You are on page 1of 8

Fuel 89 (2010) 677684

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Variables affecting the in situ transesterication of microalgae lipids


E.A. Ehimen, Z.F. Sun *, C.G. Carrington
Physics Department, University of Otago, 730 Cumberland Street, Dunedin 9016, New Zealand

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

Article history: This paper describes the effect of important reaction variables on the production of biodiesel from non-
Received 17 February 2009 edible microalgae lipids, using the acid-catalysed in situ transesterication process. The specic gravity
Received in revised form 22 September of the biodiesel product was used to monitor the conversion progress. The results indicate that increasing
2009
the reacting alcohol volume and the temperature lead to improved fatty acid methyl ester (FAME) con-
Accepted 13 October 2009
Available online 31 October 2009
versions. With the exception of in situ transesterication carried out at room temperature (23 C), the
equilibrium FAME conversions appear to approach asymptotic limits for reaction times greater than
8 h for all temperatures investigated. Stirring the reaction vessel had a signicant positive inuence on
Keywords:
Microalgae
the rate of biodiesel formation. Increasing the moisture content of the microalgae biomass had a strong
Biodiesel negative inuence on the equilibrium FAME yield, and in situ transesterication was inhibited when the
In situ transesterication biomass water content was greater than 115% w/w (based on oil weight).
2009 Elsevier Ltd. All rights reserved.

1. Introduction acidic (hydrochloric acid, HCl) and alkaline (sodium hydroxide,


NaOH). These authors showed that the use of acidic catalysts for
The use of microalgae biomass for the production of biodiesel the production of fatty acid methyl esters (FAME) from microalgae
(fatty acid alkyl esters, FAAE) has been described by Chisti [1] as oil resulted in higher product yields than alkaline catalysts under
one of the most promising biomass feedstocks with the potential the same reaction conditions. The inuence of the concentration
to meet petro-diesel replacement targets without encroaching on of the catalyst, the reaction time and the temperature, on the
arable land suitable for food production. The production of biodie- transesterication reaction was also investigated in Ref. [2], with
sel from microalgae oil has previously been demonstrated in the an optimal equilibrium yield of biodiesel from microalgae oil ob-
literature using the conventional route [2,3], which involves the tained using 0.6 N hydrochloric acid in methanol for 0.1 h at a reac-
extraction of the lipids from the microalgae biomass followed by tion temperature of 70 C. The conventional transesterication of
its conversion to FAAE and glycerol. Most of the current research microalgae lipids was also demonstrated by Miao and Wu [3],
on the application of the conventional transesterication scheme who showed that the use of a 100% acid catalysts (on the basis of
on biomass oils has focused on the use of alkaline catalysts and vir- the microalgae oil weight), a molar ratio of methanol to oil of
gin biomass oils with a free fatty acid (FFA) content of <0.5% w/w 56:1, a temperature of 30 C and a reaction time of 4 h, gave the
(based on oil weight). However, in considering the production of best biodiesel yields within the experimental conditions tested.
biodiesel from microalgae, the use of the alkaline catalysed transe- The commercial production of biodiesel from microalgae still
sterication technology would not be suitable, due to the charac- faces hurdles however, mainly due to the high costs associated
teristically high FFA content of microalgae lipids. This is because with the present biomass production and fuel conversion routes.
the use of alkaline catalysts with high FFA containing oils would Various schemes to reduce the cost requirement, in order to im-
result in a partial saponication reaction, leading to the soap for- prove the competiveness of this biomass source, are currently
mation [4] and difculties in the biodiesel separation and purica- being investigated. These include research into improvements in
tion downstream. The use of inorganic acids, such as sulphuric and the biomass productivity and yield, photo-bioreactor optimisation
hydrochloric acids, as reaction catalysts have therefore been con- and alternative microalgae-biodiesel conversion technologies.
sidered for microalgae lipid transesterication, due to its insensi- One of the options for microalgae lipid conversion to biodiesel is
tivity to the FFA content of this oil feedstock, since both biodiesel the supercritical transesterication process, which is facilitated at
producing transesterication and esterication reactions are facil- high reaction temperatures and pressures (>240 C and >8 MPa,
itated via acidic catalysis. Nagle and Lemke [2] examined the use of respectively) without the requirement of catalysts [5]. This tech-
the two groups of catalysts for the conversion of microalgae oil: nology has been demonstrated to have the shortest reaction times
to reach equilibrium conversion of the oils to biodiesel (120240 s)
* Corresponding author. Tel.: +64 3 4797812; fax: +64 3 4790964.
as well as higher fuel product yields compared to the conventional
E-mail address: zhifa@physics.otago.ac.nz (Z.F. Sun). process, and a simplied fuel purication step [6]. However, this

0016-2361/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.fuel.2009.10.011
678 E.A. Ehimen et al. / Fuel 89 (2010) 677684

method is currently disadvantaged due to the adverse process eco- ing information on the operating conditions that provide the best
nomics as well as safety concerns related to the reaction conditions yield while also having the least material and energy requirements,
[7]. The supercritical method is not explored further in this paper, and consequently lowest process costs. In addition, the effect of the
as the authors are not primarily concerned with the application of moisture content of the microalgae biomass on the conversion pro-
this scheme for reducing microalgae biodiesel production costs. cess will be studied since this is of importance energetically and
Another alternative to the conventional process, which is con- cost-wise, especially in regions where drying cannot be achieved
sidered to have potential of reducing the processing units and costs using low cost sun drying or other solar heating methods. This is
of the fuel conversion process, is the in situ transesterication especially important since biomass drying has been demonstrated
method, the main focus of this paper. The in situ process facilitates to be one of the most important economic steps in the microalgae
the conversion of the biomass oil to FAAE directly from the oil production process, reported to account for up to 30% of the total
bearing biomass, thereby eliminating the solvent extraction step production costs [14]. With the microalgae biomass containing
required to obtain the oil feedstock as in the conventional method. up to 80% moisture content after the harvesting process (for exam-
This biodiesel production scheme could therefore aid in the simpli- ple, via centrifugation), this study therefore also aims to evaluate
cation of the fuel conversion process, potentially reducing the the extent to which moisture containing microalgae biomass can
overall process cost, hence lowering the nal fuel product costs be utilised for the in situ transesterication process in order to re-
[8] as well. This method may be especially advantageous for use duce the biomass drying requirements.
with microalgae, since the extraction of microalgae lipids is usually
accomplished via solvent extraction and not with the use of cheap-
er physical extraction methods (for example, expellers) as utilised
for conventional oil crops. The alcoholysis of the oil in the biomass 2. Materials and methods
directly has been shown to result in increased biodiesel yields,
compared to the conventional route [9]. Process wastes and pollu- 2.1. Biomass production and microalgae oil characterisation
tion could also be reduced [8] by this method. This paper aims to
provide useful preliminary information on the application of the Local strains of the microalgae species Chlorella obtained from
in situ transesterication process for the production of biodiesel the algal collection of Landcare Research, Nelson, New Zealand
using microalgae. Due to the FFA content of microalgae lipids, acid were used. The microalgae was cultivated as in Ref. [15] using
catalysis is considered for this conversion scheme. BG 11 media consisting of 1.49 g l1 NaNO3, 74.9 mg l1 MgSO4,
The application of the in situ transesterication process using 36.0 mg l1 CaCl22H2O, 6.0 mg l1 citric acid, 11.2 lg l1 Na-EDTA
homogenous acid catalysts for biodiesel production from oil bear- (pH 8.0, 0.25 M), 2.86 mg l1 H3BO3, 1.81 mg l1 MnCl24H2O,
ing biomass is not a novel one. The method was rst demonstrated 0.22 mg l1 ZnSO47H2O, 0.39 mg l1 Na2MoO42H2O, 79.0 lg l1
by Harrington and DArcy-Evans [9,10] with sunower seeds as CuSO45H2O, 49.4 lg l1 Co(NO3)26H2O, 1.0 ml l1 Fe ammonium
feedstock. Using the in situ method these authors achieved an in- citrate (1000x), 1.0 ml l1 Na2CO3 (0.19 M), and 1.0 ml l1 K2HPO4
crease in biodiesel yields of up to 20% compared to the conven- (0.175 M). The Chlorella biomass was grown autotrophically at
tional process. This improvement in the biodiesel yields was 25 C (1 C) in 5 l asks in an environmental growth chamber
considered by these authors to be attributable to the improved with a light intensity of 80.15 lmol m2 s1 and a 16 h photope-
accessibility of the oil in the biomass by the acidic medium [10]. riod per day. Aeration and agitation of the cultures investigated
The in situ transesterication of macerated sunower seeds was were achieved by bubbling with air. The biomass sample was then
also studied by Siler-Marinkovic and Tomasevic [11] who investi- harvested by centrifugation and dried to a constant weight in a
gated two temperature levels of 30 C and 64.5 C and a range of forced draft oven at 80 C. Dried microalgae biomass was ground
test reaction conditions: the alcohol (methanol) to oil molar ratio in a mortar and its oil content obtained using the extraction meth-
varied from 100:1 to 300:1, the sulphuric acid catalysts concentra- ods described by Zhu et al. [16]. After re-extraction with hexane,
tion ranged from 16% to 100% (on the basis of the oil) and a reac- followed by vacuum evaporation and drying with anhydrous so-
tion time of 14 h. Under the conditions studied, the best FAME dium sulphate, the extracted oil was weighed and the total oil con-
yields (98.2%) based on the oil content of the sunower seeds tent per dry mass of algal biomass was determined.
was obtained at a molar ratio of methanol to oil of 300:1, an acid The specic gravity (SG) of the extracted microalgae oil at 25 C
catalyst concentration of 100% and a reaction time of 1 h [11]. was determined using a pycnometer using the AOCS ofcial meth-
The use of the acid in situ transesterication scheme has also been od Cc10a-25 [17]. The fatty acid composition of the Chlorella oil
demonstrated for FAAE production from soybean and high FFA was determined via gas chromatography (GC) to provide informa-
containing rice bran oil [12,13]. tion on the concentration of the component fatty acids and aid in
Due to expected process differences arising from biomass mac- the estimation of the molecular mass of the oil sample. The fatty
eration, as well as the perceived differences in oil solubility and acid methyl esters for GC analysis were prepared via saponication
conversion, the ndings from these studies cannot be applied to followed by boron triuoride/methanol esterication using meth-
the in situ process for microalgae biomass. Consequently informa- ods described in Ref. [18]. The samples were then run on an Agilent
tion is required on the effects that important reaction variables 6890A GC system equipped with a 7683 series auto sampler and a
have on the in situ microalgae transesterication process. Since ame ionisation detector (FID). Separation was performed on a

the main driver for the use of the in situ transesterication process BPX70 column (0.32 mm ID, 50 m length and 0.25 lm lm thick-
is its possible application as a cost reducing scheme for microalgae ness, SGE) with hydrogen as the carrier gas (ow rate-2 ml/min).
biodiesel, this paper considers primarily reaction variables that The temperatures were as follows: injector, 250 C; detector,
might potentially inuence the process costs of the in situ process. 250 C; oven, 205 C (programmed to start at 35 C, held at this
Accordingly this investigation includes the following inuences on temperature for 5 min and heated at a rate of 2.5 C/min to
the in situ transesterication of microalgae lipids: (i) the reacting 205 C). The reference FAME standard was FAMQ005 by AccuStan-
alcohol volume (the greater the alcohol volume, the more the dard (USA), diluted 1:10 in dichloromethane. The resulting data

material costs input), (ii) the temperature (process energy require- was then analysed using Chemstation software by Agilent.
ment increasing with temperature increases), (iii) the reaction The acid value as well as the free fatty acid content of the mic-
time, and (iv) process mixing (energy requirement for the reaction roalgae oil was titrimetrically determined using the AOCS ofcial
agitation). This study has been carried out with the aim of provid- methods [17].
E.A. Ehimen et al. / Fuel 89 (2010) 677684 679

2.2. The in situ transesterication set up showing a relationship between the recorded SG and FAME con-
version was then obtained.
The in situ transesterication process was carried out using
dried microalgae, and various investigated volumes of the reacting 2.3. Effect of alcohol volume and temperature on in situ
alcohol (methanol) containing sulphuric acid as the transesterica- transesterication
tion catalyst. A xed sulphuric acid concentration of 0.04 mol,
which relates to a 100% acid catalyst concentration (on the basis HPLC grade methanol (99.9% purity) was used as the reacting
of the microalgae oil content mass), was used throughout this alcohol in this study. Chlorella powder (15 g) was mixed with var-
study. The vessels containing the reaction mixtures were then ious methanol volumes (20.0, 40.0, 60.0, 80.0 and 100.0 ml) con-
heated and maintained at the temperatures of interest for specied taining 2.2 ml of sulphuric acid (98% purity) in screwed cap
periods. The major in situ transesterication reaction and product reaction vessels as described in Section 2.2. A minimum volume
purication steps used are shown in Fig. 1ae. After the transeste- of 20.0 ml methanol was selected since it was the minimum
rication step (Fig. 1a) the reaction vessel was allowed to stand for amount that facilitated a complete submersion of 15 g of the mic-
1 h to enable its contents to settle. The reaction mixture was l- roalgae powder. The experiment involved heating the reaction
tered and the residues washed twice by re-suspension in methanol mixtures in screwed cap vessels for 8 h, with each trial at one of
(30 ml) for 10 min to recover any traces of FAME product left in the four different temperatures (23, 30, 60 and 90 C) with continuous
residues (Fig. 1b). Water (50 ml) was added to the ltrate, to facil- stirring using a hot plate with a magnetic stirrer. The in situ reac-
itate the separation of the hydrophilic components of the extract, tion was studied at a local average room temperature of 23 C, i.e.
and then transferred to a separating funnel (Fig. 1c). Further the process was not heated externally. Due to the expected build
extraction of the FAME product was achieved by extracting three up of pressure in the reaction vessels when investigating temper-
times for 15 min using 30 ml of hexane (Fig. 1c). The pooled hex- atures greater than the boiling point of methanol (65 C) at atmo-
ane extracts were washed with water (to remove left-over traces spheric pressure, the experiments at 90 C were carried out in a
of the acidic catalyst and methanol), separated, and then dried over pressurised environment (3 bar) to ensure the reacting species re-
anhydrous sodium sulphate (Fig. 1d). The FAME product was then mained in a liquid state. The respective FAME products at different
ltered and evaporated to obtain the FAME yields (Fig. 1e). Dupli- investigated variable levels were then obtained as in Section 2.2
cate experiments were carried out for each parameter investigated. and their SGs determined.
The SG of the extracted FAME products was then determined and
compared with that of the microalgae oil to monitor the extent 2.4. Effect of reaction time on in situ transesterication
of the conversion process, as in Ref. [19]. With the forward reaction
resulting in FAME production and the process nearing completion, At each of the four temperature levels, the in situ transesteri-
the SG of the puried reaction product (after the removal of the cation of 15 g microalgae biomass was repeated in duplicate (as
glycerol co-product) is expected to decrease until constant, signify- described in Section 2.2) with reaction times of 0.25, 0.5, 1, 1.5,
ing an equilibrium conversion of the microalgae lipids to the 2, 4, 8 and 12 h with 60 ml methanol containing 2.2 ml sulphuric
methyl esters. The terms FAME product and biodiesel are used acid. This was carried out to provide a greater insight on the pro-
interchangeably in this paper to represent the mixture of FAME gression of the transesterication process with time with respect
and any remaining unreacted glycerides after the glycerol purica- to the various investigated reaction temperatures. The purication
tion step. of the FAME product and its SG determination was carried out as
To further elucidate the extent of oil to FAME conversion by described above.
the in situ transesterication process, a more quantitative rela-
tionship between the measured SG and the percentage FAME 2.5. Effect of moisture content of microalgae biomass on in situ
conversion was made by subjecting selected puried FAME prod- transesterication
ucts, obtained for different experimental levels (as described in
Section 2.3) with dissimilar measured SGs, to the GC analysis Chlorella biomass, initially dried in a draft oven at 80 C to a con-
as described in Section 2.1. The concentration of the FAME stant mass, was designated to represent a dry (0.0% moisture)
species for each sample was acquired and a calibration curve sample, and was used as the basis for calculating the moisture

Fig. 1. Schematic representation of the in situ transesterication steps used in this study.
680 E.A. Ehimen et al. / Fuel 89 (2010) 677684

content of the microalgae samples. To study the effect of moisture ment, transesterication was carried out as before using 15 g bio-
on the in situ transesterication process, freshly harvested Chlorella mass with 60 ml of methanol containing 0.04 mol sulphuric acid
paste was air dried (fan aided) at 26 C (1 C) to obtain samples with a reaction time of 8 h and a temperature of 60 C. The reaction
with biomass moisture contents of 72.5%, 56.0%, 40.9%, 31.7%, products were puried and the SGs determined.
19.5% and 8.7% (based on the dry sample weight). Using a draft
oven at 80 C, samples of 3.2%, 1.4% and 0.7% moisture content
3. Results and discussion
(dry basis) were also obtained. The in situ transesterication pro-
cess was then investigated in duplicate for different moisture lev-
3.1. Lipid content and SG of pure oil
els on a basis of a 15 g dry reacting sample (for example, 25.88,
23.40 and 17.95 g were used for the 72.5%, 56.0% and 19.5% mois-
With the culture conditions used in this study, the Chlorella
ture levels, respectively) at varied reaction times up to 6 h using a
samples were determined to have a total transesteriable lipid
xed reaction temperature of 60 C and a methanol volume of
fraction of 0.276 g of microalgae oil/g of Chlorella biomass. It must
60 ml containing 0.04 mol of sulphuric acid. The purication of
however be noted that the biomass oil content observed in this
the product was carried out and the respective SGs determined.
study is not only inuenced by the microalgae specie, but is also
highly dependent on the specic growth conditions. As mentioned
2.6. Effect of stirring on in situ transesterication
in Section 2.1, increases in the oil content of the biomass, and
hence increases in the biodiesel production potential of the micro-
To investigate the effect of stirring, the reaction vessels used for
algae feedstock, can be achieved by manipulating the microalgae
the in situ transesterication process were subjected to four differ-
culture conditions, such as the culture nutrients and light intensity
ent stirring treatments: (1) continuously stirred, (2) stirred for only
[20].
1 h, (3) stirred intermittently, 1 h on and 1 h off, and (4) not stirred
The SG of the extracted oil, which was used to characterise the
throughout the experiment. The reaction stirring was carried out
reacting oil at the start of the transesterication reaction, was
using a magnetic stirrer system with a rotation speed of 500 rpm
determined to be 0.914 at 25 C (0.1 C). With all the constituent
kept constant throughout the duration for the agitated samples.
microalgal oils converted to their respective FAME using the
This speed was used since it was observed to facilitate a complete
saponication and transesterication methods described in Sec-
suspension of the particles in the reaction vessels. For each treat-
tion 2.1, the results for the percentage principal fatty acids of the
extracted microalgae oil, as elucidated via GC analysis of the result-
ing FAME mixture, are shown in Table 1. This data was used to
Table 1 determine the average molecular mass of the Chlorella oil. Fatty
Fatty acid composition of the Chlorella biomass. acids detected only in trace amounts (<0.1%) were, however, not
Fatty Molecular mass (g/ % in Molecular mass contribution
included in these results. Since the microalgae oil (assumed to be
acid mol) (MMFA) sample (g/mol) (MMc) primarily composed of triglycerides) consists of different fatty
acids, their respective contributions to the overall molecular mass
C16:0 256.42 4.37 11.21
C16:1 254.41 0.44 1.12 of the microalgae lipid is used (as seen in the MMc column in Ta-
C18:1 282.46 61.81 174.59 ble 1) to estimate the mean molecular mass of the constituent lipid
C18:2 280.45 19.94 55.92 fatty acids (MMFA). With the formation of the triglyceride molecule
C18:3 278.43 12.22 34.02
facilitated by the combination of fatty acid molecules and a mole-
C20:1 310.52 1.22 3.79
Average molecular mass of constituent 280.65 cule of glycerol with the condensation of three molecules of water,
fatty acids (MMFA) the average molecular mass of the microalgae oil (MM oil) can be
calculated using Eq. (1).

0.92

0.915

0.91
SG of purified FAME product

0.905

0.9

0.895

0.89

0.885

0.88
0 10 20 30 40 50 60 70 80 90 100
% FAME conversion

Fig. 2. Calibration curve showing the relationship the measured SG and the % FAME conversion obtained using GC.
E.A. Ehimen et al. / Fuel 89 (2010) 677684 681

MMoil 3MMFA MMglycerol   3MMwater 1 investigated reacting methanol volumes and temperatures, using
a xed reaction time of 8 h and a xed acidic catalyst molar
where, MMglycerol and MMwater represent the molecular masses of
concentration (0.04 mol sulphuric acid), was determined. This is
glycerol and water, respectively. The average molecular weight of shown in Fig 3. Using the SG of the extracted oil (0.914) to indicate
the Chlorella lipid was calculated to be 880 g/mol.
the start of the transesterication reaction, the conversion of the
As described in Section 2.1, following the purication of the constituent microalgae oils to biodiesel can therefore be moni-
FAME products and measurement of the SG for different FAME
tored, with lower asymptotic SG values of the puried transesteri-
yields, the concentration of the FAME species in the fuel product cation products indicating improved equilibrium FAME
were obtained by GC analysis and used to quantify the percentage
production. Eqs. (2)(4) show that the transesterication reaction
of Chlorella oil converted to FAME products. The calibration curve consists of a series of reversible steps in which the oils (repre-
obtained, showing the relationship between the measured SG
sented as triglycerides) are sequentially converted to di- and
and the percentage FAME conversion, is shown in Fig. 2. The curve mono-glycerides, and eventually to glycerol, with one mole of fatty
can therefore be used to predict the corresponding percentage
acid alkyl esters (RCOOR1) liberated at each step.
FAME conversions in this study since it covers a range from a 0%
conversion using the pure extracted oil (SG, 0.914) to 92.22% which Trigyceride R1 OH $ diglyceride RCOOR1 2
was for the lowest measured SG (0.8826).
The acid value of the microalgae oil was determined to be
Diglyceride R1 OH $ monoglyceride RCOOR1 3
10.21 mg KOH/g Chlorella oil. Using the estimated molecular mass
of 280.65 for the constituent fatty acids, the FFA content of the
microalgae oil was determined to be 5.11% (on the basis of the Monogyceride R1 OH $ glycerol RCOOR1 4
oil weight). Due to the high FFA content (>0.5% w/w) of the micro- 1
With methanol as the reacting alcohol (R OH), the equilibrium
algae oil, the choice of acidic over alkaline catalysts for the in situ transesterication products range from a mixture of FAME, glyce-
transesterication process is justied. rides and glycerol (when the reaction is incomplete) to a FAME
and glycerol mixture (on completion). Since the heavier glycerol
3.2. Effect of alcohol volume and temperature product of the reaction is eliminated from the nal product via
purication steps, the SG of the puried FAME product (containing
Using the average molecular mass of the Chlorella oil, as deter- glycerides) as expected will be less than that of the extracted oil,
mined in Section 3.1, and with the knowledge of the transesteri- and would vary with the respective percentage composition of
cation stoichiometry involving 3 mol of reacting alcohol per mole the FAME specie in the puried product. For example, an increase
of oil, the methanol volumes investigated in this study represent in the lighter FAME concentration would result in lower measured
a reacting molar alcohol to oil ratio range of 105:1524:1 (calcu- product SGs.
lated with a methanol density of 0.7918 g/cm3). This range includes The results obtained indicate an improvement of the microalgae
and exceeds that of a similar investigation of the in situ transeste- oil conversion to FAME with increasing temperature and increasing
rication of sunower oil by Siler-Marinkovic and Tomasevic alcohol volume, with the lowest FAME equilibrium conversions ob-
[11], described in Section 1. The SG of the FAME products for the served with the reacting molar ratios of the methanol to oil at

20 ml methanol 40 ml methanol 60 ml methanol 80 ml methanol 100 ml methanol


0.8895

0.8885
Specific gravity (SG) of the extracted FAME product

0.8875

0.8865

0.8855

0.8845

0.8835

0.8825

0.8815
15 25 35 45 55 65 75 85 95

In-situ transesterification temperature (C)

Fig. 3. Effect of the alcohol volume and reaction temperature on the specic gravity of the FAME products after a reaction time of 8 h using 15 g microalgae biomass and 100%
acidic catalyst (on basis of oil).
682 E.A. Ehimen et al. / Fuel 89 (2010) 677684

105:1 (that is, a methanol volume of 20 ml) for all the temperature 4 h for temperatures of 60 and 90 C. Although faster conversion
levels studied. However, with the use of alcohol volumes over rates can be observed by use of reaction temperatures greater than
60 ml (i.e. a reacting molar ratio of alcohol to microalgae oil greater the boiling point of the reacting methanol (for example, 90 C), the
than 315:1) for the in situ transesterication of 15 g of microalgae process heating and pressure requirements may inhibit the use of
biomass, no signicant trends were observed for the measured SGs such temperature levels. The use of a reaction temperature of 60 C
determined for the reaction temperature levels of 60 and 90 C. may therefore prove more benecial, if we consider the total en-
ergy consumption and operation cost of the whole biodiesel con-
3.3. Effect of reaction time and temperature version system.

To investigate the inuence of reaction time and temperature, a 3.4. Effect of the moisture content of the microalgae biomass
methanol volume of 60 ml was used since it was found (Section 3.2)
that no appreciable differences in the equilibrium FAME conver- As discussed in Section 1, the inuence of the microalgae mois-
sion were obtained with the use of higher alcohol volumes for ture levels is important in this study due to its potential impact on
reactions carried out at 60 and 90 C. Fig. 4 shows the progression the overall biomass and fuel production costs. The different mois-
of the microalgae oil to biodiesel conversion process with time and ture content levels of the samples investigated were used to repre-
at different temperature levels, using the measured SG of the puri- sent biomass samples, ranging from freshly harvested (after
ed FAME product as a conversion indicator. For the samples centrifugation) with moisture content of 72.5% to completely
investigated at room temperature (no process heating), asymptotic dried. It must be noted, however, that the water content of the
FAME conversion values were not reached within the time bound- dried sample (0.0%), used as the basis for determining the moisture
aries of this study. When the in situ transesterication process was content of the microalgae samples investigated, refers to free bio-
carried out at 90 C under the same experimental conditions the mass water, not the more tightly bound water. The measured SG of
highest equilibrium conversion levels (lowest FAME SG) were the puried FAME yields obtained from Chlorella biomass with dif-
reached within the shortest time, with 70% and 92% of the overall ferent moisture content levels at different reaction time intervals
equilibrium FAME conversion values attained after a reaction time are shown in Fig. 5. The general trends of the results obtained indi-
of 15 min and 1 h, respectively. Using a reacting molar ratio of cate that increases in the biomass moisture content have a signif-
methanol to oil of 315:1, and continuously stirring the reaction icant negative effect on the equilibrium conversion of oil to
at 500 rpm, a faster conversion phase was observed between biodiesel using the acid-catalysed in situ transesterication pro-
15 min and 2 h for the 15 g microalgae samples investigated at cess. The oil to FAME conversion was observed to be similar, and
23 and 30 C, as shown in Fig. 4. This phase is considered to be almost inhibited, for the microalgae samples with moisture con-
due to the emulsifying effects of the initial fatty acid esters formed tents greater than 31.7% (dry basis) i.e. containing 4.755 g of water
after the mass transfer controlled phase, as discussed in Ref. [21]. per sample, or 114.85% w/w when compared with the microalgal
This effect was however not signicantly observed for the 60 and oil weight of 4.14 g. Hence, for the conditions investigated, these
90 C experiment runs, probably due to the fact that the elevated results indicate that the process would be inhibited in moisture
temperatures (and pressures) improved the initial miscibility of containing microalgae samples with water levels greater than
the reacting species, leading to a signicant reduction in the reac- 115% of the reacting oil weight for the acid-catalysed in situ transe-
tion times, as observed in Fig. 3. Within the experimental condi- sterication process. Obvious differences were only observed after
tions investigated, equilibrium FAME conversions were observed a 73% removal of water from the freshly harvested samples, i.e.
to reach similar asymptotic values after a reaction time of 2 and with the in situ transesterication of the 19.5% (dry basis) samples.

0.915

0.91
Specific gravity (SG) of extracted FAME product

0.905 23C

30C

60C
0.9 90C

0.895

0.89

0.885

0.88
0 2 4 6 8 10 12

In-situ transesterification time (h)

Fig. 4. Inuence of varying the reaction time on the in situ transesterication of 15 g microalgae at reaction temperatures of 23, 30, 60 and 90 C using 60 ml methanol and
100% acidic catalyst (on basis of oil).
E.A. Ehimen et al. / Fuel 89 (2010) 677684 683

72.5% 56.0% 40.9% 31.7% 19.5% 8.7% 3.2% 1.4% 0.7% 0.0%
0.915

Specific gravity (SG) of the extracted FAME product


0.91

0.905

0.9

0.895

0.89

0.885

0.88
0 1 2 3 4 5 6 7
In-situ transesterification time (h)

Fig. 5. Inuence of microalgae moisture content (dry basis) on the FAME conversion using 100% acid catalyst (on basis of oil) and a xed reaction alcohol volume of 60 ml and
temperature of 60 C, respectively.

A reduction of the biomass moisture levels to 0.7% (dry basis), methyl esters acting as a dual solvent for the oil and alcohol, there-
which relates to a reacting water mass content of 0.1 g, or 2.4% by facilitating further FAME production. However, after a reaction
w/w (reacting oil basis), resulted in equilibrium conversions of time of 8 h, the samples stirred for 1 h only were observed to
81.7% of the values obtained with completely dried samples after achieve only 91.3% of the equilibrium FAME conversion achieved
a reaction time of 6 h. Biomass drying cannot therefore be over- by the samples which were continuously stirred. Furthermore,
looked in the in situ transesterication process, with improved intermittent stirring at 1 h intervals achieved equilibrium FAME
equilibrium conversions observed when using completely dried yields which were close to, but nevertheless lower than, those ob-
samples compared with moisture containing samples. served in the vessels stirred continuously.
It should be noted that the FAME product SG for the trial with-
out stirring might have had a higher SG than indicated in Table 2 in
3.5. Effect of stirring on in situ transesterication
the absence of any stirring at all, since all the samples were initially
swirled by hand before being placed on the hot plate and stirring
The effect of stirring on the in situ process was investigated as a
system. The purpose of this action was to prevent clumping and
potential process cost saving strategy. When the in situ transeste-
ensure that the biomass samples were adequately exposed to the
rication process was conducted without stirring, the equilibrium
acidic methanol mixture. However, it could have also promoted
conversion of the microalgae oil content to biodiesel, compared to
some initial phase mixing of the reactants in the not stirred sam-
that for the continuously stirred sample, was signicantly reduced
ple. The results in Table 2 may therefore overstate the equilibrium
(Table 2). This indicates that stirring is required to some extent to
conversion yields of the unstirred samples.
enhance the reaction progress, evidently by aiding the initial mis-
cibility of the reacting species. The stirring of the reaction vessels
for only 1 h (from the start of the reaction) was investigated. This
time was chosen because 87% of the equilibrium FAME conversion 4. Conclusion
was achieved after 1 h with continuous stirring under the experi-
mental transesterication conditions, using 15 g microalgae bio- This study investigated the inuence of potentially important
mass with 60 ml methanol (containing 0.04 mol acid catalyst), at economic reaction factors on the progress of the conversion of mic-
a reaction temperature of 60 C. It was therefore envisaged that, roalgae oil to biodiesel using the acid-catalysed in situ transesteri-
without further stirring, the equilibrium conversion might ap- cation process. These variables include the inuence of reacting
proach that of the continuously stirred system with the produced alcohol volumes, temperature, reaction time, biomass moisture
content and stirring. The results show that increases in the reaction
temperature and alcohol volume favour the production of biodie-
Table 2 sel. However, within the experimental conditions, an increase in
Inuence of stirring on the in situ transesterication of microalgae lipids with a xed the reacting alcohol volume of more than 60 ml did not show sig-
acid catalyst concentration (100%, on basis of oil), methanol volume of 60 ml and
nicant changes in the yield of the FAME products obtained, when
reaction temperature at 60 C.
using 15 g of Chlorella biomass. The application of the in situ acid-
SG of extracted FAME product catalysed process has been discussed in the literature [9] to offer
No stirring 0.9032 conversion rate improvements when compared with the conven-
Stirring for 1 h 0.8859 tional process. However, the energetic cost of the recovery of the
Stirred intermittently (1 h off, 1 h on) 0.8845
excess alcohol reactants in this process might, as well as an in-
Stirred continuously 0.8831
crease in downstream FAME purication requirements, might limit
684 E.A. Ehimen et al. / Fuel 89 (2010) 677684

the achievable cost reduction, one of the main reasons which References
would favour the use of the in situ transesterication method.
Biodiesel yields by the in situ transesterication of microalgae [1] Chisti Y. Biodiesel from microalgae. Biotechnol Adv 2007;25:294306.
[2] Nagle N, Lemke P. Production of methyl ester fuels from microalgae. Appl
were shown to improve signicantly with process stirring. How- Biochem Biotechnol 1990;24/25:35561.
ever, some savings may be achieved in the stirring energy since [3] Miao X, Wu Q. Biodiesel production from heterotrophic microalgal oil.
the reactors can be stirred intermittently without a major reduc- Bioresour Technol 2006;97:8416.
[4] Al-Zuhair S. Production of biodiesel: possibilities and challenges. Biofuels
tion in the yield. Biomass drying was observed to play an impor- Bioprod Biorening 2007;1:5766.
tant role, with an increase in the moisture content of the [5] Demibras A. Production of biodiesel from algae oil. Energy Sources Part A
biomass resulting in signicant reductions of the equilibrium Recov Util Environ Eff 2009;31:16974.
[6] Saka S, Kusidana D. Biodiesel fuel from rapeseed oil as prepared in supercritical
FAME conversion yields. Therefore extensive biomass drying may
methanol. Fuel 2001;80:22531.
be needed prior to biodiesel production by the in situ process. [7] Marchetti JM, Errazu AF. Techno-economic study of supercritical biodiesel
Although not covered in this study, continuous removal of the production plant. Energy Conv Manag 2008;49:21604.
water product from the reactor, as well as an increase in the acidic [8] Haas MJ, Scott KM, Foglia TA, Marmer WN. The general applicability of in situ
transesterication for the production of fatty acid esters from a variety of
catalyst concentration, are schemes which may be utilised to drive feedstocks. J Am Oil Chem Soc 2007;84:96370.
the forward reaction when microalgae biomass with high moisture [9] Harrington KJ, DArcy-Evans C. A comparison of conventional and in situ
content is used in the in situ transesterication process. methods of transesterication of seed oil from a series of sunower cultivars. J
Am Oil Chem Soc 1985;62:100913.
In summary, the production of biodiesel from microalgae is still [10] Harrington KJ, DArcy-Evans C. Transesterication in situ of sunower seed oil.
in the research and development stages, with laboratory scale Ind Eng Chem Prod Res Dev 1985;24:3148.
transesterication trials (in situ, in this study) carried out aimed [11] Siler-Marinkovic S, Tomasevic A. Transesterication of sunower oil in situ.
Fuel 1998;77:138991.
at providing useful information to aid the design and modelling [12] Kildiran G, zgl-Ycel S, Trkay S. In-situ alcoholysis of soybean oil. J Am Oil
of industrial sized biodiesel production processes. Further process Chem Soc 1996;73:2258.
design, optimisation and integration investigations on unit opera- [13] zgl-Ycel S, Trkay S. Variables affecting the yields of the methyl esters
derived from the in-situ esterication of rice bran oil. J Am Oil Chem Soc
tions such as biomass drying, ltration, evaporation, extraction, 2002;79:6114.
adsorption, as well as the chemical processing steps are still re- [14] Becker EW. Microalgae: biotechnology and microbiology. Cambridge studies in
quired to optimise the use of this biomass feedstock. biotechnology. Cambridge: Cambridge University Press; 1994. 293pp.
[15] Ehimen EA, Carrington CG, Sun ZF, Eaton-Rye J. Biodiesel from microalgae:
investigating reaction conditions for in-situ transesterication. In: Proceedings
of the 2nd international symposium on energy from biomass and waste,
Venice, November 2008.
Acknowledgements [16] Zhu M, Zhou PP, Yu LJ. Extraction of lipids from Mortierella alpine and
enrichment of arachidonic acid from the fungal lipids. Bioresour Technol
2002;84:935.
The authors are most grateful to Assoc Prof. Julian Eaton-Rye [17] AOCS. The ofcial methods and recommended practices of the AOCS. 5th ed.
and Jackie Shand of the Photosystem II lab, Biochemistry Depart- American Oil Chemists Society, 2005.
[18] Van Wijngaarden D. Modied rapid preparation of fatty acid esters from lipids
ment, as well as Dr John Birch and Michelle Leus of the Food Sci- for gas chromatographic analysis. Anal Chem 1967;39:8489.
ence Department, University of Otago, Dunedin for the use of [19] Al-Widyan MI, Al-Shyoukh AO. Experimental evaluation of the
their facilities, instruction and assistance. We wish to also thank transesterication of waste palm oil into biodiesel. Bioresour Technol
2002;85:2536.
Dr Phil Novis of Landcare Research for providing the Chlorella cul- [20] Qiang H, Sommerfeld M, Jarvis E, Ghiradi M, Posewitz M, Siebert M, Darzins A.
ture used in this study and Mr & Mrs P. Ehimen for their kind nan- Microalgal triacyglycerols as feedstock for biofuel production; perspectives
cial support. The authors would also like to acknowledge the and advances. Plant J 2008;54:62139.
[21] Freedman B, Buttereld RO, Pryde EH. Transesterication kinetics of soybean
anonymous reviewers whose useful comments and criticisms were oil. J Am Oil Chem Soc 1986;63:137580.
most helpful in editing this paper.

You might also like