and
environmentalists.

Correa and Arbilla (2007) studied use of biodiesel for seven carbonyl emissions
(formaldehyde, acetaldehyde, acrolein, acetone, propionaldehyde, butyraldehyde, and
benzaldehyde) in heavy-duty diesel engine fueled with pure diesel (D) and biodiesel
blends (v/v) of 2% (B2), 5% (B5), 10% (B10), and 20% (B20). Tests were conducted using
a six cylinder heavy-duty engine, typical under 1000, 1500, and 2000 rpm. The exhaust
gases were diluted nearly 20 times and the carbonyls were sampled with SiO2–C18
cartridges, impregnated with acid solution of 2,4-dinitrophenylhydrazine. The chemical
analyses were performed by high performance liquid chromatography using UV
detection. It was reported that by using average values for the three modes of operation
(1000, 1500, and 2000 rpm) benzaldehyde showed a reduction on the emission (3.4%
for B2, 5.3% for B5, 5.7% for B10, and 6.9% for B20) and all other carbonyls showed a

35
significant increase: 2.6, 7.3, 17.6, and 35.5% for formaldehyde; 1.4, 2.5, 5.4, and 15.8%
for acetaldehyde; 2.1, 5.4, 11.1, and 22.0% for acrolein+acetone; 0.8, 2.7, 4.6, and 10.0%
for propionaldehyde; 3.3, 7.8, 16.0,and 26.0% for butyraldehyde.

An experimental program examining performance and emissions from spark and


compression-ignition engines, running on a variety of bio-fuels including simulated bio-
gas and commercial seed oil was presented in engines. Both engines used were single-
cylinder laboratory type engines of comparable power output having variable speed and
load capability, the spark-ignition engine additionally having variable compression ratio.
For bio-gas, containing carbon dioxide, emissions of oxides of nitrogen were reduced
relative to natural gas while un-burnt hydrocarbons were increased. Brake power and
specific fuel consumption changed little and carbon monoxide was predominantly
affected by air: fuel ratio. As a result similar effects were demonstrated with nitrogen
replacing carbon dioxide in the simulated bio-gas; similar trends were evident as
compression ratio was increased. It was observed that seed-oil biofuel gave similar
performance to diesel fuel without major disadvantages, other than an increase in
specific fuel consumption. Tests with cetane and rape-seed methyl ester bio-diesel are
also presented for comparison.

Crookes (2006) reported that specific fuel consumption was about the same and
specific NOx emissions were lower with bio-fuel than results from the spark-ignition
engine tests running on biogas.

Supercritical methanol has a high potential for both trans esterification of


triglycerides and methyl esterification of free fatty acids to methyl esters for a diesel
fuel substitute as reported by Demirbas (2006). Biodiesel production increased the yield
percentage by 95% in 10 min when supercritical methanol was used for trans
esterification. Viscosity values of vegetable oils were found between 27.2 and 53.6
mm2/s, whereas those of vegetable oil methyl esters was between 3.59 and 4.63 mm2/s.

Flash point values of vegetable oil methyl esters were found to be much lower
than those of vegetable oils in studies conducted by Demirbas (2006). An increase in

36
density from 860 to 885 kg/m3 for vegetable oil methyl esters or biodiesels increases the
viscosity from 3.59 to 4.63 mm2/s.

Demirbas (2007) reported that exhaust emissions of carbon monoxide (CO) from
biodiesel were 50% lower than CO emissions from petro diesel. Exhaust emissions of
particulate matter (PM) from biodiesel were 30% lower than over all particulate matter
emissions from petro diesel. It was also observed that biodiesel emission may have a
slight increase or decrease in nitrogen oxides depending on engine family; there was a
decrease in the levels of polycyclic aromatic hydrocarbons (PAH) compounds along with
nitrited PAH compounds (Demirbas, 2007) identified as potential cancer causing
compounds.

One of the largest studies of biodiesel in both on-road, off-road uses and the
testing was conducted for the military and encompassed a wide range of application
types including two medium-duty trucks, two Humvees, a heavy-duty diesel truck, a bus,
two stationary backup generators (BUGs), a forklift and an airport tow vehicle by Durbin
et al (2007). The full range of fuel testing included a California ultra-low sulfur diesel
(ULSD) fuel, different blend ratios of two different yellow-grease biodiesels, one soy-
based biodiesel, JP-8 and yellow-grease biodiesel blends with two different NOx
reduction additives. The B20-YGA, B20-YGB and B20-Soy did not show trends relative to
ULSD. Higher biodiesel blends, tested only in one vehicle, showed a tendency for higher
total hydrocarbons (THC), carbon monoxide (CO) emissions and lower particulate matter
(PM) emissions.

Methyl and ethyl esters from the oil of Jatropha curcas seeds were prepared and
the fuel properties of both ester fuels were determined according to existing standards
for biodiesel by Foidl et al (1996). Jatropha oil and blends of Jatropha biodiesel with
diesel in proportions of 97.4%/2.6%; 80%/20%; and 50%/50% by volume were tested on
a single-cylinder direct-injection engine by Forson (2004). The results covered a range of
operating loads on the engine. Brake specific fuel consumption, brake power, brake
thermal efficiency, engine torque, concentrations of carbon monoxide, carbon dioxide
and oxygen in the exhaust gases were tested as mechanical properties and were found
37
similar for all fuels. 97.4% diesel/2.6% Jatropha fuel blend was observed as ideal since it
was lower net contributor to the atmospheric level by producing highest cetane
number, maximum values of the brake power and brake thermal efficiency as well as
minimum values of the specific fuel consumption. The trend of carbon monoxide
emissions was similar for the fuels but diesel fuel showed slightly lower emissions to the
atmosphere. The test showed that Jatropha oil could be conveniently used as a diesel
substitute in a diesel engine. The test further showed increase in brake thermal
efficiency, brake power and reduction of specific fuel consumption for Jatropha oil and
its blends with diesel. It was concluded that biodiesel can be used as an ignition-
accelerator additive for diesel fuel and showed even better engine performance than
the diesel fuel.

Hebbal et al (2006) selected Deccan hemp oil and non-edible vegetable oil for
the test on a diesel engine and its suitability as an alternate fuel. The viscosity of Deccan
hemp oil was reduced by blending with diesel in 25/75%, 50/50%, 75/25%, 100/ 0% on
volume basis; then analyzed and compared with diesel. Further blends were heated and
effect of viscosity on temperature was studied. The performance and emission
characteristics of blends were evaluated at variable loads at a constant rated speed of
1500 rpm and their results were compared with diesel. The thermal efficiency, brake
specific fuel consumption, brake specific energy consumption (BSEC) was comparable
with diesel; however, emissions were a little higher for 25% and 50% blends. At rated
load, smoke, carbon monoxide (CO) and un-burnt hydrocarbon (HC) emissions of 50%
blend were higher compared with diesel by 51.74%, 71.42% and 33.3%, respectively.
Pure Deccan hemp oil results were compared with the results of Jatropha and Pongamia
oil for similar works available in the literature and were well comparable. From
investigation it was suggested that up to 25% of blend of Deccan hemp oil without
heating and up to 50% blend with preheating can be substituted for diesel engine
without any engine modification.

Corrosion characteristics of biodiesel are important for long term durability of


engine parts and very little information is available on this aspect. Kaul et al (2007)

38
assessed corrosion of synthesized biodiesel from the non-edible oils of Jatropha curcas,
Madhuca indica and Salvadora oleoides. They found that use of biodiesel from above
mentioned oil will lead to drastic reduction in sulphur content and increase in cetane
number which, in turn, will adversely affects the lubricity characteristics of the diesel
fuel. Using long duration static immersion test method corrosion studies on engine parts
like piston metal and piston liner were carried out by them with neat diesel procured
from one of the Indian refinery. Biodiesel from Salvadora biodiesel showed marked
corrosion on both metal parts of diesel engine whereas biodiesel from other oils
showed little or/no corrosion as compared to neat diesel (Kaul et al, 2007).

Lapuerta et al (2007) analyzed the effect of biodiesel fuels on diesel engine


emissions. The comparison was to maintain engine performance by analyzing the effect
of biodiesel fuel on engine power, fuel consumption and thermal efficiency. Highest
consensus lies in an increase in fuel consumption in approximate proportion to the loss
of heating value. Engine emissions from biodiesel and diesel fuels were compared, for
emissions: nitric oxides and particulate matter. They reported a sharp reduction in
particulate emissions.

Nabi et al (2006) investigated neat diesel fuel and diesel–biodiesel blends in a


four stroke naturally aspirated (NA) direct injection (DI) diesel engine. Comparison with
conventional diesel fuel, diesel–biodiesel blends showed emission of lower carbon
monoxide (CO) and smoke emissions but higher oxides of nitrogen (NO x) emission.
However, comparison with the diesel fuel, NOx emission with diesel–biodiesel blends
was found slightly reduced when exhaust gas recirculation was applied.

High viscosity of J. curcas oil has been considered as a potential advantage for
compression ignition (C.I.) engine Pramanik (2003) studied blends of Jatropha curcas oil
with diesel. Blends of varying proportions of J. curcas oil diesel were analyzed and
compared with diesel fuel. Effect of temperature on viscosity of Jatropha oil, blends and
biodiesel was studied and performance of the engine was evaluated in a single cylinder
C.I. engine compared with the performance obtained with diesel. Significant
improvement in engine performance was observed compared to vegetable oil alone.
39
The specific fuel consumption and the exhaust gas temperature were reduced due to
decrease in viscosity of the vegetable oil. Acceptable thermal efficiencies of the engine
were obtained with blends containing up to 50% volume of Jatropha oil. From the
properties and engine test it was recommended that 40–50% of oil can be substituted
for diesel without any engine modification and preheating of the blends.

Sufficient amount of trans-esterified oil esters was prepared from Karanj oil
which was used to run the farm engines (3.73 kW) for at least 8 h and composition of
fatty acids of Karanj oil studied by Raheman and Phadatare (2004). They compared the
diesel engine emissions and performance of Karanj methyl ester and diesel and
recommended Karanj oil as a good substitute.

A 5.2 kW diesel engine with alternator was used to test J. curcas biodiesel and its
blends with conventional commercial diesel fuel. A biodiesel pilot plant was developed
and used for biodiesel production from Jatropha oil. The fuel properties of Jatropha
biodiesel were found to be similar to the diesel fuel. In the case of Jatropha biodiesel
alone, the fuel consumption was about 14 per cent higher than that of diesel as
investigated by Ramesh and Sampathrajan (2008). The percent increase in specific fuel
consumption ranged from 3 to 14 for B20 to B100 fuels. The brake thermal efficiency for
biodiesel and its blends was found to be slightly higher than that of diesel fuel at tested
load conditions and there was no difference found between the biodiesel and its
blended fuel efficiencies. For Jatropha biodiesel and its blended fuels, the exhaust gas
temperature increased with increase in load and amount of biodiesel.

Carbon monoxide reduction by biodiesel was 16, 14 and 14 percent, respectively


at 2, 2.5 and 3.5 kW load conditions. The NOx emission from biodiesel increased by 15,
18 and 19 percent higher than that of the diesel fuel at 2, 2.5 and 3.5 kW load
conditions respectively.

Methyl ester (ME) of Pongamia (P), Jatropha (J) and Neem (N) were derived
through trans esterification process. Experimental investigations were carried out by
Rao et al (2008) to examine properties, performance and emissions of different blends

40
(B10, B20, and B40) of PME, JME and NME in comparison to diesel. Results indicated
that B20 have closer performance to diesel and B100 had lower brake thermal efficiency
mainly due to its high viscosity compared to diesel. However, its diesel blends showed
reasonable efficiencies, lower smoke, CO and HC. Pongamia methyl ester gave better
performance as compared to Jatropha and Neem methyl esters in their study.

A single cylinder, constant speed, direct injection diesel engine was operated on
neat Jatropha oil. Injection timing, injector opening pressure, injection rate and air swirl
level were changed to study their influence on performance, emissions and combustion.
Results have been compared with neat diesel operation in a study undertaken by Reddy
and Ramesh (2006). The injection timing was varied by changing the position of the fuel
injection pump with respect to the cam; and injection rate was varied by changing the
diameter of the plunger of the fuel injection pump. A properly oriented masked inlet
valve was employed to enhance the air swirl level. Advancing the injection timing from
the base diesel value and increasing the injector opening pressure increase the brake
thermal efficiency and reduce HC and smoke emissions significantly. Enhancing the swirl
had only a small effect on emissions. The ignition delay with Jatropha biodiesel was
always found higher than that of diesel under similar conditions. Improved premixed
heat release rates were observed with Jatropha when the injector opening pressure was
enhanced. When the injection timing was retarded with enhanced injection rate, a
significant improvement in performance and emissions was noticed. In this case
emissions with Jatropha biodiesel were even lower than diesel. At full output, the HC
emission level is 532 ppm as against 798 ppm with diesel. NO level and smoke with
Jatropha biodiesel were found to be 1162.5 ppm and 2 BSU while they were 1760 ppm
and 2.7 BSU with diesel as reported by them.

It was observed that methyl ester of Karanj oil had slightly reduced thermal
efficiency as compared to diesel. The brake specific fuel consumption, exhaust gas
temperature and HC, CO and NO emission of methyl ester of Karanj oil was slightly
higher as compared to diesel in a study undertaken by Srivastava and Verma (2007). It
was observed that almost all properties of the methyl ester of Karanj oil were found

41
quite closer to those of the diesel oil. Therefore, they concluded that methyl ester of
Karanj oil can be used as an alternative renewable source of energy.

Sahoo et al (2007) reported that 0.65% by volume H2SO4 and a molar ratio of 6:1
gave maximum conversion efficiency of free fatty acids to triglycerides and thereby
reducing the acid value of the product below 4 mg KOH/g in acid based reaction and in a
same manner molar ratio of 9:1 and the 1.5% by weight of potassium hydroxide was
found to give the maximum ester yield for reaction duration of 4 h in the case of alkali
trans-esterification reaction. This diesel was examined for engine performance without
any engine hardware modifications. The 100% biodiesel was found to be ideal since it
improved the thermal efficiency, brake specific energy consumption of the engine by
0.1% and the exhaust emissions were reduced. Smoke emissions also reduced by 35%
for B60 as compared to neat petro-diesel. The objective of this study was to ascertain
suitability of these fuels for engine application. Based on the exhaustive engine tests, it
was concluded that polanga based biodiesel can be adopted as an alternative fuel for
the existing conventional diesel engines without any major hardware modifications in
the system because it has acceptable physico-chemical as well as mechanical properties
with the diesel oil.

The density and viscosity of the Polanga oil methyl ester formed after triple
stage trans esterification were found to be close to those of petroleum diesel oil. The
flash point of all the blends of Pongamia or polanga oil methyl ester was found higher
than that of diesel oil. Based on the exhaustive engine tests undertaken by Sahoo et al
(2007), it was concluded that polanga based biodiesel can be adopted as an alternative
fuel for the existing conventional diesel engines without any major modifications in the
engine system. Particularly, 100% biodiesel showed higher flash point than petroleum
diesel oil. All these tests for characterization of biodiesel demonstrated that almost all
the important properties of biodiesel were in very close agreement with the diesel oil
making it a potential fuel for the application in compression ignition engines for
complete replacement of diesel fuel.

42
Sarin et al (2007) reported that the Jatropha biodiesel blended with palm methyl
ester leads to a composition having efficient and improved low temperature property as
well as good oxidation stability. They reported that Jatropha biodiesel itself has poor
oxidation stability with good low temperature properties but the palm biodiesel has
good oxidative stability and poor low temperature properties and the combinations of
Jatropha and palm gave an additive effect on these two critical properties of biodiesel.
The stability of biodiesel was very critical and biodiesel requires antioxidant to meet
storage requirements. In order to meet specification, some concentration of antioxidant
was required for biodiesel (except palm biodiesel), which is much higher than that
required for petroleum diesel. In order to minimize the dosage of antioxidant
appropriate blends of Jatropha and palm biodiesel were made. According to their
studies antioxidant dosage could be reduced by 80–90%, if palm oil biodiesel is blended
with Jatropha biodiesel at around 20–40% concentration. Since palm biodiesel has poor
low temperature properties like cloud point and pour point, the blending of Jatropha
biodiesel improves the same. Therefore, optimum mixture of Jatropha biodiesel with
palm biodiesel can lead to a synergistic combination with improved oxidation stability
and low temperature property.

An 80,000 km durability test was performed on two engines using diesel and
biodiesel (made up of waste cooking oil) as fuel in order to examine emissions by the
use of biodiesel by Yang et al (2007). The testing biodiesel (B20) was blended with 80%
diesel and 20% methyl ester derived from waste cooking oil. The emissions of CO, HC,
NOx, particulate matter (PM) and polycyclic aromatic hydrocarbons (PAHs) were
measured using identical-model engines installed on a standard dynamometer equipped
with a dilution tunnel used of regulated air pollutants. It was observed that HC, CO and
PM emission levels were lower in engine with the (B20) blend than those for diesel in 0
km of the durability test after running for 20,000 km and longer, they were higher.
However, the deterioration coefficients for these regulated air pollutants were not
statistically higher and did not increase significantly after 80,000 km of driving. Total
(gaseous+particulate phase) PAH emission levels for both B20 and diesel decreased as

43
the driving mileage accumulated. However, for the engine using B20 fuel, particulate
PAH emissions increased as engine mileage increased. Finally result showed that B20
use can reduce both PAH emission and its corresponding carcinogenic potency.

Technologies for Biodiesel


Aliske et al (2007) studied development and consolidation of automotive ethanol
technology in the last thirty years. They presented a method of measurement to cover
the full range of mixtures (0–100%) of biodiesel in diesel oil using mid infrared
spectroscopy. According to them the carbonyl absorption peak, which was present only
in biodiesel presents power law proportionality with mixture percentile of 0.889 for
peak area and 0.841 for peak height. They reported that it is possible to measure
mixture percentiles of biodiesel in diesel oil even at very low concentrations allowing
quality control and law enforcement as studied by them.

Demirbas (2007) reviewed the production of modern biomass based


transportation fuels such as fuels from Fischer–Tropsch synthesis, bioethanol, fatty acid
methyl ester, bio methanol and bio hydrogen as a result he observed that production of
different types of biofuel (biodiesel, fatty acid methyl ester, bio methanol and bio
hydrogen) can be produced by using biomass and production may become easier by
adopting Fischer–Tropsch synthesis pathway.

Fernando et al (2007) identified possibilities of blending glycerol and glycerol


based co-products, such as propanediol and propanol with gasoline as oxygenates. The
study revealed that there was a possibility to use glycerol and its derivatives with
gasoline as an automotive fuel because with the increasing production of biodiesel a
glut of glycerol will be created, causing market prices to plummet. This situation
warrants finding alternative uses for glycerol. Octane number and heating value of
different mixtures of gasoline, ethanol, glycerol and its derivatives were also
determined.

Integrated utilization, suitability of rapeseed oil for biodiesel production as well


as parallel use of its solid residues for energy and second generation biofuels production

44
has been studied by Zabaniotou et al (2007). Fast pyrolysis at high temperature and
fixed bed air gasification of the rapeseed residues was studied by them. They carried out
thermo gravimetric analysis and kinetic studies and reported that high temperature
pyrolysis could produce higher yields of syngas and hydrogen production comparing to
air fixed bed gasification.

Biodiesel obtained from vegetable oils has been considered as a promising


option. Barnwal and Sharma (2005) have reported on biodiesel production from
vegetable oils, utilization, resources available, process developed, performance in
existing engines and environmental considerations, the economic aspect, advantages in
and barriers to the use of biodiesel.

Biodiesel in developing countries is usually produced from food-grade vegetable


oils that are more expensive than diesel fuel. Therefore, biodiesel produced from food-
grade vegetable oil is currently not economically feasible. Waste cooking oils, restaurant
grease and animal fats are potential feedstock for biodiesel. These inexpensive
feedstock represent one-third of the US total fats and oil production but are currently
devoted mostly to industrial uses and animal feed. Characteristics of feedstock are very
important during the initial research and production stage. Free fatty acids and moisture
reduces the efficiency of trans esterification in converting these feedstock into
economically feasible biodiesel.

Canakci (2007) determined level of contaminants in feedstock samples. He found


levels of free fatty acids varied from 0.7% to 41.8% and moisture from 0.01% to 55.38%
in different feedstock. These wide ranges indicate that an efficient process for
converting waste grease and animal fats must tolerate wide range of feedstock
properties.

Cvengros and Cvengrosovia (2004) studied available capacity used for frying oils
or fats (UFO) that they found an attractive raw material for the production of methyl
esters (ME) of higher fatty acids as alternative fuels for diesel engines. They developed
some simplified methods for the quality control of UFO and ME. The conversion of acyl

45
glycerols to ME was monitored by GLC with a packed column, where the peak areas of
ME were found within an effective range.

Haas et al (2006) developed a computer model to estimate the capital and


operating costs of a moderately sized industrial biodiesel production facility. The major
process operations in the plant were continuous process vegetable oil trans-
esterification for ester and glycerol recovery. The model was designed using
contemporary process stimulation software, current reagent, equipment, supply costs,
following current production practices. An analysis of the dependence of production
costs on the cost of the feedstock indicated a direct linear relationship between the
two, with a change of US$0.020/l ($0.075/gal) in product cost per US$0.022/kg
($0.01/lb.) change in oil cost. Process economics included recovery of co-product
glycerol generated during biodiesel production and its sale into the commercial glycerol
market as an 80% w/w aqueous solution, which reduced production costs by 6%. The
production cost of biodiesel was found to vary inversely and linearly with variations in
the market value of glycerol increasing by US$0.0022/l ($0.0085/gal) for every US
$0.022/kg ($0.01/lb.) reduction in glycerol value. The model was designed in such a way
and rather flexible way that it can be modified to calculate the effects on capital and
production costs of changes in feedstock cost, changes in the type of feedstock
employed, changes in the value of the glycerol co-product and changes in process
chemistry and technology.

Fuel stability and fuel transportation


Demirbas (2007) studied and reviewed modern biomass-based transportation
fuels such as fuels from Fischer–Tropsch synthesis, bioethanol, fatty acid methyl ester,
bio methanol and bio hydrogen. Here, the term biofuel is referred to as liquid or
gaseous fuels for the transport sector that are predominantly produced from biomass.
They included energy security reasons, environmental concerns, foreign exchange
savings and socioeconomic issues related to the rural sector. They opined that modern
biomass can be used for generation of electricity and heat. Bioethanol, biodiesel and
diesel produced from biomass by Fischer–Tropsch synthesis are the most modern
46
biomass-based transportation fuels. Bio-ethanol is a petrol additive/substitute. It is
possible that wood, straw and even household wastes may be economically converted
to bio-ethanol. Currently crops generating starch, sugar or oil are the basis for transport
fuel production. According to his research low blends of biodiesel can be used in any
normal internal combustion diesel engine with no modifications. Higher blends of
biodiesel (over 20%) may require minor modifications, since biodiesel properties are
close to that of diesel fuels. He characterized biodiesel by determining its viscosity,
density, cetane number, cloud and pour points, characteristics of distillation, flash and
combustion points and higher heating value (HHV) according to ISO norms. He opined
that biodiesel use can reduce carbon dioxide, carbon monoxide, PAHs, nitrated PAHs
emissions. Use of biodiesel also decreases solid carbon fraction of particulate matter
and reduces the sulfate fraction, while the soluble, or HC, fraction stays the same or
increases. Emissions of nitrogen oxides increase with the concentration of biodiesel in
the fuel. Some biodiesel do produce more nitrogen oxides than others, and some
additives have shown promise in modifying the increases.

Wood (2005) illustrated that transport fuel almost exclusively from finite
reserves of fossil fuels, currently accounts for around 25% of greenhouse gas emissions
in many European countries. The introduction of low carbon biodiesel could help to
deliver the necessary emissions cuts however EU partner countries can only grow a
portion of the feedstock required to deliver a secure, low cost supply of biodiesel.
Europe is looking towards the developing world for its biodiesel supplies and that this
could offer a significant opportunity for agriculture in developing economies.

47
Chapter 3

Materials and Methods

Seed material
Approximately one kilogram seeds capsule/fruits were collected from each
candidate plus trees to study and provide potentially useful genetic variation. Seeds of
collected fruit capsules were separated through manual threshing and cleaned through
winnowing. Damaged seeds were discarded before seeds in good condition were
cleaned, de-shelled and dried at high temperature of 100–105˚C for 35 min. All moist
seed lots were dried under similar temperature and humidity conditions to reach
constant weight and stored in muslin bags at ambient conditions.

Samples of collections from North Indian region 1) Rajasthan 2) Uttaranchal


were submitted to NBPGR along with passport datasheet for IC number allotment.
Mean of triplicate values were recorded for each accession for seeds morphological
parameters (seed wt., kernel wt., seed coat wt., seed length and width.

Oil extraction
The seed kernels were ground, using a mechanical grinder, and defatted in a
Soxhlet apparatus, using hexane (boiling point of 40–60˚C). The extracted lipid was
obtained by filtrating the solvent lipid to remove solid from solvent before the hexane
was removed using rotary evaporator apparatus at 40˚C. Extracted seed oil was stored
in freezer at −2˚C for subsequent physicochemical analysis.

Quantitative and qualitative (physico-chemical properties (Refractive index, Sp.


gravity, Density) of oil were analyzed.

Chemical and Physical analysis of seed oil

Oil Content
The weight of oil extracted from 10 g of seeds powders was measured to
determine the lipid content.

Result was expressed as the percentage of oil in the dry matter of seed powders.

48
Acid value, % FFA
Acid value of seed oil was determined according to AOAC Official Method Cd 3a-
63. Percentage free fatty acids (FFAs) were calculated using oleic acid as a factor.

Iodine value
Iodine value of seed oil was determined according to AOAC Official Method
993.20.

Saponification value
The saponification value was determined according to MPOB Official Test
Method 2004.

Peroxide value
The peroxide value was determined according to AOAC Official Method 965.33

Viscosity
Viscosity of seed oil was carried out using Brookfield RV-I. Spindle of S03 was
used at 10 rpm at room temperature.

Density
The density of the samples was determined at 25 °C by using density meter.

Method for study of developmental stages


J. curcas plants were grown in the research field of National Botanical Research
Institute, Lucknow. In order to study different developmental stages, developing
embryos were collected 7 days after fertilization at an interval of 5 days. Six such stages
were thus collected. The seventh lot was collected after the fruits matured on the plant
and were nearly dry. The first stage represents fertilized embryos that were dissected
out of the developing fruits. The developing seeds were separated from fruits in all the
seven stages (Stage I-VII). Three such sets were collected as three replications. Seed
area was measured using area meter (Delta-T Devices Ltd., Burwell, Cambridge,
England). Seed size, fresh weight and moisture content were monitored. Seeds of all the

49
seven stages were dried using (Heto LyoPro 6000) freeze drier for removing the
moisture. Moisture content of the seeds was determined by drying the seeds in an oven
at 60 °C till constant weight. Dried samples were weighed and extracted thrice with
hexane using tissue homogenizer (Kinematica Polytron Homogenizer PT 6100).
Combined extracts were filtered using Whatman filter paper No.1 and concentrated
under reduced pressure using rotovap (Laborota 4000, Heidolph, Germany). Oil content
in all the stages was calculated on the basis of dry weight of the seeds.

Optimization of oil extraction from Jatropha seeds:


The effect of five main factors: type of solvent, temperature, solvent to solid
ratio, particle size

of the meal and reaction time were investigated to optimize the extraction operating
conditions for achieving maximum oil yield.
% Oil content = weight of extracted oil (g) / weight of sample (g) × 100

Twenty g of grinded meal was extracted with different non polar solvent namely
n-hexane and petroleum ether. The extraction temperature was varied from room
temperature (26°C) to boiling point of the solvent while the reaction time was varied
between 4-8 h. The solvent to solid ratio was investigated from 3:1-6:1 and seed size
was fixed at three sizes namely below 0.5, 0.5-0.75 and above 0.75 mm. These
parameters were varied one at a time to identify the optimum conditions for each type
of solvent. At the end of the extraction, the micelle was filtered using a vacuum
filtration to remove suspended solids. Subsequently, the solvent was separated from
the oil using rotary vacuum evaporator (Buchi Evaporator) and was collected in the
receiving flask. The oil that remained in the sample flask was weighed after the process
was completed. The percentage of extracted oil was calculated by dividing the amount
of obtained oil by the amount of the seeds multiply by 100. All experiments were
repeated thrice.

50
Fatty Acid Compositions
Fatty acid composition of seed oil was determined using Agilent 6890 series gas
chromatography (GC) equipped with flame ionization detector and capillary column
(30m×0.25mm×0.25mm). About 0.1 ml oil was converted to methyl ester using 1ml
NaOMe (1 M) in 1ml hexane before being injected into the GC. The detector
temperature was programmed at 240ºC with flow rate of 0.8 ml/min. The injector
temperature was set at 240ºC. Hydrogen was used as the carrier gas. The identification
characteristic and composition of Jatropha Curcas oil was achieved by retention times
by means of comparing them with authentic standards analyzed under the same
conditions.

TAGs Composition
TAGs profile of Jatropha oil was determined by using high-performance liquid
chromatography (HPLC) equipped with ELSD 800 detector (Altech). The TAGs of the oil
was separated using commercially column, Inertsil ODS 3 (250mm x 4.6mm) The mobile
phase was a mixture of acetonitrile: diclhrohometane (60:40) set at a flow rate of 0.8
ml/min, with pressure 2.3 bar. TAG peaks were identified based on the retention time of
available commercial TAGs standard.

Details of methods are also given in respective chapters.

51
Chapter 4

Screening of germplasm

Collection of germplasm
An extensive survey was conducted during January to March 2006 and January
to February 2007 to select candidate plus trees (CPTs) of J. curcas from different
locations of Uttaranchal and Rajasthan (India). The selection was made on phenotypic
assessment of characters of economic interest i.e., yield potential, number of fruits per
cluster, crown spread, girth, disease resistance etc. Total 30 morphologically superior
trees from Rajasthan and Uttaranchal provenances were selected. 20 accessions were
collected from different regions of Rajasthan and 10 from Uttaranchal region.
Morphological observations were recorded. Collection code was allotted for each
germplasm before quantitative and qualitative analysis of collected germplasm seeds as
shown in Table 4.1. Accessions for the collections were obtained from NBPGR, New
Delhi by depositing the collections who then provided Indian Collection (IC) numbers
and are reported in Table 4.1.

Table 4.1: Collection of accessions of Jatropha curcas


PROVENANCES PLACE OF COLLECTION I.C. NO CODE
RAJASTHAN SAIRA (2) COMMERCIAL 471332 NBRI JB-1
SAIRA 3 471333 NBRI JB-2
DILWADA 471339 NBRI JB-3
MOUNT ABU 471319 NBRI JB-4
CHOTI UNDRI 471301 NBRI JB-5
SAIRA1 471331 NBRI JB-6
MOUNT ABU 471309 NBRI JB-7
SHISHIRMA GAON 468911 NBRI JB-8
BADI UNDRI 471303 NBRI JB-9
THOOR GAON 468912 NBRI JB-10
BAGHPURA 468913 NBRI JB-11
UDAIPUR 468906 NBRI JB-12
UDAIPUR 471122 NBRI JB-13
UDAIPUR 468908 NBRI JB-14
SIROHI 468907 NBRI JB-15
GOGUNDA 471314 NBRI JB-16
GOGUNDA 471315 NBRI JB-17
ODAN 471316 NBRI JB-18
KHAMNOR 471318 NBRI JB-19
ODAN 471317 NBRI JB-20

52
UTTARANCHAL PANTNAGAR-1 NBRI JB-21
PANTNAGAR-2 NBRI JB-22
PAURI 415764 NBRI JB-23
HAPPRC,GARHWAL 415765 NBRI JB-24
SHERGARH 415772 NBRI JB-25
GARHWAL,PAURI 415766 NBRI JB-26
GARHWAL 415767 NBRI JB-27
BACHELIKHAL 415782 NBRI JB-28
TEHRI 415783 NBRI JB-29
HNB,GARHWAL 415784 NBRI JB-30

Seed morphological traits


Seed weight (Table 4.2) ranged from 0.42 g to 0.76 g in various accessions with
an average of 0.63 g (standard deviation 0.08). It was a very wide range. The ratio of
lightest seed to the heaviest was nearly 1:1.8. Similarly kernel weight amongst 30
accessions ranged from 0.28 to 0.47 g with an average of 0.39 (standard deviation 0.06).
The ratio of lightest kernel to the heaviest was nearly 1:1.67. There was no correlation
of seed and kernel weight. The slope was negative with a value of -0.24. Heaviest seeds
were collected from Pauri of Uttaranchal while the lightest ones were procured from
market sources of Saira, Rajasthan. It is interesting to observe that seeds from Pauri of
Uttaranchal were not only heaviest; their kernel too was among the heaviest ones. The
seed coat weight in this lot was also one of highest (0.35 g). This lot can be selected on
the basis of seed weight and kernel weight. High seed coat weight is not a desirable
character.

Table 4.2: Seed phenology


# Collection site I.C. # Code Seed Kernel Seed coat Seed / Seed Seed Ratio
wt./seed wt./seed wt./seed kernel length breadth length:
(g) (g) (g) ratio (mm) (mm) breadth

1 SAIRA 3 471333 NBRI 0.54 0.28 0.23 1.89 16.96 10.43


JB-2 1.6
2 KHAMNOR 471318 NBRI 0.68 0.43 0.24 1.59 18.06 8.63
JB-19 2.1
3 SHERGARH 415772 NBRI 0.56 0.31 0.24 1.76 16.73 8.03
JB-25 2.1
4 MOUNT ABU 471319 NBRI 0.59 0.32 0.23 1.83 17.03 10.70
JB-4 1.6
5 BADI UNDRI 468906 NBRI 0.55 0.32 0.21 1.73 17.13 10.80
JB-9 1.6
6 ODAN 471316 NBRI 0.69 0.43 0.23 1.59 15.16 10.76
JB-18 1.4
7 SIROHI 468907 NBRI 0.60 0.38 0.24 1.56 17.03 10.66
JB-15 1.6
8 HAPPRC, 415764 NBRI 0.72 0.46 0.25 1.56 17.43 9.20 1.9

53
GERHWAL JB-24
9 THOOR GAON 468912 NBRI 0.66 0.47 0.27 1.41 17.60 11.13
JB-10 1.6
10 SAIRA1 471331 NBRI 0.76 0.44 0.23 1.65 18.70 11.06
JB-6 1.7
11 MOUNT ABU 471309 NBRI 0.71 0.44 0.25 1.59 17.26 11.13
JB-7 1.6
12 SAIRA (2) 471332 NBRI 0.42 0.34 0.16 1.24 15.43 10.73
COMMERCIAL JB-1 1.4
13 TEHRI 415783 NBRI 0.66 0.41 0.23 1.58 16.66 10.03
JB-29 1.7
14 GOGUNDA 471315 NBRI 0.59 0.36 0.23 1.67 18.90 10.46
JB-17 1.8
15 DILWADA 471339 NBRI 0.51 0.33 0.21 1.50 18.76 11.26
JB-3 1.7
16 GOGUNDA 471314 NBRI 0.55 0.32 0.22 1.70 18.23 11.40
JB-16 1.6
17 BACHELIKHAL 415782 NBRI 0.53 0.30 0.25 1.75 17.60 10.60
JB-28 1.7
18 CHOTI UNDRI 471301 NBRI 0.54 0.30 0.22 1.76 17.10 11.33
JB-5 1.5
19 SHISHIRMA 468913 NBRI 0.66 0.39 0.30 1.66 16.53 10.80
GAON JB-8 1.5
20 PANTNAGAR- NBRI 0.73 0.47 0.35 1.55 19.43 10.93
1 JB-21 1.8
21 GERHWAL 415767 NBRI 0.71 0.46 0.35 1.52 18.20 11.23
JB-27 1.6
22 PANTNAGAR- NBRI 0.54 0.34 0.19 1.59 18.86 10.60
2 JB-22 1.8
23 UDAIPUR 468908 NBRI 0.69 0.43 0.28 1.60 18.43 12.56
JB-14 1.5
24 PAURI 415764 NBRI 0.76 0.47 0.35 1.59 19.26 11.20
JB-23 1.7
25 HNB,GERHWA 415784 NBRI 0.66 0.44 0.24 1.49 19.30 11.73
L JB-30 1.6
26 GERHWAL,PA 415766 NBRI 0.69 0.42 0.24 1.65 19.30 10.93
URI JB-26 1.8
27 UDAIPUR 468906 NBRI 0.67 0.42 0.23 1.57 18.73 10.23
JB-12 1.8
28 BAGHPURA 468913 NBRI 0.58 0.36 0.21 1.61 15.86 10.06
JB-11 1.6
29 ODAN 471317 NBRI 0.64 0.42 0.23 1.52 19.43 11.30
JB-20 1.7
30 UDAIPUR 471122 NBRI 0.69 0.43 0.23 1.60 17.16 10.30
JB-13 1.7

Accession from Saira 3, Rajasthan had very light kernel with a weight of only 0.29
g. The lightest seed coat was observed in the seeds from commercial sample of Saira,
Rajasthan (0.16 g).

The seed coat though considered a very trivial trait but is crucial for determining
potential of oil percentage in a given accession. It ranged from 0.16 g to 0.35 g with an
average of 0.24 (standard deviation 0.04). In some accessions the seed coat was nearly

54
as heavy as kernel weight and hence could be a negative trait for selection of
germplasm (Figure 4.1).

Table 4.3: Statistical analysis of seed phenology


Ratio of
Seed
Seed Kernel Seed / Seed Seed seed
coat
Function wt./seed wt./seed kernel length breadth length
wt./seed
(g) (g) ratio (mm) (mm) to
(g)
breadth
Max 0.76 0.47 0.35 1.89 19.43 12.56 2.1
Min 0.42 0.28 0.16 1.24 15.16 8.03 1.4
Avg 0.63 0.39 0.24 1.61 17.74 10.67 1.7
STD 0.08 0.06 0.04 0.13 1.20 0.88 0.2

Figure 4.1: Seed, kernel & seed coat weight (g/seed)

55
Figure 4.2: Ratio of seed to kernel weight

Figure 4.3: seed length, breadth and their ratio

Figure 4.4 Variations in seed coat of 30 accessions

0.6

0.5
Seed coat wt. (gm)

0.4

0.3

0.2

0.1

0.0

56
There was a very wide variation in seed to kernel (weight) ratio (Figure 4.2). It ranged
from 1.24 to 1.89 with an average of 1.61 (standard deviation 0.13). Highest seed kernel
ratio was observed in Saira 3 samples of Rajasthan; while the lowest was in Saira
commercial samples. The same from other selection from Pauri, Uttaranchal was 1.6.
Low ratio will indicate better kernel size and hence higher probability of good oil
content.

Seed weight of some accessions was so low that it was less than the kernel
weight as found in some other accessions. Case in study is that of Saira commercial,
Rajasthan where seed weight was only 0.42 g while kernel weight in many accessions of
Uttaranchal like that of Pauri, HNB University, Udaipur of Rajasthan and many others
was more than 0.4 g. Selections should be made of heavy kernel lots.

Earlier, variability in seed traits of 24 accessions of Jatropha curcas collected


from different agro climatic zones of Haryana state, India were reported by Kaushik, et
al. (2007). They reported significant differences (P<0.05) in seed size and 100-seed
weight between accessions. High heritability and genetic gain were reported for seed
weight (96.00% and 18.00%), indicating the additive gene action. Seed weight had
positive correlation with seed length, breadth, thickness and oil content. On the basis of
non-hierarchical Euclidian cluster analysis they reported six clusters with highest
number of accession falling under cluster III.

Seed length varied from 8 mm to 12.6 mm amongst the accessions investigated.


It had an average of 10.67 with a standard deviation of 1.2. Seeds could be classified
arbitrarily into various frequency classes. There were only 10 and 13 % seeds with seed
length of 15 and 16 mm, respectively. Those with 17 and 18 mm were 30 % each; while
those in 19 mm range were 17 %. Variability in seed width was not as wide and
distributed. Majority of seeds fell in the frequency class of seeds with 10 to 11 mm
width (53%) while 33 % had seed width of 11 to 12 mm. There was only one lot with
more than 9 or 12 mm width, each. Average width of the seeds was 10.67 mm with a
maximum of 12.56 and a minimum of 8.03 mm (standard deviation of 0.88). The

57
average seed length to width ratio was 1.7 with a maximum of 2.1 and minimum of 1.4
(standard deviation of 0.2).

Figure 4.3 shows graphical presentation of seed length, breadth and their ratio.
It can be observed that there was no direct correlation between seed length and
breadth. Variation in seed length was relatively high as compared to seed width. Longer
seeds were not necessarily wide too. This is an interesting observation since it implies
that only seed length or only seed width cannot be taken as a parameter to evaluate
variation. Seed area may be a better unit to evaluate variability while making selections
among accessions. Seed coat variation (Figure 4.4) is also an important trait useful for
genetic enhancement of Jatropha curcas.

There were not wide differences in seeds of Uttaranchal as compared to


Rajasthan. Average seed weight in both cases was 0.76 g; similarly there were no
significant differences in kernel weight.

Work on collection, characterization and evaluation of germplasm of Jatropha


curcas for growth, morphology, seed characteristics and yield traits is still at preliminary
stage. Species and provenance trials contribute fundamental information for further
breeding and genetic improvement (Burley & Wood 1976). Systematic screening from
different locations and provenance trials have not been carried out with the physic nut
to the necessary extent. Certain provenances may differ relatively from others if
cultivated at different sites, due to genetic and environmental interaction (Namkoong et
al. 1988). Priority should be given to assess intra- and inter-accessional variability in the
available germplasm, selection of pure lines and then multiplication. Existence of natural
hybrid complexes is reported in the genus Jatropha, such as, J. curcas-canascens
complex in Mexico (Dehgan & Webster 1978), J. integerrima–hastata complex in Cuba
and West Indian islands (Pax 1910) and J. curcas–gossypifolia (J. tanjorensis) in India
(Prabakaran & Sujatha 1999). Analysis of phenotypic diversity in collections can facilitate
reliable classification of accessions and their identification with future utility for specific
breeding purposes. There are very few studies on phenotypic diversity. Sunil et al.
(2008) developed a methodology for identification of superior lines by assessing the
58
phenotypic traits of plants recorded in situ. Germplasm exhibiting gross morphological
differences should be subjected for pollen studies and lines with pollen abnormality or
poor seed set should be investigated in detail before drawing conclusions about the
distinctness. Makkar et al. (1997) and Saikia et al. (2009) studied variations in
physiological parameters; Wani et al. (2006), Pant et al. (2006), Kaushik et al. (2007) and
Rao et al. (2008) in oil content & 100 seed weight of Jatropha curcas.

Very little work on genetic improvement aspects of this species has been taken
Systematic provenance trials at different locations have been initiated through a
network sponsored by the Department of Biotechnology, India. Efforts at other places
in the world have not been carried out to the necessary extent in the world (Dehgan &
Webster 1979). At the global level the information on genetic improvement of J. curcas
is restricted to few publications. Srivastava (1999) observed morphological differences
among 42 accessions. Adaptive trials on J. curcas were at the North East Institute of
Science and Technology, Jorhat, Assam to determine source variation in J. curcas
accessions collected from 17 states (34 locations) of India to identify the best sources on
the basis of phenological variations (i.e. plant height, stem girth, branches per plant and
100-seed weight). Ginwal and coworkers (2005) evaluated the seed sources of J. curcas
for growth performance from nursery stage to 2-yr-old plantation and oil content. Das
et al. (2010) evaluated sixteen J. curcas genotype on the basis of 12 different characters
from four Indian states .

The present investigation compares various accessions from Rajasthan and


Uttrakhand to evaluate range of variations that can be useful for further improvement
in Jatropha curcas. The study evaluates the magnitude of genetic variability present in
the existing base population and identifies important yield attributing characters to
provide useful information for developing high yielding Jatropha genotypes.

Tree breeding strategy largely depends upon extent of variability in base


population which may be measured by different parameters viz. ‘genotypic, phenotypic
variances’ and ‘genotypic, phenotypic coefficient of variation’.

59
Chemical and Physical Properties

Seed oil percentage


The data collected from the study of the physical and chemical properties of the
test samples is shown in Table 4.4. The oil content of Jatropha curcas seeds ranged from
17 % to 38.1 %. We have observed some seeds with oil more than 40 % oil too but the
samples selected for this study had highest oil % of 38 %. Average oil percentage was
31.36 % with a standard deviation of 4.28.46 % seed lots had 30 to 35 % oil while 33%
had less than 30 % oil. However 20 % seed lots had more than 35 % oil in the seeds
Figure 4.5.

Table 4.4: Oil percentage in 30 accessions


# Collection site I.C. # Code Oil (%)
1 SAIRA 3 471333 NBRI JB-2 17.01
2 KHAMNOR 471318 NBRI JB-19 24.66
3 SHERGARH 415772 NBRI JB-25 26.30
4 MOUNT ABU 471319 NBRI JB-4 27.32
5 BADI UNDRI 468906 NBRI JB-9 27.60
6 ODAN 471316 NBRI JB-18 28.16
7 SIROHI 468907 NBRI JB-15 28.45
8 HAPPRC, 415764 NBRI JB-24 29.35
GERHWAL
9 THOOR GAON 468912 NBRI JB-10 29.57
10 SAIRA1 471331 NBRI JB-6 29.88
11 MOUNT ABU 471309 NBRI JB-7 30.20
12 SAIRA (2) COMMERCIAL 471332 NBRI JB-1 30.37
13 TEHRI 415783 NBRI JB-29 30.52
14 GOGUNDA 471315 NBRI JB-17 31.03
15 DILWADA 471339 NBRI JB-3 31.52
16 GOGUNDA 471314 NBRI JB-16 32.23
17 BACHELIKHAL 415782 NBRI JB-28 32.97
18 CHOTI UNDRI 471301 NBRI JB-5 33.24
19 SHISHIRMA GAON 468913 NBRI JB-8 33.26
20 PANTNAGAR-1 NBRI JB-21 33.30
21 GERHWAL 415767 NBRI JB-27 33.83
22 PANTNAGAR-2 NBRI JB-22 34.02

60
23 UDAIPUR 468908 NBRI JB-14 34.06
24 PAURI 415764 NBRI JB-23 34.49
25 HNB,GERHWAL 415784 NBRI JB-30 35.02
26 GERHWAL,PAURI 415766 NBRI JB-26 35.12
27 UDAIPUR 468906 NBRI JB-12 35.97
28 BAGHPURA 468913 NBRI JB-11 36.19
29 ODAN 471317 NBRI JB-20 37.07
30 UDAIPUR 471122 NBRI JB-13 38.08

On the basis of kernel the oil was determined at 57 to 63 %. Oil content of


Jatropha curcas kernel was found higher than linseed, soybean, and palm kernel which
is 33.33%, 18.35% and 44.6%, respectively (Gunstone, 1994). This percentage is much
higher than those recorded for most oil-rich seeds (Esuoso and Bayer, 1998; Esuoso et
al., 1998). This would therefore be an advantage in terms of the exploitation of the oil.
High oil content of Jatropha Curcas indicated that Jatropha Curcas is suitable as non-
edible vegetable oil feedstock in oleochemical industries (biodiesel, fatty acids, soap,
fatty nitrogenous derivatives, surfactants and detergents).

Earlier, variability in oil content of 24 accessions of Jatropha curcas collected


from different agro climatic zones of Haryana state, India were reported by Kaushik, et
al. (2007). They reported significant differences (P<0.05) in oil content between
accessions. Oil variability ranged from 28.00% to 38.80%. In general phenotypic
coefficient of variation was higher than the genotypic coefficient of variation indicating
the predominant role of environment. High heritability and genetic gain were reported
for oil content (99.00% and 18.90%), indicating the additive gene action. Seed weight
had positive correlation with seed length, breadth, thickness and oil content.

61
Figure 4.5: Oil percentage in 30 accessions

Triacylglycerol was the dominant lipid specie (88.2%). The high unsaponifiable
matter (3.8%) is an advantage for use as natural insecticide. This is because
unsaponifiable matter contains sterols and triterpene alcohols which are responsible for
the insecticidal properties of fixed oils (Haftmann, 1970).

Table 4.5: Chemical and Physical Properties


Parameter Value
Seed kernel oil % 63.16 to 66.4%
% FFA as oleic acid 2.23 to 3.34
Unsaponifiable 3.8 %
Hydrocarbons 4.8 %
Diglycerols 2.5 %
Monoacyglycerols 1.7 %
Sterols 2.2 %
Iodine value 103.62±0.07
Saponification 193.55±0.61
value Polar lipids 2.0 % ±0.6
Peroxide value 1.93±0.012
Percentage oil content 63.16±0.35
(kernel) Density at 20° C 0.90317
(g/ml) 42.88
Viscosity at room Liquid
temperature (cp) Physical
state at room temperature
A Values are mean ± standard deviation of triplicate determinations

62
Chemical & physical properties of oil
Chemical and physical properties of oil are given in table 4.5.

Iodine
The high iodine value (103 to 116 mg.l2.g-1) indicates a preponderance of
unsaturated fatty acid. The iodine value is a measure of the unsaturation of fats and oils.
Higher iodine value indicated higher unsaturation of fats and oils (Knothe, 2002;
Kyriakidis and Katsiloulis, 2000). The iodine value of Jatropha curcas oil was determined
at 103.62g I2/100g, standard iodine value for biodiesel is 120 as per European EN 14214
specification. The limitation of unsaturated fatty acids is necessary due to the fact that
heating unsaturated fatty acids results in polymerization of glycerides. This can lead to
the formation of deposits or to deterioration of the lubricating (Mittelbach, 1996). Fuels
with this characteristic (such as sunflower oil, soybean oil and safflower oil) are likely to
produce thick sludge in the sump of the engine, when fuel seeps down the sides of the
cylinder into crankcase (Gunstone, 2004). The iodine values of Jatropha curcas place it in
the semi-drying oil group. High iodine value of Jatropha is due to high content of
unsaturation fatty acid such as oleic acid and linoleic acid. Earlier, Akintayo (2004)
suggested that Jatropha curcas oil can be used for production of alkyd resin, shoe polish,
varnishes etc. because of its iodine value.

Peroxidase value
Jatropha oil seed oil consists of 78.5% unsaturated fatty acid as shown in Table
below. The usual method of assessment hydroperoxides (primary oxidation products) is
by determination of peroxide value (Gunstone, 2004). Peroxide value of Jatropha oil
seed was low (as crude seed oil) at 1.93 meq/kg, indicating its oxidative stabilities. The
high iodine value and oxidative stability suggests that the seed oil has good qualities of
semidrying oil (Eromosele et al., 1997). Saponification value of the samples studied was
193.55. High saponification value indicated that oils are normal triglycerides and very
useful in production of liquid soap and shampoo.

63
FFA
Experimental result showed that Jatropha oil seeds have FFA content 2.23%. The
FFA and moisture contents have significant effects on the transesterification of
glycerides with alcohol using catalyst (Goodrum, 2002). The high FFA content (>1% w/w)
will result in soap formation and the separation of products will be exceedingly difficult,
and as a result, it will have a low yield of biodiesel product. The acid-catalyzed
esterification of the oil is an alternative (Crabbe et al., 2001), but it is much slower than
the base-catalyzed transesterification reaction. Therefore, an alternative process such as
a two-step process has been suggested for feedstock having high FFA content (Veljkovic
et al., 2006).

Viscosity
Viscosity provides resistance to the liquid to flow. Viscosity increases with
molecular weight but decreases with increasing unsaturated level and temperature
(Nouredini et al 1992). The viscosity of Jatropha oil seed must be reduced for biodiesel
application since the kinematic viscosity of biodiesel were very low compared to
vegetable oils. High viscosity of the Jatropha oil seed is not suitable if it has to be used
directly as engine fuel, it will often results in operational problems such as carbon
deposits, oil ring sticking, and thickening and gelling of lubricating oil as a result of
contamination by the vegetable oils. Different methods such as preheating, blending,
ultrasonically assisted methanol transesterification and supercritical methanol
transesterification are being used to reduce the viscosity and make it suitable for engine
applications (Pramanik, 2003; Banapurmath, 2008).

Density
The density of a material is a defined as a measure of its mass per unit volume
(e.g. in g/ml). The density of vegetable oil is lower than of water; and the differences
amongst vegetables oils are quite small, particularly amongst the common vegetable
oils. Generally, the density of oil decreases with molecular weight, yet increase with

64
unsaturation level (Gunstone, 2004). In the current study, the density of Jatropha seed
oil was 0.90317 g/ml (Table 4.6).

Table 4.6:- Physico-chemical properties of extracted oil


GERMPLASM COLLECTION CODE OIL SEED FFA ACID VALUE R.I DENSITY
I.C. No. THICKNESS
(%) (mm) (%) (mgKOH/g) (nD) (g/ml)

471333 NBRI JB-2 17.0100 8.0333 1.6133 2.3133 1.4664 0.9100

471318 NBRI JB-19 24.6667 6.6333 0.5167 1.1333 1.4677 0.9200


415772 NBRI JB-25 26.3033 7.9667 0.4533 1.9067 1.4665 0.9200
471319 NBRI JB-4 27.3200 8.6667 0.7833 1.5667 1.4659 0.9100
468906 NBRI JB-9 27.6033 8.7000 1.1867 2.0067 1.4662 0.9100
471316 NBRI JB-18 28.1667 7.8333 0.7700 1.5433 1.4664 0.9100
468907 NBRI JB-15 28.4500 8.1000 0.9767 1.9267 1.4665 0.9100
415764 NBRI JB-24 29.3533 7.7667 0.7267 1.4533 1.4668 0.9100
468912 NBRI JB-10 29.5767 9.4667 1.5367 3.1433 1.4661 0.9200
471331 NBRI JB-6 29.8867 8.8667 0.9700 1.9733 1.4668 0.8900
471309 NBRI JB-7 30.2033 8.3667 0.4233 0.8467 1.4663 0.8800
471332 NBRI JB-1 30.3733 6.4333 1.8833 3.4300 1.4664 0.9100
415783 NBRI JB-29 30.5233 9.6000 0.6833 1.3667 1.4679 0.8900
471315 NBRI JB-17 31.0267 8.4667 0.9467 1.8933 1.4671 0.9100
471339 NBRI JB-3 31.5200 8.6333 0.5467 2.2500 1.4667 0.9100
471314 NBRI JB-16 32.2300 8.5000 0.8633 1.7267 1.4669 0.9200
415782 NBRI JB-28 32.9733 9.4000 0.5533 1.1067 1.4673 0.9000
471301 NBRI JB-5 33.2433 8.3000 0.9067 1.8167 1.4664 0.9100
468913 NBRI JB-8 33.2567 9.9333 0.9900 1.9800 1.4665 0.9000
NBRI JB-21 33.3033 9.2000 1.2867 2.5733 1.4657 0.9100
415767 NBRI JB-27 33.8267 9.8333 0.7733 1.5467 1.4667 0.9100
NBRI JB-22 34.0200 9.3333 0.7400 1.4800 1.4660 0.9200
468908 NBRI JB-14 34.0633 7.9667 0.9833 1.9667 1.4667 0.9100
415764 NBRI JB-23 34.4933 9.9667 0.8867 1.7733 1.4656 0.9100
415784 NBRI JB-30 35.0267 10.1000 1.9167 1.5533 1.4663 0.9100
415766 NBRI JB-26 35.1200 9.9333 1.1433 2.2733 1.4667 0.9100
468906 NBRI JB-12 35.9767 7.6000 1.1267 2.2533 1.4671 0.9100
468913 NBRI JB-11 36.1933 7.3667 1.4533 3.9133 1.4664 0.9100
471317 NBRI JB-20 37.0700 9.4333 0.5067 1.0133 1.4668 0.9100
471122 NBRI JB-13 38.0867 7.7333 1.2033 2.4067 1.4670 0.9200

Saponification Value & Insoluble impurities


Saponification value ranged from 71.6 (in # 415772) to 129.5 (in # 468908).
There was no correlation between saponification value and insoluble impurities. The
insoluble impurities ranged from to 0.53 in #415764 to 2.37 in # 471332. Average
impurities were 1.1 mg/kg while average saponification value was 102.3 (Table 4.7).

65
Table 4.7: Oil physico-chemical properties contd.
Sap. value Insoluble impurities
Code IC # mg/kg

NBRI JB-25 415772 71.61 1.19


NBRI JB-26 415766 76.31 0.72
NBRI JB-29 415783 78.84 0.77
NBRI JB-20 471317 86.77 0.74
NBRI JB-19 471318 87.92 0.66
NBRI JB-16 471314 88.62 1.11
NBRI JB-10 468912 89.89 1.43
NBRI JB-9 471303 90.63 1.22
NBRI JB-23 415764 91.16 0.98
NBRI JB-28 415782 91.23 0.68
NBRI JB-24 415764 93.44 0.53
NBRI JB-13 471122 96.65 1.54
NBRI JB-2 471333 98.54 1.42
NBRI JB-21 - 98.72 0.53
NBRI JB-12 468906 98.84 1.63
NBRI JB-11 468913 99.73 1.32
NBRI JB-22 - 100.91 0.78
NBRI JB-1 471332 101.4 2.37
NBRI JB-17 471315 102.41 1.08
NBRI JB-8 468911 111.1 0.98
NBRI JB-7 471309 112.41 1.65
NBRI JB-30 415784 112.45 1.05
NBRI JB-18 471316 116.49 1.19
NBRI JB-5 471301 119.29 1.64
NBRI JB-15 468907 121.03 0.98
NBRI JB-27 415767 123.17 0.55
NBRI JB-3 471339 125.99 0.88
NBRI JB-6 471331 126.26 1.32
NBRI JB-4 471319 128.6 0.65
NBRI JB-14 468908 129.51 1.22

66
Correlation of seed oil, breadth and oil characteristics
While the oil percentage had a wide range from 17 to 38 with an average of
31.36 (standard deviation of 4.3), seed thickness ranged from 6.4 to 10.1 with an
average of 8.6 (standard deviation of 0.98). free fatty acid percentage content ranged
from 0.42 to 1.92 with an average of 0.98 (standard deviation of 0.4). Acid value of the
oil ranged from 0.85 to 3.91 with an average of 1.94 (standard deviation of 0.68).
Similarly, density ranged from 0.92 to 0.88 with an average of 0.91 (standard deviation
of 0.1). refractive index was nearly uniform in all the samples.

Seed oil content was correlated with other traits like seed thickness, free fatty
acid percentage, acid value (mg/KOH/g) and refractive index and density (Table 4.6). The
oil content range was from 17 to 38 %. Pearson Correlation coefficient was calculated
for the said factors. Pearson value was 0.33 for correlation between seed oil and seed
width. Apart from this, very poor correlation was observed between other traits when
compared with seed oil. It was 0.02 for seed oil and FFA content, 0.09 for seed oil and
acid value; zero value for seed oil and refractive index; and 0.01 for seed oil against
density of oil. It suggests that those factors that have poor correlation with seed oil are
not determined by genetical makeup of the plant rather are governed by environmental
factors.

Free fatty acid percentage was also poorly correlated with refractive index and
density. Pearson correlation value was negative (-0.32) for free fatty acid percentage
and refractive index, while it was 0.2 for free fatty acid percentage and density.
However, it was positively correlated with acid value with correlation factor of 0.71. free
fatty acid percentage is controlled by environmental factors and can change with
moisture and exposure to air.

Fatty Acid Composition


Fatty acid composition determination was another important characteristic
carried out in this study. The properties of the triglyceride and the biodiesel fuel are
determined by the amounts of fatty acids present in the molecules. Chain length and

67
number of double bonds determine the physical characteristics of both fatty acids and
triglycerides (Mittelbach and Remschmidt, 2004). Transesterification does not alter the
fatty acid composition of the feedstock and this composition plays an important role in
some critical parameters of the biodiesel, as cetane number and cold flow properties
(Ramos et al., 2008). Fatty acid composition of studied oil shown in table is compared
with other vegetable oils such as palm oil, sunflower oil, palm oil and soybean oil.

There are three main types of fatty acids that can be present in a triglyceride
which is saturated (Cn:0), monounsaturated (Cn:1) and polyunsaturated with two or
three double bonds (Cn:2,3). Various vegetable oil is a potential feedstock for the
production of a fatty acid methyl ester or biodiesel but the quality of the fuel will be
affected by the oil composition. Ideally the vegetable oil should have low saturation and
low polyunsaturation i.e. be high in monounsaturated fatty acid (Gunstone, 2004).

Vegetable oils that rich in polyunsaturated such as linoleic and linolenic acids,
such as soybean, sunflower (table), tend to give methyl ester fuels with poor oxidation
stability. Vegetable with high degree unsaturation tend to have high freezing point.
Those oils that have poor flow characteristic, may become solid (e.g. palm oil) at low
temperatures though they may perform satisfactorily in hot climates (Gunstone, 2004).

The predominant fatty acids in the present study were monounsaturated (45.4%,
followed by polyunsaturated fatty acid (33%) and saturated fatty acid (21.6%). Mon
unsaturation of Jatropha seed oil was higher than other vegetable oil as palm kernel,
sunflower and palm oil (Table 4.8). The major fatty acids in Jatropha seed oil were the
oleic, linoleic, palmitic and the stearic fatty acid.

In majority of samples, Oleic acid showed the highest percentage of composition


of 42.8% followed by linoleic acid with 32.8%. However, in other samples Linoleic acid
was the dominant one with 47.3 % while oleic acid was at 22.8 %. Similarly, palmitic acid
ranged from 11.3 to 14.2 in different samples. Stearic acid ranged from 7.0 to 17.0 % in
different samples. There were similar differences in other fatty acids too but the Thus,
Jatropha seed oil can be classified as oleic–linoleic oil. Compared to others vegetable

68
differences were not as significant except that of Arachidic acid (20:0) where it ranged from 0.2
to 4.7 %. The most characteristic differences were between oleic and linoleic acids (Table 4.8).

Oils (table), Jatropha oil seed has highest oleic as compared to palm oil, palm kernel,
sunflower, coconut and soybean oil. According to the European standard the
concentration of linolenic acid and acid containing four double bonds in FAMEs should
not exceed the limit of 12% and 1%, respectively. Jatropha oil seed contains only 0.2%
linolenic acid, which is lower as compared to sunflower oil and palm oil (Table 4.8).

Table 4.8: Fatty acid composition (%)


Fatty Acid Jatropha curcas oil seed Palm kernel oila Sunflower oila Soybean oila Palm oila
Oleic 18:1 22.8 to 44.7 15.4 21.1 23.4 39.2
Linoleic 18:2 32.8 to 47.3 2.4 66.2 53.2 10.1
Palmitic 16:0 11.3 to 14.2 8.4 - 11.0 44.0
Stearic 18:0 7.0 to 17.0 2.4 4.5 4.0 4.5
Palmitoliec 16:1 0.7 to 1.0 - - - -
Linolenic 18:3 0.2 to 1.2 - - 7.8 0.4
Arachidic 20:0 0.2 to 4.7 0.1 0.3 - -
Margaric 17:0 0.1 - - - -
Myristic 14:0 0.1 16.3 - 0.1 1.1
Caproic 6:0 - 0.2 - - -
Caprylic 8:0 - 3.3 - - -
Lauric 12:0 - 47.8 - - 0.2
Capric 10:0 - 3.5 - - -
Saturated 21.6 82.1 11.3 15.1 49.9
aFrom Edem, D.O. (2002)

TAGs Composition
The TAGs profile of Jatropha seed oil, was characterized by reversed phase HPLC
where the mechanism in separating the TAGs involves the chain length and degree of
unsaturation of the fatty acids (Gutierrez and Barron,1995). The identified TAGs of
Jatropha seed oil were concluded by comparing the retention time of standard TAGs
peak chromatographs obtained under same analytical condition. Triacyglycerol content
of Jatropha oil seed is showed at Table 4.9. From the chromatograph obtained, it is
evident that the most prominent polyunsaturated TAG was OOL (22.94%), followed by
OLL (17.9%), POL (14.95%), PLL +MOL 7.08, SOO (2.48%). Monounsaturated ones
detected were MPP+OOO (16.65%) POO (9.72%), PLP+MOP (1.85%) and POP (0.91%).

69
OOL was highest at 22.9 % while OLL was next at 17.9 %. Apart from MPP+OOO
and POL which were at 16.65 and 14.95 %, respectively, all others were less than 10%.

Table 4.9: TAGs composition (%)


Triacylglycerol Relative Composition (%) Triacylglycerol Relative Composition (%)
OOL 22.94 SOO 2.48
OLL 17.90 PLP+MOP 1.85
MPP+OOO 16.65 POP POS 0.91
POL 14.95 MM 0.59
POO 9.72 CC 0.48
PLL + MOL 7.08 PP 0.44
Nd 3.60 0.38

Characteristics of Jatropha methyl esters, diesel & ASTM and DIN


standards
All the characters studied for biodiesel methyl esters were compared with that
of diesel and also compared to the ASTM D6751-02 and DIN EN 14214 standards
(German standards). Density at 150C was 922 for Jatropha oil while it was 855 for
Jatropha methyl esters as compared to 850 for diesel. This density was within range for
biodiesel though crude Jatropha oil did not meet the standard. Similarly Viscosity at 40
0
C was 35.8 for Jatropha oil as compared to 4.82 for Jatropha methyl esters and only 2.6
for Diesel. It is 1.9 to 2.6 for ASTM D6751-02 standard and 3.5 to 5.0 for DIN EN 14214
standard.

Flash point of Jatropha methyl esters was 165 much lower than that of Jatropha
oil while that for Diesel is only70 though ASTM D6751-02 standard and DIN EN 14214
standard are set at >130 and >120, respectively. It shows that Flash point is slightly
higher for Jatropha methyl esters. Pour point was -6 0C, same as that of Jatropha curcas
oil and the same for Diesel was -20 0C. There was no issue of water content since there
was no water in Jatropha methyl esters or oil and up to <0.03 is permitted in ASTM
D6751-02 standard and <0.05 in DIN EN 14214 standard. Similarly ash content is
permitted up to <0.02 % in both ASTM D6751-02 standard and DIN EN 14214 standard;
the same as was observed in Diesel while Jatropha methyl esters and Jatropha curcas
oil had none.

70
Carbon residue was within the limits of standards. The ASTM D6751-02 standard
and DIN EN 14214 standard permit up to <0.3 % while the same in Diesel is 0.17 % and
Jatropha methyl esters was 0.026 and it was the same in Jatropha curcas oil. Sulphur
content was nil in Jatropha methyl esters and oil was nil and so was in Diesel while
ASTM D6751-02 standard permit up to 0.05 % (Table 4.10).

Acid value, considered to be another important character was 0.27 mg/KOH/g in


Jatropha methyl esters and its oil. Diesel had 0.35 while the ASTM D6751-02 standard
has it at <0.8 and DIN EN 14214 standard at <0.5 mg/KOH/g. Iodine value of Jatropha
methyl esters and its oil was at 107. The same for Diesel could not be ascertained.

Calorific value was slightly lower at 33.5 JJ/kg for Jatropha curcas oil and 36.7
MJ/kg for Jatropha methyl esters. It was lower than that of Diesel where it was 42
MJ/kg. Cetane value was 50.6 in Jatropha curcas oil while it was 52.0 in Jatropha methyl
esters and 46 in Diesel.

Table 4.10: Characteristics of Jatropha methyl esters, diesel & ASTM and DIN
standards

Properties Units Jatropha Jatropha Diesel ASTM DIN EN


curcas oil curcas D6751-02 14214
methyl
ester
kg/m3
Density at 922 885 850 875-900 860-900
150C
Viscosity at Mm2/s 35.8 4.82 2.60 1.9-6.0 3.5-5.0
40 0C
0
Flash point C 190 165 70 >130 >120
0
Pour point C -6 -6 -20 - -
Water % Nil Nil 0.02 <0.03 <0.05
content
Ash content % Nil Nil 0.02 <0.02 <0.02
Carbon % 0.026 0.026 0.17 - <0.3
residue
Sulphur % Nil Nil - 0.05 -
content
Acid value Mg/KOH/g 0.27 0.27 0.35 <0.8 <0.5

71
Iodine value - 107 107 - - -
Saponification - 188 190 - - -
value
Calorific value MJ/kg 33.5 36.7 42 - -
Cetane value - 50.6 52.0 46 - -
ASTM = American Society for testing & Materials
DIN – German Institute for Standardization

Fatty acid composition


The major fatty acids in Jatropha seed oil were the oleic acid, linoleic acid ,
palmitic acid and the stearic acid. The most prominent TAGs of Jatropha seed oil were
OLL and OOL. The oil extracts exhibited good physicochemical properties and could be
useful as biodiesel feedstock and industrial application.

The fuel properties of Jatropha curcas oil, Jatropha curcas methyl ester and
diesel were compared as given in the table below. Calorific value of these methyl esters
are around 37 mj/kg, which are low, compared to diesel fuels (42.0 mj/kg). The methyl
ester, however, has higher cloud and pour points than conventional diesel. This is
important for engine operation in cold or cooler environments. Kinematic viscosity and
cetane values are slightly higher than the diesel, which is favorable for combustion.
Higher flash point is advantageous for fuel transportation. Various properties of
Jatropha curcas oil methyl ester were found to be comparable with that of diesel fuel
(Singh & Padhi, 2009).

Optimization and Standardization of oil extraction process


Protocol has been optimized for determination of oil yield from J. curcas seeds.
Five accessions were selected for protocol optimization. Optimization for recovery of oil
yield has been done by comparing seeds from different growth stages using non-polar
solvents (hexane, petroleum ether, cyclohexane and n-pentane). The extraction
methods evaluated were expeller, Soxhlet extraction, extraction using Polytron and
SCFE methods. Physicochemical properties of extracted oil were also analyzed according
to BIS and ASTM standards.

72
Oil synthesis at development stages of Jatropha curcas
Seeds of 10 accessions taken for the present study were collected at various
stages of development, starting from a week after fertilization and five days thereafter
till maturity. These were classified as stage I to stage VII. The seeds collected from
mature black coloured fruits, ready for dehiscence were classified as Stage VII. Seeds
from various stages of development were examined for phonological traits, moisture
and oil content. Fresh weight of young seeds was recorded. Mean of triplicate reading
has been taken for each experiment to correlate the analysis.

Seeds were collected at various stages of development listed as below:

Sequential maturation stages of seeds Seeds maturation stages


I Very small seeds fruit
II Immature seeds fruit
III Immature green seed fruit
IV Greenish yellow seeds fruit
V Green and Yellow seed fruit
VI Brown fruits
VII Mature Jatropha dried black
fruit seeds

There was loss of moisture from the seeds as these matured. Increase in fresh
weight from stage I to III was nearly ten times (from 32.4 to 317 mg) while it was only
three times from stage III to stage V. Fruits as well as seeds lost great deal of moisture
when the fruits turned black and were ready to fall down. The weight of the seeds and
seed area are given in enclosed figures. Increase in seed area corresponded to increase
in fresh weight of the seeds. There was an increase in seed area from young stage I to
stage VI. However, the seed area shrunk at stage VII. Moisture content of the seeds
ranged from nearly 8 to 90 %; the lowest being in mature seeds of stage VII and the
highest in stage I.

73
Figure 4.6:- Sequential maturing stages of J. curcas fruits and seeds

Very small seeds fruit Immature seeds fruit Immature green seed fruit

Greenish yellow seeds fruit Green and Yellow seed fruit

Brown fruits Jatropha dried seeds

Figure 4.7: Comparative study of maturation stages of J. curcas seeds of 5 accession

35

NBRI JB 1
30
Oil content of seeds according to stages

NBRI JB 2
NBRI JB 3
25 NBRI JB 4
NBRI JB 5

20

15

10

0
ds ds ds ds ds ds ed
l se e ture see fruit see t see fru it see fru it see rie d se
ys mal a n n frui w n d
V er Im m
G re e
G r e e
Y e llo
B r o w
r e and
e e atu
atur Ma t
ur enis
h
ow
a nd
lly m
I mm Gre Ye ll Tot a

Different maturing stages of Jatropha germplasm

74
Figure 4.8: Comparative study of maturation stages of J. curcas seeds in 5 accessions

40

NBRI JB 6

Oil cotent of seeds according to stages


NBRI JB 7
NBRI JB 8
30 NBRI JB 9
NBRI JB10

20

10

0
ds ds ds ds ed s ds ds
see see see see it se see see
a ll fruit re fruit en fruit g reen ow fru n fruit d fruit
u e e ll w rie
y sm Immat e gr Ma t
ur h ye nd b
ro
and
d
Ver atur eni s wa
I mm Gr e ure
yello lly mat
F u

Different maturing stages of Jatropha seeds

Freeze dried seeds were extracted with hexane using homogenizer and the
solvent was evaporated under reduced pressure. Percent oil content (hexane soluble)
was measured on dry weight basis of seeds. Oil content was the lowest being at stage I,
and the highest at stage VI. There was a slight decrease in total oil percentage in dry
mature seeds in stage VII.

Fatty acid composition of hexane extracts of J. curcas seeds showed free fatty
acids (FFA), methyl ester of fatty acids (FAME) and triglycerol esters (TAG), along with
small quantity of sterols. The fatty acids had both saturated fatty acids (SFA) as well as
unsaturated fatty acids (USFA). The latter had a combination of monounsaturated fatty
acids (MUFA), di-unsaturated fatty acids and tri-unsaturated fatty acids, commonly
known as polyunsaturated fatty acids (PUFA).

Moisture content of the seeds ranged from 10.1 to 89.7%; the lowest in mature
seeds in stage VII and highest in stage I. Seed fresh weight, seed area and oil
percentages of all seed developments were measured.

The study was carried out for 10 seeds lots (NBRI 1 to 10) and shown in enclosed
figures. Highest oil percentage at stage VII of different seed lots was 30.3, 17.0, and

75
31.5. 27.3, 23.2, 33.3, 34.0, 34.4, 29.3, 26.3 in NBRI 1,2 ,3, 4, 5, 6, 7, 8, 9, and NBRI 10,
respectively. Detailed discussion of NBRI 1 and 2 is given below.

Growth & oil % at various stages of development in 10 accessions


Growth and oil percentage in various accessions (NBRI 1 to 10) with respect to
seed weight, seed area and oil content in Figures 4.9 to 4.18.

Figure 4.9: Increase in seed wt., seed area and oil content in NBRI1

76
Figure 4.10: Increase in seed wt., seed area and oil content in NBRI2

Figure 4.11: Increase in seed wt., seed area and oil content in NBRI3

77
Figure 4.12: Increase in seed wt., seed area and oil content in NBRI4

Figure 4.13: Increase in seed wt., seed area and oil content in NBRI5

78
Figure 4.14: Increase in seed wt., seed area and oil content in NBRI6

Figure 4.15: Increase in seed wt., seed area and oil content in NBRI7

79
Figure 4.16: Increase in seed wt., seed area and oil content in NBRI8

Figure 4.17: Increase in seed wt., seed area and oil content in NBRI9

80
Figure 4.18: Increase in seed wt., seed area and oil content in NBRI10

The oil percentage was 4.9 % at Stage III, it gradually increased to 10.7 and 23.7
at Stages IV and V, respectively. There was a gradual increase to next stage where oil
percentage was 30.2 %, however, further increase to final stage was insignificant at only
30.4 as compared to 30.2 in the previous stage in the first set of experiment.

These changes corresponded to increase in seed weight (seed weight of 10 seeds


shown in the figure) that gradually increased from Stage I onwards, however, seed
weight (10 seeds) at final stage was lower than that in Stage VI (4.23 at Stage VII as
compared to 4.27 in Stage VI). Although the seed weight decreased in the final stage
due to less moisture, apparently but the oil percentage increased though slightly.

Another lot with relatively low oil percentage was also investigated following the
same protocol. The increase in oil percentage followed the same pattern. There was no
oil at Stage I while only 1.46 % was recorded at Stage II. It increased gradually to 16.8 %
at Stage VI and then a very minor increase was recorded at Stage VII (17.01 as compared
to 16.8 % at Stage VI).

In this set of experiment, seed weight increase pattern was the same as in earlier
case. The seed weight kept on increasing up to the VI stage. There was a minor decrease
in the final stage (5.39 g at Stage VII as compared to 5.45 g for 10 seeds at Stage VI).

81
Seed area, however kept on increasing although the increase at the final stage was only
marginal. Maximum seed area increase was observed at Stage VI as compared to Stage
V (1.63 at Stage V as compared to Stage VI where it was 2.5 cm2 for 10 seeds (Figure 4.9
to 4.18)

Earlier. Annarao et al (2008) studied seed development in Jatropha curcas L. with


respect to phenology, oil content, lipid profile and concentration of sterols. They
reported that the seed area increased as the seed grew from stage I to stage VI (0.2–
10.2 mm2 per seed), however, the seed area shrunk at stage VII. Increase in seed area
1
corresponded to increase in fresh weight of the seeds. They also reported H NMR
spectroscopy of hexane extracts made at different stages of seed development revealed
the presence of free fatty acids (FFA), methyl esters of fatty acids (FAME) and triglycerol
esters (TAG), along with small quantity of sterols. The young seeds synthesized
predominantly polar lipids. Lipid synthesis was noticed nearly three weeks after
fertilization. From the fourth week the seeds actively synthesized TAG. Stage III is a
turning point in seed development since at this stage, the concentration of sterols
decreased to negligible, there was very little FAME formation, accumulation of TAG
increased substantially, and there was a sudden decrease in FFA concentration. Their
findings can be helpful in understanding the biosynthesis and in efforts to improve
biosynthesis of TAG and reduce FFA content in the mature seeds. It also indicates the
reasons for the variation near seed maturity.

Earlier, 1H-NMR spectra were evaluated for determination of percentage


contribution of TAG, FAME and FFA present in the hexane extract of seeds of J. curcas at
various stages of growth using standard methods (Khetrapal and Lakshminarayana,
1984; Lakshminarayana et. al., 1984; Gunstone and Shukla, 1995).

Study of seed development from one week after fertilization to maturity was
studied with respect to phenology, oil content, lipid profile and concentration of sterols.
While there was a gradual increase in seed size, its fresh weight, moisture content and
oil concentration, significant changes were observed in fatty acid profiles. The
development of seed can be grouped in two classes: early development and maturity.
82
Stages I to III in the present study can be grouped as “early developmental stages”.
Similar observations have been made by Annarao et al (2008) who supported this with
their the observations that sterols were detected only in the initial stages of Jatropha
curcas whereas these decreased substantially in the later stages. They quoted that
sterols have an essential role at the cellular level of hormonal signaling, organized
divisions and embryo altering (Hartmann et al., 2002). Sterols are primary metabolites
of lipidic nature. There is a positive correlation between sterols and fatty acid
composition in the initial (development) stages of J. curcas seeds.

Annarao (2008) also reported that high FFA content in the initial stages was in
correspondence with that of sterols. They detected methyl esters only in the initial
stages when high concentration of sterols was present. Occurrence of sterols only in
early stage is in agreement with its main function at that development stage i.e., to
reinforcement of cell walls, and development of bilayers and not the synthesis of oil
which occurs later. Akintayo (2004) and Adebowale and Adedire (2006) reported
variations in campesterol, stigmasterol and sitosterol in different provenances of J.
curcas.

Maturing J. curcas seeds synthesize lipids through nearly three weeks after
fertilization. The young seed forms predominantly polar lipids, specific to membranes.
From the fourth week after fertilization, the seed acquires the properties to actively
synthesize TAG. The process took another two weeks. Triki, et. al., (2000) investigated
the role of specific enzymes in triggering synthesis of TAG. Measurements of the key
enzymes of fatty-acid synthesis in cell-free extracts of seeds of different maturities from
Cuphea wrightii showed that malonyl-CoA synthesis may be a triggering factor for the
observed high capacity for fatty-acid synthesis (Deerberg et al., 1990).

The study suggests that stage III of seed development can be identified as a
turning point for development since at this stage concentration of sterols decreases to
negligible, there is very little methyl ester formation. The findings shall be useful for
silviculturists for fixing harvest rotations, providing field inputs at seed maturation stage
and in understanding of synthesis of metabolites including that of specific triglycerides.
83
The findings can be helpful for understanding of biosynthesis and in efforts to improve
biosynthesis of TAG and reducing FFA content in the mature seeds.

The present study also includes analysis of two lots of seeds, one with low oil
content and the other one with moderate to high oil percentage. The results indicate
that there is a possibility of early analysis of oil content at young stage. This can be
helpful for the breeders to select the individuals by analyzing the oil content at early
stage.

Optimizing of oil extraction

Solid liquid extraction, sometimes called leaching, involves the transfer of a


soluble fraction (the solute orleachant) from a solid material to a liquid solvent. The
solute diffuses from the solid into the surrounding solvent. Normally, solid liquid
extraction is dependent on the nature of the solvent and oil, reaction time between
solvent and seeds, temperature of the process, particle size of the meal and the ratio of
solvent to the meal.

The effect of solvent, temperature, solvent to solid ratio, particle size of the meal
and reaction time were investigated to optimize the extraction operating conditions for
achieving maximum oil yield. Protocol for the extraction of oil from the seeds were
studied following extraction using Soxhlet apparatus, Super Critical Fluid Extraction,
Polytron homogenizer, and by conventional expeller. The results are presented for 10
accessions to determine the ideal oil extraction method. The results as presented in
(Figures 4.19-4.20).

84
Figure 4.19: Oil yield in different extraction methods of #1-5

35

NBRI JB 1
30 NBRI JB 2
NBRI JB 3
NBRI JB 4
25 NBRI JB 5

20
Oil %

15

10

0
Expeller Polytron SCFE Soxhlet

Different oil extraction methods

Figure 4.20: Oil yield in different extraction methods of #6-

40

NBRI 6
NBRI 7
NBRI 8
30 NBRI 9
NBRI 10
Oil %

20

10

0
Expeller Polytron SCFE Soxhlet

Different oil extraction methods

It was observed that Soxhlet extraction gave highest oil percentage although the
seed meal was extracted in solvent over heat (Figure 4.19 and 4.20). Polytron
homogenizer extraction was next best and rather quite close to the results obtained
with Soxhlet extraction. Polytron homogenizer method is recommended since there was
no heating of the solvent that saved energy and c footprints. It was also quick method
and oil could be extracted in less than 15 minutes as compared to up to six hours of
cooking in solvent in Soxhlet method.

85
Super Critical Fluid Extraction procedure gave lowest oil percentage, although
there was no difference in quality. In fact, in terms of moisture, free fatty acid
percentage content and other traits, oil extracted by Super Critical Fluid Extraction was
better as compared to other methods. This method, however, is not recommended
since it consumed relatively high power which may not be feasible for commercial
purpose.

Although Soxhlet method also gave relatively high yield of oil but looking the
advantages of time and energy, the difference can be factored and assessment can be
made using Polytron homogenizer extraction method.

Various researches and studies have been conducted to describe the kinetics and
mechanism of the extraction processes (Forson, et al., 2004; So et al., 1986; Wisese et
al., 1987; Meziane et al., 2006; and Meziane et al., 2008). A solid liquid extraction
process is found to be most appropriately fitted by a second order model (Forson, et al.,
2004; So et al., 1986; Wisese et al., 1987; Meziane et al., 2006; and Meziane et al., 2008;
Rakotondramasy et al., 2007). According to the literature, it is typical of a second order
process to take place in two subsequent stages. First, the major part of the solute gets
extracted quickly because of the scrubbing and dissolution caused by driving force of
the fresh solvent and then comes the next stage where the extraction process gets
much slower accomplished by external diffusion of the remainder solute into the
solution.

Sayyar et al (2009) reported results of their study on extraction of Jatropha oil


from seeds using organic solvent. The kinetics of extraction was also investigated and its
parameters were determined based on a second order model. They investigated impact
of five parameters: type of solvents, temperature, and solvent to solid ratio, processing
time and particle size of the meal. The kinetics was optimized in view of these
parameters. The kinetics of extraction was assumed based on a second order
mechanism. The initial extraction rate, the saturated extraction capacity, the rate
constant of extraction and the activation energy were calculated using the proposed
model. They found that 8 h of reaction time at a temperature of around 68°C, with
86
coarse particle size (0.5-0.75 mm) and extraction solvent as hexane used in a solid ratio
of 6:1 to hexane was ideal for extraction. The activation energy was reported to be 8021.9 J
moL.

Earlier, ultra sonication was used as a pretreatment before aqueous oil


extraction by Shah et al (2005). They reported that aqueous enzymatic oil extraction
was found to be useful in the case of extraction of oil from the seeds of Jatropha curcas
L. The use of ultra-sonication for 10 min at pH 9.0 followed by aqueous oil extraction
gave a yield of 67%. However, the maximum yield of 74% was obtained by ultra-
sonication for 5 min followed by aqueous enzymatic oil extraction using an alkaline
protease at pH 9.0. Use of ultra-sonication also resulted in reducing the process time
from 18 to 6 h. However, this method has not been investigated in the present study for
its costs.

Effect of type of solvent:


The determination of oil content in biodiesel feed stocks can be performed using
several methods, including mechanical press, solvent extraction, and Nuclear Magnetic
Resonance. For the feedstock quality control in terms of oil content, it is important that
the applied method is universally accepted so as to obtain results that can be compared
with those reported from alternate sources.

Different solvents were used to evaluate their extraction efficiency. The oil
extraction capabilities of n-hexane, n-Pentane and petroleum ether are shown in Figure.
Highest oil was extracted in Petroleum ether as compared to the other three. Next best
solvent was n-Hexane. Since it is not advisable to use Petroleum ether, further
experiments were done with hexane only. The extraction yield with n-hexane was found
to be about 1.1% more than that of petroleum ether (37.6% and 35.4.0%, respectively
under similar conditions.

Out of various Hydrocarbon solvents, usually hexane is used instead of hexane,


heptane, and pentane. Extraction-grade hexane has an n-hexane content between 48 –
98% and a narrow distillation range. It is free of nitrogen and sulphur and unsaturated

87
compounds and sufficiently stable to be used indefinitely. Commercial heptane might be
preferred for the extraction of castor oil, which is not freely miscible with hydrocarbons
except at elevated temperature.

Halogenated solvents (trichloroethylene, dichloromethane) are not used


because of environmental concerns in commercial production. Although water (with
and without enzymes)is used as a solvent in the extraction of oil from palm, olive and
coconuts. Its drawbacks are more than 10% residual oil in cake and energy intensive
separation and drying. Enzymes can be used to digest the cell wall to set the contents
free. Although enzymatic assisted extraction currently is economical only in olive oil
processing (mild conditions, high oil quality), it has potential for future use once the cost
of enzymes can be brought down.

From these hexane is the only solvent used commercially on a large scale. Both
enzyme-assisted aqueous extraction, ethanol and isopropylalcohol seem promising for
application on the shorter term in developing countries. Supercritical CO2 is interesting
because of its non-toxic nature. More research will have to be done before these
alternative solvents can be used commercially.

Bhagat et al (2010) extracted oil by the Soxhlet method and reported an average
yield of 38% in Jatropha curcas seed in their study of several accessions collected by
them. Oil content of collected accessions was determined by nuclear magnetic
resonance and Soxhlet extraction. Oil variability ranged from 45.5% to 11.5% in different
accessions. The level of unsaturated fatty acid ranged from 85% to 75.5% in. Saturated
fatty acids ranged from 24% in to 15%, the oxidative stability index was highest at 2.1
and lowest at 0.68.

The European Norm (EN) has specified two methods for the determination of oil
content in oil seed crops: conventional Soxhlet extraction and NMR imaging. However,
these methods have several disadvantages. Both methods are time, labor, and cost
intensive, and both require highly skilled labor and a significant amount of sample.

88
These methods are also unfriendly to the environment. We, in our study have
recommended Polytron homogenization to extract oil for laboratory studies.

Dionex systems (2009) reported their experiments where the extraction was
carried out on Jatropha seeds using both European Norm (EN ISO 659:1998) and ASE
100 system methods. The Accelerated Solvent Extraction (ASE®) system showed much
higher oil percentage in relatively less energy intensive process. The reported system
developed by Dionex had great potential to overcome these constraints. Furthermore,
they reported that it has a high potential for application to oil-content testing of the
third generation of biodiesel feedstock (e.g., microalgae). Oil extraction using an ASE
system would require only one to one-and-a-half hours as compared to nine hours
consumed by the Soxhlet extraction. It was also emphasized that the oil extraction
method using the ASE system requires less than one hour as compared to ten hours
consumed by the Soxhlet extraction method, permitting a fully automated extraction
process that can be performed in minutes for fast and easy extraction with low solvent
consumption.

Effect of extraction temperature:


The effect of extraction temperature on the amount of extracted oil is shown in enclosed

figure. At room temperature, the oil extracted in two solvents was 30.3 and 29.8 %,
nearly 20 % less than optimum conditions. It increased to 33.6 % and 34.7 % at 45° C
and 55° C, respectively when extracted with hexane. Similar trend was observed when
the seeds were extracted with petroleum ether. Optimum condition was found to be
the boiling point of solvent 65 °C for hexane, however, optimization was achieved at 55°
C with petroleum ether.

When using hexane as the solvent, the amount of extracted oil increased around
20 % by increasing the extraction temperature from room temperature to 45; there was
a further increase of nearly 35 for hexane and 7 % for petroleum ether when the
temperature was further increased to 55° C.

89
Table 4.11: Optimization of solvent conditions
Oil extraction at different temperatures
Solvent Optimum Room 45 °C 55 °C 65 °C
conditions temperature
Hexane % 37.6 30.3 33.6 34.7 37.6
Pet ether % 35.4 29.8 33 35.3 35.4

Oil extraction at different time intervals


Solvent Optimum 2 hrs of 4 hrs of 6 hrs of 8 hrs of
conditions extraction extraction extraction extraction
Hexane % 37.6 33.8 35.3 37.3 37.6
Pet ether % 35.4 31.2 33.7 35.1 35.4

Oil extraction with seed meal at different mesh size


Optimum More
conditions 0.5 mm or 0.5 to 7.5 7.5 to 9.0 than 9
less mm mm mm
Hexane % 37.6 37.6 35.9 35.1 33.7
Pet ether % 35.4 35.4 34.8 33.9 32.8

Oil extraction with different ratio of solvent to solid


Optimum 2:01 3:01 4:01 5:01
conditions
Hexane % 37.6 37.1 37.2 37.5 37.6
Pet ether % 35.4 34.8 35 35.4 35.4

Effect of extraction time:


Oil extracted varied when the extraction time varied from 2 hours to 8 hours. It
was 33.8 and 31.2 with the two solvents when extracted for 2 hrs. It was 10 to 12 % less
than the optimum when extracted for only 2 hours. Percentage of oil extracted
increased to 35.3 and 33.7 for hexane and petroleum ether, respectively when extracted
for 4 hrs. Optimum extraction was achieved when extracted for 8 hrs., however
differences between 6 and 8 hrs. were insignificant. The amount of extracted oil by
hexane and petroleum ether did not change significantly after 6 h. Most of the oil is
extracted after 6 h although maximum extracted oil is achieved after 8 h (Table 4.11).

90
Oil extraction with seed meal at different mesh size
Size of the seed meal is one of the traits that can change oil extracted at
otherwise optimal conditions. Different seed meal sizes were experimented: 0.5 mm or
less, 0.5 to 7.5 mm, 7.5 to 9 mm, and more than 9 mm. Optimum extraction as achieved
with very small mesh size of feedstock (0.5 mm or less). It was similar to optimum
conditions. The extracted oil was only 33.7 for hexane and 32.8 for petroleum ether
when the seed meal size was more than 9 mm, suggesting this is not an optimum
condition for extraction. Oil extraction at intermediate stage of 7.5 to 9 mm was 35.1 5
for hexane and 33.9 % for petroleum ether. 4 to 6 % less oil was extracted at these
conditions.

Less oil was extracted from the larger particles (>0.9 mm) compared to the
smaller size particles. The reason is that larger particles with smaller contact surface
areas are more resistant to solvent entrance and oil diffusion. Therefore, less amount of
oil will be transferred from inside the larger particles to the surrounding solution in
comparison with the smaller ones. Earlier, Tzia, and Liadakis (2003) & Tzia & Liadakis
(2003) have reported that smaller size particles can cause agglomeration of the fine
particles which may reduce the effective surface area available for the free flow of
solvent to solid. However this was not observed in the present study since the oil is
extracted at high temperature with sufficient solvent. In case the solvent is not
sufficient, this problem may reduce optimum oil extraction. This problem may be more
apparent when oily Jatropha seeds are in use since the oily particles stick together easily
to form paste material which may prevent free interactions between solid and solvent in
case the solvent is not sufficient. . It can be concluded that the small to medium particle
size meal between 0.5 and 0.75 mm are more suitable for solid liquid extraction of
Jatropha seeds.

Oil extraction with different ratio of solvent to solid


Different ratio of solvent to solid is an important factor in extraction procedure.
Different ratios of solvent to solid: 2:1, 3;1, 4:1 and 5:1 were investigated. 4:1 and 5:1
were found to be ideal conditions although differences between 3:1 and 4:1 were very

91
small. It was 37.2 % at 3:1 and 37.55 at 4:1 when extracted with hexane. Similar
differences were observed for petroleum ether. Overall 2:1 and 5:1 did not show wide
variability (Figure 4.21).

Increasing solvent to solid ratio up to a specific limit will increase the yield since the
concentration

gradient between the solid and the liquid phase becomes greater which favors
good mass transfer. Based on the results, the solvent to solid ratio of 4:1 would be
sufficient for both solvents to extract the maximum amount of oil and only marginal
increment of oil amount can be obtained by increasing the ratio to 5:1. Therefore ratio
4:1 is the optimal solvent to solid ratio.

Free fatty acid percentage at various storage conditions


The results show that on keeping the Jatropha seeds for long time, oil content of
the seeds declines. Also, on keeping the Jatropha oil for long intervals, Free Fatty Acid
(FFA) content of the oil increases, resulting in more wastage of oil as well as chemicals
for it.

No water No water No water No water


0 days 15 days 30 days 45 days
FFA 1.1 1.8 2.9 4.6

1 % water 1 % water 1 % water 1 % water


26˚C 30˚C 35˚C 45˚C
FFA 2.4 3.6 4 6.9

Interpretation of biodiversity among accessions

Cluster formation by dendrogram interpretation was drawn to see relations


amongst the accessions collected for the present study. The dendrogram shows
diversity among different accessions within inter and intra-regions. Six clusters were
formed. Cluster1 showed maximum genetical diversity of 20-25%. One (23rd) accession
was found distinct with minimum genetical diversity and therefore it formed its own
cluster (Figure 4.21).

92
Clusters (20%-25%) diversity variation
1. Cluster 1 -Accession number 1,18,12,13,11,14,22
2. Cluster 2- Accession number 9,10,24,29,25,17,21,20,2
3. Cluster 3- Accession number 27,30,26
4. Cluster 4- Accession number 23
5. Cluster 5- Accession number 3,6,4,15,5,16,28
6. Cluster 6 - Accession number 9,8,18

Figure 4. 21: Interpretation of biodiversity among accessions

93
Statistical analysis of various factors
Various morphological and chemical parameters of oil were analyzed Variations
amongst these traits have been statistically analyzed Significant value has been shown
in asterisks.

Table 4.12: Correlation between morphological & physicochemical parameters of 30


accessions

SOUR df MSS value (MSS)**


CE
Oil Seed Kern Seed Mois Seed Seed Seed FFA Aci Ref. Den S. Sap. Insol
% wt. el coat ture leng brea thick d valu sity Grav valu .
wt. wt. % th th ness valu e ity e imp.
e
Treat 30 55. 0.02 0.01 0.06 0.54 4.31 2.30 2.88 0.47 1.3 0.00 0.0 0.00 780 780
ment ** ** ** ** ** ** ** ** ** ** ** ** ** ** **
Repli 2 0.1 0.00 0.00 0.00 0.00 0.41 0.21 0.29 0.00 0.0 0.00 0.0 0.00 0.06 0.65
catio 08 2 0 4 0 6 7 4 0 16 0 00 0 60 0
n
Error 60 0.0 0.00 0.00 0.00 0.00 0.34 0.50 0.17 0.00 0.0 0.01 0.0 0.00 0.01 0.00
52 287 201 243 014 539 274 67 019 193 439 000 000 0376 0166

Conclusions
The fact that Jatropha has adapted itself to a wide range of edaphic and
ecological conditions suggest that there exist considerable amount of genetic variability
to be exploited for potential realization (Rao et al. 2008). Environmental factor in
combination with genetic and physiological factors play important role in determination
of plant potential for seed quality. These characters appear to be under strong genetic
control (Roy et al. 2004). The key for success of any genetic improvement program lies
in the availability of genetic variability for desired traits (Heller 1996). Genetic resource
through global exploration, introduction, characterization and evaluation will provide
strong base for development of elite varieties by various improvement methods. Wide
variation in seed traits and oil percentage were observed in the present study.

94
Chapter 5

Comparative study of phenology, oil and methyl ester properties


Oil and transesterified oil were analyzed for their physico-chemical and fatty acids
(methyl esters) composition. Comparative study of non-edible seed oil can help to
select, screen and identify suitable feedstock for biodiesel production. Data are shown
in Tables 5.1 to 5.6.

Table 5.1: Comparative study of Pongamia, Neem and Jatropha oil and their methyl
ester properties with commercial diesel

Properties Neem seed Pongamia Jatropha Neem Pongamia Jatropha Diesel


oil seed oil seed oil biodiesel biodiesel biodiesel
Seed
wt.(g)/seed 0.22±0.98 2.81±0.88 0.81±1.08

Seed kernel 1.58±0.63 0.54±1.02


wt. g/seed 0.12±0.77
Seed coat wt.
g/seed 0.09±0.33 1.33±0.25 0.26±0.22

Seed length
(mm) 14.40±1.01 46.83±1.16 17.23±1.86

Seed breath
(mm) 6.40±0.26 24.70±1.31 10.57±0.50

Seed thickness
(mm) 7.27±0.21 9.90±0.72 8.27±0.21

Seed Moisture 5.55±0.01 4.52±0.006 6.32±0.005


%
Oil content % 26.87±0.59 33.37±0.04 34.20±0.57
Specific gravity 0.92±0.001 0.93±0.001 0.92±0.001 0.872±0.002 0.894±0.001 0.89±0.001 0.83
Density @15ºC
(g/ml) 0.92±0.001 1.05±0.01 1.13±0.001 0.89±0.006 0.89±0.006 0.88±0.006 0.827
Oil moisture % 1.32±0.01 1.12±0.01 2.07±0.01 0.31±0.006 0.51±0.006 0.22±0.006 -
Insoluble
impurities % by
wt. 8.97±0.006 1.64±0.006 3.32±0.57 1.13±0.006 0.97±0.010 0.81±0.008 -
Acid value (mg
KOH/g) 16.61±0.10 5.71±0.06 6.3±0.06 0.51±0.0047 0.16±0.0047 0.31±0.006 -
F.F.A % 8.25±0.010 2.91±0.047 3.12±0.058 0.26±0.005 0.09±0.005 0.16±0.005 -
Saponification
value (mg
KOH/g) 115.61±0.26 168.93±0.15 174.31±0.09 123.11±0.02 139.21±0.02 148.28±0.05 -
Refractive
Index 1.51±0.058 1.52±0.02 1.52±0.04 1.44±0.015 1.45±0.015 1.44±0.006 1.4759
Color Value
(1''cell) 25.41±0.1 112.52±0.08 31.43±0.10 19.94±0.058 32.81±0.058 12.31±0.06 -
Peroxide Value
(milliequi/1000
g) 12.45±0.14 4.22±0.12 10.8±0.06 2.81±0.06 0.44±0.05 10.81±0.07 -

95
Kinematic
viscosity (cst) 51.13±0.15 49.61±0.26 45.72±0.15 5.81±0.06 5.23±0.05 4.91±0.068 3.08
Flash Point(ºC) 277±1.00 265±1.00 239±1.00 222±0.577 216±0.577 191±0.577 76
Degradability(
days) (Shelf
life) 15-22 90-110 45-60 45-50 >140 130-140 -
*Readings are the mean of triplicate value with standard deviation

Table 5.2:Fatty acids composition of methyl esters of Neem, Pongamia & Jatropha
FATTY ACIDS STRUCTURAL
COMPOSITION FORMULA SYSTEMATIC NAME NEEM PONGAMIA JATROPHA
Palmitic acid (16:0) C16H32O2 Hexadecanoic 13.5±0.100 7.9±0.058 17.2±0.058
Stearic acid (18:0) C18H36O2 Octadecanoic 14.4±0.058 8.9±0.058 9.7±0.058

Oleic acid (18:1) C18H34O2 cis-9-Octadecenoic 49.2±0.100 71.3±0.058 63.1±0.058


Linoleic acid (18:2) C18H32O2 cis-9,cis-12-Octadecadienoic 7.5±0.058 - 0.80±0.012
Arachidic acid (20:0) C20H40O2 Eicosanoic 0.2±0.058 4.65±0.058 -
Behenic acid (22:0) C22H44O2 Docosanoic -_ 4.3±0.058 -_
 All readings are the mean of triplicate value with standard deviation

Table 5.3: Correlation between morphological characters of Jatropha, neem and


Pongamia seeds and oil.

(PHENOTYPICAL MATRIX)
Properties Seed wt.(g)/seed Seed Kernel wt. Seed coat wt. Seed length Seed breath Seed thickness Seed
Moisture Oil Specific gravity Density @15ºC Oil moisture
(g) (g) (g) (mm) (mm) (mm) (%)
(%) (g/ml) (%)

Seed 1.8462 0.9901 1.0000 0.9863 0.0000 0.5979 -


wt.(g)/seed 0.9981 0.9956 0.6020 0.3646 0.4834
(g)
Seed 0.9796 0.9981 0.9946 0.0000 0.6459 0.5520 -
kernel 0.5628 0.9880 0.4209 0.4289
wt.(g)/seed
(g)
Seed coat 0.9989 0.9957 0.9665 0.0000 -
wt.(g)/seed 0.4570 0.5200 0.6743 0.2755 0.5634
(g)
Seed 322.6848 0.9902 0.9533 0.7084 -
length 0.0000 0.4792 0.2299 0.6017
(mm)
Seed 0.9861 0.0000 0.6029 -
breath 92.0004 0.5970 0.3635 0.4844
(mm)
Seed 0.0000 0.7218 0.5130 -
thickness 1.7670 0.4622 0.3325
(mm)
Seed 0.0000
Moisture 0.0000 0.0000 0.0000 0.0000
(%)
Oil (%) 16.0932 - 0.9644 -
0.2801 0.9483
Specific - -

96
gravity 0.0000 0.5240 0.9900
Density -
@15ºC 0.0004 0.6389
(g/ml)
Oil
moisture 0.2508
(%)

(PHENOTYPICAL MATRIX)

Table 5.4: Correlation between physico-chemical properties of Jatropha, neem and


Pongamia seeds and oil
Insoluble Acid F.F.A Saponification Refractive Color Peroxide Kinematic Flash
impurities value (%) value Index Value Value viscosity Point
(g)% by (mg (mg KOH/g) (nD) (1''cell) (m (Cst) (ºC)
wt. KOH/g) eq./1000g)
Insoluble 14.7500 0.9863 0.9845 -0.9546 -0.9759 -0.7211 0.7973 0.5390 0.5790
impurities (g)
% by wt.
Acid value 37.5644 0.9999 -0.9907 -0.9985 -0.5970 0.6869 0.6705 0.7056
(mg KOH/g)
F.F.A (%) 9.2053 -0.9920 -0.9990 -0.5886 0.6792 0.6783 0.7129
Saponification 1052.4444 0.9966 0.4821 -0.5814 -0.7654 -0.7955
value (mg
KOH/g)
Refractive 0.0004 0.5525 -0.6463 -0.7099 -0.7430
Index (nD)
Color 2366.6033 -0.9931 0.1949 0.1473
Value(1''cell)
Peroxide 18.5137 -0.0786 -0.0304
Value
(m eq./1000g)
Kinematic 7.7970 0.9988
viscosity (Cst)
Flash Point 377.3333
(ºC)

Table 5.5: Correlation between physico-chemical properties of methyl ester of


Jatropha, Neem and Pongamia seeds and their oil content

(PHENOTYPIC MATRIX)
Specific Density Oil Insol. Acid F.F.A Sap. R. I. Color Peroxide Kinematic Flash
gravity @15ºC moisture Impurities value (%) Value (nD) Value Value (m viscosity Point
(g/ml) (%) (g) % by (mg (mg (1''cell) eq./1000g) (Cst) (ºC)
wt. KOH/g) KOH/g)
Specific 0.0001 0.0180 0.6744 -0.4737 - - 0.6174 0.7330 0.6392 -0.2387 -0.6004 -0.1647
gravity 0.9961 0.9912
Density 0.0000 0.7505 0.8720 0.0700 0.1147 -0.7754 0.6934 0.7804 -0.9752 0.7887 0.9832
@15ºC(g/ml)
Oil moisture 0.0229 0.3308 - - -0.1645 0.9966 0.9989 -0.8780 0.1856 0.6173
(%) 0.6068 0.5705

97
Insol. 0.0256 0.5494 0.5863 -0.9853 0.2519 0.3744 -0.7421 0.9887 0.9467
Impurities (g)
% by wt.
Acid value 0.0320 0.9990 -0.6842 - -0.5691 0.1523 0.6684 0.2508
0.6703
F.F.A % 0.0070 -0.7163 0.6363 -0.5316 0.1078 0.7012 0.2940
Sap. value 162.5006 - -0.2103 0.6165 -0.9998 -0.8776
0.0826
Refractive 0.0000 0.9917 -0.8356 0.1039 0.5503
Index (nD)
Color Value 107.5226 -0.8994 0.2312 0.6532
(1''cell)
Peroxide 29.4493 -0.6333 -0.9185
Value
(meq./1000g)
Kinematic 0.2381 0.8877
viscosity (Cst)
Flash 270.3333
Point(ºC)

Table 5.6: ANOVA Statistical analysis of seed morphological and oil physico-
chemical parameters of Neem, Pongamia and Jatropha
S.NO SOURCE df MSS value (MSS)**

Seed wt. Seed Seed Seed Seed Seed Seed Oil % Specific Density
kernel coat length breath thickness Moisture gravity
wt. wt. %
1. Treatment 2 11.077** 3.37** 2.74** 1936.11** 552.002** 0.202** 0.000** 96.559** 0.000** 0.000**
2. Replication 2 0.021 0.000 0.00 7.48 0.535 10.602 0.001 0.889 0.000 0.000
3. Error 4 .0161 .0006 .0007 .0037 .8877 .2527 .000 .0064 .0000 .0000

** INDICATES THAT THE STASTICAL MEAN VALUES WAS FOUND SIGNIFICANT AT


P≤ 0.09 LEVEL ANALYSED BY INDOSTAT ANOVA

98
Table 5.7: ANOVA Statistical analysis of seed morphological and oil physico-
chemical parameters of Neem, Pongamia and Jatropha

# SOURCE d MSS value (MSS)**


f

Sp. Dens Oil Insol. A. F.F.A Sap. R. Color Peroxid Kin. Flash
gravi ity mois impu value % value Inde Value e Value viscosi Point(ºC)
ty ture rities x ty
% (g)

1. Treatm 2 0.000 0.154 0.192 0.153 0.192 0.042 0.004 0.000 645.135 176.695 1.428* 1622.00
ent ** ** ** ** ** ** ** ** ** ** * 0**

2. Replicat 2 0.000 0.000 0.000 0.000 0.000 0.000 975.0 0.000 0.0155 0.002 0.0155 2.000
ion 03

3. Error 4 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.001 0.004 0.001 0.000

99
Table 5.8: ANOVA Statistical analysis of methyl esters of neem, Pongamia and Jatropha

S.NO SOURCE df MSS value (MSS)**

Oil Insol. Acid F.F.A Sap. value Refra Color Peroxide Kin. Flash
moisture impurities value % ctive Value Value viscosit Point(ºC)
% (g) Index y

4. Treatment 2 0.066** 88.500** 225.386* 55.232 6314.666* 0.002 14199.62 111.082** 46.782* 2264.000*
* ** * ** 0** * *

5. Replicatio 2 0.000 0.000 0.020 0.007 0.107 0.002 0.047 0.015 0.135 0.666
n

6. Error 4 0.0000 0.0000 0.003 0.0015 0.0266 0.001 0.0033 0.0077 0.0244 1.333
1

** INDICATES THAT THE STASTICAL MEAN VALUES WAS FOUND SIGNIFICANT AT P≤


0.09 LEVEL ANALYSED BY INDOSTAT ANOVA

Study of seed samples


Statistical analysis of oil %, morphological properties (like seed wt., kernel wt., seed
coat wt., moisture percentage, seed length, seed breath and seed thickness) and other
physico-chemical parameters (refractive index, specific gravity, density, saponification value
and insoluble impurities) for Jatropha, neem and Pongamia seed was undertaken. Statistical
analysis, correlation and ANOVA test are shown in the Table 5.3 and 5.5.

All the three non-edible seed oils showed significant variability for 100 weighed seeds
and oil content.

Statistical analysis
Statistics for Correlation and Anova analysis has been carried out to find relationships
between different traits such as physicochemical properties and fatty acids composition. The
values are shown in enclosed Tables. Statistical analysis of seed morphological and other
physic chemical properties of neem, Pongamia pinnata and Jatropha curcas indicate that
statistical mean values were significant at P< 0.09 level.

100
Different accessions of J. curcas and neem and Pongamia pinnata investigated to
compare the seed traits and methyl ester traits exhibited considerable amount of variation in
growth and yield traits. Analysis of variance showed highly significant variations among the
experimental accessions for all traits. It suggests presence of considerable amount of
variability in material which can be used in improvement for different characters including
seed yield provided the germplasm has been collected with proper selection.

101
Chapter 6
Transesterification to manage free fatty acid percentage and viscosity

Introduction
Biodiesel is a fuel obtained from renewable biomass feedstock that can be used in
diesel engines as neat fuel or blended at various proportions with conventional fossil diesel
fuel (Arzamendi et. al., 2007; Ayhan, D. A., 2009; Zeng, et. al., 2008; Van Gerpen, 2005.
It consists of mono-alkyl esters usually produced by transesterification of vegetable oils,
animal fats and cooking oils with low-molecular weight alcohols, most c o m m o n l y
m e t ha nol o r ethanol. Biodiesel is an excellent substitute for conventional diesel fuel
because of being renewable, nontox ic and biodegradable. The energy content, cetane
number and viscosity of biodiesel are similar to those of petroleum-based diesel fuel.
Moreover, it is essentially sulfur-free and emits significantly fewer particulates, unburnt
hydrocarbons and less carbon oxides as compared with conventional fossil fuels (Xie, & Li,
2006; Meher & Dharmagadda, 2006; Michael, et. al., 2006; Shu, et. al., 2007; Phan &
Phan, 2008; Bhatti, et. al., 2008; Pramanik , 2003; Gao et. al., 2009; Meher, et. al.,
2006). At present, the high cost of biodiesel is the major obstacle for its commercialization.
Approximately 70% of biodiesel cost is attributed to raw feedstock (Zhang and Jiang, 2008;
Zhang, et. al., 2003; Vyas, et. al., 2009). Use of cheap and non-edible vegetable oils,
animal fats and waste oils as raw feedstock for bio­ diesel production is an effective way to
reduce the cost particularly in developing countries. Jatropha curcas L. oil is a potential
cheap feedstock for biodiesel production as compared with refined and edible-grade oils
such as rapeseed oil, soybean oil and sunflower oil that are common feedstock in USA and
Europe.

Biodiesel, which is made from renewable sources, consists of the simple alkyl esters of
fatty acids. As a future prospective fuel, biodiesel has to compete economically with
petroleum diesel fuels. One way of reducing the biodiesel production costs is to use the less
expensive feedstock containing fatty acids such as inedible oils, animal fats, waste food oil and
byproducts of the refining vegetables oils (Veljkovic, et. al., 2006). The availability and
102
sustainability of sufficient supplies of less expensive feedstock will be a crucial determinant
delivering a competitive biodiesel to the commercials filling stations. Fortunately, inedible
vegetable oils, mostly produced by seed-bearing trees and shrubs can provide an alternative.
With no competing food uses, this characteristic turns attention to Jatropha curcas, which
grows in tropical and subtropical climates across the developing world (Openshaw, 2000).

Earlier, oil contents, physicochemical properties, fatty acid composition and energy
values of Jatropha species were investigated and reported by several workers (Banerji, et. al.,
1985; Kandpal and Madan, 1995; Kumar et al., 2003; Pramanik, 2003; Akintayo, 2004; Shah,
et. al., 2004).

It is considered that Jatropha has toxic substance (Hirota, et. al., 1988; Gandhi, et. al.,
1995; Makkar, et. al., 1998; Haas, and Mittelbach, 2000; Abdel Gadir, et. al., 2003). Presence
of some anti-nutritional factors such as curcin render this oil unsuitable for cooking
purposes has been reported by many workers (Tamlampudi, et. al., 2008; Tapanes, et.
al., 2008; Sarin, et. al., 2007).

Transesterification
The conventional industrial production of biodiesel is via transesterification of
crude oil with a homogeneous strong base catalyst (e.g., NaOH, KOH or NaOCH3) or
acid catalyst ( e.g., H 2S04) (Xie, and Li, 2006; Tapanes, et. al., 2008). Other methods such

as transesterification with solid catalysts, biocatalysts and non-catalytic supercritical


methanol were studied extensively (Arzamendi, et. al., 2007; Ayhan, D. A., 2009; Zeng,
et. al., 2008; Van Gerpen, 2005; Zhang, et. al., 2003; Tamlampudi, et. al., 2008; Xi and
Davis, 2008; Xiw & Huang, 2006; Hawash, et. al., 2009; Kumari et. al., 2009;
Demirbas, 2009; Yin, et. al., 2009).

Vyas, et. al., (2009) reported 84 % biodiesel yield w h e n Jatropha curcas oil with high
free fatty acids (FFAs) w a s u s e d a s crude feedstock with KN0 3 /Ah03 solid catalyst at 70
°C for 6 h. Maximum biodiesel yield of 94% with immobilized lipase catalyst at 55° C for
48 h was achieved by Kumari, et. al. (2009). 100% biodiesel yield could be obtained at
320° C and 8.4 MPa when Jatropha curcas oil was transesterified using supercritical

103
methanol w it ho ut c a t a l y s t by Hawash, et. al. (2009). The above methods need either
long reaction times (up to 48 h) or high temperatures and pressures ( e.g., 320°C and 8.4
MPa) that will increase cost and consume a lot of energy.

Trans-esterification & conditions to maximize biodiesel yield


The present study was undertaken to optimize transesterification conditions to
maximum biodiesel production of Jatropha curcas oil. Different variables have been evaluated
to optimize the reaction process. This optimized reaction process of transesterification has
been taken as a standard process for further study. The properties of oil and biodiesel were
also analyzed and compared with Indian standards (prescribed by the Bureau of Indian
Standards, BIS).

Process variables for optimization of transesterification reaction


Relevant variables and conditions that can affect transesterification

a) - Raw oils
b)- Nature of alcohols
c)- Catalysts as variable
d)- Period of reaction
f)- Reaction temperature
g)- Concentration and ratio of alcohol for reaction

High FFA in Jatropha & Neem oil


Jatropha oil is reported to have high free fatty acid percentage ( up to 15%) (Berchmans &
Hirata, 2008). In many cases crude Jatropha curcas oil quality deteriorates gradually due to
improper handling and inappropriate storage condition. Deterioration is usually caused due to
improper handling of crude Jatropha curcas oil and due to increase in water content of oil. In
addition, exposing the oil to open air and sunlight for long time would also affect the
concentration of free fatty acid percentage increase significantly to high level of FFA above
1%. The free fatty acid percentage of crude Jatropha curcas oil will vary and depend on the
quality of feedstock.

104
The free fatty acid percentage and moisture contents have significant effects on the
transesterification of glycerides with alcohol using catalyst (Goodrum, 2002). The high free
fatty acid content (>1% w/w) will result in soap formation and the separation of products will
be exceedingly difficult, and as a result, yield of biodiesel is affected. Crabbe, et. al. (2001)
recommended that the acid-catalyzed esterification of the oil can be an alternative but it is
much slower than the base-catalyzed transesterification reaction. Therefore, an alternative
process such as a two-step process was investigated for feedstock having the high FFA content
(Ghadge and Raheman, 2005; Veljkovic´ et al., 2006). When a base homogeneous catalyst is
used, free fatty acid content reacts with the catalyst to produce emulsified soap that will
inhibit biodiesel production.

Shekhar, et. al. (2009) reported that the viscosity of vegetable oil is about ten times
higher than that of diesel, thus the vegetable oil cause poor fuel atomization,
incomplete combustion and carbon deposition on the injector and valve seats resulting in
serious engine fouling. This necessitates the reduction in viscosity of the vegetable oils for
use as fuel in CI engines.

Methods
Transesterification (alcoholysis) is the chemical reaction between triglycerides
and alcohol in the presence of catalyst to produce mono-esters. The long and branched chain
triglyceride molecules are transformed to mono-esters and glycerin.
Transesterification process consists of a sequence of three consecutive reversible
reactions. That is, conversion of triglycerides to diglycerides, followed by the
conversion of diglycerides to monoglycerides. The glycerides are converted into glycerol
and yielding one ester molecule in each step. The properties of these esters are
comparable to that of diesel. The overall transesterification reaction can be
represented by the following reaction scheme.

Triglyceride + ROH Diglyceride + RICOOR Diglyceride + ROH

Monoglyceride + RIICOOR Monoglyceride + ROH Glycerol + RIIICOOR

105
Method of analysis of free fatty acid percentage
For titration first 0.1 to 10 g of oil was weighed and dissolved in about 50 ml of a
suitable solvent. Methanol, ethanol and ether are some normally used solvents; in this case
methanol was used as the solvent. It was heated gently for some time. A small drop of
indicator was added. Phenolphthalein was used as indictor. Then the solution was titrated
with KOH. The amount of KOH required, in milligram (mg) to neutralizing the free
fatty acid in one gram of oil expressed as a number is known as acid number. From acid
number the free fatty acid present in the oil can be calculated.

Acid number calculation for the selected sample = 56.1×N×V/M

Where, V is the number of ml of KOH, N is the normality of KOH, M is the mass (in g) of
sample

Esterification procedure

Conditions
The experiments were carried out with the two different reactors. First reaction was
done in glass made vessel in laboratory scale and secondly reaction has been carried out in
reactor to compare the productivity of the biodiesel. There was a simple strategy for
conducting the experiments that each reaction was repeated in triplicate form to assess the
transesterification process and rate of yield.

Transesterification Method
The transesterification method comprises of several stages of reaction, isolating
phases, weighing and washing the final product to eliminate any un-reacted components. The
underlying assumption of the process was that the reaction is complete and the oil was
converted to biodiesel. Typical reaction times was 45 minutes. When ninety five to ninety nine
percent of the reaction get completed in the 45 min it was reasonable to assume that the
reaction has gone to completion.

106
Acid Esterification:
The firsts step reduces the free fatty acid value of crude oil to about 2% using
acid catalyst

Alkaline Esterification:
After removing the impurities of the product of first step, it is transesterified
to mono-esters of fatty acids using alkaline catalyst. The parameters affecting the process
such as alcohol to oil molar

A round bottom flask is used as laboratory scale reactor for these experimental
purposes. A hot plate with magnetic stirrer arrangement is used for heating the
mixture in the flask. The mixture is stirred at the same speed for all test runs. The
temperature range of 50–60 °C is maintained during this experiment.

Acid Esterification
One liter of crude oil requires 250 ml of methanol for the acid esterification
process. The oil is poured into the flask and heated to about 50 °C. The ethanol is
added with the preheated oil and stirred for a few minutes. 1% of sulphuric acid is also
added with the mixture. Heating and stirring is continued for 30 min at atmospheric
pressure. On completion of this reaction, the product is poured into a separating funnel for
separating the excess alcohol. The excess alcohol, with sulphuric acid and impurities moves to
the top surface and is removed. The lower layer is separated for further processing (alkaline
esterification).

Reaction temperature
At room temperature the conversion efficiency is noted to be very low, even
a fter 2 hrs. of stirring. With increase in temperature the conversion takes place at a faster
rate. The optimum temperature for this reaction is found to be in the range of 50±5
°C. At higher reaction temperatures, there is a chance of loss of methanol and increase in
darkness of the product. High reaction temperature increase the production cost of biodiesel
also.

107
Alkaline Esterification
Alkaline catalyzed esterification process uses the experimental setup of acid catalyzed
pretreatment process. The products of first step are preheated to the required reaction
temperature of 50±5 °C in the flask. Meanwhile, 5 g of KOH is dissolved in 250 ml methanol
and is poured into the flask. The mixture is heated and stirred for 30 min. The reaction is
stopped, and the products are allowed to separate into two layers. Glycerin which is
heavier deposits at the bottom and the esterified neem oil is obtained at the top
portion. The esterified oil is separated from the funnel and now it is ready for blending with
diesel.

Reaction temperature
The maximum yield of ester is obtained at the temperatures of 50±5 °C. The decrease
in yield is observed when the reaction temperature goes above 55 °C. The reaction
temperatures greater than 60 °C should be avoided, in the case of oil, because they tend to
accelerate saponification of the glycerides by the alkaline catalyst before completion of the
alcoholysis.

Acid value
The acid value of the reaction mixture in the first stage was determined by the acid
base titration technique (ASTM, 2003). A standard solution of one mol potassium
hydroxide solution was used.

GC method
The compositions of Jatropha oil and biodiesel were analyzed by Gas Chromatography
(GC, Shimadzu) with a flame ionization detector and a capillary column. Oxygen
free nitrogen was used as a carrier gas at a flow rate of 1.4 mL/min, initial oven
temperature of 170 °C (2 min), increased at the rate of 5.0 °C per min to increase the final
temperature of 220 °C (3 min); injector temperature of 250 ° C, detector temperature
of 280 C and the split ratio of 39/1. Analysis was carried out by injecting 1 m i c r o
l i t e r sample solution (0.25 mL Jatropha oil or biodiesel dissolved in 9.75-mL
dichloromethane. The analysis conditions for biodiesel were similar to Jatropha oil, but

108
carrier gas at a flow rate of 1.0 mL/min and the split ratio of 20:1. The fatty acids and their
esters as well as their compositions of Jatropha oil and biodiesel were identified and
quantified by comparing their retention times to the standard retention times of fatty
acids and their esters. The methyl esters were identified by comparing retention time to
the retention time of standard methyl ester of fatty acid. Quantitative analysis of methyl
esters/biodiesel was determined following Nguyen, et. al (2005).

HPLC method
Fatty acids in the methanol extracts from the oils were analyzed by using high
performance liquid chromatography (HPLC) o f Shimadzu equipped with Capcell Pak C18,
25 cm in length · 4.6 mm column, ultraviolet detector at 220 nm operated at room
t e m pe rat ure with 1 ml/min flow rate of 85% acetonitrile containing 20% of 0.1%
H3PO4 as a carrier solvent. 2 5 m i c r o l i t e r sample was injected. The peaks were
compared with that of standards.

Methods, units & standards


Property unit Standard Method

Density kg/m3 D 4052-96 Relative density bottle

Viscosity D 445-03 Redwood viscometer


mm2/s
Calorific value MJ/kg D 240-02 Bomb calorimeter

Flash point °C D 93-02a

Results & discussion


Table 6.1: Fatty acid composition of crude Jatropha curcas oil
Fatty acid Formula Systemic name Structure wt.%
Myristic C14H28O2 Tetradecanoic 14:0 0-0.1
Palmitic C16H32O2 Hexadecanoic 16:0 14.1-15.3
Palmitoleic C16H30O2 cis-9-Hexadecenoic 16:1 0-1.3
Stearic C18H36O2 Octadecanoic 18:0 3.7-9.8
Oleic C18H34O2 cis-9-Octadecenoic 18:1 34.3-45.8
Linoleic C18H32O2 cis-9,cis-12-Octadecedianoic 18:2 29.0-44.2
Linolenic C18H30O2 cis-6,cis-9,cis-12-Octadecatrienoic 18:3 0-0.3

109
Arachidic C20H40O2 Eicosanoic 20:0 0-0.3
Behenic C22H44O2 Docosanoic 22:0 0.2

Transesterification using alkali a s a c a t a l y s t


Stoichiometrically, three moles of alcohol are required for each mole of
triglyceride, but in practice a higher molar ratio is employed in order to displace the
equilibrium for getting greater ester production. Though esters are the desired
products of the transesterification reactions, glycerin recovery also is important due to its
numerous applications in different industrial processes. Commonly used short chain
alcohols are methanol, ethanol, propanol and butanol. The yield of esterification is
independent of the type of alcohol used. Therefore, the eventual selection of one of these
three alcohols will be based on cost and performance considerations. The methanol is
used commercially because of its low price. Alkaline hydroxides are the most effective
transesterification catalysts as compared to acid catalysts. Potassium hydroxide and sodium
hydroxide are the commonly used alkaline catalysts. Alkaline catalyzed transesterification of
vegetable oils is possible only if the acid value of oil is less than 4. Higher percentage of FFA in
the oil reduces the yield of the esterification process.

One-step alkali-base catalyzed transesterification was carried out for methyl ester
production process from crude Pongamia pinnata oil, crude neem oil, and crude Jatropha
curcas oil at reaction temperature of 70 °C for 3 hrs. of reaction time in tubes immersed in
water bath and stirred at 400 rpm. Different catalyst NaOH-to-oil ratios (0.5%, 1.0%, 1.5%,
2.0%, 2.5% and 3.0% w/w) and different methanol-to-oil ratios (10%, 15%, 20%, 25%, 30%
and 40% w/w) were used to investigate their influence on the methyl ester yields of the oils.
The reaction mixture was allowed to settle for 2 h to overnight before separating the glycerol
layer and the top layer including methyl ester fraction was removed in a separated bottles,
weight and analyzed by GC. The methyl esters were washed with water to remove impurities.

It was established that transesterification depends on several basic variables, namely,


catalyst type, alcohol type, catalyst-to-oil ratio, alcohol-to-oil ratio, reaction temperature,
reaction time, agitation rate, FFA, and water content of oils (Ma, and Hanna, 1999).

110
Reaction Procedure
Once the reaction was completed, the rubber stopper and magnetic stirrer were
removed and the reactants weighed. The conical flask was periodically weighed until the loss
of weighed mass was equal to the approximate mass of excess methanol /ethanol that was
calculated theoretically. The contents of the conical flask was placed in a separating funnel
and allowed to settle overnight. The glycerol was then separated and weighed. The remaining
contents of the separating funnel were also weighed and returned to the funnel and washed
with warm water. Water and moisture was removed by rotary evaporator, final product was
weighed and calculated for the biodiesel yield percentage.

It was found that adding and washing of ester layer with cold water, will enhance the
washing step and remove the impurities that seemed to be responsible for formation of an
emulsion. Emulsion formation was a major handicap in the process. Emulsion tends to be
form from reactions with alkali and formed when the mixture was washed and agitated with
water. The emulsions are hard to break once formed; these need to be avoided. To minimize
their formation, warm water was used for washing.

Figure 6.1:- Formation of emulsion during washing of methyl ester; and its removal

The reactions were carried out in two reactors. Laboratory scale experiments were
done using a conical flask, while others in the reactor. Gravimetric method involved washing

111
of the final ester phase to remove any residual glycerol or unreacted/ partially reacted oil.
Washing and rigorous agitation resulted in emulsion formation. The ensuing emulsion seemed
to be unbreakable, but on the addition of table salt which was used in the soap Industry, the
emulsion dissipated to some extent.

Use of various alcohols


Different methylating agent has been used for the reaction to study suitability of
different alcohol for the reaction. Reaction was carried out in presence of methylating agent
mixed catalyst in 65-70ºC temperature for 30 minutes in presence of NaOH as a catalyst.
Various alcohols viz. methanol, ethanol and butanol were used along with crude oil in the
ratio of 45% w.r.t. the volume of oil in order to check their suitability for the reaction.

Table 6.2: Comparison between catalytic alcohols for transesterification


Methylating MeOH EtOH BuOH
Agent:-
Amount (ml) 45 45 45
Amount of oil (ml) 100 100 100
Catalyst NaOH NaOH NaOH
Catalyst amount 1 ml 1 ml 1 ml
Reaction time 30 min. 30 min. 30 min.
Reaction temp. 65-70ºC 65-70ºC 65-70ºC
Methyl ester yield 70.5 ml 65.2 ml 50.73 ml
Byproducts MeOH, catalyst, glycerol EtOH, catalyst, glycerol, Saponified and
and saponified products and saponified products jelly like products
Recovery of 30 approx. 20 approx. 10 approx.
Alcohol (ml):-
Result Highest yield of Biodiesel Relatively low yield Not suitable

Maximum biodiesel yield was found with Methanol (70.5 %) and maximum recovery of
alcohol was observed when methanol was used as methylating agent. It was found that

112
methanol in presence of catalyst was most suitable for the maximum yield of biodiesel.
Ethanol was close to methanol but butanol was not found to be an alcohol of choice (Table).

Suitable catalyst
0.5 ml of different catalysts viz. NaOH, KOH, Na2CO3, KCl, NaHCO3, K2CO3 and Na2SO4
were used. The Reaction was carried out in the presence of methylating agent (45% methanol)
mixed with catalysts at 60-65º temperature for 30 minutes to check suitability of catalysts. As
shown in Table 28 nil amount of biodiesel was formed with catalysts NaOH, KOH, Na2CO3, KCl,
NaHCO3, K2CO3 and Na2SO4 hence, were discarded. No biodiesel was formed with most of the
catalysts, even NaOH and KOH had poor yield of 52 to 58 % methyl esters. Reactions with
NaOH & KOH yielded 55.0-58.0 ml and 52.6-54.2ml of biodiesel, respectively.

Table 6.3: Comparative study of type and amount of catalyst for transesterification
Methylating MeOH MeOH MeOH MeOH MeOH MeOH MeOH
Agent
Amount of 45 45 45 45 45 45 45
methylating
Agent(ml)
Amount of oil 100 100 100 100 100 100 100
(ml)
Catalyst NaOH KOH Na2CO3 KCl K2CO3 NaHCO3 Na2SO4
Amount of 0.5 0.5 0.5 0.5 0.5 0.5 0.5
catalyst (ml)
Reaction time 30 min. 30 min. 30 min. 30 min. 30 min. 30 min. 30 min.
Reaction temp 65-70°C 65-70°C 65-70°C 65-70°C 65-70°C 65-70°C 65-70°C
Methyl ester 55.0-58.0 52.6-54.2 0.0 0.0 0.0 0.0 0.0
yield (ml)
Byproducts Biodiesel, Biodiesel, No biodiesel No No No No
MeOH, MeOH, biodiesel biodiesel biodiesel biodiesel
catalyst catalyst
glycerol glycerol
Alcohol 38.5 ml 25 ml 20.5 ml 10.2 ml 30.1 ml 10.0 ml 5.2 ml
Recovery
Result Applicable Applicable Not Not Not Not Not
applicable applicable applicable applicable applicable

113
However, when 1 ml of catalyst was used, the recovery improved. 70 % methyl esters
were formed with NaOH. It was 65.8 % with KOH as catalyst. Recovery with other catalysts
was too low.

Table 6.4: Comparative study of catalyst for transesterification


Methylati MeOH MeOH MeOH MeOH MeOH MeOH MeOH
ng Agent:-
Amount of 45 45 45 45 45 45 45
methylatin
g
Agent
(ml):-
Amount of 100 100 100 100 100 100 100
oil (ml):-
Catalyst:- NaOH KOH KCl NaHCO3
Na2CO3 K2CO3 Na2SO4

Amount of 1 1 1 1 1 1 1
catalyst(m
l):-
Reaction 30 min. 30 min. 30 min. 30 min. 30 min. 30 min. 30 min.
time:-
Reaction 65-70ºC 65-70ºC 65-70ºC 65-70ºC 65-70ºC 65-70ºC 65-70ºC
temp.:-
Methyl 70.0 65.8 0.68 0.78 0.85 1.0 1.0
ester yield
(ml):-
Byproduct MeOH, MeOH, Jelly like Jelly like Jelly like Jelly like Jelly like
s catalyst, catalyst, formation formation formation formation formation
glycerol and glycerol take place take place take place take place take place
saponified and during during during during during
material saponifie reaction reaction reaction reaction reaction
d
material
Recovery 30.5 28.5 20.5 10.5 30.2 10.5 15.3
of Alcohol
(ml):-
Result Successfully Applicabl Not Not Not Not Not
applicable e for applicable applicable applicable applicable applicable
for Biodiesel Biodiesel
production producti
on

114
Time required for reaction
Experiment was carried out to study the effect of reaction time for transesterification
in presence of various catalysts. As shown in Table, reactions were carried out for 30 and 45
minutes, biodiesel (ester) yield increased substantially when the duration was increased to 45
minutes. NaOH and KOH catalyst based reaction yielded 80.5 -82.0 ml and 73.6-77.3 ml
biodiesel, respectively.

Reaction period
Study was carried out to study effect of reaction time of transesterification. Increasing
the time from 30 minutes to 1 hour increased production of biodiesel (methyl esters). All
reactions were carried out in the presence of NaOH and KOH.

Table 6.5: Comparative study of duration of reaction for transesterification


Methylating MeOH MeOH MeOH MeO MeOH MeOH MeOH
Agent:- H
Amount of 45 45 45 45 45 45 45
methylating
Agent(ml):-

Amount of oil 100 100 100 100 100 100 100


(ml):-
Catalyst (ml):- NaOH KOH KCl
Na2CO3 K2CO3 NaHCO3 Na2SO4

Amount of 1 1 1 1 1 1 1
catalyst (ml):-
Reaction 45 min. 45 min. 45 min. 45 45 min. 45 min. 45
time:- min. min.
Reaction 65-70ºC 65-70ºC 65-70ºC 65- 65- 65- 65-
temp:- 70ºC 70ºC 70ºC 70ºC
Methyl ester 80.5 -82.0 73.6-77.3 0.0 0.0 0.0 0.0 0.0
yield (ml):-
By products MeOH, MeOH, Jelly formation takes place
catalyst, catalyst,
glycerol and glycerol and
biodiesel biodiesel
Recovery of 30 approx. 20 20 10 30 10 15
Alcohol(ml):- approx. approx. approx. approx. approx. approx.

115
Result Successfully Successfully Not applicable
applicable for applicable
Biodiesel for
production Biodiesel
production

Table 6.6: Comparative study of time for transesterification production


Methylating MeOH MeOH MeOH MeOH MeOH MeOH
Agent:-
Amount of 45 45 45 45 45 45
methylating
Agent(ml):-
Amount of oil 100 100 100 100 100 100
(ml):-
Catalyst:- NaOH KOH NaOH KOH NaOH KOH
Amount of 1 1 1 1 1 1
catalyst(ml):-
Reaction 30 min. 30 min. 45 min. 45 min. 1 hour 1 hour
time:-
Reaction 65-70ºC 65-70ºC 65-70ºC 65-70ºC 65-70ºC 65-70ºC
temp.:-
Methyl ester 70.0 65.8 80.5 -82.0 73.6-77.3 80.5-82.9 73.6-77.3
yield (ml):-
Byproducts catalyst , glycerol & saponified material

Recovery of 30.5 28.5 30.5 20.8 20.5 20.1


Alcohol(ml):-

Methyl ester formation was 70 % when reaction time was 30 minutes with NaOH as a
catalyst. It increased to 80.5 to 82 %. There was no substantial increase in methyl ester
formation when the reaction period was increased to 60 minutes (80.5 to 82.9 of methyl
esters. Similar response was observed with KOH.

Effect of reaction temperature


The reaction temperature variations were 60-65°, 65-70°, 75-80°C. The reaction time
was kept constant at 45 min. Similarly, methanol (45 ml) and NaOH (1 ml) concentrations
were kept constant. 70-75°C temperature was found to be ideal with 85-88.2 ml of methyl
esters.

116
Table 6.7: Comparative study of reaction temperature on transesterification
Amount of methylating 45 45 45 45
Agent(ml):-
Amount of oil (ml):- 100 100 100 100
Catalyst:- NaOH NaOH NaOH NaOH

Amount of catalyst (ml):- 1(g) 1(g) 1(g) 1(g)

Reaction time:- 45 min. 45 min. 45 min. 1 hour

Reaction temp.:- 60-65°C 65-70°C 70-75°C 75-80°C

Methyl ester yield (ml):- 72.5-73 80.5-82 85-88.2 85-88.2

Byproducts MeOH, catalyst glycerol and saponified material

Recovery of Alcohol (ml):- 38.5 ml approx. 30.5 ml 25.0 ml approx. 10.5 ml


approx. approx.

Amount of alcohol required for transesterification

Table 6.8: Comparative amount of alcohol required for transesterification


Methylat MeOH MeOH MeOH MeOH MeOH MeOH MeOH MeOH MeOH
ing
Agent:-
Amount 25 35 45 55 65 75 85 95 105
(ml):-
Amount 100 100 100 100 100 100 100 100 100
of oil
(ml):-
Catalyst: NaOH NaOH NaOH NaOH NaOH NaOH NaOH NaOH NaOH
-
Amount 1g 1g 1g 1g 1g 1g 1g 1g 1g
of
catalyst:-
Reaction 45 min. 45 min. 45 min. 45 min. 45 min. 45 min. 45 min. 45 min. 45 min.
time:-
Reaction 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C
temp.:-
Methyl 60.0- 68.0- 95.2- 65.0– 70.0- 75.0- 82.0- 80.0- 92.0-94.5
ester 63.2 72.0 97.8 68.7 73.0 80.0 85.0 82.0
yield
(ml):-
Byprodu MeOH catalyst glycerol and saponified material
ct
Recover 15.0 20.0 25.0 30.5 35.5 40.2 45.1 50.3 75.5
y of
Alcohol:-

117
Result Highly suitable
for high yield

Amount of methanol required for reaction


The molar requirement of methanol has been reported to be 30% - 90%. In order to
optimize the amount of methanol required for the reaction, experiments were conducted
with 35%, 45%, 55%, 65%, 75%, 85%, 95% and 105% methanol. The concentration of NaOH,
reaction temperature and reaction time were kept constant at 1 ml alkali at 70-75°C and 45
min, respectively.

The results clearly indicate that the optimum concentration of methanol required for
effective transesterification of Jatropha oil was 45%. Moreover, it was found that when the
concentration of methanol was increased above or decreased below the optimum, there was
significant increase in the biodiesel production, with the excess or short falling concentration
of methanol only contributed to the increase in recovery of methanol and emulsion also.

Table shows that the maximum ester yield of 95.2-97.8 % was obtained using 45 %
methanol. Apparently, 105 ml methanol catalyzed reaction yielded 92-94.5% of biodiesel but
it was not found appropriate for its effect on recovery of methanol.

Table 6.9: Amount of alcohol required for transesterification


Methylating Agent:- MeOH MeOH MeOH MeOH MeOH MeOH MeOH MeOH MeOH
Amount of methylating 25 35 45 55 65 75 85 95 105
Agent(ml):-

Amount of oil (ml):- 100 100 100 100 100 100 100 100 100
Catalyst:- NaOH NaOH NaOH NaOH NaOH NaOH NaOH NaOH NaOH
Amount of catalyst:- 1g 1g 1g 1g 1g 1g 1g 1g 1g
Reaction time:- 45 45 min. 45 min. 45 min. 45 min. 45 min. 45 min. 45 min. 45 min.
min.
Reaction temp.:- 70- 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C 70-75°C
75°C
Methyl ester yield (ml):- 60.0- 68.0- 95.2- 65.0 – 70.0- 75.0- 82.0- 80.0- 92.0-
63.2 72.0 97.8 68.7 73.0 80.0 85.0 82.0 94.5
Removal for purification:- MeOH catalyst glycerol and saponified material
(Byproducts)

Recovery of Alcohol:- 15.0 20.0 25.0 30.5 35.5 40.2 45.1 50.3 75.5

118
Concentration of reaction catalyst
Study on catalyst ratio was undertaken to optimize its requirement for optimum
transesterification. Three levels: 0.25 ml, 0.50 ml and 1.0 ml of catalyst were studied with
other factors of alcohol (45 ml), temperature (70-75°C ) and duration (45 minutes) kept
constant.

It was found that methyl esters formation increased from 23.0 - 97.5 percentage and
thus maximum yield of esters was formed with 1 ml of alkali as shown in table. Optimum alkali
required for the reaction was thus found to be 1ml of NaOH for Jatropha oil.

Comparison of transesterification reaction in crude Jatropha curcas oil, crude


Pongamia pinnata oil and crude Neem oil
Having standardized optimum alcohol ratio, catalyst, reaction temperature and
reaction time, these conditions were applied on three common non edible vegetable oils:
Neem, Pongamia pinnata and Jatropha curcas.

It was observed that highest biodiesel was produced from Jatropha crude oil (97.0-
98.5 ml) and minimum biodiesel was produced from neem crude oil (75.5-78.9 ml) at
optimum conditions.

TABLE 6.10: Effect of different concentration of catalyst on the yield of Biodiesel


Methylating MeOH MeOH MeOH
Agent:-
Amount of 45 45 45
methylating
Agent(ml):-
Amount of oil 100 100 100
(ml):-
Catalyst:- NaOH NaOH NaOH
Amount of catalyst 0.25 0.50 1.0
(ml):-
Reaction time:- 45 min. 45 min. 45 min.
Reaction temp.:- 70-75°C 70-75°C 70-75°C
Methyl ester 23.0 40.2 95.2-97.5
yield:-
Byproducts Saponified material MeOH catalyst glycerol and saponified material
formed

119
Recovery of 40.0 45.5 25.0
alcohol(ml):-
Result Not applicable for Successfully applicable for Biodiesel production
conversion

Table 6.11: Comparison of transesterification with Jatropha, Pongamia & Neem oil
Non-edible oil Jatropha Pongamia Neem
Methylating Agent:- MeOH MeOH MeOH
Amount of methylating 45 45 45
Agent(ml):-
Amount of oil (ml):- 100 100 100
Catalyst:- NaOH NaOH NaOH
Amount of catalyst 1.0 1.0 1.0
(ml):-
Reaction time:- 45 min. 45 min. 45 min.
Reaction temp.:- 70-75°C 70-75°C 70-75°C
Methyl ester yield:- 97.0-98.5 82.5-85.5 75.5-78.9
Removal for Saponified material MeOH catalyst glycerol and saponified
purification:- formed material
(Byproducts)
Recovery of 65.5 63.5 60.3
alcohol(ml):-
Result Successfully applicable for Biodiesel production

The study suggests that optimum temperature of transesterification reaction should


be 45 minutes in high temperature and in 45 ml of alcohol to get highest yield as biodiesel.

Transesterification using acid a s a c a t a l y s t


Transesterification using acid as catalyst was undertaken followed by base-catalyzed
transesterification for converting crude Jatropha curcas oil to methyl esters of fatty acids. The
first step was acid esterification and pretreatment for removing free fatty acid in the oil,

120
which is mainly a pretreatment process, which could reduce the free fatty acid
percentage. The reaction temperature was 50 °C.

The process would convert free fatty acid to esters using acid catalyst (H2SO4 1%
w/w) thus reducing the free fatty acid concentration of crude Jatropha curcas oil to below
2%. Second step was alkali base catalyzed transesterification.

Acid pretreatment
The reaction glass tubes with crude Jatropha curcas oil w a s heated with
concentration H2SO4 acid s olution (1.0% based on the oil weight) in methanol at 50 °C.
Different methanol to oil ratios: 0.10, 0.15, 0.20, 0.25, 0.30, 0.40, 0.50, 0.60 and 0.70 by
weight were used to investigate their influence on the acid value of crude Jatropha curcas
oil. The acid values of the reaction mixtures were measured. The reaction mixtures were
a na ly ze d us ing high performance liquid chromatography (HPLC) and by gas
chromatography (GC). After one hour of reaction, the mixture was allowed to settle for
2 h and the methanol–water fraction at the top layer was removed. The optimum condition
having the lowest acid value was used for the main transesterification reaction.

Base catalyzed transesterification


In the second step, optimum NaOH to oil ratio and methanol to oil ratio were
investigated. T he pretreated oil from the first step was heated at 50 oC. NaOH & methanol
solution at 0.5%, 1.0%, 1.5%, 2.0% and 3.0% w/w of the oil were heated to 50 oC. The
reaction mixture was heated and stirred again at 65.0 ± 0.5 oC and 400 rpm for 2 h. The
mixture was allowed to settle for overnight. Glycerol layer w a s s e p a r a t e d to get the
methyl ester layer of fatty acids on the top. Methyl esters were analyzed by GC. Same
procedure was conducted to investigate optimum methanol to oil ratio (NaOH amount in
various methanol to oil ratio by weight were 0.1, 0.2, 0.35, 0.5, and 0.6).

Table 6.12; Percent FAME yields of alkali base transesterification of crude Jatropha curcas
oil
NaOH to oil ratio Percent FAME Methanol to oil ratio Percent FAME
%w/w yield %w/w yield

121
0.5 0 10 10
1 0 20 24
1.5 24 30 32
2 32 40 63
2.5 65 50 78
3 48 60 58
3.5 0 70 33
4 0 80 12

Table 6.13: Percent FAME yields of alkali base transesterification of crude Pongamia
pinnata oil
NaOH to oil ratio Percent FAME Methanol to oil ratio Percent FAME
%w/w yield %w/w yield
0.5 0 10 10
1 22 20 35
1.5 68 30 65
2 51 40 85
2.5 34 50 32
3 22 60 22
3.5 0 70
4 0 80

Table 6. 14: Percent FAME yields of alkali base transesterification of crude Neem oil
NaOH to oil ratio Percent FAME Methanol to oil ratio Percent FAME
%w/w yield %w/w yield
0.5 0 10 0
1 0 20 0
1.5 12 30 12
2 12 40 35
2.5 30 50 40
3 67 60 61
3.5 65 70 45
4 18 80 32

FFA in crude Jatropha curcas oil


Usually crude Jatropha curcas oil will be stored for long time after extraction,
the free fatty acid percentage is bound to increase. Without proper handling and
storage, the process w o u l d cause various chemical reactions such as hydrolysis,
polymerization, and oxidation. Therefore, the physical and chemical properties of the

122
crude Jatropha curcas oil w i l l change during handling and storage. free fatty acid
percentage of crude Jatropha curcas oil stored sample was found to be 6.1%.

Earlier researches have s h o w n s i m i l a r c h a n g e s i n physical and chemical


properties of other vegetable oils (Monyem, and Van Gerpen, 2001; Leung, et. al., 2006;
Oversen, et. al., 1998). Percentage of free fatty acid has been found to increase due to
the hydrolysis of triglycerides in the presence of moisture and oxidations. Degradation
of the crude Jatropha curcas oil and crude Neem oil results higher concentration of FFA
as compared to crude Pongamia pinnata oil. Comparing to crude Pongamia pinnata oil a n d
c r u d e J a t r o p h a c u r c a s o i l , crude Neem oil contains higher concentration of
unsaturated fatty acids, which are mainly linoleic acid (18:2) and oleic acid (18:1). The
oxidation of the unsaturated fatty acids component leads to degradation of the oil
(Canakci, 2007). The reason for auto oxidation is due to the presence of double bounds in
the chains of unsaturated fatty acids compounds.

The free fatty acids in the extracts with methanol from the oils were analyzed by GC.
Oleic acid and linoleic acid were detected in the FFA extract from crude Jatropha curcas
oil, and their concentrations were much higher than FFA in the methanol extracts from
crude Pongamia pinnata oil.

Table 6.15: Free fatty acid % of crude Jatropha curcas, Pongamia pinnata and neem oil
Species Free fatty acid percentage

Crude Jatropha curcas oil 6 to 12

2.1 to 6
Crude Pongamia pinnata oil

Crude neem oil 14.6 to 21.3

Since Jatropha oil contained high free fatty acid percentage, they would react with
catalyst sodium hydroxide to form sodium soap according to the following reaction:

123
Free fatty acid+ sodium hydroxide = sodium soap+ water
In order to avoid the emulsion by saponification reaction, after the
transesterification reactions, 0.1% aqueous citric acid was used as a washing solution to
remove catalyst for the biodiesel purification. In the following biodiesel production, sulfuric
acid was only used as catalyst to avoid saponification reaction by sodium hydroxide.

Table 6.16: Properties of Jatropha oil and biodiesel produced


Properties Jatropha curcas oil Biodiesel Biodiesel standard ASTM
Specific gravity (g/mL) 0.912 (15 C) 0.882 (15 C) 0.870–0.900
Viscosity (mm2/s) 8.72 (40°C) 3.96 (40 °C) 1.90–6.00
Acid value (mg KOH/g) 10.47 0.32 0.8 max
Cetane number 57 57 47 min
Pour point (C) -2 6 10
Water content (wt.%) 0.135 0.047 0.05 max
Flash point (C) 125 133 100 min

Transesterification using alkali a s a c a t a l y s t

Sodium hydroxide as catalyst


When sodium hydroxide was used, at the optimized condition as described above,
biodiesel yield was only 47.2%. After the reaction, biodiesel was refined for acid value
examination and biodiesel analysis.

Transesterification using alkali a s a c a t a l y s t o f crude Pongamia pinnata oil,


crude neem oil, and crude Jatropha curcas oil by changing catalyst NaOH to oil ratios
(% w/w) and catalyst to oil ratios (% w/w) is shown in enclosed table. The
transesterification reactions with sodium hydroxide catalyst alone were performed at
the optimized condition (Encinar, et. al., 1999), i.e., methanol/oil ratio of24 vol.%
(molar ratio 6:1), catalyst concentration of 1 wt.% (sodium hydroxide solid to Jatropha oil)
and reaction time w a s 1 h. The mixture used was 24 mL methanol, 1-wt.% sodium
hydroxide catalyst and 100 ml Jatropha oil.

The high conversion was obtained with 1.5 % of catalyst NaOH to oil ratio
and 40 % of methanol to oil ratio for Pongamia pinnata oil. The FAME yield was 68%
and 85 % for crude Pongamia pinnata oil . Pongamia pinnata oil had only 2.1. free fatty acid

124
percentage. On the contrary 2.5 % of catalyst NaOH to oil ratio and 50 % of methanol
to oil ratio for crude Jatropha curcas oil. The FAME yield was 65% and 78 % for crude
Jatropha curcas oil . Jatropha curcas oil had 6.1 free fatty acid percentage. The results for
crude Jatropha curcas oil indicate that more catalyst and methanol were required for alkali
base catalyzed transesterification. Optimum NaOH to oil ratio and the optimum
methanol to oil ratio were 2.5% w/w and 50 to 60% w/w. The reason of higher
consumption of catalyst NaOH and methanol in the transesterification process for
crude Jatropha curcas oil could be higher free fatty acid percentage. T h e a b o v e f a c t
was further supported from the observations as shown in Table
s h o w i n g r e q u i r e m e n t o f consumption of catalyst NaOH and methanol in the
transesterification process for crude neem oil. More than 3 % of catalyst NaOH to oil
ratio and 60% of methanol to oil ratio for transesterification of Neem oil. The FAME yield
was 67 % and 71%, respectively.

Acid number for neem oil is 52, generally free fatty acid value is half of the acid
value so percentage of free fatty acid (FFA) in neem oil is 26%. It is very high so cannot be
used directly f o r alkaline esterification u n l e s s acid pretreatment i s d o n e .

The crude neem oil, however, was found to have much higher values of fuel
properties especially viscosity, way above any of these standard limits –thus
restricting its direct use as a fuel for diesel engines.

Flash points of crude neem oil and neem biodiesel were determined to be 214 and
120 0C, respectively and were quite high compared to 55 0C for the diesel. Thus overall
flammability hazard of both crude neem oil and neem biodiesel is much less than that of
conventional diesel.

The densities of crude neem oil and neem biodiesel were observed to be about 12%
and 5.5% higher than that of diesel. Thus, density of crude neem oil was reduced about
6.5% on its conversion to biodiesel. The higher densities of crude neem oil and neem
biodiesel as compared to diesel may be attributed to the higher molecular weights of
triglyceride molecules present in them.

125
The kinematic viscosity of crude neem oil was found to be near 44, which is
twenty times more than that of diesel and it reduced to 4.38 after
transesterification. The kinematic viscosity values are decreased with increase in
temperature.

The gross calorific values of crude neem oil and neem biodiesel were found to be 34.1
and 35.2 MJ/kg, respectively, which are 19% and 16.5% lower than 42.2 MJ/kg for diesel. This
could be due to the difference in their chemical composition from that of diesel or the
difference in the percentage of carbon and hydrogen content, or the presence of oxygen
molecule in the molecular structure of neem oil and biodiesel.

A high FFA in the oil deactivates the catalyst NaOH, and the addition of excess
amount of NaOH as compensation gave rise to the formation of emulsion, which increases
viscosity and results in formation of gels. This causes problem with glycerin separation and
loss in ester yield. Although it is shown that the maximum FAME yields for crude Pongamia
pinnata oil was 85 % while that for Jatropha curcas oil was 78%. The least FAME output was
obtained with crude neem oil. This was proportional to free fatty acid percentage in the three
crude oils as shown in enclosed table.

Transesterification process was a sequence of three equivalents, consecutive and


reversible reactions, in which diglyceride and monoglyceride were formed as intermediates
that were converted further to final products methyl esters as biodiesel and glycerol
(Tapanes, et. al., 2008). After reaction, glycerol was separated spontaneously by gravity.
Biodiesel was purified by distilling the unreacted methanol at 80 C and by absorbing water
with anhydrous Na2S04.

Preparation of biodiesel using concentrated sulfuric acid as catalyst


Similar to the above experiments, 4-mL high-concentrated sulfuric acid was added
to the flask with a reflux condenser filled with 100 mL Jatropha oil and 40 mL anhydrous
methanol. In this case, free fatty acid percentage in the Jatropha oil formed methyl ester as
biodiesel rather than s o a p .

126
A two-step acid–base transesterification of crude Jatropha curcas oil
In order to resolve the above problems of saponification phenomenon, slow
reaction rate and biodiesel stability, a two-step process, acid-esterification
pretreatment and followed by base­ transesterification process, were selected for
Jatropha oil transesterification.

Since crude Jatropha curcas oil had high free fatty acid percentage; and it resulted in
higher consumption of catalyst and methanol (using an alkaline catalyst) due to formation of
fatty acid salts (soap). The soap prevented separation of the methyl ester layer from the
glycerin fraction. An alternative method is to use acid catalysis, which can esterify free fatty
acids.

Canakci and Van Greppen (1999) reported that the esterification reaction stops in
many cases due to the effect of the water produced when the FFA react with methanol to form
esters. Therefore, the two steps process, acid-catalyzed esterification process followed
by base-catalyzed transesterification process is recommended.

First, only sulfuric acid was used as catalyst for acid-esterification pretreatment to
remove the free fatty acid content. In this step, different catalyst concentrations (1%, 2%, 3%, 4%,
5% and 6%; sulfuric acid/ ]atropha-oil, vol.%) and methanol to oil ratios (10%, 20%, 30%,40%,
50%; vol.%) were used to study their influence on reduction of the acid value of Jatropha oil.
The flask with a water-cooled condenser was filled with 200 mL Jatropha oil and
appropriate volume of anhydrous methanol and sulfuric acid. The mixture was vigorously
stirred and refluxed for the required reaction times.

After the reaction, the mixture was filtered and the unreacted methanol was
separated from the liquid phase via rotary evaporation. After pretreatment, the acid
value of Jatropha oil could be reduced to as low as 1.2 mg KOH per g.

At the second step, sodium hydroxide was used as catalyst because it had a higher
catalytic activity for the transesterification reactions. In this process, different catalyst
concentrations of sodium hydroxide solid to pre­treated-oil and methanol to pretreated-oil
ratios were used to investigate their influence on biodiesel yield.

127
Concentrated sulfuric acid as catalyst
The objective of acid pretreatment was to reduce the acid value or free fatty acid
percentage of crude Jatropha curcas oil. Important variable affecting the acid value in the
esterification process was the methanol to oil ratio, the acid to oil ratio, reaction
temperature, and reaction time.

Ghadge, and Raheman, ( 2005); Veljkovic, et. al., ( 2006) reported reaction
temperature 50 °C, the reaction time one hour, and the acid H2SO4 to oil ratio 1% w/w
as suitable conditions for complete esterification o f f r e e f a t t y a c i d in some vegetable
oils.

In order to avoid saponification phenomenon, high-concentrated sulfuric acid


as catalyst was used for the biodiesel production at the optimized condition (Goff, et. al.,
2004), i.e., methanol/Jatropha oil volume ratio 0.4:1(V/V), sulfuric acid catalyst amount 1.0
wt. % (sulfuric acid/Jatropha oil) and reaction time of 4 h. Biodiesel yield was up to 92.8%
and its acid value was reduced to 1 mg KOH/g. However 4 h reaction time was needed as
compared to only 1 h for the above experiment with sodium hydroxide. Goff, et. al.
(2004) reported that it will be much shorter when ultrasonic approach was used.

The biodiesel was not stable, when stored for longer than 30 days. The newly
produced biodiesel stored for 2 year had increased acid value, also ester composition of
linoleic acid (C18:2) decreased while that of oleic acid (C18:1) increased. The probable
reason was that one of unsaturated carbon in linoleic acid ester was hydrolyzed or oxidized
to form oleic acid ester r e s u l t i n g i n precipitation.

Similar conditions were adopted in the present study to find out methanol to oil
ratios for crude Jatropha curcas oil. The results indicated that the acid value or FFA
concentration was influenced by the quantity of methanol. The free fatty acid concentration
reduced sharply to 3% at 10% w/w of methanol to oil ratio and then decreased gradually to 1%
at 70% w/w of methanol to oil ratio. Increasing the methanol amount had no significant effect
on acid value or free fatty acid concentration. Water is a limiting factor, water is produced

128
during the esterification of free fatty acid concentration. If water is completely removed, the
esterification process might be improved.

It is crucial to optimize quantity of methanol required to complete the esterification


process of all free fatty acid percentage in crude Jatropha curcas oil. The optimum
methanol to oil ratio was selected 60% w/w at the free fatty acid concentration less than 1%,
acid value of 2 mg KOH/g-oil.

Two major factors affecting the rate and conversion efficiency of acid esterification
were catalyst concentrations and methanol/ Jatropha-oil volume ratios as reported by
Ramadhas, et. al. (2005); Veljkovic, et. al. (2006); Ghadge, and Raheman, (2005); &
Canakci & Gerpenj, (2001). So, in the acid pretreatment process, different catalyst
concentrations (1-6 vol.%, sulfuric acid/Jatropha-oil) and methanol/}atropha-oil ratios (16-
48 vol.%) were used to investigate their influence on reduction acid value for 1 h reaction
time.

At a fixed methanol/Jatropha oil ratio of 40 vol.%, when sulfuric acid concentration


increased from 1 to 4 vol.%, the acid value was found to decrease from more than 10
to a minimum value of nearly 1 mg KOH/g . However, as the sulfuric acid concentration
increased further, the acid value rose again. The probable reason was that at high acid
concentrations, triglyceride was hydrolyzed to form FFAs and low-molecular weight
alcohols.

Base catalyzed transesterification


The pretreated oil produced by the above first step was further transesterified to
biodiesel catalyzed by sodium hydroxide. When sodium hydroxide concentration
increased from 0.8 to 1.4 wt.%, biodiesel yield rose from 63 % to the maximum value of 96 %
and its corresponding acid value decreased from 1 to the minimum value of 0.32 mg
KOH/g. With further increase of sodium hydroxide concentration, biodiesel yield decreased
to 87.4% and its acid value increased to 0.44 mg KOH/g, which was possibly due to the
hydrolysis reaction that inhibited biodiesel formation. So, 1.4 wt. % sodium hydroxide
concentration was selected for biodiesel production.

129
Table 6.17: Transesterification in Jatropha curcas

Fatty acids Crude Jatropha oil (%) Biodiesel (esters, %) Biodiesel (esters, %)
( by two-step) (with acid catalysis)

Fresh oil 2-yr-old Fresh oil 2-yr-old Fresh oil 2-yr-old


Palmitic acid (C16:0) 16.09 15.18 13.79 13.72 14.43 14.32
Palmitoleic acid1.09 0.99 0.95 1.09 0.93 0.96
(C16:1)
Stearic acid ( C18:0) 6.44 6.25 6.33 6.44 6.2 6.65

Oleic acid ( C18:1) 42.10 41.17 42.61 42.81 40.4 43.90


Linoleic acid (C18:2) 28.72 31.25 26.34 26.40 31.1 28.10

Linolenic acid (C18:3) 0.81 0.08 0.06 0.20 0.34 0.64

Others 4.75 5.08 9.92 9.34 6.52 6.92


Acid value (mg KOH/g) 4.60 10.45 0.32 0.36 1.21 1.84

Crude Jatropha oil and biodiesel


• Biodiesel was produced from stored Jatropha oil.

The volume ratio of methanol/ pretreated-oil was one of the important factors
that could affect biodiesel yield. Stoichiometrically, three moles of methanol were
required for each mole of Jatropha oil. However, in practice, the methanol/oil volume
ratio should be higher than that of stoichiometry in order to drive the reaction towards
completion and produce more methyl esters.

Successful alkali base catalyzed transesterification process requires lower free fatty
acid content in crude Jatropha curcas oil, preferably less than 2%. However, free fatty
acid content in crude Jatropha curcas oil could be 4 to 15%. Result of direct
transesterification of crude Jatropha curcas oil using basic catalyst, by direct
transesterification, methyl ester yield of crude Jatropha curcas oil was only 55%, which
was very low yield comparing to methyl ester yield of other vegetable oils as reported in
literature. Thus, it is clear that high free fatty acid content in the oil affected the methyl
ester yield.

130
The two-stage transesterification process of crude Jatropha curcas oil showed
higher methyl ester yield than single stage or direct transesterification process. After acid
pretreatment/acid catalyst esterification, free fatty acid percentage in the oil moved into
methanol phase. The cleaned oil was trans-esterified with only NaOH to oil ratio at 1%
w/w. Optimum catalyst to oil ratio was f o u n d t o b e 1.4% w/w and optimum methanol to
oil ratio was 24% w/w. At this optimum condition, the methyl ester yield was 90%, which was
higher than the methyl ester yield of direct transesterification.

Transesterification of high free fatty acid content Jatropha oil with methanol to
biodiesel catalyzed directly by NaOH and high-concentrated H 2S04 or by two-step
process were studied by X in a t a l. , 2010) in an ultrasonic reactor at 60 °C. They
reported that when NaOH was used as catalyst, biodiesel yield was only 47.2% with
saponification problem, however, with H2S04 as catalyst, biodiesel yield was increased to
92.8%.However, longer reaction time (4 h) was needed and the biodiesel was not stable.
They proposed a two-step, acid-esterification and base-transesterification pro­cess with
first-step pretreatment with H 2S04 for 1 h, where the acid value of Jatropha oil was
reduced from 10.45 to 1.2 mg KOH /g, and subsequently, NaOH was used for the second-
step transesterification. 96.4% yield after reaction for 0.5 h was reported in a total
production time of only 1.5 h that was just half of normal. The two-step process with
ultrasonic radiation was reported to be effective and time-saving for biodiesel production
from Jatropha oil.

Shekhar, et. al. (2009) reported manufacturing process of biodiesel from neem oil by
esterification by analyzing reaction temperature, reaction rate & catalyst. They could bring
the viscosity of biodiesel oil nearer to that of diesel but the calorific value was about 16%
less than that of diesel. Their process could overcome free fatty acid percentage issue of
neem oil.

131
Conclusions
 Either NaOH or H 2S04 alone was not suitable as catalyst for biodiesel production.

 Saponification phenomenon occurred when o n l y NaOH catalyst was used


and resulted in low biodiesel yield.

 H 2S04 catalyst needed long reaction time, more than4 h and produced unstable
biodiesel.

 Both H 2S04 and NaOH used as catalysts in a two-step (acid esterification and
base-transesterification) process produced economical and stable biodiesel.

 It was found that the acid value of Jatropha oil was reduced to ne a rl y 1
from m ore than 10 mg KOH/g in 1 h.

 At the subsequent second step, 96 % biodiesel yield was obtained with 0.32
mg KOH/g acid value in half an hr. transesterification reaction. The two-step
process only needed n e a r l y 2 h o u r s to achieve 96 % diesel yield

 A two-step transesterification process is developed to convert the high FFA non


edible oils to its esters. The first step (acid catalyzed transesterification) reduces
the FFA content of the oil to less than 2%. The alkaline catalyst
transesterification process converts the products of the first step to its mono-
esters and glycerol.

 Factors effecting the biodiesel production (reaction temperature, reaction rate &
catalyst) were standardized.

 The fuel properties of Jatropha curcas biodiesel were within the limits and
comparable with the conventional diesel. Except calorific value, all other fuel
properties of biodiesel were found to be higher as compared to diesel.

132
Chapter 7

Mechanical Properties of Neem, Pongamia and Jatropha as a biodiesel and


comparison with Commercial Diesel
Mechanical testing has been done with the engine specification as given below:-
 Engine Specifications
 Test engine specifications
 Items Specification
 Model S 195
 Type Single cylinder
 Bore · stroke 95 · 115 mm
 Rated output 9.8 kW/2000 rpm
 Compression ratio 20
 Type of cooling Water evaporative
 Injection pressure 13.5 MPa

Blends of biodiesel vs. Kinematic viscosity


Investigations were carried out to evaluate different blends of Neem, Jatropha and
Pongamia methyl esters (biodiesel). Performance was evaluated and compared with diesel in
diesel engine specification. Kinematic Viscosity (at room temperature of 40°C) of different
blends of methyl esters B40 and B100 were higher than the viscosity of commercial diesel.
Viscosity of B20 biodiesel was very close to the viscosity of commercial diesel. B5, B10, B15
and B20 biodiesel blends can be used without any heating arrangement and engine
modification. There were minor differences between kinematic viscosity of Neem, Jatropha
curcas and Pongamia pinnata at various blends with commercial diesel.

133
Figure 7.1: Effect of blends with diesel in engine kinematic viscosity

Neem
Jatropha
25 Pongamia
Diesel

KV10-6 (sqm/s)
20

15

10

Blend of biodiesel

Brake power vs thermal efficiency


There was a slight drop in efficiency with methyl esters (biodiesel) when compared
with commercial diesel. This drop in thermal efficiency can be attributed to poor combustion
characteristics of methyl esters due to high viscosity. It was observed that the brake thermal
efficiency of B10 and B20 are very close to brake thermal efficiency of commercial diesel. B20
methyl ester had similar efficiency with commercial diesel. Pongamia methyl ester (PME) had
better brake thermal efficiency than that of methyl esters of Jatropha curcas and Neem. So
B20 can easily be suggested as best blend for biodiesel preparation with Pongamia pinnata
and Jatropha curcas oil.

134
Figure 7.2: Brake power vs. thermal efficiency for 10 % blends with diesel
10% Brake power vs Thermal eff.
20
Neem
Pongamia
Jatropha
Diesel

Brake Thermal Efficiency


15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Brake Power(KW)

Figure 7.3: Brake power vs. thermal efficiency for 20 % blends with diesel

20% Brake power vs Thermal effi.


20

Neem
Pongamia
Jatropha
Diesel
Brake Thermal Efficiency

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Brake Power (KW)

Figure 7.4: Brake power vs. thermal efficiency for 40 % blends with diesel
100% Brake power vs thermal effi.
20

Neem
Pongamia
Jatropha
Diesel
Brake Thermal Efficiency

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Brake power (KW)

135
Figure 7.5: Brake power vs. thermal efficiency for 100 % blends with diesel

100% Brake power vs thermal effi.


20

Neem
Pongamia
Jatropha
Diesel

Brake Thermal Efficiency


15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Brake power (KW)

Blends of biodiesel vs density


Density of different blends of methyl esters increased with increase in blend
percentage as shown in enclosed figure. B10, B20, B40 blends and B100 of Neem, Pongamia
pinnata and Jatropha curcas methyl esters were found to be closer to the viscosity of
commercial diesel. The same can be reduced by heating of fuel. It suggests that Neem,
Pongamia pinnata and Jatropha curcas and methyl esters can be a good alternative fuel for
diesel. The highest density of methyl esters was observed in B100, which was very high and
not suitable for engine with 100% consumption in engine as a substitute of diesel.

It was observed that Jatropha curcas methyl esters have about the same density as
commercial diesel up to B40 blending and suddenly density increased in the case of B100
blending. The latter was true for both Pongamia pinnata and Neem methyl esters as well.

136
Figure 7.6: Different blends vs. density of blended biodiesel of Neem, Pongamia and
Jatropha

Different blends vs density

Neem
900
Pongamia
Jatropha
Diesel

880

Density kg/m3
860

840

820
B10 B20 B40 B100 D

Brake power vs smoke density


Smoke density was calculated by opacity test for various blends of biodiesel and
commercial diesel. Biodiesel gives less smoke density compared to petroleum diesel. When
percentage of blend of biodiesel increased, smoke density decreased as shown in enclosed
figure, but smoke density increased for B80 and B100 due to insufficient combustion. It
required changes in injection pressure and combustion chamber design. Smoke density also
decreased when load increased.

Figure 7.7: Smoke density in different Brake power with 10% blends of Neem, Pongamia
and Jatropha

Brake power vs smoke density 10% blends


100
90
80
70
60 Neem
50 Pongamia
40
Jatropha
30
20 Diesel
10
0
0 0.588 1.177 2.345

137
Figure 7.8: Smoke density in different Brake power with 20% blends of Neem, Pongamia
and Jatropha

Brake power vs smoke density 20% blends


100
90
80
70
60 Neem
50 Pongamia
40
Jatropha
30
20 Diesel
10
0
0 0.588 1.177 2.345

Smoke density

Figure 7.9: Smoke density in different Brake power with 40% blends of Neem, Pongamia
and Jatropha

Brake power vs Smoke Density blends 40%


100
90
80
70
60 Neem
50 Pongamia
40
Jatropha
30
20 Diesel
10
0
0 0.588 1.177 2.345

Smoke density

Brake power vs carbon monoxide


Carbon monoxide was calculated by emission test for various blends of biodiesel and
commercial petroleum diesel. Biodiesel gave less carbon monoxide as compared to petroleum
diesel. When percentage of blend of biodiesel increased, carbon monoxide decreased. But

138
carbon monoxide increased for B60, B80 and B100 due to insufficient combustion. It required
changes in injection pressure and combustion chamber design.

Figure 7.10: Brake power vs. emission of CO for B10 blends of Neem, Pongamia and
Jatropha

Figure 7.11: Brake power vs. emission of Carbon monoxide for B20 blends of Neem
Pongamia and Jatropha

Figure 7.12: Brake power vs. emission of Carbon monoxide for B40 blends of Neem,
Pongamia and Jatropha

139
Brake power vs hydrocarbon
Hydrocarbons were calculated by emission test for various blends of biodiesel and
commercial diesel. Biodiesel left fewer hydrocarbons as compared to petroleum diesel. When
percentage of blend of biodiesel increased, hydrocarbons parentage decreased. But
hydrocarbons increased for B60, B80 and B100 due to insufficient combustion. It required
changes in injection pressure and combustion chamber design. Hydrocarbons also increased
when load increased.

Figure 7.13: Brake power vs. Hydrocarbons for B10 blends of Neem (N), Pongamia(P) and
Jatropha (J)

140
Figure 7.14: Brake power vs. Hydrocarbons for B20 blends of Neem (N), Pongamia(P) and
Jatropha (J)

Figure 7.15: Brake power vs. Hydrocarbons for B40 blends of Neem (N), Pongamia(P) and
Jatropha (J)

Biodiesel blends vs flash point


The flash points of different blends of methyl esters increased with increase in methyl
ester percentage as shown in enclosed figure. It was also observed that the flash points of raw
and esterified oils were higher as compared to diesel. Thus, the biodiesel blends can be used
as a fuel without any fire accidents.

141
Figure 7.16: Biodiesel blends vs. flash point temperature for Jatropha (J), Neem (N) and
Pongamia (P)

Table 7.1: Various biodiesel blends with mixed feedstock

CODE MIXING OF OIL BIODIESEL


J P N (RESULT)
(ml) (ml) (ml)
DIESEL - - - 100 %
A) 80 10 10 68-70 %
B) 10 80 10 60-64 %
C) 60 20 20 62-65 %
D) 20 60 20 57-60 %
E) 40 40 20 62-65 %
F) 40 20 40 42-45 %
G) 20 40 40 38-40 %
H) 20 20 60 32-35 %
I) 10 10 80 27-30 %
J) 80 20 0 85-88 %
K) 60 40 0 80-82 %
L) 40 60 0 75-78 %
M) 20 80 0 72-75 %
N) 80 0 20 62-65 %
O) 0 80 20 55-60 %
P) 60 0 40 67-70 %
Q) 0 60 40 62-64 %
R) 40 0 60 69-72 %
S) 0 40 60 63- 66 %
T) 20 0 80 70-73 %
U) 0 20 80 68-70 %

142
V) 0 0 100 74-76 %
w) 0 100 0 88-90 %
x) 100 0 0 96-99 %

Co
mparison of mixed feedstock combinations
Physico-chemical properties of biodiesel produced by mixed feedstock of non-edible
oils is shown in enclosed tables. Reactions were carried out by the standardized procedure of
transesterification with standardized amount of catalyst, temperature, alcohol etc. Out of
many combinations tried, A, J, K and L combinations had best biodiesel production yield and
physico-chemical properties then the other feedstock combinations and also as compared to
pure Neem, Pongamia and Jatropha biodiesel. Therefore, these four selective combinations
were applied in the engine with the diesel in different blending percentages for mechanical
testing. Testing was done with respect to kinematic viscosity, density and flash point for
various biodiesel blends with the feedstock as shown in enclosed figure.

Mixed feedstock biodiesel vs kinematic viscosity


Table 7.2: Physico-chemical and mechanical properties of mixed feedstock produced from
the blending of Jatropha, Neem and Pongamia oil of combination

Code Sp. Densit R.I FFA Acid Sap. Color Flash Pour Clou Kinemati Carbo Calorifi Viscosi Sulph
Gravit y (nD) (%) Value Value value point point d c n c value ty ated
y (g/ml) (KOH (mg (1”ce (°C) (°C) point Viscosity residu (MJ/Kg (25°C) Ash
2
) KOH/g ll) (°C) (mm /s) e ) Cst (%wt
) (wt.%) .in g)
Diesel 0.83 0.827 1.47 - - - - 76 -16 -10 3.06 86.8 44.46 6.8 0.042
59
A) 0.87 0.86 1.46 0.06 0.10 150.32 12.4 182 -9 +18 4.9 75.1 38.9 8.5 0.009
B) 0.91 0.91 1.47 0.08 0.15 139.50 33.9 220 +6 + 11 5.3 76.2 40.9 8.9 0.004
C) 0.92 0.91 1.48 0.06 0.06 141.72 34.1 222 +6 +9 5.4 76.1 40.2 9.2 0.006
D) 0.93 0.92 1.49 0.11 0.23 139.51 35.3 227 +8 + 16 5.6 77.6 42.1 9.2 0.009
E) 0.93 0.93 1.45 0.12 0.24 141.52 35.7 228 +6 + 13 5.3 76.2 45.7 9.1 0.012
F) 0.93 0.93 1.48 0.14 0.27 130.52 38.7 235 +3 +9 5.7 77.7 46.4 9.2 0.016
G) 0.93 0.94 1.48 0.17 0.32 137.12 39.2 237 +6 + 12 6.1 77.5 47.1 9.3 0.017
H) 0.89 0.89 1.48 0.17 0.35 150.12 13.2 203 +7 + 13 5.1 75.1 41.1 9.4 0.010

I) 0.90 0.88 1.51 0.13 0.35 127.52 20.1 225 +8 +2 6.1 77.1 39.3 8.2 0.003
J) 0.86 0.85 1.46 0.02 0.20 149.91 13.1 172 -5 +1 4.1 74.3 33.5 8.3 0.010
K) 0.87 0.86 1.44 0.04 0.29 146.12 16.3 125 +2 -4 4.4 72.3 38.1 8.0 0.005
L) 0.85 0.84 1.45 0.05 0.16 138.20 31.0 110 -18 -12 5.0 73.3 35.7 9.3 0.005
M) 0.88 0.88 1.44 0.07 0.15 135.10 32.5 215 +3 +10 5.2 72.8 37.5 9.5 0.005
N) 0.88 0.88 1.46 0.15 0.31 148.32 12.2 199 + 14 +5 4.7 75.3 39.7 9.3 0.006
O) 0.90 0.89 1.45 0.19 0.39 139.51 34.8 219 +4 + 10 6.3 81.1 39.4 9.5 0.005
P) 0.89 0.88 1.43 0.16 0.31 146.32 12.3 190 +9 + 16 5.7 77.5 39.7 9.4 0.007
Q) 0.87 0.87 1.44 0.20 0.41 140.19 12.5 239 +3 +7 5.5 73.5 39.6 9.7 0.008
R) 0.89 0.89 1.43 0.22 0.45 125.13 18.9 199 +5 + 11 5.1 77.4 35.1 9.5 0.009
S) 0.88 0.87 1.42 0.25 0.51 119.71 19.5 222 +4 + 11 5.9 73.2 34.1 9.4 0.003

143
T) 0.88 0.86 1.45 0.36 0.75 127.57 14.9 200 +9 +16 5.2 76.9 35.9 9.3 0.005
U) 0.87 0.86 1.43 0.15 0.30 132.78 15.7 201 +10 +8 5.2 75.2 36.5 9.6 0.005
V) 0.87 0.89 1.44 0.51 0.26 123.11 19.9 222 +10 +7 5.8 76.7 41.2 8.9 0.009
w) 0.89 0.89 1.45 0.16 0.31 139.21 32.8 216 +5 +10 5.2 75.2 40.2 8.8 0.005
x) 0.89 0.88 1.44 0.09 0.16 148.28 12.3 199 +4 +8 4.9 75.1 44.3 8.6 0.005

An experimental investigation was carried out for different blends made with
combination of feedstock from Neem, Pongamia pinnata and Jatropha curcas methyl esters.
Highly suitable blends with promising properties were evaluated in engines. Performance was
evaluated and compared with commercial diesel in diesel engine specification as given above.
Kinematic Viscosity (at room temperature of 40°C) of different blends of methyl esters of B10,
B20, B40 and B100 were higher than the viscosity of commercial diesel. But up to B20 the
viscosity of biodiesel was very close to the viscosity of commercial diesel. So, B5, B10, B15
and B20 biodiesel blends can be used without any heating arrangement and engine
modification.

It was observed that A as a biodiesel feedstock had highest kinematic viscosity.

Brake power vs thermal efficiency with mixed feedstock biodiesel


A slight drop in efficiency was found with mixed feedstock biodiesel methyl esters
when compared with commercial diesel. This drop in thermal efficiency can be attributed to
the poor combustion characteristics of methyl esters due to high viscosity. It was observed
that the brake thermal efficiency of B10 and B20 were very close to brake thermal efficiency
of commercial diesel. B20 methyl ester had similar efficiency to that of commercial diesel. A, J
and K combination of mixed feedstock methyl esters had better brake thermal efficiency as
compared to the methyl esters of L combination. So B10, B20 can be suggested as best blend
for biodiesel preparation with A,J and K combination of biodiesel.

144
Figure 7.17: Effect of biodiesel blends on engine vs. kinematic viscosity

Figure 7.18: Brake power vs brake thermal efficiency for B20 of mixed feedstock biodiesel
Brake power Vs Brake Thermal Efficiency for 20 % blend of mixed feedstock biodiesel
20

A
J
K
Brake Thermal Efficiency

15 L
DIESEL

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Brake Power (KW)

145
Figure 7.19: Brake power vs. brake thermal efficiency for B40 of mixed feedstock
biodiesel

Brake power Vs Brake Thermal Efficiency for 40%Blends of mixed feedstock biodiesel

20

A
J
K
Brake Thermal Efficiency 15 L
DIESEL

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Brake Power (KW)

Mixed feedstock biodiesel vs density


Comparative study of density of mixed feedstock blends with diesel was investigated.
The density of different blends of methyl esters increased with increase in blend percentage
as shown in enclosed figure. B10, B20, B40 and B100 blends of Pongamia pinnata, Jatropha
curcas and Neem methyl esters were found closer to the viscosity of diesel. These can be
reduced by heating of fuel. Accordingly, mixed feedstock of Pongamia, Jatropha and Neem
methyl esters can be an alternative fuel for diesel. The highest density of methyl esters was
observed in B100 which was very high and not suitable for engine which 100% consumption in
engine as a fuel. it was observed that K and L feedstock had about the same density as
commercial diesel up to B20. However, density suddenly increased in case of B40 and B100
blends of A, J, K and L biodiesel combinations of feedstock of Pongamia pinnata. Jatropha
curcas and Neem methyl esters.

146
Figure 7.20: Density of different mixed feedstock blends

Mixed feedstock biodiesel vs flash point


Flash point temperature of mixed feedstock blends versus diesel was observed. The
flash points of methyl esters were found to increase with increase in different combinations
made of mixed feedstock biodiesel (methyl ester) blends. Blends of A, J, K, and L as shown in
enclosed figure. It was observed that the flash points of raw and esterified oils were relatively
high when compared to commercial diesel.

Figure 7.21 Mixed feedstock biodiesel vs. flash point temperature

Ignition temperature of mixed feedstock biodiesel was between 50 to 200 ºC. Highest
ignition temperature was with B100 and lowest was observed for B20. Combination A had
nearly same properties as commercial diesel. The range of flash point for B10 and B20 was
found close to commercial diesel and showed good performance as engine fuel; and is
recommend for the engine. Thus, it can be commercially used as a biofuel.

147
Chapter 8

Storage conditions for Biodiesel:

Fuel stability
A fuel is considered unstable when it undergoes chemical changes that produce
undesirable consequences such as deposits, acidity, or a bad smell. There are three different
types of stability commonly described in the technical literature. Thermal stability addresses
fuel changes that occur due to elevated temperature. These changes may occur at conditions
encountered in modern fuel injection systems as fuel is recirculated through the engine
cylinder head and back to the fuel tank. Oxidative stability refers to the tendency of fuels to
react with oxygen at temperatures near ambient. These reactions are much slower than those
that would occur at combustion temperatures, and they produce varnish deposits and
sediments. Storage stability is also a frequently used term and refers to the stability of the
fuel while it is in long-term storage. These terms are not necessarily exclusive terms. For
example, oxidative attack is probably one of the primary concerns of storage stability but
storage stability might also involve issues of water contamination and microbial growth.

Biodiesel can be stored for long periods of time in closed containers with little head
space. The containers should be protected from weather, direct sunlight and low
temperatures. Avoid long term storage in partially filled containers, particularly in damp
locations like dock boxes. Condensation in the container can contribute to the long term
deterioration of the petroleum diesel or biodiesel. Low temperatures can cause the biodiesel
to gel, but the biodiesel will quickly liquefy again as it warms up. In cold weather (near or
below freezing), additives can be used to prevent gelation.

Petroleum diesel and biodiesel are both susceptible to growing microbes when water
is present in the fuel, but the solvent action of the biodiesel can also cause microbial slime to
detach from the inside of the tank. The accumulation of the newly released slime and
sediment can be very dangerous if it clogs the fuel filters and causes the engine to suddenly

148
stop. It is very important to monitor the filters on a diesel engine that has been switched over
to biodiesel, particularly if the tank is old and has not been cleaned.

Introduction of degradation
Biodiesel produced from vegetable oils and other feedstocks can be more prone to
oxidation than a typical petroleum diesel unless modified or treated with additives. The
purpose of this study was to provide data that define how stable biodiesel or a biodiesel blend
should be to prevent the formation of acids or sediment during transportation, storage, and
use.

There are several key properties that need to be considered in the transportation and
storage of biodiesel. These properties include exposure temperature, oxidative stability, fuel
solvency, and material compatibility. Exposure temperature refers to the temperature at
which the biodiesel is being stored or transported. Oxidative stability concerns the ability of
the biodiesel to be stored for extended periods without degradation. Fuel solvency refers to
the solvent properties of the biodiesel. The interaction of the biodiesel with materials of
construction for storage tanks, seals, and gaskets is the material compatibility. Each of these
properties will be discussed in more detail and within the context of storage.

Accelerated oxidation stability test


The objectives of this work were to identify a practical accelerated oxidative stability
test method and to define a reasonable, data-based, stability minimum requirement.

Storage stability refers to the ability of the fuel to resist chemical changes during long
term storage. These changes usually consist of oxidation due to contact with oxygen from the
air. The fatty acid composition of the biodiesel fuel is an important factor in determining
stability towards air. Generally, the more unsaturated fatty acids (C18:2, linoleic acid; C18:3
linolenic acid) are more susceptible towards oxidation. The changes can be catalyzed by the
presence of certain metals (including those making up the storage container) and light. If
water is present, hydrolysis can also occur. The chemical changes in the fuel associated with
oxidation usually produce hydroperoxides which can, in turn, produce short chain fatty acids
and aldehydes. Under the right conditions, the hydroperoxides can also polymerize.

149
Therefore, oxidation is usually denoted by an increase in the acid value and viscosity of the
fuel. Often these changes are accompanied by a darkening of the biodiesel color from yellow
to brown and the development of a “paint” smell.

When water is present, the esters can hydrolyze to long chain free fatty acids which
also cause the acid value to increase. There is currently no generally accepted method for
measuring the stability of biodiesel. The techniques generally used for petroleum-based fuels,
such as ASTM D 2274, have been shown to be incompatible with biodiesel. Other procedures,
such as the Oil Stability Index or the Rancimat apparatus, which are widely used in the fats
and oils industry, seem to be more appropriate for use with biodiesel. However, the engine
industry has no experience with these tests and acceptable values are not known. Also, the
validity of accelerated testing methods has not been established or correlated to actual
engine problems. If biodiesel’s acid number, viscosity, or sediment content increase to the
point where they exceed biodiesel’s ASTM limits, the fuel should not be used as a
transportation fuel.

All fuels are subject to degradation over time when they are stored. This degradation
may be due to microbial action, water intrusion, air oxidation, etc.

Materials & Methods


Originally, two B100 samples were obtained from biodiesel industry while other three
were developed in the laboratory. These five samples were selected for more detailed study
(samples I to V). The samples were stored in a dark room at room temperature. All B100
samples were tested initially for total acid number or acid value (ASTM D664) and free and
total glycerin (ASTM D6584). More detailed characterization were also performed on a subset
of samples for flash point (ASTM D93); peroxide value (ASTM D3703); and Karl Fischer
moisture (ASTM D6304). Petroleum diesel was purchased from petrol station for making B5
and B20.

Blends of petroleum diesel and biodiesel were prepared volumetrically. Blends are
labeled with the diesel fuel, followed by the biodiesel and biodiesel concentration.

150
Peroxide was measured as per methods detailed earlier in the thesis. Peroxide
contents were expressed in the range of 1−5 ppm. In this method, the peroxides are reduced
with potassium iodide and an equivalent amount of iodine is formed. This is titrated with
sodium thiosulfate. Results are reported as mg peroxide per kg (ppm).The absolute accuracy
of the test for biodiesel is currently unknown, but we believe it is adequate for showing trends
over time.

B100 and blends were analyzed using two accelerated oxidation stability tests.
Induction time was measured using test method EN14112. In this method, air is surged
through the sample and then through a water trap. Volatile oxidation products (primarily
formaldehyde and short-chain acids) are absorbed by the water, causing an increase in
conductivity. The water conductivity is monitored to determine the onset of oxidation and to
define an induction time.

All samples were tested in duplicate and the average induction period was reported.

Results
Biodiesel and biodiesel blends were studied for degradation over 12 weeks (storage
for 12 weeks at 43 °C).

The study of degradation of biodiesel and biodiesel blends suggests that it degrades
over period. The simulated experiment reveals that biodiesel will degrade over a period of 12
months.

B100 Stability
The B100 samples appear to cover the full range of feedstocks currently used. All
samples met the specification requirements for acid value and for free and total glycerin.

Storage stability was assessed for five B100 samples using ASTM D4625 protocol
(storage for 12 weeks at 43 °C).

All samples showed degradation. The peroxide value over the test indicates that all of
the samples had clearly degraded, even at 4 weeks. Acid value also changed over the test.
Some stable samples showed a small increase in acidity at 4 weeks, but one of the samples

151
had high acid value indicating degradation. Other samples had acid value less than 0.8 which
was in acceptable limits. By 8 weeks (simulating 8 months) acidity had increased to above the
0.80 mg KOH/g limit for all samples. It was above 0.8 initially and showed a large increase in
acidity over the test. The results are consistent with basic mechanistic understanding showing
initial peroxide formation followed by formation of acid and oligomer in parallel.

Storage stability of B 20 samples


Figures 8.1 to 8.3 show induction time results for the B20 samples. There are different
results for various biodiesel blends. Some samples showed significant stability degradation as
compared to others. Results for peroxide value are shown in the enclosed Figure 8.1. All
samples were degraded though to different levels in this test as was indicated from changes in
peroxide value. One of the samples starting to degrade at 6th week. The acid value increase
for B20 blends are also shown in the Figure 8.2. The samples showed a significant increase in
acidity. Another sample showed increase that was significantly higher than the other two.
Changes in total insolubles formation is shown in the Figure 8.3.

Figure 8.1:- Change in peroxide value of 20% biodiesel blend in 12 weeks

152
Figure 8.2:- Change in acid value of 20% biodiesel blend in 12 weeks

Figure 8.3:- Change in insoluble impurities of 20% biodiesel blend in 12 weeks

Storage of B5
Except for one sample very little degradation was observed in B5. At 8 weeks all
samples showed low peroxides, and at 12 weeks except for one sample the peroxide value

153
was within limits (Figure 8.4). Peroxides can form and then decompose to form acids or
oligomers. The acid value followed the pattern of peroxide value (Figure 8.5). Acid values of
0.05 mg KOH/g or lower are below detection limit of this method indicating no degradation.

Total insoluble formation (Figure 8.6) also followed the pattern of peroxide and acid
value. On the basis of the apparent formation and decomposition of peroxide between 4 and
8 weeks, an increase in acid and insolubles might have been expected but was not observed.
Variation, if any, were below experimental limits.

Figure 8.4:- Change in peroxide value of 5% biodiesel blend in 12 weeks

154
Figure 8.5:- Change in acid value of 5% biodiesel blend within 12 weeks

Figure 8.6:- Change in insoluble impurities of 5% biodiesel blend within 12 weeks

Trans-esterified biodiesel (B100) samples show a broad distribution of stability in


accelerated tests, with EN14112 induction time results ranging from less than 1 h to as much
as 12 h and total insolubles ranging from less than 1 mg/100 mL to nearly 18 mg/100 mL. The
long-term storage results for B100 indicate that most biodiesel samples, regardless of initial
induction time, will begin to oxidize immediately during storage. If induction time is near or

155
below the 3 h limit, the B100 will most likely go out of specification for either stability or acid
value within 4 months (4 weeks on the test). Even B100 with induction times longer than 7 h
will be out of specification for oxidation stability at only 4 months, although these samples
may not have shown a significant increase in acidity or in deposit formation. The 3 h B100
induction time limit appears to be adequate to prevent oxidative degradation for B5 blends in
storage for up to 12 months and B20 blends for up to 4 months. The results indicate that B100
stability is the main factor that affects the stability of B5 and B20 blends, independent of
diesel fuel aromatic content, sulfur level, or stability. Synthetic antioxidants had been
reported to be highly effective at preventing acid and insoluble formation during storage.

The general mechanism of fat (or lipid) oxidation is reasonably well understood
(Frankel, 2005; Waynick, 2005). For biodiesel made from common feedstocks such as oils from
soya, rapeseed, and palm as well as lard and tallow the fatty acid chains contain primarily 16
or 18 carbon atoms and from 0 to 3 double bonds denoted as C18:x, where x is the number of
double bonds in the chain. When multiple double bonds are present, they are in allelic
configuration, separated by a single methylene or bis-allylic carbon (Cosgrove et. al., 1987).
They reported that the relative oxidation rates for these C18 esters are C18:3 > C18:2 C18:1
because the di- and tri- unsaturated fatty acids contain the most reactive sites for initiating
the auto-oxidation chain reaction sequence. The oxidation rate correlates with the total
number of bis-allylic sites, not with the total number of double bonds (Knothe, 2002). The
allelic position (a methylene adjacent to a single double bond) is much less reactive, explaining
the much lower oxidation rate of oleic acid ester (C18:1). The bis-allylic sites have high
reactivity for the formation of free radicals that immediately isomerize to form a more stable
conjugated structure and then can react directly with oxygen to form peroxide radicals.
Peroxide can form via this route even at ambient temperature. The peroxide radicals can
cleave to form acids and aldehydes, or they can react with another fatty acid chain to form a
dimer. Thus, peroxides and acids form in a sequential process, while oligomers (and ultimately
insolubles) form in parallel with acid formation via a slower, multistep process.

Fang and McCormick (2006) showed that the oxidation and dimerization of the
peroxide species are not the only mechanisms for molecular weight growth and insoluble

156
formation in biodiesel, and they identified several other mechanisms by which biodiesel can
degrade. In particular, aldehydes formed by peroxide decomposition could also polymerize via
aldol condensation. However, all pathways to deposit formation involved peroxide formation
as the initial step, highlighting the importance of preventing peroxide formation at the point
of biodiesel manufacture and throughout the biodiesel distribution chain. Fang and
McCormick (2006) also showed an antagonistic or synergistic effect on deposit formation for
biodiesel blends. The formation of deposits under oxidative stressing was higher than
predicted for biodiesel blends based on the weight per cent biodiesel and the deposits formed
from the diesel fuel or biodiesel alone. This effect was most significant for blends in the
20−30% range.

Westbrook (2005) used the ASTM D4625 test, in which the subject fuel or blending
component is held for 12 weeks at 43 °C in a capped but vented glass bottle, and oxidation
effects (the amount of sediment formed, acidity, and change in viscosity) were periodically
measured. A wide variation of insolubles formation, acid number, and viscosity increase were
observed; the least stable samples exhibited unacceptable levels of insolubles and acidity 4−8
weeks into the test. To more fully understand the causes of these differences, McCormick and
co-workers (2007) examined the stability characteristics of commercial biodiesel samples and
showed that the stability range results primarily from differences in fatty acid makeup and
natural antioxidant content. However, samples containing high out-of-specification levels of
glycerides (unconverted or partially converted feedstock) also tend to form higher levels of
deposits in oxidation tests. Thus, for real-world samples containing impurities, the correlation
of oxidation rate or tendency with the number of bis-allylic sites may be skewed or
overshadowed by antioxidant content and the effects of other minor components.

Vegetable oils contain naturally occurring antioxidants that can cause oxidation
stability to vary over a wide range, even for samples with similar fatty acid makeup
(McCormick, 2007). The most common antioxidants are tocopherols. For example, soy oils
contain 500−3000 ppm tocopherols, along with other antioxidants such as sterols and
tocotrienols (Tatandjiiska, 1996) which may not be affected by the ester preparation process
(Van Gerpenet et al., 1997). However, some production processes include steps to purify the

157
methyl esters that can remove antioxidants. Biodiesel produced using distillation to purify the
product, for example, typically contains few or no natural antioxidants, and it is less stable
than biodiesel that does contain antioxidants. The stability of biodiesel prepared by
distillation, as well as biodiesel that contains natural antioxidants, can be improved by adding
synthetic antioxidants. Westbrook (2007) has shown that several antioxidants, including
proprietary antioxidants developed for petroleum fuels, can improve biodiesel oxidation
stability.

Bondioli and co-workers (2003) stored several drum-quantity samples of biodiesel at


ambient conditions for 1 year. One sample was “shaken” once per week to promote intimate
contact with air. Over this period the quiescent samples exhibited little or no change in
properties, including only minor reductions in EN14112 induction time. This contrasts strongly
with the results of studies conducted at higher temperatures (for example, 43 °C, as in ASTM
D4625), which have shown large changes in acid value and other parameters for some
biodiesel samples (Bondioliet. al., 2002). The agitated sample exhibited significant increases in
peroxide and acid values, however, and a large reduction in induction time because of the
increased exposure to dissolved oxygen.

Mittelbach and Gangl also stored biodiesel produced from rapeseed oil and used frying
oil under different conditions for up to 200 days (Mittelbach and Gangl, 2002). Degradation
caused by oxidation began immediately, as shown by the formation of peroxides and
reduction in induction time. However, even at the end of the storage period, limits for
viscosity and acidity were not exceeded.

High levels of oxidation can cause operational problems for engine fuel system
components. Terry and co-workers (2006) showed that, at very high levels of oxidation,
biodiesel blends can separate into two phases to cause fuel pump and injector operational
problems or lacquer deposits on fuel system components. Blassnegger performed 500 h fuel
injector bench tests with rapeseed methyl ester (RME) (B100) samples that had a range of
stability (Blassnegger, 2003). Lower stability B100, with a EN14112 induction time of 1.8−3.5
h, produced injector deposits and reduced the amount of fuel injected per stroke. Tsuchiya et
al. demonstrated that the formation of high levels of acids, which can be generated from
158
unstable biodiesel during extended fuel recirculation, can corrode metal vehicle fuel tanks
(Tsuchiya et. al., 2006). Thus, there are strong arguments for ensuring the stability of biodiesel
and biodiesel blends. Because of the various factors that can affect the stability of biodiesel
and biodiesel blends, the study was conducted with the goals of identifying a practical
accelerated stability test method and of defining a reasonable, data-based, minimum stability
requirement.

Fuel tanks should be kept as filled as possible (regardless of whether they contain
biodiesel), particularly during rainy winter months or periods of inactivity, to minimize the
condensation of moisture. Condensed moisture accumulates as water in the bottom of tank
and can contribute to the corrosion of metal fuel tanks, especially with petroleum diesel that
also contains sulphur. The condensed water in the fuel tank can also support the growth of
bacteria and mold that use the diesel and biodiesel hydrocarbons as a food source. These
hydrocarbon-degrading bacteria and molds will grow as a film or slime in the tank and
accumulate as sediment over long periods of time. These hydrocarbon-degrading microbes
are frequently referred to incorrectly as "algae" in advertisements for fuel treatments,
perhaps because the colonies often have a reddish orange color and tend to form mats.

Biocides can treat diesel fuels suspected of having bio-growth. The biocides are
chemicals that kill bacteria and molds growing in fuel tanks without interfering with the
combustion of the fuel or the operation of the engine. Used in very dilute concentrations, the
biocides can inhibit the growth of microbes over long periods of time. These products are very
toxic and should be used only as directed by the manufacturers. Precautions should be taken
to avoid any contact with the products (wear gloves and eye protection) and to prevent any
spills or drips. It is important to remember that the biocides may kill the microbes, but they do
not remove the accumulated sediment, so expect to replace fuel filters often as the debris is
drawn from the tank. In some cases, it may be necessary to have the fuel filtered.

As with gelling, all diesel fuels have the potential for oxidative stability issues.
Oxidative stability concerns the degradation of diesel fuel due to reaction of the fuel with
oxygen and catalysts. As such, the reactivity of the fuel is important. This reactivity can be
related to the presence of C=C (olefinic) bonds in the fuel with increased content of the C=C
159
bonds correlated to decreased oxidative stability of the diesel fuel. The increase in instability
of a given diesel fuel molecule is generally directly proportional to the increase in the number
of C=C bonds in the molecule (i.e., a molecule containing two C=C bonds has half the stability
of a molecule containing one C=C bond). The oxidative stability of a diesel fuel can be
estimated using the iodine number (ASTM D 1510). The longer term stability of a diesel fuel
can be evaluated using an accelerated stability test (ASTM D 2274). Oxidative degradation of
diesel fuel is manifest by higher acid numbers, increased viscosity, and the formation of gums
and sediments.

While biodiesel is only a mild solvent, it does have higher solvency properties than
diesel fuel. Due to this property, residual sediments in diesel storage tanks or vehicle fuel
tanks can be solvated by biodiesel. The solvent properties of biodiesel are mitigated by the
use of biodiesel in blends with diesel fuel. Typically, 20% or less blends of biodiesel in diesel
will nearly completely dilute the solvency effect. However, this solvency effect should be
considered if pure biodiesel is stored in a tank that was previously used for standard diesel.

Conclusion
All types of diesel fuels can gel at low temperatures. The temperature required to
induce gelling (freezing) in biodiesel is generally higher than that of conventional diesel fuel
and depends on the composition of the biodiesel. In particular, saturated methyl esters will
have higher freezing points than unsaturated methyl esters. The composition of the biodiesel
is, of course, dictated by the oil feedstock used for its production.

Biodiesel produced from vegetable oils and other feedstocks can be more prone to
oxidation than a typical petroleum diesel unless modified or treated with additives. Moisture,
temperature, oxidative stress can lead to deterioration of biodiesel. The purpose of this study
was to provide data that define how stable biodiesel or a biodiesel blend should be to prevent
the formation of acids or sediment during transportation, storage, and use.

160
Chapter 9
Summary
Study entitled “Application of Jatropha curcas oil as a feedstock for biodiesel” was
undertaken to evaluate issues related to application of Jatropha curcas as a feedstock for
biodiesel viz. improvement of Jatropha curcas, oil extraction, transesterification and use in
stationery engines. The objectives of the study were:

 To screen germplasm of J. curcas for oil quantity and quality from different
provenances of India (Rajasthan and Uttaranchal).

 To standardize the protocols to maximize oil yield.

 To assess of physico-chemical properties of J. curcas and other non-edible oils.

 To develop protocols for esterification and purification so that the crude oil can
be improved for use as diesel substitute for stationary motors.

 To evaluate application of biodiesel in stationary engines.

 To evaluate and compare Neem oil, Pongamia oil with that of J. curcas oil for
developing the biodiesel from mixed feedstock.

World oil demand growth in 2011 will be higher than initially expected, the
International Energy Agency (IEA), an adviser to 28 industrialized countries on energy policy. It
raised its 2011 oil demand growth forecast for this year by 80,000 barrels per day (bpd) to
1.41 million bpd on signs of buoyant economic growth, especially in emerging markets.
According to Oil Chief of OPEC, the global oil demand is expected to rise during this year by
1.5 to 1.8 million barrels per day" (mbd), based on the global economic growth, mainly from
three major regions -- Asia, particularly China and India, the Middle East and Latin America.

India is a developing economy and is set to become the world’s fastest growing major
economy with its GDP growth increasing from 5.7 % in 2009 to 8.3 % in 2010 and set to reach
8.6% in 2011. China, on the other hand, had its GDP growth of 8.7 % and 9.5 % in 2009 and
2010, respectively and may drop to 8.5 and 8.2 % in the following years, thus matching that of
India. However, India’s growth engine is a fuel guzzler. Total primary consumption (Btu per

161
dollar) is very high, next only to China; and is three times that of USA (and six times that of
Japan). A large percent of its energy demands are met from crude while the crude production
has been stagnant at 33.69 million tonne (mt) in 2009-10 (34 mt in 2006-7; 34.1 mt in 2007-8
and 33.5 mt in 2008-9). It leaves India heavily reliant on imports. Net crude imports have
increased from 99 mt in 2005-6 to 153.2 mt in 2009-10; and is expected to rise tremendously
in the coming years.

The Government of India has outlined ambitious capacity expansion and investment
plans for the eleventh five year plan period (FY 2007- FY 2012). It has proposed an addition of
15,000 MW of Renewable Energy generation capacities during the period7. The Government
of India has set an indicative target of a minimum 20 per cent ethanol-blended petrol and
diesel across the country by 2017.

India has opted Jatropha curcas as a feedstock since it can grow on degraded and non-
agriculture sites. With Jatropha curcas, issue of food versus fuel is not there. Western
countries are using rape seed, corn and sunflower as feedstock which has a direct conflict with
food. Seeds of Madhuca indica (Mahua) produce oil that can be converted to biodiesel by
transesterification. The cake left after extraction of oil can be used as a fertilizer. However,
Madhuca indica is also edible seed oil and cultivation takes several years before the yield
starts. Similarly, Pongamia pinnata oil has multiple uses. It is a very good source of vegetable
oil that can be converted to biodiesel and meet diesel requirement of the country. Biodiesel
made from Karanj oil has been evaluated for its efficacy by National Botanical Research
Institute and IIP, Dehradun by Behl et al. (2007). Several other options like Salvadora oleoides
may be rich in oil but availability of feedstock is a concern.

Evaluation of germplasm
Seeds of Jatropha curcas were collected from Rajasthan and Uttrakhand. Sections was
made on phenotypic assessment of characters of economic interest i.e., yield potential,
number of fruits per cluster, crown spread, girth, disease resistance etc . All the seeds samples
were deposited at NBPGR to obtain IC #. Seed weight, area, seed and kernel weight and their
ratio, oil content were evaluated for all the accessions.

162
 Seed thickness ranged from 6.4 to 10.1 with an average of 8.6 (standard
deviation of 0.98).

 Seed coat and seed kernel weight is a conservative character and can be a
useful tool. In some accessions, the seed coat was nearly as heavy as kernel
weight and thus can be regarded as a negative trait for selection of germplasm.

 Seed weight had positive correlation with seed length, breadth, thickness and
oil content.

 Pearson Correlation coefficient was calculated for the said factors. Pearson
value was 0.33 for correlation between seed oil and seed width.

 Accessions with higher seed weight were identified for further selection and
improvement.

 The dendrogram prepared to assess diversity showed diversity among different


accessions within inter and intra-regions. Six clusters were formed. Cluster1
showed maximum genetical diversity of 20-25%. One (23rd) accession was
found distinct with minimum genetical diversity and therefore it formed its
own cluster.

Evaluation of oil from feedstock


Seed oil was assessed for qualitative and quantitative parameters.

 Oil percentage ranged from 17 to 38% with an average of 31.36 (standard


deviation of 4.3). Only a few lots had low oil. Majority of accessions had oil
percentage more than 30.

 Triacylglycerol was the dominant lipid species (88.2%).

 High unsaponifiable matter (3.8%) in Jatropha curcas seed oil is a favorable


character.

 Free fatty acid percentage content ranged from 1.9 % to more than 10. FFA is a
negative trait. High FFA demands a two-step esterification.

163
 Acid value of the oil ranged from 0.85 to 3.91 with an average of 1.94 (standard
deviation of 0.68). Similarly, density ranged from 0.92 to 0.88% with an average
of 0.91 (standard deviation of 0.1). Refractive index was nearly uniform in all
the samples.

 Seed oil content correlated with phenological traits like seed thickness as well
as acid value.

 Saponification value ranged from 71.6 (in # 415772) to 129.5 (in # 468908).
There was no correlation between saponification value and insoluble
impurities.

 The density of Jatropha seed oil was 0.90317 g/ml.

The viscosity of Jatropha oil seed must be reduced for biodiesel application since the
kinematic viscosity of biodiesel were very low compared to vegetable oils. High viscosity of
the Jatropha oil seed is not suitable if it has to be used directly as engine fuel, it will often
results in operational problems such as carbon deposits, oil ring sticking, and thickening and
gelling of lubricating oil as a result of contamination by the vegetable oils.

Free fatty acid percentage was a very variable character and would change after
storage. Free fatty acid percentage was also poorly correlated with refractive index and
density. Pearson correlation value was negative (-0.32) for free fatty acid percentage and
refractive index, while it was 0.2 for free fatty acid percentage and density. However, it was
positively correlated with acid value with correlation factor of 0.71. Free fatty acid percentage
is controlled by environmental factors and can change with moisture and exposure to air.

Fatty acid composition of oil


Fatty acid composition of hexane extracts of J. curcas seeds showed free fatty acids
(FFA), methyl ester of fatty acids (FAME) and triglycerol esters (TAG), along with small It was
observed that Jatropha oil seed has highest oleic as compared to palm oil, palm kernel,
sunflower, coconut and soybean oil.

164
 Jatropha oil seed contains only 0.2% linolenic acid, which is lower as compared
to sunflower oil and palm oil.

 The major fatty acids in Jatropha seed oil were the oleic acid, linoleic acid,
palmitic acid and the stearic acid. The most prominent TAGs of Jatropha seed
oil were OLL and OOL. The oil extracts exhibited good physicochemical
properties and could be useful as biodiesel feedstock and industrial application.

 The fuel properties of Jatropha curcas oil, Jatropha curcas methyl ester and
diesel were compared and it has been observed that calorific value of these
methyl esters are around 37 mJ/kg, which are low as compared to diesel fuels
(42.0 mJ/kg).

Synthesis of oil at various development stages


Seeds of 10 accessions were collected at various stages of development, starting from
a week after fertilization and five days thereafter till maturity. These were classified as stage I
to stage VII. The seeds collected from mature black colored fruits, ready for dehiscence were
classified as Stage VII. Seeds from various stages of development were examined for
phonological traits, moisture and oil content. Each stage was evaluated for seed weight,
moisture content and synthesis of triglycerides.

 Oil content was the lowest at stage I, and the highest at stage VI. There was a
slight decrease in total oil percentage in dry mature seeds in stage VII.

 Moisture content of the seeds ranged from 10.1 to 89.7%; the lowest in mature
seeds in stage VII and highest in stage I. Seed fresh weight, seed area and oil
percentages of all seed developments were measured.

 Variation of oil synthesis in different accessions was also evaluated so that


promising accessions could be identified.

 The study suggests that stage III of seed development can be identified as a
turning point for development since at this stage concentration of sterols
decreases to negligible; there is very little methyl ester formation. The findings

165
shall be useful for silviculturists for fixing harvest rotations, providing field
inputs at seed maturation stage and in understanding of synthesis of
metabolites including that of specific triglycerides. The findings can be helpful
for understanding of biosynthesis and in efforts to improve biosynthesis of TAG
and reducing FFA content in the mature seeds.

 The results indicate that there is a possibility of early analysis of oil content at
young stage. This can be helpful for the breeders to select the individuals by
analyzing the oil content at early stage.

Methods of extraction of oil


Extraction of Oil is extracted either usually expeller or by solvent extraction. Different
methods were evaluated. These were Soxhlet method and the same was compared with
extraction by homogenization of seed tissue using Polytron homogenizer. Physicochemical
properties of extracted oil were also analyzed according to BIS and ASTM standards.

Although Soxhlet method also gave relatively high yield of oil but looking the
advantages of time and energy, the difference can be factored and extraction using Polytron
homogenizer was recommended. Homogenization is strongly recommended as it saves
energy and takes few minutes as compared to hours in Soxhlet method. It was observed that
Soxhlet extraction gave highest oil percentage although the seed meal was extracted in
solvent over heat.

Other than Soxhlet, other extraction methods like SCFE methods were also evaluated.
Super Critical Fluid Extraction procedure gave lowest oil percentage, although there was no
difference in quality. In fact, in terms of moisture, free fatty acid percentage content and
other traits, oil extracted by Super Critical Fluid Extraction was better as compared to other
methods was observed. This method, however, is not recommended since it consumed
relatively high power which may not be feasible for commercial purpose.

166
Solvents for extraction
Different solvents were evaluated to assess the most promising one for oil extraction.
Optimization for recovery of oil yield using non-polar solvents (hexane, petroleum ether,
cyclohexane and n-pentane) was also evaluated. Hexane was found to be most promising,
safe for extraction and least expensive. The effect of solvent, temperature, solvent to solid
ratio, particle size of the meal and reaction time were investigated to optimize the extraction
operating conditions for achieving maximum oil yield.

Highest oil was extracted in Petroleum ether as compared to the other three. Next
best solvent was n-Hexane. Since it is not advisable to use Petroleum ether, further
experiments were done with hexane only. The extraction yield with n-hexane was found to be
about 1.1% more than that of petroleum ether (37.6% and 35.4.0%, respectively under similar
conditions.

At room temperature, the oil extracted in two solvents was 30.3 and 29.8 %, nearly 20
% less than optimum. It increased to 33.6 % and 34.7 % at 45˚C and 55˚C, respectively when
extracted with hexane. Optimum condition was found to be the boiling point of solvent 65˚C
for hexane; however, optimization was achieved at 55˚C with petroleum ether.

Other factors like temperature and meal size for oil extraction
Meal size was also important for oil extraction. It was concluded that small to medium
particle size meal between 0.5 and 0.75 mm were suitable for solid liquid extraction of
Jatropha seeds. Extracted oil percentage was 37.2 % at 3:1 and 37.55 at 4:1 when extracted
with hexane.

Deterioration of oil during storage


Effect of storage and oxidation of oil was evaluated in a controlled experiment. Freeze
dried seeds were extracted with hexane using homogenizer and the solvent was evaporated
under reduced pressure. The oil was kept at various conditions to study the effect of
environmentally factors on oil quality. Storage of Jatropha oil for long intervals resulted in
increase in free fatty acid (FFA) content of the oil, resulting in more wastage of oil.

167
Transesterification
Either NaOH or H 2S04 alone was not suitable as catalyst for biodiesel production
because of higher free fatty acid content in Jatropha curcas seed oil. Saponification occurred
when o n l y NaOH catalyst was used and resulted in low biodiesel yield. H 2S04 catalyst
w a s r e q u i r e d f o r a t w o - s t e p e s t e r i f i c a t i o n . It was found that the acid value
of Jatropha oil was reduced to ne a r ly 1 from more t ha n 10 mg KOH/g in 1 h. At the
subsequent second step, 96 % biodiesel yield was obtained with 0.32 mg KOH/g acid value
in half an hr. transesterification r e a c t i o n . The two-step process o n l y n e e d e d
n e a r l y 2 h o u r s to achieve 96 % diesel yield.

A two-step transesterification process is developed to convert the high FFA non


edible oils to its esters. The first step (acid catalyzed transesterification) reduces the FFA
content of the oil to less than 2%. The alkaline catalyst transesterification process
converts the products of the first step to its mono- esters and glycerol. Factors effecting the
biodiesel production (reaction temperature, reaction rate & catalyst) were standardized.
Both H 2S04 a n d NaOH were used as catalysts in a two-step (acid esterification a n d base-
transesterification) process t o h a v e economical and stable biodiesel.

Fuel properties
Fuel properties of biodiesel from neem, Pongamia and Jatropha curcas have been
evaluated. The study revealed that methyl esters had higher cloud and pour points than
conventional diesel. This is important for engine operation in cold or cooler environments.
Kinematic viscosity and cetane values are slightly higher than the diesel, which is favorable for
combustion has been observed. Higher flash point is advantageous for fuel transportation.

The fuel properties of Jatropha curcas biodiesel were within the limits and
comparable with the conventional diesel. Except calorific value, all other fuel properties of
biodiesel were found to be higher as compared to diesel. Biodiesel produced from vegetable
oils and other feedstocks can be more prone to oxidation than a typical petroleum diesel
unless modified or treated with additives. Moisture, temperature, oxidative stress can lead to
deterioration of biodiesel.

168
Performance of neat biodiesel, biodiesel mix with conventional petroleum in various
proportions (B5, B20, and B100) was evaluated for engine performance.

All types of diesel fuels can gel at low temperatures. The temperature required to
induce gelling (freezing) in biodiesel is generally higher than that of conventional diesel fuel
and depends on the composition of the biodiesel. In particular, saturated methyl esters will
have higher freezing points than unsaturated methyl esters. The composition of the biodiesel
is, of course, dictated by the oil feedstock used for its production.

Biodiesel produced from vegetable oils and other feedstocks can be more prone to
oxidation than a typical petroleum diesel unless modified or treated with additives. Moisture,
temperature, oxidative stress can lead to deterioration of biodiesel. The purpose of this study
was to provide data that define how stable biodiesel or a biodiesel blend should be to prevent
the formation of acids or sediment during transportation, storage, and use.

Jyoti
January 2011

169
Chapter 10
References

Abdel Gadir, W.S., Onsa, T.O., Ali, W.E.M., El Badwi, S.M.A., and Adam, S.E.I., 2003.
Comparative toxicity of Croton macrostachys, Jatropha curcas and Piper
abyssinica seeds in Nubian goats. Small Ruminant Research 48: 61–67.
Adebowale, K.O., and Adedire, C.O., 2006. Chemical composition and insecticidal properties
of the underutilized Jatropha curcas seed oil. African J. of Biotechnology 5(10):
901-906.
Adriaans, T., 2006. Suitability of solvent extraction for Jatropha curcas. M.Sc. thesis, Ingenia
Consultants & Engineers for FACT Foundation.
Agarwal, D., and Agarwal, A.K., 2007. Performance and emissions characteristics of
Jatropha oil (preheated and blends) in a direct injection compression
ignition engine. Applied Thermal Engineering 27: 2314-2323.
Akintayo, E.T., 2004. Characteristics and composition of Parkia biglobbossa and Jatropha
curcas oils and cakes. Bioresource Technology 92: 307–310.
Alamu, O.J., Waheed, M.A., and Jekayinfa, S.O., 2007. Effect of ethanol–palm kernel oil ratio
on alkali-catalyzed biodiesel yield. Fuel.
Aliske, M.A., Zagonel, G.F., Costa, B.J., Veiga, W., and Saul, C.K., 2007. Measurement of
biodiesel concentration in a diesel oil mixture. Fuel 86: 1461–1464.
Annarao, S., et. al., 2008. Lipid profiling of developing Jatropha curcas L. seeds using 1H
NMR spectroscopy. Bioresource Technology Vol. 99 (18): 9032-9035.
Arzamendi, G., Campo, I., Arguifiarena, E., Sanchez, M., Montes, M., and Gandia, LM., 2007.
Synthesis of biodiesel with heterogeneous NaOH/alumina
catalysts: comparison with homogeneous NaOH. J. of Chemical Engineering
134:123-30.
ASTM, American Standards for Testing of Materials., 2003. D 189-01, D 240-02, D 4052-96,
D 445-03, D 482-74, D 5555-95, D 6751-02, D 93-02a, D 95-990, D 97-02.
Augustus, G.D.P.S., JavaBeans, M., and Seiler, G.J., 2002. Evaluation and bioinduction of
energy components of Jatropha curcas. Biomass and Bioenergy 23: 161-164.
Ayhan, D., 2009. Progress and recent trends in biodiesel fuels. Energy Conversation
Management 50:14-34.
Azam M.M., Waris, A., and Nahar, N.M., 2005. Prospects and potential of fatty acid methyl
esters of some non-traditional seed oils for use as biodiesel in India. Biomass
and Bioenergy 29: 293–302.
Banapurmath, N.R., Tewari P.G., and Hosmath, R.S., 2008. Performance and emission
characteristics of a DI compression ignition engine operated on Honge, Jatropha
and sesame oil methyl esters. Renewable Energy 33: 1982–1988

170
Banerji, R., Chowdhury, A.R., Misra, G., Sudarsanam, G., Verma, S.C., and Srivastava, G.S.,
1985. Jatropha seed oils for energy. Biomass 8: 277–282.
Barakos, N., Pasias, S., and Papayannakos, N., 2008. Transesterification of triglycerides in
high and low quality oil feeds over an HT2 hydrotalcite catalyst.
Bioresource Technology 99: 5037-42.
Barnwal, B.K., and Sharma, M.P., 2005. Prospects of biodiesel production from vegetable oils
in India. Renewable and Sustainable Energy. Reviews 9 (4): 363-378.
Behl, H.M., Sidhu, O.P., and Singh,N., 2007. Agrotechnology packages of Bioenergy crops
Jatropha Curcas; DBT Govt. of India: 47-50.
Behl, H.M., Sidhu, O.P., and Singh, N., 2007. Agrotechnology packages of Bioenergy crops
Madhuca Indica; DBT Govt. of India 51-57.
Behl, H.M., Sidhu, O.P., and Singh, N., 2007. Agrotechnology packages of Bioenergy crops
Salvodora ; DBT Govt. of India 75-83.
Behl, H.M., Sidhu, O.P., and Singh, N., 2007. Agrotechnology packages of Bioenergy crops
Pongamia Pinnata; DBT Govt. of India 58-66.
Berchmans, H.J., and Hirata, S., 2007. Biodiesel production from crude Jatropha curcas L.
seed with a high content of free fatty acids. Bioresource Technology
99:1716-21.
Berrios, M., Siles, J., Martin, M.A., and Martin, A. 2007. A kinetic study of the esterification of
free fatty acids (FFA) in sunflower oil. Fuel 86: 2383–2388.
Bhagat, et. al., 2010. Physico-Chemical Analysis, NMR Spectroscopy and Gas Chromatographic
Studies of Jatropha curcas L. Germplasm. J. of American Oil Chemists Society.
DOI: 10.1007/s11746-010-1678-7.
Bhatti, HN., Hanif, MA., Qasim, M., and Rehman, AU., 2008. Biodiesel production from
waste tallow. Fuel 87:2961-6.
Bhasabatra, R., and Sutiponpeibuns, S., 1982 a. Jatropha curcas oil as a substitute for diesel
engine oil.Renewable Energy:- Review Journal. 4: 56–70.
Bhasabatra, R., Sutiponpeibuns, S., 1982 b. The study of Jatropha curcas oil as a substitute of
diesel engine oil. Thai. Verglon. Department of Agriculture, Ministry of
Agriculture and Cooperatives, Bangkok: 42.
Blassnegger, J., 2002. Biodiesel as an automotive fuel: bench tests. In stability of biodiesel −
used as a fuel for diesel engines and heating systems. Presentation of the biostab
project results. Proceedings. Graz, blt Wieselburg: Wieselburg, Austria.
Bondioli, P., Gasparoli, A., Della bella, I., Tagliabue, S., and Toso, G., 2003. Biodiesel stability
under commercial storage conditions over one year. Eur. J. Lipid sci. Technol.,
105: 735– 741.
Bondioli, P., Gasparoli, A., Della bella, l., and Tgliabue, S., 2002. Evaluation of biodiesel
storage stability using reference methods. Eur. J. Lipid sci. Technol., 104: 777–
784.

171
Bouaid, A., Martinez, M., and Aracil, J., 2007. Long storage stability of biodiesel from
vegetable and used frying oils. Fuel.
Bozbas, K., 2006. Biodiesel as an alternative motor fuel. Production and policies in the
European Union. Renewable and Sustainable Energy Reviews.
Burley, J., Wood, P.J., 1976. A manual on species and provenance research with particular
reference to the Tropics. Tropical Forestry Papers No. 10. Department of
Forestry, Commonwealth Forestry Institute, University of Oxford :226.

Canaki ,M., 2007. Jatropha –Palm biodiesel blends: An optimum mix for Asia. Fuel 86, 1365-
7.
Canakci, M., 2007. The potential of restaurant waste lipids as biodiesel feedstocks.
Bioresource Technology 98 (1): 183–190.
Canakci, M., and Gerpen, V., 2001. Biodiesel production from oils and fats with high free
fatty acids. Trans ASAE 44:1429-36.
Canakci, M., and Van Grepen, J., 1999. Biodiesel production via acid catalysis.
Transaction of the American Society of Agricultural Engineers 42 (5): 1203–
1210.
Chang, D.Y.Z., Van Gerpen, J., Lee, I., Jonson, I.A., Hammond, E.G., and Marley, S.J., 1996.
Fuel properties and emissions of soybean oil esters as diesel fuel. J. American
Oil Chemists Society 73: 1549.
Chhetri, A.B., Tango, M.S., Budge, S.M., Watts, K.C., Rafiqul, M., and Islam, M.R., 2008.
Non-edible plant oils as new sources for biodiesel production. Int. J. Mol. Sci. 9:
169-180.
Crabbe, E., Nolasco-Hipolito, C.N., Kobayashi, G., Sonomoto, K., and Ishizaki, A., 2001.
Biodiesel production from crude palm oil and evaluation of butanol extraction
and fuel properties. Process Biochemistry 37: 65–71.
Correa, S.M., and Arbilla, G., 2007. Carbonyl emissions in diesel and biodiesel exhaust.
Atmospheric Environment.
Cosgrove, J.P., Church, DF., and Pryor, Wa., 1987. The kinetics of the autoxidation of
polyunsaturated fatty acids. Lipids. 22 (5):299-304.
Crookes, R.J., 2006. Comparative bio-fuel performance in internal combustion engines.
Biomass Bioenergy. 30 (5): 461-468.
Cvengros, J., and Cvengrosova, Z., 2004. Used frying oils and fats and their utilization in the
production of methyl esters of higher fatty acids. Biomass Bioenergy 27: 173 –
181.
Deerberg, S., Twickel, J., Förster, H.H., Cole, T., and Fuhrman, J., 1990. Synthesis of medium
chain fatty acids and their incorporation into triacylglycerols by cell-free
fractions from Cuphea embryos. Planta 180: 440-444.
Dehgan, B., and Webster, G.L., 1978. Three new species of Jatropha (Euphorbiaceae) from
Western Mexico. Madrono. 25: 30–39.

172
Dehgan, B., and Webster, G.L., 1979. Morphology and intrageneric relationships of the genus
Jatropha (Euphorbiaceae). University of California Publications in Botany.
Demel, R.A., and Kruyff, De. B., 1976. The function of sterols in membranes. Biochemical and
Biophysics Acta 457: 109-132.
Demirbas, A., 2 0 0 9 . Biodiesel from waste cooking oil via base-catalytic and supercritical
methanol transesterification. Energy Conversational Management 50:923-7.
Demirbas, A., 2007. Progress and recent trends in biofuels. Progress in Energy and
Combustion Science 33: 1–18.
Demirbas, A., 2007. Importance of biodiesel as transportation fuel. Energy Policy.
Demirbas, A., 2006. Biodiesel production via non-catalytic SCF method and biodiesel fuel
characteristics. Energy Conversion and Management 47: 2271–2282.
Deng, X., et. al., 2010. Energy Conversion and Management 51: 2802-2807
Dionex, 2009. Development of Jatropha oil extraction from Biodiesel feedstocks using
accelerated solvent extraction, http://www.dionex.com/en-us/webdocs/73815-
PO-Jatropha-LPN-2258-01.pdf.
Durbin, T.D., David, R., Cocker, D.R., Sawant, A.A., Johnson, K., Miller, J.W., Holden, B.B.,
Helgeson, N.L., and Jac, J.A., 2007. Regulated emissions from biodiesel fuels
from on/off-road applications. Atmospheric Environment 41: 5647–5658.
Edem, D.O., 2002. Palm oil: Biochemical, physiological, nutritional, hematological, and
toxicological aspects: A review. Kluwer Academic Publishers. 57: 319–341.
Encinar, JM., Gonzalez, JF., Sabia, E., and Ramiro, MJ., 1999. Preparation and properties
of biodiesel from Cynara cardunculus L. oil. Ind. Eng. Chern. Res. 38: 2927-
31.
Eromosele, I.C., Eromosele, C.O., Innazo, P., and Njerim, P., 1997. Short communication:
studies on some seeds and seed oils. Bioresource Technology 64: 245–247.
Esuoso, KO., and Bayer, E., 1998. Chemical composition and potentials of some underutilized
tropical biomass II. Andenopus breviflorus and Cucumeropsis edulis. La Rivista
italiana Delle Sostanze Grasse 75: 191 – 196.
Esuoso, KO., Lutz, H., Kutubuddin, M., and Bayer, E., 1998. Chemical composition and
potentials of some understerilised tropical biomass Telfaria occidentalis. Food
Chemistry 61 (4): 487-492.
Fang, H., and Mccormick, R. L., 2006. Spectroscopic study of biodiesel degradation pathways.
Sae technical paper search.
Fernando, S., Adhikari, S., Kota, K., and Bandi, R., 2007. Glycerol based automotive fuels
from future bio refineries. Fuel 86: 2806–2809.
Fernando, S., Karra, P., Hernandez, R., and Jha, S.K., 2007. Effect of incompletely converted
soybean oil on biodiesel quality. Energy 32 (5): 844-851.
Foidl, N., Foildl, G., Sanchez, M., and Mittelbach, M., 1996. Jatropha curcas L. as a source for
the production of biofuel in Nicaragua. Bioresource Technology 58: 77-82.

173
Forson, F.K., Oduro, E.K., and Donkoh, E.H., 2004. Performance of Jatropha oil blends in a
diesel engine. Renewable Energy 29 (7): 1135-1145.
Frankel, E. N., 2005. Lipid oxidation, 2nd ed.; the oily press ltd: Bridgewater, England.
Freedman, B., Butterfield, R.O., and Pryde, E.H., 1986. Transesterification kinetics of soybean
oil. J. American Oil Chemists Society 63: 1357-1380.
Fukuda, H., Kondo, A., and Noda, H., 2001. Biodiesel fuel production by transesterification of
oils: review. J. Bioscience and Bioenergy 92:5-16.
Gandhi, V.M., Cherian, K.M., and Mulky, M.J., 1995. Toxicological studies on Ratanjyot oil.
Food and Chemical Toxicology 33 (1): 39–42.
Gao, YY., Chen, WW., Lei, HW., Liu, YH., Lin, XY., and Ruan, R., 2009. Optimization of
transesterification conditions for the production of fatty acid methyl
ester ( FAME) from Chinese tallow kernel oil with surfactant coated lipase.
Biomass and Bioenergy 33:277-82.
Gerpen, J.V., 2005. Biodiesel processing and production. Fuel Processing Technology 86,
1097– 1107.
Gerpen, J.V., Shanks, B., Pruszko, R., Clements, D., and Knothe, G., August 2002–January
2004. Subcontractor report on biodiesel production technology. National
Renewable Energy Laboratory.
Ghadge, S.V., and Raheman H., 2006. Process optimization for biodiesel production from
mahua (Madhuca indica) oil using response surface methodology. Bioresource
Technology 97: 379–384.
Ghadge, S.V., and Raheman, H., 2005. Biodiesel production from mahua (Madhuca indica) oil
having high free fatty acids. Biomass and Bioenergy 28: 601–605.
Ginwal, HS., Phartyal, S.S., Rawat, P.S., Srivastava, R.L., 2005. Seed source variation in
morphology, germination and seedling growth of Jatropha curcas Linn. in
Central Asia. Silvae Genetica 54: 76-80.
Ginwal, H.S., Rawat, P.S., Srivastava, R.L., 2004. Seed Source Variation in Growth
Performance and Oil Yield of Jatropha curcas Linn. in Central India. Silvae
Genetica 53: 186-192.
Grandmougin-Ferjani, A., Schuller-Muller, I., and Hartmann, M.A., 1997. Sterol modulation of
the plasma membrane H+-ATPase activity from corn roots reconstituted into
soybean lipids. Plant Physiolology 113: 163-174.
Goff, MJ., Bauer, NS., Lopes, S., Sutterlin, WR., and Suppes, GJ., 2004. Acid-catalyzed
alcoholysis of soybean oil. J. Am. Oil Chern. Soc. 81:415-20.
Goering, C.E., Schwab, A.W., Daugherty, M.J., Pryde, E.H., and Heakin, A.J., 1982. Fuel
properties of eleven vegetable oils. Trans ASAE 85: 1472-1483.
Goodrum, J.W., 2002. Volatility and boiling points of biodiesel from vegetable oils and tallow.
Biomass Bioenergy 22: 205–211.
Gubitz, G.M., Mittelbach, M., and Trabi, M., 1999. Exploitation of tropical oil seed plant
Jatropha curcas L. Bioresource Technology 67: 73–82.

174
Gutierrez, V.R., and Barron, L.J.R., 1995. Method for analysis of triacylglycerols. J. of
Chromatogrphy. Biomedical Application: 671 133–168.
Gupta, P.K., Kumar, R., and Panesar, B.S., 2008. Storage studies on plant oil based bio-
diesel fuels. Agricultural Engineering International: the CIGR Ejournal. Vol.
X.
Gunstone, F.D., 2004. Rapeseed and Canola Oil: Production, Processing, properties and
uses. London: Blackwell Publishing Ltd.
Gunstone, F.D., and Shukla, V.K.S., 1995. NMR of lipids. Annu. Rep. NMR Spectroscopy 31:
219-237.
Gunstone, F.D., 1994. The chemistry of oils and Fats: Sources, composition, properties and
uses. London: Blackwell Publishing Ltd.
Haas, M.J., Mc Aloon, A.J., Yee, C.W., and Foglia, T.A., 2006. Process model to estimate
biodiesel production costs. Bioresource Technology 97: 671–678.
Haas, W., and Mittelbach, M., 2000. Detoxification experiments with the seed oil from
Jatropha curcas L. Industrial Crops and Products 12:111–118.
Harding, K.G., Dennis, J.S., Blottnitz, H.V., and Harrison, S.T.L., 2007. A life-cycle
comparison between inorganic and biological catalysis for the production of
biodiesel. J. of Cleaner Production.
Hartmann, M. A., Perret, A.M., Carde, J.P., Cassagne, C., and Moreau, P., 2002. Inhibition of
the sterol pathway in leek seedlings impairs phosphatidylserine and
glucosylceramide synthesis but triggers an accumulation of triacylglycerols.
Biochemical and Biophysics Acta 1583: 285-296.
Hawash, S., Kamal, N., Zaher, F., Kenawi, O., and Diwani, G., 2009. Biodiesel fuel from
Jatropha oil via non-catalytic supercritical methanol transesterification. Fuel
88:579-82.
He, H., Wang, T., and Zhu, S., 2007. Continuous production of biodiesel fuel from vegetable
oil using supercritical methanol process. Fuel 86: 442–447.
Hebbal, O.D., Reddy, K.V., and Rajagopal, R., 2006. Performance characteristics of a diesel
engine with deccan-hemp oil. Fuel 85 (10): 2187–2194.
Heller, J., 1996. Physic nut Jatropha curcas L. Promoting the conservation and use of
underutilized and neglected crops. Institute of Plant Genetic and Crop Plant
Research, Gatersleben/International Plant Genetic Resource Institute, Rome,
Italy. <http://www.ipgri.cgiar.org/Publications/pdf/161.pdf>.
Heftmann, E., 1970. Recent advances in Photochemistry; Runeckensad, V. Steehnk, C. (ed).
Appleton Century New York. Vol.3.
Hikwa, D., 1995. Jatropha curcas L. Agronomy Research Institute, Department of Research
and Specialist Services, Harare, Zimbabwe p. 4.
Hirota, M., Suttajit, M., Suguri, H., Endo, Y., Shudo, K., Wongchai, V., Hecker, E., and Fujiki,
H., 1988. A new tumor promoter from the seed oil of Jatropha curcas L., an

175
intramolecular diester of 12-deoxy-16-hydroxyphorbol. Cancer Research 48:
5800–5804.
Issariyakul, T., Kulkarni, M.G., Lekha, C.M., Dalai, A.K., and Bakhshi, N.N., 2007. Biodiesel
production from mixtures of canola oil and used cooking oil. J. Chemical
Engineering.
Juan, L., Fang, Y., Lin, T., and Fang, C., 2003. Antitumor effects of curcin from seeds of
Jatropha curcas. Acta Pharmacology Sin. 24(3): 241-246.
Kandpal, J.B., and Madan, M., 1995. Jatropha curcas: a renewable source of energy for
meeting future energy needs. Renewable Energy 6 (2):159–160.
Karmee, S.K., and Chadha, A., 2005. Preparation of biodiesel from crude oil of Pongamia
pinnata. Bioresource Technology 96:1425–1429.
Karmee, S.K., Mahesh, P., Ravi, R., and Chadha, A., 2004. Kinetics study of the base catalyzed
transesterification of monoglycerides from Pongamia oil. J. of American Oil
Chemists Society 81: 425-30.
Kaul, S., Saxena, R.C., Kumar, A., Negi, M.S., Bhatnagar, A.K., Goyal, H.B., and Gupta, A.K.,
2007. Corrosion behavior of biodiesel from seed oils of Indian origin on diesel
engine parts. Fuel Processing Technology 88 (3): 303-307.
Kaushik, N., Kumar, K., Kumar, S., Kaushik, N., and Roy, S., 2007. Genetic variability and
divergence studies in seed traits and oil content of Jatropha (Jatropha curcas L.)
accessions. Biomass and Bioenergy 31: 497-502.
Khetrapal, C.L., and Lakshminarayana, M.R., 1984. Application of NMR in Agriculture food
Stuff, , soil science and allied fields. Quarterly Chemistry Reviews 1: 1-13.
Knothe, G., 2007. Some aspects of biodiesel oxidative stability. Fuel Processing
Technology 88: 669–677.
Knothe, G., 2006. Analyzing biodiesel: standards and other methods. J. of American Oil
Chemistry Society 83:823–833.
Knothe, G., 2002. Structure indices in FA chemistry. How relevant is the iodine value. J. of
American Oil Chemical Society 9: 847–853.
Knothe, G. J., 2002. Structure indices in fa chemistry. How relevant is the iodine value. J. of
American oil chemical society 79: 847– 854.
Kumari, A., Mahapatra, P., Garlapati, VK., Banerjee, R., 2009. Enzymatic
transeseterification of Jatropha oil. Biotechnol. Biofuels 2:1-7.
Kumar, A., a n d Sharma, S., 2008. An evaluation of mulipurpose oil seed crop for
industrial uses (Jatropha curcas): A review. Industrial Crops and Products.
Kumar, N., and Sharma, P.B., 2005. Jatropha curcas-A sustainable source for production of
biodiesel. J. of Scientific Industrial Research 64: 883-889.
Kumar, M.S., Ramesh, A., a n d Nagalingam, B., 2003. An experimental comparison of
methods to use methanol and Jatropha oil in a compression ignition engine.
Biomass and Bioenergy 25: 309–318.

176
Kyriakidis, N.B., and Katsiloulis, T., 2000. Calculation of iodine value from measurements of
fatty acid methyl esters of some oils: comparison with the relevant American
oil chemists society method. J. of American Oil Chemical Society 77:1235–
1238.
Lakshminarayana, M.R., Giriraj, Ramanathan, K.V., and Khetrapal, C.L., 1984. Oil build up
and quality in developing sunflower seeds. La Ravista Italiana Delle Sostanze
Grasse 487-490.
Lapuerta, M.N., Armas, O., Fernandez, J.R., 2007. Effect of biodiesel fuels on diesel engine
emissions. Progress in Energy Combustion Science.
Leung, D.Y.C., Koo, B.C.P., and Guo, Y., 2006. Degradation of biodiesel under di erent
storage conditions. Bioresource Technology 97: 250–256.
Liauw, M.Y., Natan, F.A., Widiyanti, P., D. Ikasari, D., Indraswati, D., Soetaredjo, F.E.,
2008. Extraction of Neem Oil (Azadiarchta indica) using n-Hexane and
Ethanol: Studies of oil quality, kinetic and thermodynamic. J. of Engineering
and Applied sciences vol. 3 (3).
Ma, F., a n d Hanna, M.A., 1999. Biodiesel production: a review. Bioresource Technology 70:
1–15.
Makkar, H.P.S., Aderibigbe, A.O., and Becker, K., 1998. Comparative evaluation of non-toxic
and toxic varieties of Jatropha curcas for chemical composition, digestibility,
protein degradability and toxic factors. Food Chemistry 62: 207–215.
Makkar, HPS., Becker, K., and Schmook, B., 2001. Edible provenances of Jatropha curcas
from QuintnaRoo state of Mexico and effect of roasting on antinutrient and toxic
factors in seeds.Institute for Animal Production in the Tropics and Subtropics
(480), University of Hohenheim, D-70593 Stuttgart, Germany.
Makkar, HPS., Aderibigbe, A.O., and Becker, K., 1998. Comparative evaluation of non-toxic
and toxic varieties of Jatropha curcas for chemical composition, digestibility,
protein degradability and toxic factors. Food Chemistry 62: 207-215.
Makkar, HPS., Becker, K., and Schmook, B., 1997. Edible provenances of Jatropha curcas
from QuintnaRoo State of Mexico and effect of roasting on anti-nutrient and
toxic factors in seeds. Institute for Animal Production in the Tropics and
Subtropics, University of Hohenheim, Germany p. 6.
Makkar, HPS., Becker, K., Sporer, F., and Wink, M., 1997. Studies on nutritive potential and
toxic constituents of different provenances of Jatropha curcas. J. Agri Food
Chem. 45: 3152–3157.
Martinez-Herrera J., Siddhuraju P., Francis, G., Davila-Ortiz., G., and Becker, K., 2006.
Chemical composition, toxic/antimetabolic constituents, and effects of different
treatments on their levels in four provenances of Jatropha curcas L from
Mexico. Food Chemistry 96: 80-89.
Mazhar, H., Quayle, R., Fido, R.J., Stobart, A.K., Napier, J.A., and Shewry, P.R., 1998.
Synthesis of storage reserves in developing seeds of sunflower. Phytochemistry
48: 429-432.

177
Mccormick, R. L., Ratcliff, M., Moens, l., and Lawrence, R., 2007. Several factors affecting
the stability of biodiesel in standard accelerated tests, fuel processing technology
88:651– 657
Mittelbach, M., and Gangl, S., 2001. Long storage stability of biodiesel made from rapeseed
and used frying oil, J. Am. Oil Chem. Soc., 78: 573– 577.
Meher, LC., Dharmagadda, VS., 2006. Optimization of alkali-catalyzed transesterification
of Pongamia pinnata oil for production of biodiesel. Bioresour
Technol;97:1392-7.
Meher LC, Sagar DV, Naik SN. Technical aspects of biodiesel production by
transesterification - a review. Renew Sustain Energy Rev 2006; 10:248-
68.
Meziane, S., Kadi, H., and Lamrous, O., 2006. Kinetic study of oil extraction from olive foot
cake. Grasasy Aceites, 57: 175-179.
Meziane, S., and H. Kadi., 2008. Kinetics and thermodynamics of oil extraction from olive
cake. J. of American Oil Chemists Society 85: 391-396.
Michael, JH., Andrew, JM., Winnie, CY., and Thomas, AF., 2006. A process model to estimate
biodiesel production costs. Bioresource Technology 97: 671-8.
Mittelbach, M., and Remschmidt, C., 1996. Biodiesel: The Comprehensive Handbook.
Mittal, K.F., Norris, F.A., Stirton, A.J., and Swern, D., 1964. Bailey’s Industrial Oil and
Fat products.

Minami, E., Saka, S., 2006. Kinetics of hydrolysis and methyl esterification for biodiesel
production in two-step supercritical methanol process. Fuel 85: 2479–2483.
Morin, P., Hamad, B., Sapaly G., Carneiro, M.G.R., Oliveira, P.G., Pries, De., Gonzalez,
W.A., Sales, E.A., and Essayem, N., 2007. Transesterification of rapeseed
oil with ethanol I. Catalysis with homogeneous Keggin heteropolyacids.
Applied Catalysis A: General 330: 69–76.
Monyem, A., and Van Gerpen, J.H., 2001. The effect of biodiesel oxidation on engine
performance and emissions. Biomass and Bioenergy 20:317–325.
Munshi, S.K., and Sukhija, P. S., 2006. Compositional changes and biosynthesis of lipids in the
developing kernels of almond (Prunus amygdalus batsch). J. of Science Food
Agriculture 35: 689-697.
Nabi, Md.N., Akhter, Md. Shamim., Shahadat, and Mhia Md. Zaglul., 2006. Improvement of
engine emissions with conventional diesel fuel and diesel–biodiesel blends.
Bioresource Technology 97 (3): 372-378.
Nambisan, P., 2007 Biotechnological intervention in Jatropha for biodiesel production.
Current Science 93(10).
Namkoong, G., Kang, HC., and Brouard, JS., 1988. Tree breeding: principles and
strategies.Monographs on theoretical and applied genetics, vol. 11. New York:
Springer Verlag, New York.

178
Ngamcharussrivichai, C., Wiwatnimit, W., Wangnoi, S., 2007. Modified dolomites as
catalysts for palm kernel oil transesterification. Journal of Molecular
Catalysis A: Chemical 276: 24–33.
Nguyen, M.T., Minowa, T., Hanaoka, T., and Hirata, S., 2005. Final report for Biodiesel
production from palm oil by transesterification. Biomass Technology
Research Centre, AIST Chugoka, Kure, Hiroshima, Japan.
Nourredini, H., Teoh, B.C and Clements, L.D., 1992a. Viscosities of vegetable oils and fatty
acids. J. of American Chemical Society 69:1184-1188.
Nouredini, H., Teoh, B.C and Clements, L.D., 1992b. Viscosities of vegetable oils and fatty
acids. J. of American Chemical Society 69: 1189-1191.
Openshaw, K., 2000. A review of Jatropha curcas: an oil plant of unfulfilled promise. Biomass
and Bioenergy 19: 1-15.
Oversen, L., Leth, T., and Hansen, K., 1998. Fatty acid composition and contents of trans
monounsaturated fatty acids in frying fats, and in margarines and shortenings
marketed in Denmark. J. of the American Oil Chemists Society 75 (9): 1079–
1083.
Pant, KS., Khosla, V., Kumar, D., and Gairola, S., 2006. Seed oil content variation in Jatropha
curcas Linn. in different altitudinal ranges and site condition in H.P. India.
Lyonia, 11(2):31-34.
Parida, A., and Aganathan, P., 2007. Agrotechnology packages of Bioenergy crops Jatropha
Curcas; DBT Govt. of India 47-50.
Patil, SVR., 2004. Jatropha oil – As an alternative to diesel. Agriculture and Industry Survey,
13: 17-22
Perry, H. J., Bligny, R., Gout, E., and Harwood, J.L., 1999. Changes in Kennedy pathway
intermediates associated with increased triacylglycerol synthesis in oil-seed
rape. Phytochemistry 52: 799-804.
Phan, AN., and Phan, TM., 2008. Biodiesel production from waste cooking oils. Fuel 87:
3490-6.
Pax, F., 1910. Euphorbiaceae–Jatropheae. In: Engler A, editor. Das Pflanzenreich IV, vol.
147(42). Leipzig: Verlag von Wilhelm Engelmann.
Prabakaran, AJ., and Sujatha, M., 1999. Jatropha tanjorensis Ellis and Saroja, a natural
interspecific hybrid occurring in Tamil Nadu, India.Genet Resour Crop Evol 46:
213–218.
Pramanik, K., 2003. Properties and use of Jatropha curcas oil and diesel fuel blends in
compression ignition engine. Renewable Energy 28: 239-248.
Proctor, A., and Bowen, D.J., 1996. Ambient-temperature extraction of Rice Bran oil with
Hexane and Isopropanol. J. of American Oil Chemists Society 73 (6): 811-813.
Raheman, H. a n d Phadatare, A.G., 2004. Diesel engine emissions and performance from
blends of karanja methyl ester and diesel. Biomass and Bioenergy 27: 393–397.

179
Raffaele, S., Francesco, A., and Livio P., 1997. 1H and 13C NMR of Virgin Olive Oil. An
Overview. Magnetic Resonance Chemistry 35: 133-145.
Rakotondramasy, R.L., Havet, J.L., Porte, C., and Fauduet, H., 2007. Solid-liquid extraction of
protopine from fumaria officinalis L-Analysis determination, kinetic reaction
and model building. Separat. Purification Technology 54: 253-261.
Ramadhas, A.S., Muraleedharan, C., and Jayaraj, S., 2005. Performance and emission
evaluation of a diesel engine fueled with methyl esters of rubber seed oil. J.
Renewable Energy 30: 1700–1789.
Ramadhas, AS., Jayaraj, S., and Muraleedharan, C., 2005. Biodiesel production from high
FFA rubber seed oil. Fuel 84: 335-40.
Ramadhas, A.S., Jayaraj, S., and Muraleedharan, C., 2004. Use of vegetable oils as I.C.
engine fuels—a review. Renewable Energy 29: 727–742.
Ramesh D., and Sampathrajan, A., 2008. Investigations on performance and emission
characteristics of diesel engine with Jatropha biodiesel and its blends. Vol. (10).
Ramos, M.J., Fernández, C.M., Abraham, C., and Ángel, R.L.P., 2008. Influence of fatty
acid composition of raw materials on biodiesel properties. Bioresource
Technology. doi:10.1016/j.biortech.2008.06.039.
Rao, T.V., Rao, G.P. and Reddy, K.H.C., 2008. Experimental investigation of Pongamia,
Jatropha and Neem methyl esters as Biodiesel on C.I. Engine. Jordan J. of
Mechanical and Industrial engineering (vol.2), 117-122.
Rao, GR., Korwar, GR., Shanker, AK., and Ramakrishna, YS., 2008. Genetic associations,
variability and diversity in seed characters, growth, reproductive phenology and
yield in Jatropha curcas (L.) accessions. Trees 22: 697–709.
Rathore, V., and Madras, G., 2007. Synthesis of biodiesel from edible and non-edible oils in
supercritical alcohols and enzymatic synthesis in supercritical carbon dioxide.
Fuel.
Reddy, J.N., and Ramesh, A., 2006. Parametric studies for improving the performance of a
Jatropha oil-fuelled compression ignition engine. Renewable Energy 31 (12)
1994-2016.
Richardson, J.F., and Harker, J.H., 2002. Coulson and Richardson’s Chemical Engineering-
Particle Technology and Separation Processes. Butterworth-Heinemann, ISBN:
0750644451:502-542.
Ross, J.H.E., and Murphy, D.J., 1992. Biosynthesis and localization of storage proteins,
oleosins and lipids during seed development in Coriandrum sativum and other
Umbelliferae. Plant Science 86: 59-70.
Roy, SM., Thapliyal, RC., and Phartyal, SS., 2004. Seed source variation in cone, seed and
seedling characteristic across the natural distribution of Himalayan low-level
pine Pinusroxburghii Sarg. Silvae Genetica, 53(3):116–123.

180
Sahoo, P.K., Das, L.M., Babu, M.K.G., and Naik, S.N., 2007. Biodiesel development from high
acid value polanga seed oil and performance evaluation in a CI engine. Fuel 86
(3): 448-454.
Saikia, SP., Bhau, BS., and Rabha, A, 2009. Study of accession source variation in morpho-
physiological parameters and growth performance of Jatropha curcas Linn.
Current Science 96 (12): 1631-1636.
Sarathy, S.M., Gaıl, S., Syed, S.A., Thomson, M.J.A., and Dagaut, P., 2007. A comparison of
saturated and unsaturated C4 fatty acid methyl esters in an opposed flow
diffusion flame and a jet stirred reactor. Proceedings of the Combustion Institute
31: 1015–1022.
Sarin, R., Sharma, M., Sinharay, S., and Malhotra, R.K., 2007. Jatropha–Palm biodiesel
blends: An optimum mix for Asia. Fuel 86: 1365–1371.
Sayyar, S., Abidin, Z. Z., Yunus, R., and Muhammad, A., 2009. Extraction of Oil from
Jatropha Seeds-Optimization and Kinetics. American J. of Applied Sciences 6
(7): 1390-1395.
Schaller, H., 2004. New aspects of sterol biosynthesis in growth and development of higher
plants. Plant Physiology and Biochemistry 42: 465-476.
Sekhar, M.C., Mamilla, V.R., Mallikarjun, M.V., and Reddy, K.V.K., 2009. Production of
Biodiesel from Neem Oil, International Journal of Engineering Studies, Vol. 1
(4): 295–302.
Shah, S., Sharma, A., and Gupta, M. N., 2005. Extraction of oil from Jatropha curcas L. seed
kernels by combination of ultrasonication and aqueous enzymatic oil extraction,
Bioresource Technology Vol. 96(1): 121-123.
Shah, S., Sharma, A., and Gupta, M.N., 2004. Extraction of oil from Jatropha curcas L.
seed kernels by enzyme assisted three phase partitioning. Industrial Crops and
Products 20: 275–279.
Shu, Q., Yang, BL., Yuan, H., Qing, S., and Zhu, G., 2007 Synthesis of biodiesel from soybean
oil and methanol catalyzed by zeolite beta modified with La3. Catal
Commun., 8: 2159-65.
Singh, R. K., and Padhi, S. K., 2009. Characterization of Jatropha oil for the preparation of
biodiesel. Natural Products radiance, Vol. 8(2): 127-132.
So, G.C., and Macdonald, D.G., 1986. Kinetics of oil extraction from Canola (rapeseed).
Canadian J. of Chemical Engineering 64: 80-86.
Srivastava, P.K., and Verma, M., 2007. Methyl ester of karanja oil as alternative renewable
source energy. Fuel.
Srivastava, A., and Prasad, R., 2000. Triglycerides-based diesel fuels. Renewable
Sustainable Energy Rev 4: 111–133.
Srivastava, R., 1999. Study in variation in morpho-physiological parameters with reference to
oil yield and quality in Jatropha curcas Linn.Ph. D. thesis, Forest Research
Institute (Deemed University), Dehradun, India.

181
Stamenkovic, O.S., Lazic, M.L., Todorovic, Z.B., Veljkovic, V.B., and Skala, D.U., 2007. The
effect of agitation intensity on alkali-catalyzed methanolysis of sunflower oil.
Bioresource Technology 98: 2688–2699.
Sunil, N., Varaprasad, K.S., Kumar, S.T., Abrahama, B., and Prasad, R.B.N., 2007. Assessing
Jatropha curcas L. germplasm in-situ—A case study. Biomass and Bioenergy.
Talens, L., Villalba, G., and Gabarrell, X., 2007. Exergy analysis applied to biodiesel
production. Resources, Conservation and Recycling 51: 397–407.
Tamlampudi, S., Talukder, MR., Hama, S., Numata, T., Kondo, A., and Fukuda, H., 2007. A
comparative study of immobilized whole cell and commercial lipases as
biocatalyst. J. Biochem. Eng. 39:185-9.
Tapanes, NCO., Aranda, DAG., Carneiro, JWD., and Antunes, OAC., 2008. Transesterification
of Jatropha curcas oil glycerides: theoretical and experimental studies of
biodiesel reaction. Fuel 87:2286-95.
Tatandjiiska, R. B., Marekov, I. N., and Nikilova-damyanova, B. M. J., 1996. Determination of
triacylglycerol classes and molecular species in seed oils with high content of
linoleic and linolenic fatty acids. Sci. Food. Agric. 72: 403– 410
Terry, B., mccormick, R. L., and Atarajan, M., 2006. Impact of biodiesel blends on fuel system
component durability. Sae tech. Pap. Ser.
Tiwari AK, Kumar A, and Raheman H., Biodiesel production from Jatropha oil with high
free fatty acids: an optimized process. Biomass and Bioenergy
2007: 31:569-75.
Tiwari, A.K., Kumar, A., and Raheman, H., 2007. Biodiesel production from Jatropha oil
(Jatropha curcas) with high free fatty acids: An optimized process. Biomass
Bioenergy 31: 569–575.
Tsuchiya, T., Shiotani, H., Goto, S., Sugiyama, G., and Maeda, A., 2006. Japanese standard for
diesel fuel containing 5% of fame; investigation on oxidation stability of fame
blended diesel. Sae tech. Pap. Ser.
Triki, S., Hamida, J. B., and Mazliak, P., 2000. Diacylglycerol acyltransferase in maturing
sunflower seeds. Biochemical Society Trans. 28: 689–692.
Tzia, C., and Liadakis, G., 2003. Extraction Optimization in Food Engineering. CRC Press,
USA., ISBN: 10: 0824741080: 170-171.
Van, G.J., 2005. Biodiesel processing and production. Fuel 86: 1097-107.
Van Gerpen, J. H., Hammond, E. G., YU, l., and Monyem, A., 1997. Determining the
influence of contaminants on biodiesel properties. Sae Tech. Pap. Ser., 971685.
Veljkovic, V.B., Lakicevic, S.H., Stamenkovic, O.S., Todorovic, Z.B., a n d Lazic, K.L.,
2006. Biodiesel production from tobacco (Nicotiana tabacum L.) seed oil with a
high content of free fatty acids. Fuel 85: 2671–2675.
Vicente, G., and Martınez, M., A, J., 2007. Optimisation of integrated biodiesel production.
Part I. A study of the biodiesel purity and yield. Bioresource Technology 98,
1724–1733.

182
Vicente, G., Martínez, M., and Aracil, J., 2007. Optimisation of integrated biodiesel
production. Part II: A study of the material balance. Bioresource Technology 98
(9): 1754-1761.
Visvanathan, R., Palanisamy, P.T., Gothandapani, L., and Sreenarayanan, V.V., 1996. Physical
Properties of neem nut. J. Agricultural Engineering Research 63 (1): 19-25.
Vyas, AP., Subrahmanyam, N., and Patel, PA., 2009. Production of biodiesel through
transesterification of Jatropha oil using KN03/Al 2 03 solid catalyst.
Fuel 88:625-8.
Wani, SP., Osman, M., D’silva, E., Sreedevi, TK., 2006. Improved livelihoods and
environmental protection through biodiesel plantations in Asia. Asian
Biotechnol Develop Rev, 8(2): 11–29.
Waynick, J. A., 2005. Characterization of biodiesel oxidation and oxidation products: crc
project no. Avfl-2b, nrel/tp-540−39096; National renewable energy laboratory:
golden, colorado, USA.
Westbrook, S. R., 2005. An evaluation and comparison of test methods to measure the
oxidation stability of neat biodiesel, nrel/sr-540−38983; National renewable
energy laboratory: golden, colorado, USA.
Wiese, K.L., and Snyder, H.E., 1987. Analysis of the oil extraction process in soybeans: A new
continuous procedure. J. of American Oil Chemical Society 64: 402-406.
Wood, P., 2005. Could Jatropha vegetable oil be Europe's biodiesel feedstock. Biofuels out of
Africa. Refocus 6 (4): 40-44.
Xie, WL., and Li, HT., 2006. Alumina- supported potassium iodide as a heterogeneous
catalyst for biodiesel production from soybean oil. J. Mol. Catal. A.
Chern., 255:1-9.
Xin, D., Zeng, F., Z., and Yun-hu, L., 2010. Ultrasonic transesterification of Jatropha
curcas L. oil to biodiesel by a two-step process. Energy Conversion and
Management 51: 2802-2807.
Xi, WL.,and Huang, XM., 2006. Synthesis of biodiesel from soybean oil using
heterogeneous KF/ZnO catalyst. Catal. Lett 107:53-9.
Xi, YZ., and Davis, RJ., 2008. Influence of water on the activity and stability of activated Mg­
Al hydrotalcites for the transesterification of tributyrin with methanol. J.
Catal. 254:190-7.
Yang, H.H., Chien, S.M., Lo, M.Y., Lan, J.C.W., Lu, W.C., and Ku, Y.Y., 2007. Effects of
biodiesel on emissions of regulated air pollutants and polycyclic aromatic
hydrocarbons under engine durability testing. Atmospheric Environment 41:
7232–7240.
Yin, J.Z., Xiao, M., and Song, JB., 2008. Biodiesel from soybean oil in supercritical
methanol with co-solvent. Energy Convers. Manage.49:909-12.
Zabaniotou, A., Ioannidou, O., and Skoulou, A., 2007. Rapeseed residues utilization for energy
and 2nd generation biofuels. Fuel.

183
Zeng, HY., Feng, Z., Deng, X., and Li, YQ., 2008. Activation of Mg-Al hydrotalcite catalysts
for transesterification of rape oil. Fuel 87:3071-6.
Zhang, JH., and Jiang, LF., 2008. Acid-catalyzed esterification of Zanthoxylum bungeanum
seed oil with high free fatty acids for biodiesel production.
Bioresource Technology 99:8995-8.
Zhang, Y., Dube, MA., McLean, DD., and Kates, M., 2003. Biodiesel production from waste
cooking oil: 1. Process design and technological assessment. Bioresource
Technology 89:1-16.

184

S-ar putea să vă placă și