You are on page 1of 8

Fuel Processing Technology 102 (2012) 53–60

Contents lists available at SciVerse ScienceDirect

Fuel Processing Technology


journal homepage: www.elsevier.com/locate/fuproc

Direct production of biodiesel from rapeseed by reactive extraction/in


situ transesterification
Rabitah Zakaria a,⁎, Adam P. Harvey b
a
Department of Process and Food Engineering, Faculty of Engineering, University Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
b
School of Chemical Engineering and Advanced Materials, Newcastle University, NE1 7RU, UK

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

Article history: Biodiesel is a fuel derived from renewable resources such as edible and inedible oil-bearing seed, algae, and
Received 27 October 2011 waste cooking oil. The conventional biodiesel process involves oil extraction, refining and transesterification.
Received in revised form 21 March 2012 Alternatively, transesterification can actually be performed directly from the oil-bearing materials without
Accepted 18 April 2012
prior extraction. This route which is often termed “reactive extraction” or “in situ transesterification” has
Available online 15 May 2012
the advantages of simplifying the biodiesel production process as well as potentially reducing production
Keywords:
cost. In this study, the reactive extraction of rapeseed with methanol has been characterised. The effects of
Biodiesel process parameters on the yield, conversion and reaction rate differ substantially from conventional trans-
Rapeseed esterification due to the dependence on both extraction and reaction. The rate of ester formation is mainly
Canola affected by the catalyst concentration, temperature and particle size while the equilibrium yield largely de-
Reactive extraction pends on the solvent to oil molar ratio. A high yield of ester (>85%) can only be achieved at high solvent
In situ transesterification to oil molar ratios (> 475:1). Parametric studies and light microscope images of reactively extracted seed
suggested that reactive extraction occurs by transesterification of the oil inside the seed, followed by diffu-
sion of the products into the bulk solvent.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction The conventional method for production of biodiesel requires the oil
to be extracted from the biomass before it can be transesterified into
The exploration of new energy resources has become increasingly im- ester. The transesterification reaction occurs in a liquid state. The oil ex-
portant in recent years due to the dwindling fossil fuel reserve and the traction process typically involves the use of a solvent (usually hexane)
adverse environmental effects of traditional fuel combustion [1]. Conse- in a counter current or percolation extractor, or a mechanical method
quently, the development of sustainable and “greener” energy resources such as screw press extractor. However, solvent extraction plant is ex-
has taken centre stage. One of the alternatives of renewable energy is pensive, complex and poses health and safety hazard due to the han-
fuel derived from plant and animal matter. These so-called “biofuels” dling of flammable and explosive solvent [4]. On the other hand,
can be produced by various processes, such as fermentation, trans- mechanical pressing often yields less oil than solvent extraction
esterification, pyrolysis, gasification, liquefaction, and hydrotreatment. resulting in a high residual oil in the waste biomass. Alternatively, the
Among these processes, transesterification of vegetable oil or animal fat need to extract the oil from the biomass prior to transesterification
to produce biodiesel has been successfully commercialised. Biodiesel is a can be eliminated using a “reactive extraction” method to produce bio-
petroleum diesel substitute and can be readily used in most diesel engines diesel [5]. In this method, the oil is simultaneously extracted and
without any modification. Furthermore, engine performance using bio- transesterified into alkyl ester in -situ in one single process. Hence, sep-
diesel has been shown to be comparable to that of conventional diesel arate extraction and oil refining steps can be eliminated since extraction
fuel [2]. Other merits of biodiesel include reduced emission of carbon is simultaneously carried out by the reactant itself. The use of solid raw
monoxides, particulates and hydrocarbon from the engine [2] and en- material rather than refined oil could certainly reduce the feedstock
hanced engine lubrication [3]. There are abundant types of resources cost. Since the method directly uses oil-bearing materials rather than
such as inedible oilseed (i.e. Jatropha, algae, and pongamia) or waste bio- pre-extracted oil, it eliminates processes such as crushing, solvent ex-
mass (i.e. waste cooking oil, animal fats, sewage sludges) that can be traction and degumming, and perhaps drying, which are key processes
utilised to produce biodiesel, directly contributing to the efficient use of in conventional biodiesel production. Furthermore, the elimination of
biomass and land resources. hexane from the overall process provides an added health and environ-
mental benefit as hexane is classified as a hazardous air pollutant and
contributes to the production of smog and global warming [6].
⁎ Corresponding author. Tel.: + 60 389464301; fax: + 60 389464440.
E-mail addresses: rabitah@eng.upm.edu.my (R. Zakaria), adam.harvey@ncl.ac.uk Previous studies have demonstrated that biodiesel production via re-
(A.P. Harvey). active extraction (also referred to as in situ transesterification) is capable

0378-3820/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.fuproc.2012.04.026
54 R. Zakaria, A.P. Harvey / Fuel Processing Technology 102 (2012) 53–60

of producing high yields of fatty acid esters from a variety of oil-bearing prepared by dissolving a known amount of sodium hydroxide in meth-
raw materials [7–9]. Increase in extraction yield is also possible depending anol, was first heated to the desired temperature using a heated circu-
on the type of oil in the biomass as certain oil contains high amount of lating water bath. Once heated, the alcohol was transferred to the
polar components (i.e. phospholipids), which can be extracted by the al- round-bottom flask and the reaction was carried out at the desired tem-
cohol during reactive extraction, whereas they are non-extractable by perature and reaction periods. The stirrer speed was maintained at
conventional methods [8]. The application of this process to substances 200 rpm for all of the experiment. It was important to ensure that
with low oil content, such as distillers dried grains with soluble (DDGS) there was no escape of methanol from the flask as losses of methanol
and meat and bone meal (MBM) seems to offer economic advantages to the atmosphere would distort the outcome of the parametric study.
due to the higher wastage associated with a conventional two-step pro- After operation for the desired time, a specific amount of glacial acetic
cesses [9]. Furthermore, some plant materials, such as algae, are difficult acid was added to the mixture to neutralize the catalyst and prevent fur-
to crush: preparing the biodiesel via reactive extraction may overcome ther reaction. The liquid was then separated from the seed using vacuum
this limitation [10]. filtration. The solid residue was washed repeatedly with methanol to re-
Reactive extraction substantially differs from the conventional bio- cover any product that adhered to the seed and the excess methanol was
diesel process in many ways. Notably, the system is heterogeneous, in- removed using a rotary evaporator. After evaporation, two layers of liquid
volving biological material with an array of biological structures and were formed. The upper layer contained the ester phase, while the bot-
components. As pointed out by Poirot et al. [11], extraction efficiency tom layer contained the glycerol phase. The layers were separated using
is influenced by the microstructure of the vegetable seed matrix, the a separating funnel. The masses of the phases were recorded and the
molecular structure and size of the solute, localization and link with ester content was analysed using a gas chromatograph.
other components. Hence, the efficiency of reactive extraction could The yield of ester extracted after the reaction completion was calcu-
be very different for different biological materials. This is evident in lated as:
the differing optimum parameters found in several studies [9,12,13]. Al-
though the method has been shown to be viable by early researchers, Esteryieldðwt:% Þ ¼ mass of ester phase  0:995  100
ð1Þ
considerable effort needs to be put into fully defining the characteristics mass of triglycerides in the seed  100
of the process. Most studies reported the yield of ester, but to our
knowledge none provided time profiles of the reaction intermediates where the factor 0.995 is to correct for the molecular weight difference
(mono- and diglycerides) as well as triglyceride. This information is es- between the triglyceride and ester. The concentration of mono-, di-
sential in order to identify the substance extracted and to fully describe and triglycerides were quantified using a Gas Chromatograph/Mass
the process kinetics. This work aims to investigate the factors affecting spectrometer.
reactive extraction of oil- bearing biomass to produce biodiesel. The pa-
rameters that can influence the reactive extraction process were stud- 2.4. Progress of reaction
ied in considerable depth in order to understand their effects on the
equilibrium yield and the kinetics. In depth study of the characteristics To obtain the time profile of ester and the reaction intermediates,
of the process may enable us to understand the process mechanism, 0.5 mL samples of the reaction mixture were taken at various time in-
which is critical in predicting the effects of various process parameters tervals using a syringe and a tube from the sampling pots. 3 μL of glacial
and in providing insight into possible further improvements. The oil- acetic acid was added to stop the reaction. A syringe filter was used to
seed chosen was rapeseed (or “canola”) as it is the biodiesel feedstock separate the liquid from the rest of the seed particles. A small portion
in Europe. of the liquid, which contains methanol and the reaction products, was
analysed using a gas chromatograph in order to quantify the amount
2. Material and methods of ester in the bulk methanol phase. The rest of the liquid was evaporat-
ed to remove the methanol until the sample reached constant weight.
2.1. Reagents and material The product was then allowed to settle into two layers and the ester
phase was withdrawn from the top and the concentration of mono-,
The rapeseed was kindly provided by a local farm near Newcastle di- and triglycerides were quantified using a gas chromatograph/mass
upon Tyne, UK. The hexane used was of 99.9% purity. The methanol and spectrometer (GC-MS).
ethanol used throughout this study were of 99.97% and 99.95% purity re-
spectively. All solvents were purchased from Fisher Scientific. Methyl 2.5. Effect of water
heptadeacanoate and tricaprin internal standards were obtained from
Sigma-Aldrich. To study the effect of seed moisture, reactive extraction was con-
ducted using dried seed. Seeds were dried in an oven at 104 °C until
2.2. Oil extraction they achieved constant weight. The yield of dry and wet seed was com-
pared by performing the reactive extraction experiment using 25 g of
The oil seed was first extracted using 4 different solvents namely seed, 250 ml methanol, 0.1 m catalyst concentrations and at 60 °C. The
methanol, ethanol, hexane and a 50:50 mixture of ethanol and metha- experiments were carried out in duplicate. The effect of water in meth-
nol. 20 g of the seed were pulverised using a coffee grinder for 1 min. anol was also studied using methanol containing 1, 2 and 4 wt.% water.
The ground seed were transferred into a cylindrical cup, covered with The experiment was performed using 25 g of both dried and wet seed,
glass wool, and placed in the Soxhlet extraction unit. 230 ml of solvent 100 ml methanol, and 0.1 m catalyst concentrations and at 60 °C. The
was placed in the round bottom flask and heated. The Soxhlet unit was water concentration in the bulk methanol phase during reactive extrac-
allowed to run for a pre-determined time i.e. 1, 3, 6, 8, and 12 h until tion was also analysed in duplicate using the Karl Fisher Volumetric
constant oil yield has been reached. Solvent was removed from the Analysis Method (Metrohm 701 KT Titrino). The sample (about 0.2 g)
extracted oil using a vacuum rotary evaporator. was taken from the bulk methanol phase using a clean, dry syringe.

2.3. Reactive extraction 2.6. Ester analysis

A known amount of ground and sieved seed was placed in a 500 mL Gas chromatography analysis for the determination of individual
round-bottom flask equipped with a condenser and an overhead stirrer and total esters was based on the British Standard BS EN 14103:2003.
and placed in a constant temperature water bath. Alkaline alcohol, The internal standard used was methyl heptadecanoate (MHDN). The
R. Zakaria, A.P. Harvey / Fuel Processing Technology 102 (2012) 53–60 55

stock solution of internal standard was prepared by adding 5.00 g of samples were examined using an Olympus BX41 light microscope and
methyl heptadecanoate to 322.6 g of analytical grade heptane. About digital images were collected using an Altra20 Soft Imaging System.
0.625 g of sample and 1.25 g of the stock solution was weighed accu-
rately to 4 decimal places on an AND HR-200 analytical balance. The 3.0. Results and discussions
GC used was a HP 5890 Series II, with a Varian CP wax column with di-
mensions of 0.53 mm i.d, 10 μm coating thickness and 25 m length. The 3.1. Extraction with different solvents
carrier gas was helium at a flow rate of 2 mL/min and the column, injec-
tor and detector temperature was maintained at 220 °C. The column The mass of extract from Soxhlet extraction of 20 g of seed for 6 h,
separated the esters and the methyl heptadecanoate (MHDN) into indi- using methanol, ethanol, 50/50 mixtures of ethanol/methanol and
vidual peaks. hexane yield different amount of oil can be seen in Table 1.
To analyze the ester concentration in the bulk methanol phase Methanol extracts some amount of material from the seed but very
obtained at various reaction times during the experiment, a modified little is triglyceride (TG). This was similar to the findings of several pre-
GC method was employed. The stock solution of the internal standards vious studies [14–16]. For example, Ozgul and Turkay [14] found that
was prepared by adding 5 g of methyl heptadecanoate in 322.6 g meth- methanol dissolves free fatty acid from rice bran oil leaving triglycerides
anol instead of hexane. A known amount of sample (1–2 g range) was in the bran. The poor triglyceride solubility in methanol is as expected,
weighed accurately into a sample vial with 1.25 g of the stock solution. since methanol is a very polar solvent, whereas most triglycerides are
The injection volume used was 0.2 μl. The gas chromatograph tempera- non-polar long chain hydrocarbon molecules. However, various other
ture program used was the same as described above. compounds in the seed can potentially dissolve in methanol, e.g. phos-
pholipids, sterols, phenols, and vitamins. Ethanol, on the other hand, is a
2.7. Gas chromatography/mass spectroscopy (GCMS) method for glycer- much better solvent for triglycerides, although it extracts other polar
ide analysis materials as well. The ability of ethanol to extract considerable amount
of oil has been well-documented [17–19]. A mixture of methanol and
A GCMS was used to quantify the monoglycerides, diglycerides and ethanol extracts some oil, but not as much as when using ethanol
triglycerides in the ester phase. The method was based on BS EN 14105 alone. Despite the poor solubility of triglycerides in methanol, its use
with modification to the calibration procedure in order to account for in simultaneous extraction and transesterification results in extraction
the higher glyceride contents, which occurred early in the reaction. of most of the oil and transesterification into ester [14–16]. The degree
Tricaprin (1,2,3-tricaproylglycerol) were used as a standard to calibrate of extraction, however, depends on the process parameters, which is
the glycerides. the subject of the following sections.
For the analysis of monoglyceride, about 20 mg of sample was accu-
rately weighed and placed in a sample vial. 20 μl of tricaprin stock solu- 3.2. Methanol to oil molar ratio
tion (12.6 mg/mL) and 20 μL of MSTFA was added to the sample vial.
The solution was left to silylate for 15 min after which 2 mL of heptane As can be seen in Fig. 1a and b, increasing the methanol to oil molar
was added. For identification of TG and DG about 20 mg of sample was ratio increases the equilibrium ester yield. The yield of ester (as defined
accurately weighed and place in a sample vial. 100 μL of tricaprin stock by Eq. (1)) improves greatly as the methanol amount is increased from
solution (0.5 mg/ml) and 20 μL of MSTFA were added to the sample vial. 0 to 300 molar ratio, but after that, increasing the methanol volume in-
The solution was left to silylate for 15 min after which 0.5 mL of heptane creases the yield only slightly until a maximum of about 82% of the tri-
was added. The GCMS was fitted with Perkin Elmer Col-elit column (PE- glycerides content in the seed is reached at about 600 molar ratio.
5HT) of 15 m length, 0.25 i.d and 0.1 μm film thickness. The detector However, increasing the methanol ratio to more than 900:1 seems to
type in the mass spectrometer was electron impact positive. The flow cause a reduction in the yield of ester. At these molar ratios, the
rate of helium was 1 mL/min. The oven temperature protocol was resulting product consists of a relatively smaller ester phase but a
50 °C hold for 1 min, heat to 180 °C at 15 °C/min, then at 7 °C/min to much larger glycerol phase. When the glycerol phase was extracted
230 °C and at 10 °C/min to 370 °C. The temperature was then held at with hexane, it was found that it does contain a significant amount of
370 °C for 10 min for a total run time of 31.5 min. The inlet line to the ester and the total ester yield is similar to that of the lower molar
MS was kept at 270 °C while the MS source temperature was kept at ratio. Hence, at higher methanol to oil molar ratio, it appears that the
250 °C. The standard used to calibrate the MS was 1.5,7 Triazabicyclo full separation of ester and glycerol phase becomes increasingly diffi-
(4,4,0) dec-5‐ene purchased from Fluka. cult. It should be noted that at high methanol to oil molar ratio, there
is also a high amount of catalyst in the solvent since the catalyst concen-
2.8. Light microscope Image tration in the solvent is kept constant. Therefore, a high amount of cat-
alyst may have increased the solubility of the ester in the glycerol phase
For light microscope observation, the rapeseed particles were rapid- due to increasing amount of water that is unavoidably produced upon
ly fixed in 1% (wt:vol) osmium tetroxide and then transferred to 2.5% dissolution of the catalyst. The time to reach final yield was similar for
(vol) glutaraldhyde in 0.1 M sodium cacodylate buffer (pH 7.2) and high and low methanol to oil molar ratio as shown in Fig. 1b, indicating
left overnight. Subsequently, all the samples were washed twice, each that the rate of extraction was not significantly affected by the solvent
time for 30 min, in 0.1 M sodium cacodylate or phosphate buffer and amount. Similar finding was observed by the study of Mondala et al.
then post-fixed in 1% osmium tetroxide for 2 h. Samples were [20] on in situ transesterification of municipal sludges, who found that
dehydrated in graded ethanol serial dilutions for 30 min for each solu- methanol to oil molar ratio had a weak effect on the overall extraction
tion and then finally in 100% ethanol for 30 min 3 times. All samples rate.
were impregnated with Spurr resin and embedded in moulds and poly-
merized at 60 °C. Sections were cut on Reichert Ultracut ultramicro-
Table 1
tome (Lecia Microsystems Ltd, Milton Keynes, UK), mounted on glass
Soxhlet extraction for 20 g of rapeseed for 6 h using various solvents.
slides. These procedures were performed by the Newcastle University
Electron Microscopy Research Services. Prior to microscopic examina- Methanol Ethanol 50/50 ethanol/methanol Hexane
tion, the slide containing the fixed sample was immersed in Sudan Total extract 3.12 10.21 6.32 9.14
black B for 5 min to stain the lipids. The reagents were then removed Triglycerides phase 0.10 7.44 1.60 9.14
by washing with 90% ethanol and air-dried. Sudan black B and periodic Other non-triglyceride 3.02 2.77 4.72 0.00
material
acid-Schiff reagent was purchased from Sigma Aldrich. The stained
56 R. Zakaria, A.P. Harvey / Fuel Processing Technology 102 (2012) 53–60

a increasing ethanol [23]. In another study, the extraction of oil from


100 Jatropha seed in a batch reactor was demonstrated to increase with
90 increasing hexane [24].
80 The high methanol to oil molar ratio requirement has been observed
70 in most studies for reactive extraction involving different oil-bearing
Yield ( wt %)

60 materials [25]. For example Haas et al. [12] found that 226:1 molar
50 ratio is needed to reactively extract soybeans with methanol. A 673:1
40 molar ratio is required to obtain maximum yield for cottonseed [26]
30 while 476:1 is required for sunflower [27]. The high energy required
20 for methanol recovery causes the production cost of biodiesel by this
10 method to be higher than for the conventional method [28].
0 Fig. 2 shows the concentration of ester in the ester phase versus
0 200 400 600 800 1000 1200 1400 methanol to oil molar ratio. The purity of the ester phase indicates the
Molar Ratio extent of conversion of the extracted material into ester in the bulk
ester hexane extracted ester methanol. The data shows that the conversion to ester is similar at
b low or high molar ratio. Hence even at low molar ratio (e.g. at 200:1),
100 the amount of methanol is adequate to convert the extracted material
90 into fatty acid methyl ester. This shows that the low equilibrium ester
80 yield at lower molar ratio is not the result of incomplete conversion of
Yield (wt%)

70
triglyceride to ester in the bulk methanol phase, but is probably due
60
50
to the low concentration gradient between the seed and the bulk liquid
40 as explained earlier.
30
20 3.3. Catalyst concentration
10
0 In conventional transesterification, the amount of catalyst is usually
0 50 100 150 200
0.5 to 1.0 wt.% of the oil. Higher concentrations cause excessive saponifi-
Time (min)
cation [29], while lower concentrations cause low conversion of triglycer-
270 Molar Ratio 475 Molar Ratio 150 Molar Ratio
ides to biodiesel [22]. However, this range may not be applicable to in situ
Fig. 1. a: Effect of methanol to oil molar ratio on the ester yield for 1 h reaction. Reac- transesterification due to the physical difference in the reaction phase,
tion conditions: 60 °C, 0.1 m catalyst, 300–1000 μm particle size, 25 g seed. b: Effect of such as the presence of solid and the vast amount of solvent used. The ef-
methanol to oil molar ratio on equilibrium yield. Reaction conditions: 60 °C, 0.1 m cat- fect of catalyst concentration ranging from 0.03 m (equivalent to 2.1 wt.%
alyst, 300–500 μm particle size, 25 g seed. oil) to 0.1 m is shown in Fig. 3, below. It can be seen that increasing the
catalyst concentration increases both the final ester yield and the rate of
The minimum amount of methanol needed for in situ trans- ester formation. Hence, an adequate amount of catalyst is essential in en-
esterification using batch type reaction is presumably the amount suring complete transesterification of the oil which consequently allows
that is needed to simply cover the seed and to render the solid for the extraction of triglycerides to take place. Increasing the amount of
phase workable for agitation. For 25 g of seed used in the study, the catalyst causes higher rates of ester extraction, which indicates that the
amount of methanol needed to adequately cover the seed is 40 ml rate of reaction controls the reactive extraction process.
which is a methanol to oil ratio of 76:1 for the rapeseed used. This is Fig. 4 shows the interactive effect of molar ratio and catalyst concen-
much higher than the optimum required for pre-extracted oil trans- tration. At a catalyst concentration of 0.05 m, the maximum equilibrium
esterification of 6:1 (as suggested by Freedman et al. [21]). The amount yield is achieved at a 1200:1 methanol to oil molar ratio whereas for
of catalyst used here is 1.1 wt.% of the oil, which is within the range 0.1 m, a similar maximum equilibrium yield is achieved at a 600:1
suggested for conventional transesterification of 0.5 to 1.0 wt.% [22]. methanol to oil molar ratio. Increasing the catalyst concentration fur-
Hence these amounts should be able to provide adequate reactants ther to 0.15 m produces a similar amount of extracted materials but
and catalyst for a significant conversion to be achieved. However, this the ester yield is lower than at 0.1 m. This is probably due to the adverse
is not the case for in situ reaction, as using a methanol to oil ratio of effects of the saponification reaction or dissolution of methyl ester in
95:1 and below does not result in any significant biodiesel being
produced.
Ester concentration in ester phase (wt %)

The fact that a large amount of solvent is needed in in situ trans-


esterification is probably due to the amount required to drive the ex- 100
traction rather than the reaction. Solvent extraction of oil from seed 95
can be explained partly by Fick's Law of diffusion, in which the yield
90
of extraction depends on the concentration gradient between the
bulk liquid and the internal seed, as well as the oil solubility in the 85
solvent. Early in the reaction, the yield increases rapidly due to the 80
high concentration gradient between the inside of the seed and
the bulk liquid. As more products are extracted, the concentration 75
gradient starts to decrease, thereby decreasing the extraction rate. 70
Extraction will stop once the concentration inside the seed is in equi-
65
librium with that in the bulk liquid. Therefore, the higher the amount
of solvent used, the more dilute the bulk liquid will be, so the higher 60
0 200 400 600 800 1000 1200
the concentration gradient, which will allow greater amounts of oil to
Molar Ratio
be extracted. This effect has been observed in several oilseed extrac-
tion studies with various solvents. In the extraction of oil with ethanol Fig. 2. Concentration of ester in the ester phase at different molar ratios. Reaction con-
for example it was found that the equilibrium yield increased with ditions: 1 h, 60 °C, 0.1 m catalyst, 300–1000 μm particle size, 25 g seed.
R. Zakaria, A.P. Harvey / Fuel Processing Technology 102 (2012) 53–60 57

100 Table 2
90 Ester yield and purity at different catalyst concentrations. Reaction conditions: 1 h,
Ester yield (wt%)

80 60 °C, 475:1 methanol to oil molar ratio, 300–500 μm particle size, 25 g seed.
70
60 Catalyst concentration (m) Ester concentration (wt.%) Ester yield (%)
50 0.1 90.3 ± 1.3 88.8 ± 0.1
40 0.05 89.7 ± 1.4 59.5 ± 0.8
30
20
10 3.4. Particle size
0
0 100 200 300 400
The effect of particle size on rate of reactive extraction can be seen
Time (min)
in Fig. 5.
0.03 molal 0.05 molal 0.1 molal For particle sizes between 300 and 500 μm, the time to reach maxi-
mum ester yield was about 1 hour but it increased to more than 3 h
Fig. 3. Effect of catalyst concentration on the ester yield. Reaction conditions: 1 h, 60 °C,
475:1 methanol to oil molar ratio, 300–500 μm particle size, 25 g seed. when a particle size of 1000 to 1400 μm were used. Despite the longer
extraction time, the final yield of the larger particles was essentially
the same as that of the smaller particles. Reducing the particle size
shortens the diffusion path of the oil and solvent, thereby increasing
the glycerol phase associated with excess catalyst. Clearly, increasing the extraction rate. It is also possible that the more extensive mechani-
the catalyst concentration reduces the methanol requirement to cal grinding needed to reduce the particle size may also cause a higher
achieve maximum yield. However, at 0.05 catalyst concentration and degree of distortion of the cell wall causing an increase in the diffusion
1200:1 molar ratio, the catalyst to oil ratio is the same as at 0.1 m and coefficient and faster extraction. Although there are several mathemat-
600:1 molar ratio, which explains how at lower catalyst concentration ical formulae in the literature that relate extraction time to particle size
a similar yield can be achieved by increasing the solvent amount. There- (see: [30,31], the extent of improvement in extraction rate may vary be-
fore, it appears that it is the catalyst to oil ratio rather than the catalyst tween particles of the same size since different seed preparation
concentration which dictates the amount of triglyceride extracted from methods can result in different degrees of cell wall disruption.
the seed. Low catalyst to oil ratio causes low conversion of the triglycer- The smaller the particle size, the greater the proportion of broken
ides to ester and hence low extraction yield. On the contrary, too much surface cell walls releasing more oil at the particle surface. The oil at
catalyst will reduce the yield presumably due to the excessive saponifi- the surface is not, of course, released to the bulk solvent by internal dif-
cation reaction. The optimum amount of catalyst appears to be in the fusion, rather it is solubilised or washed off at a much quicker rate. It is
region of 7 wt.% of the oil, which is higher than that required for con- postulated that if most of the oil accumulates at the particle surface, less
ventional transesterification. Adequate methanol to molar ratio on the solvent is needed for extraction, since the diffusion process is mini-
other hand is still important in order to provide the concentration gra- mized. Comparing the ester yield at different methanol to oil molar ra-
dient for extraction and even with enough catalyst the yield will still be tios for a low particle size of 300 to 500 μm and a particle size of 1000 to
low if the methanol to oil molar ratio is low. 1400 μm, it can be seen from Fig. 6 that similar amounts of solvent are
Table 2 shows the concentration of ester and the equilibrium yield at needed to achieve maximum yield. This indicates that for particle
0.1 and 0.05 m catalyst concentration. The data shows even though the sizes of 300–500 μm, the extraction of the oil still depends largely on
ester yield is lower at 0.05 m, similar concentration of ester is achieved diffusion rather than solubilisation.
for both catalyst concentration. This indicates that even at the low cat-
alyst concentration sufficient catalyst is available in the bulk phase to 3.5. Temperature
convert the extracted material into ester. Therefore, it could be inferred
that the lower ester yield obtained with the lower catalyst concentra- In general, temperature affects both the rate of diffusion and reac-
tion is not the result of incomplete reaction in the bulk phase. Converse- tion. Examination of Fig. 7 shows that increasing the reaction tempera-
ly, it is possible that there is insufficient conversion inside the seed ture from 30 to 60 °C increases the rate of ester formation particularly
which prevents further extraction of the triglycerides. early in the reaction, but the final equilibrium yield is the same. A
number of researchers have shown that transesterification can be
adequately carried out at temperatures as low as 20 °C without a
great deal of compromise on the rate and conversion [22,29]. Oil extrac-
18
tion, on the other hand, proceeds faster at higher temperatures, as the
16 viscosity of the oil and the solvent is reduced and diffusivity is increased.
14
12 90
Mass (g)

10 80
Ester yield (wt%)

8 70
60
6
50
4 40
2 30
0 20
0 200 400 600 800 1000 1200 1400 1600 10
Methanol to oil molar ratio 0
0 50 100 150 200 250 300 350
ester yield 0.1 molal ester yield 0.05 molal Time (min)
ester yield 0.15 molal total extract 0.1molal
total extract 0.05 molal total extract 0.15 molal 300-500 micron 800-1000 micron 1000-1400 micron

Fig. 4. Effect of catalyst concentration at different molar ratio on ester yield and total Fig. 5. Effect of particle size at 571:1 methanol to oil molar ratio on ester yield. Reaction
extract. Reaction conditions: 1 h, 60 °C, 300–1000 μm particle size, 25 g seed. conditions: 1 h, 60 °C, 25 g seed, 0.1 m catalyst concentration.
58 R. Zakaria, A.P. Harvey / Fuel Processing Technology 102 (2012) 53–60

100 90
90 80

Ester yield (wt %)


80 70
Ester yield (wt %)

70 60
60 50
50 40
40 30
30 20
10
20
0
10 190 380 571 666
0 Molar Ratio ( Methanol to Oil)
571 380 285 190
Wet Seed Dry Seed ( 0 % Moisture)
Molar Ratio
300-500 micron 1400-1000 micron
Fig. 8. Effect of drying on the yield of ester at different methanol to oil molar ratios. Re-
action conditions: 1 h, 25 g seed, 60 °C, 0.1 m catalyst concentration, 300–500 μm par-
Fig. 6. Effect of particle size on final ester yield at different molar ratios. Reaction con-
ticle size.
ditions: 1 h, 60 °C, 25 g seed, 0.1 m catalyst concentration.

The solubility of the oil also increases at high temperature, which conse- in the bulk solvent, which might make the process apparently less toler-
quently increases the final extraction yield, notably at low solvent to oil ant to water than in situ transesterification using rapeseed.
ratio [32]. However, in reactive extraction, the increase in rate is only Water analysis of the bulk methanol phase (25 g of rapeseed and
significant during the early part of the extraction, as a similar time is 250 mL methanol) shows that for the wet seed, the water content of
needed to achieve maximum yield for lower and higher temperatures. methanol increases quickly to about 1.3 wt.% after 1 min reaction
Other works in in situ transesterification have also shown that minor and the water level stays practically constant thereafter . The fresh al-
improvement is gained from increasing the temperature in the range kaline methanol has about 0.6 wt.% water initially, since the methanol
of 30 to 65 °C [12,13]. The extraction and subsequent transesterification is only 99.9% pure and the dissolution of sodium hydroxide in meth-
can be efficiently performed at low temperature, which is favourable for anol releases water. For dry seed, the amount of water in the bulk
the process economics. phase stays at 0.6 wt.% and throughout the reaction.
In order to identify the level of water that affects the reactive extrac-
tion, the reaction was carried out with 25 g seed and 100 mL methanol
3.6. Drying that contains 0 to 4 wt.% water. It can be seen from Fig. 9 that when
using completely dried seed, the ester yield was not affected by water
As discussed previously, the presence of water can be detrimental to until more than 2 wt.% water is used. This amount of water is equivalent
the transesterification reaction as it causes the formation of hydroxide to 6.3 wt.% of the seed weight, which is approximately the amount of
ions, which induce saponification of free fatty acids, triglycerides and es- water that is in the wet rapeseed. Apparently 6.7 wt.% of water in the
ters. With respect to extraction, water might further reduce the solubility seed is not high enough to cause a detrimental effect on the process. If
of triglycerides in methanol due to its polarity. As the rapeseed contains more than 6.7 wt.% of water in the seed is used, as in the case of meth-
6.7% water, the effect of drying the seed to 0% water was studied anol with 4% water for dry seed and 1 wt.% for wet seed, the yield of
(Fig. 8). It can be seen that drying the seed does not increase the ester ester substantially decreases. Hence, as suggested by Haas, a higher
yield, nor does it reduce the amount of methanol needed to achieve max- amount of methanol is needed in order to dilute the water concentra-
imum yield. T-test analysis (SPSS 17 Statistical Software) gave a p value of tion and to achieve similar yield as the dry seed [32].
0.715, which is greater than 0.05, indicating that the difference between
yield obtained from dried and wet seed is not significant. This is in con- 3.7. Discussion of the process mechanism
trast to a study by Haas and Scott [33] who found that drying soybeans
to 0% water substantially reduced the amount of methanol needed to It has been shown that methanol was not able to extract triglyceride
achieve maximum ester yield by 60%. However, soybeans have a lower from rapeseed to any considerable extent. However the presence of the
oil content of about 22 wt.% and hence for the same amount of soybeans alkaline catalyst allows significant production of methyl ester from tri-
the amount of methanol used is lower than that for rapeseed. The higher glycerides. The mechanism of this process is still not clear. Several stud-
seed moisture in less methanol will cause a higher concentration of water ies [12,13,16] have speculated on the mechanism although they have
not supported their theories with experimental result nor discussed it

100 70
90 60
Ester Yield (wt%)

Ester yield (wt%)

80
70 50
60
40
50
40 30
30
20 20
10 10
0
0 20 40 60 80 100 120 140 0
0 1 2 4
Time (min)
Percent water in methanol (wt%)
30 deg C 40 deg C 60 deg C
Wet seed Dry seed
Fig. 7. Effect of reaction temperature on ester yield. Reaction conditions: 1 h, 25 g seed,
0.1 m catalyst concentration, 475:1 methanol to oil molar ratio, 300–500 μm particle Fig. 9. Effect of water in methanol on the yield of ester at 270:1 methanol to oil molar
size. ratio.
R. Zakaria, A.P. Harvey / Fuel Processing Technology 102 (2012) 53–60 59

in details. One possible mechanism is that the initial presence of esters In order to have a better insight into the process mechanism, the con-
causes the remaining triglycerides in the oilseed to be more soluble in centration of mono-, di-, and triglycerides in the ester phase throughout
the methanol ester mixture. Kildiran et al. [16] suggested the extraction the reaction was observed (Fig. 11). As can be seen, there is only a maxi-
of triglycerides is possible due to the stepwise dissolution and subse- mum of about 6 wt.% of mono-, di and triglycerides in total in the ester
quent transesterification of the oil. Haas et al. [12] and Qian et al. [13] phase, which occurs very early in the reaction (1 min reaction time).
suggested that the alkaline catalyst could destroy the cell walls and in- During the reactive extraction period, the extracted material is predomi-
tracellular compartmentalization resulting in cellular solubilisation and nantly ester. It appears that triglycerides and the reaction intermediates
subsequent transesterification of triglycerides. However, our light mi- (mono- and diglycerides) do not accumulate to a significant degree in
croscope images of the seed particle stained with Sudan Black B after re- the bulk solvent. Since the extracted materials are predominantly ester,
active extraction (Fig. 10) indicate otherwise [34]. Fig. 10a shows that it could be inferred that either i) triglycerides are rapidly reacted to
the fresh seed consists of cells of about 10–30 μm in diameter and the ester once extracted into the bulk solvent or ii) the transesterification re-
cells were surrounded by cell wall membranes 0.5 to 2 μm thick. The action occurred inside the seed and the resulting ester was extracted into
lipid which can be seen as a blue colour was located inside the cells. It the bulk solvent . The latter mechanism is more likely based on the image
can be seen that most of the lipids have been removed from the cells seen in Fig. 10c earlier.
after reactive extraction (Fig. 10b). The image also shows that the cells This mechanism also fits well with the behaviour of the process as
after reactive extraction consist of a significant amount of fully intact observed from the parametric studies. As shown in Section 3.2, the
cells confirming that the addition of sodium hydroxide into the solvent yield of ester increases with an increase in the amount of methanol
does not destroy the cell wall structure. The mechanism of extraction used. If the reaction occurs first in the seed, the ester will diffuse into
would still require the diffusion of some fraction of the oil or ester out- the bulk solvent due to the concentration gradient between the ester
ward through the cell membrane and into the bulk liquid. Fig. 10c in the bulk liquid and in the seed. Increasing the amount of solvent
shows the image of particle whereby the oil has not been fully extracted will cause the concentration gradient to increase, thereby increasing
as shown by the blue colour inside the cells. The remaining oil seems to the yield of ester. On the other hand, if triglycerides were extracted
occur only in the middle of the particle rather than throughout the seed. and reacted immediately as they reach the bulk liquid, the concentra-
This indicates that the lipid does not move radially outward into the tion of triglycerides in the bulk liquid will always be zero or near it.
bulk liquid. Rather, it suggests that there was a reaction front that Therefore, there would always be a maximum triglyceride concentra-
consumes the lipid similar to a shrinking core process. Hence, the reac- tion gradient inside the seed and the bulk liquid, which does not explain
tion step involves the diffusion of methanol through the plant cell wall why higher solvent (methanol) is needed for higher yield. Also, if it is
and reaction with the intact oil bodies. Once reacted, the partial glycer- assumed that triglycerides were extracted since it becomes more solu-
ides (i.e. mono- and diglyceride) will dissolve in methanol and further ble in the mixture of ester and methanol, the rate of extraction should
reacted with methanol inside the seed to produce ester. The appearance be slower in the beginning of the reaction and should become more
of mono-and diglyceride in the bulk solvent depends on the reaction rapid when more ester is produced as the reaction progresses. In
and diffusion rate, contrast, the rate of extraction is very fast early in the reaction and

100 µm
a b

c
Fig. 10. Light micropscope images stained with Sudan Black B. a) typical seed particle prior to extraction and b) completely extracted seed particle and c) partialy extracted seed
particle after extraction.
60 R. Zakaria, A.P. Harvey / Fuel Processing Technology 102 (2012) 53–60

3.0 [6] U.S EPA, U.S. Environmental Protection Agency, The Clean Air Act Amendments of
Concentration (wt %)

1990 List of Hazardous Air Pollutants[online] Available at: http://www.epa.gov/


2.5 ttn/atw/orig189.html1990[Accessed 3 May 2010].
[7] S.H. Shuit, K.T. Lee, A.H. Kamaruddin, S. Yusup, Reactive extraction and in situ es-
2.0 terification of Jatropha curcas L. seed for the production of biodiesel, Fuel 89
monoglyceides
(2010) 527–530.
1.5 diglycerides
[8] S. Dufreche, R. Hernandez, T. French, D. Sparks, M. Zappi, E. Alley, Extraction of
triglycerides lipids from municipal wastewater plant microorganisms for production of biodie-
1.0 sel, Journal of the American Oil Chemists' Society 84 (2007) 181–187.
[9] M.J. Haas, K.M. Scott, T.A. Foglia, W.N. Marmer, The general applicability of in situ
0.5 transesterification for the production of fatty acid esters from a variety of feed-
stocks, Journal of the American Oil Chemists' Society 84 (2007) 963–970.
0.0
[10] M.B. Johnson, Z. Wen, Production of biodiesel fuel from the microalga schizochytrium
0 10 20 30 40 50 60 70
limacinum by direct transesterification of algal biomass, Energy and Fuels 23 (2009)
Time (min) 5179–5183.
[11] R. Poirot, L. Prat, C. Gourdon, C. Diard, J.-M. Autret, Fast batch to continuous solid–
Fig. 11. Concentration time profile of mono-, di-, and triglycerides in the ester phase. liquid extraction from plants in continuous industrial extractor, Chemical Engi-
neering and Technology 30 (2007) 46–51.
[12] M.J. Haas, K.M. Scott, W.N. Marmer, T.A. Foglia, In situ alkaline transesterification:
an effective method for the production of fatty acid esters from vegetable oils,
decreases as the reaction progresses. Furthermore, from Table 2 it was JAOCS, Journal of the American Oil Chemists' Society 81 (2004) 83–89.
shown that the ester yield is low at low catalyst concentration, but the [13] J. Qian, F. Wang, S. Liu, Z. Yun, In situ alkaline transesterification of cottonseed oil
for production of biodiesel and nontoxic cottonseed meal, Bioresource Technolo-
ester concentration in the bulk methanol is similar at higher and gy 99 (2008) 9009–9012.
lower concentrations. Hence, inadequate conversion of triglycerides [14] S. Özgül, S. Türkay, In situ esterification of rice bran oil with methanol and etha-
in the bulk methanol phase is not the cause of the lower extraction nol, Journal of the American Oil Chemists' Society 70 (1993) 145–147.
[15] J. Zeng, X. Wang, B. Zhao, J. Sun, Y. Wang, Rapid in situ transesterification of sun-
yield. On the other hand, it is likely that at low catalyst concentration, flower oil, Industrial and Engineering Chemistry Research 48 (2009) 850–856.
there is insufficient conversion of triglycerides to ester inside the seed, [16] G. Kildiran, S.O. Yücel, S. Türkay, In-situ alcoholysis of soybean oil, Journal of the
which reduces the yield of ester. American Oil Chemists' Society 73 (1996) 225–232.
[17] D. Franco, J. Sineiro, M. Pinelo, M.J. Núñez, Ethanolic extraction of Rosa rubiginosa
soluble substances: oil solubility equilibria and kinetic studies, Journal of Food
4. Conclusions Engineering 79 (2007) 150–157.
[18] L.A. Johnson, E.W. Lusas, Comparison of alternative solvents for oils extraction,
Journal of the American Oil Chemists' Society 60 (1983) 229–242.
The maximum yield of reactive extraction of rapeseeds using meth-
[19] J.T. Chien, J.E. Hoff, M.J. Lee, H.M. Lin, Y.J. Chen, L.F. Chen, Oil extraction of dried
anol is determined primarily by the methanol to oil molar ratio. Metha- ground corn with ethanol, Chemical Engineering Journal 43 (1990) B103–B113.
nol to oil molar ratios greater than 400:1 were required to obtain ester [20] A. Mondala, K. Liang, H. Toghiani, R. Hernandez, T. French, Biodiesel production
yields higher than 80%. This is a much higher amount of methanol than by in situ transesterification of municipal primary and secondary sludges, Bio-
resource Technology 100 (2009) 1203–1210.
that needed for conventional transesterification. In order to make the [21] B. Freedman, R.O. Butterfield, E.H. Pryde, Transesterification kinetics of soybean
process economically viable, more research is needed to identify tech- oil 1, Journal of the American Oil Chemists' Society 63 (1986) 1375–1380.
niques that can reduce the solvent consumption. [22] M. Mittelbach, B. Trathnigg, Kinetics of alkaline catalyzed methanolysis of sun-
flower oil, Fat Science. Technology 92 (1990) 145–148.
The rate of the process, on the other hand, is mainly controlled by [23] D. Franco, M. Pinelo, J. Sineiro, M.J. Núñez, Processing of Rosa rubiginosa: extrac-
the catalyst concentration and the seed particle size. Decreasing the cat- tion of oil and antioxidant substances, Bioresource Technology 98 (2007)
alyst concentration and increasing the particle size reduces the overall 3506–3512.
[24] S. Sayyar, Z.Z. Abidin, R. Yunus, A. Muhammad, Extraction of oil from Jatropha
rate of extraction. Temperatures within the range of 30 to 60 °C only af- seed-optimization and kinetics, American Journal of Applied Sciences 6 (2009)
fect the initial rate of extraction, but the time to reach the maximum 1390–1395.
yield is similar. The maximum yield of ester was achieved at a catalyst [25] F.H. Kasim, A. Harvey, R. Zakaria, Biodiesel production by in situ transesterification,
Biofuels 1 (2010) 355–365.
concentration of 0.1 mol/kg and 670:1 methanol to oil molar ratio. It ap- [26] K.G. Georgogianni, M.G. Kontominas, P.J. Pomonis, D. Avlonitis, V. Gergis, Alkaline
pears that the process is not significantly adversely affected by water conventional and in situ transesterification of cottonseed oil for the production of
when the seed moisture content is less than about 6.7 wt.%. Hence, biodiesel, Energy and Fuels 22 (2008) 2110–2115.
[27] K.G. Georgogianni, M.G. Kontominas, P.J. Pomonis, D. Avlonitis, V. Gergis, Conven-
the drying step could be removed if the moisture content does not ex-
tional and in situ transesterification of sunflower seed oil for the production of
ceed this level, which should improve the process economics. biodiesel, Fuel Processing Technology 89 (2008) 503–509.
[28] M.J. Haas, Simplifying biodiesel manufacture, Industrial Bioprocessing 27 (2005)
References 5.
[29] G. Vicente, M. Martínez, J. Aracil, A. Esteban, Kinetics of sunflower oil methanolysis,
[1] S.N. Naik, V.V. Goud, P.K. Rout, A.K. Dalai, Production of first and second genera- Industrial and Engineering Chemistry Research 44 (2005) 5447–5454.
tion biofuels: a comprehensive review, Renewable and Sustainable Energy Re- [30] H.B. Coats, M.R. Wingard, Solvent extraction. III. The effect of particle size on ex-
views 14 (2009) 578–597. traction rate, Journal of the American Oil Chemists' Society 27 (1950) 93–95.
[2] M. Canakci, J.H. Van Gerpen, Comparison of engine performance and emissions for [31] D.F. Othmer, J.C. Argawal, Extraction of soybean: theory and mechanism, Chemi-
petroleum diesel fuel, yellow grease biodiesel, and soybean oil biodiesel, Transac- cal Engineering Progress 51 (1955) 372.
tions of the American Society of Agricultural Engineers 46 (2003) 937–944. [32] J. Lajara, Solvent extraction of oil from oilseed: the real basics, Proceedings of the
[3] M. Balat, Fuel characteristics and the use of biodiesel as a transportation fuel, En- World Conference on Edible Fats and Oils Processing, 1989, pp. 49–55.
ergy Sources Part A: Recovery, Utilization and Environmental Effects 28 (2006) [33] M.J. Haas, K.M. Scott, Moisture removal substantially improves the efficiency of in
855–864. situ biodiesel production from soybeans, JAOCS, Journal of the American Oil
[4] A. Kurki, J. Bachmann, H. Hill, Oilseed processing for small-scale producersAvailable at: Chemists' Society 84 (2007) 197–204.
http://www.attra.org/attra-pub/oilseed.html2008[Accessed 10 November 2010]. [34] Y. Ren, A. Harvey, R. Zakaria, Biorefining based on biodiesel production: chemical
[5] K.J. Harrington, C. D'Arcy-Evans, A comparison of conventional and in situ methods and physical characterisation of reactively extracted rapeseed, Biobased Materials
of transesterification of seed oil from a series of sunflower cultivars, Journal of the and Bioenergy 4 (2010) 1–8.
American Oil Chemists' Society 62 (1985) 1009–1013.

You might also like