You are on page 1of 6

Renewable Energy 58 (2013) 15e20

Contents lists available at SciVerse ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Biodiesel purication using micro and ultraltration membranes


Magno Jos Alves, Suellen Mendona Nascimento, Iara Gomes Pereira, Maria Ins Martins,
Vicelma Luiz Cardoso, Miria Reis*
Universidade Federal de Uberlndia, Faculdade de Engenharia Qumica, Av. Joo Naves de vila, 2121, 38400-902 Uberlndia, MG, Brazil

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 21 May 2012
Accepted 26 February 2013
Available online 30 March 2013

Biodiesel is an environmental-friendly fuel that can replace petroleum diesel. However, after transesterication reaction of vegetable oils, the obtained crude biodiesel must be puried. The commonly
applied purication step is water washing. This step is a concern in biodiesel production, since large
quantities of clean water are used, generating a wastewater stream to be further treated. Here we propose the application of micro and ultraltration processes to purify crude biodiesel. Crude biodiesel was
ltrated in a dead-end process at different transmembrane pressures and using membranes of different
pore sizes. Flux results showed that greater transmembrane pressures, as well as greater pore sizes,
enable greater uxes. Density, viscosity and acid values of puried biodiesel (washed and ltrated) are in
accordance to the international legislation for biodiesel quality. Both processes (water washing and
membrane separation) were able to reduce the amount of soap detected in crude biodiesel. However, the
proposed microltration membranes were not efcient as the washing method to reduce the free
glycerol content. The ultraltration membrane of 30 kDa was also not able to produce a puried biodiesel
according to the international legislation for free glycerol content. Between the analyzed membranes, the
glycerol content limit (less than 0.02 wt%) was achieved only with the ultraltration membrane of
10 kDa. Water addition in the crude biodiesel improved the glycerol removal by membrane ltration. The
obtained results showed that the membrane separation process is a suitable alternative for biodiesel
purication.
2013 Elsevier Ltd. All rights reserved.

Keywords:
Biodiesel
Purication
Membrane

1. Introduction
Energetic and environmental issues have encouraged the
development of alternative fuel sources. Biodiesel is a potential fuel
for diesel engines since it is obtained of renewable sources and
combustion emissions for biodiesel are lower than for petroleum
diesel [1]. Biodiesel can be derived from vegetable oils, animal fats
or recycled greases. Choice of the raw material to be used in biodiesel production is both a process chemistry decision and an
economic decision [1]. In Brazil, soybean oil is largely applied for
biodiesel production [2].
Biodiesel is commonly produced by the transesterication reaction. In this reaction, a triglyceride reacts with an alcohol of short
chain (methanol or ethanol) in the presence of a catalyst to produce
a mixture of fatty esters (biodiesel) and glycerol. Although the 3:1
mole ratio between methanol and oil is required, extra alcohol is
added to the reactor in order to drive the reaction closer to higher
yields. Using vegetable oils as a feedstock, base catalyst systems are
* Corresponding author.
E-mail addresses: miria@feq.ufu.br, miriareis@hotmail.com (M. Reis).
0960-1481/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.renene.2013.02.035

mainly used. The most commonly used catalyst materials for converting triglycerides to biodiesel are sodium hydroxide, potassium
hydroxide, and sodium methoxide [3]. Typical catalyst loadings
range from 0.3% to about 1.5% [1]. The operating reactor temperature is usually about 60  C, although temperatures from 25  C to
85  C have been reported [1]. After the reaction, the reaction
mixture is allowed to settle in a decanter to give a separation of
ester and glycerol phases. Alcohol is removed from both the glycerol and ester stream using an evaporator or a ash unit. After these
steps, the ester phase still contains impurities such as free
dispersed glycerol particles, soap, and trace amounts of residual
catalyst and mono, di, and triglycerides. This crude biodiesel must
be treated in order to attempt standard specications that dene
the quality of biodiesel (e.g., ASTM D6751 [4] and EN 14214 [5]).
The conventional purication process of the ester phase is water
washing, being applied to remove the residual glycerol, base catalyst, and any soap formed during the reaction [6]. However, this
process generates a large amount of highly polluting wastewater,
causing environmental disposal problems.
Membrane technology can be applied for biodiesel purication
in order to avoid the water washing step. Atadashi et al. [6] stated

16

M.J. Alves et al. / Renewable Energy 58 (2013) 15e20

that the introduction of membrane technology can minimize the


difculties encountered in the separation and purication of biodiesel. However, this technology needs to be thoroughly explored
and exploited to determine its potential applications for the separation and purication of the biodiesel product mixture.
The principle of membrane separation of ester and free glycerol
is depicted in Fig. 1. Glycerol molecules must be associated with
other ones in order to increase the particle size. In this way, the
ester phase could be get in the permeate side while glycerol and
other impurities are retained in the feed side [5,6].
Some scientic papers reported the application of a membrane
reactor system to shift the reaction equilibrium to the product side.
Cao et al. [7] observed that the glycerin content in the produced
biodiesel was signicantly lower than that produced via a conventional batch transesterication reaction. Cao et al. [8] investigated the effect of membrane pore size on the performance of a
membrane reactor for biodiesel production. Baroutian et al. [9]
showed that a TiO2/Al2O3 membrane reactor was able to produce
a biodiesel with characteristics according to the ASTM standard
limits. Cheng et al. [10] simulated a membrane reactor integrated
with a prereactor and showed that this system can be used to
separate the unreacted emulsied oil from the product stream.
He et al. [11] compared membrane extraction and traditional
extraction methods for biodiesel purication. In the membrane
method, a hollow ber membrane (1 m long, 1 mm diameter, ow
rate: 0.5 mL/min, operating pressure: 0.1 MPa) was used to purify
crude biodiesel. Two types of hollow ber membranes, polysulfone
and polyacrylonitrile, which are hydrophilic and hydrophobic
respectively, were used. According to the obtained results, He et al.
[11] concluded that membrane process caused less ester loss than
washing or solvent extraction methods. The polysulfone membrane
was more efcient than the polyacrylonitrile one. However, He
et al. [11] did not compare the glycerin removal.
Cheng et al. [12] evaluated the membrane separation process
using a ceramic membrane combined a with liquideliquid extraction process for the continuous cross ow rejection of triglycerides.
Wang et al. [13] investigated rening of biodiesel by ceramic
membrane separation with different pore size membranes and at
different temperatures and transmembrane pressures. According to
Wang et al. [13], the glycerin removal was greater by the membrane
process than by the water washing process.
Saleh et al. [14] investigated the effects of soap, methanol, and
water on glycerol particle size in biodiesel purication, concluding
that the addition of small amounts of water improved the removal
of glycerol from FAME. Gomes et al. [15] showed that the water
concentration to be added to the mixture plays an important role in
glycerol separation, as well as in the permeate ux, according to the

proposed glycerol separation mechanism using a ceramic


membrane.
In this work we propose the comparison of micro and ultraltration membranes for biodiesel purication. The trade-off between permeate ux and free glycerol removal was investigated.
Besides, the effect of water addition was analyzed in order to achieve higher free glycerol removal.
2. Materials and methods
2.1. Biodiesel synthesis
Biodiesel was produced using rened soybean oil and methanol
(99.9% purity, Vetec, Brazil) as reactants and potassium hydroxide
(KOH, reagent grade, Vetec, Brazil) as a catalyst. The transesterication reaction was carried out in a 2 L three-neck roundbottomed ask tted with a thermometer to control the reaction
temperature, a reux condenser to prevent methanol loss by
cooling water, and a mechanical stirrer to keep the reaction under
constant agitation. The ask was supported on a heating mantle to
control the reaction temperature at 60  C. The oil was fed into the
ask and preheated before catalyst (KOH) and methanol addition.
The catalyst was previously dissolved into methanol until complete
dissolution. This solution was also preheated at 60  C. Methanol/oil
molar ratio was 4:1; the reaction time was 1 h; and the amount of
catalyst in relation to the oil mass was 0.75 wt%. These conditions
are well established in the literature for a satisfactory ester yield [6].
The obtained product was transferred to a rotary evaporator
under 600 mmHg vacuum at 90  C for 60 min for methanol and
water removals. This mixture was then placed on a separator funnel
and allowed to settle for 12 h, and the bottom glycerol-rich layer
was removed. A total of 4 runs were performed to generate sufcient quantities of ester-rich phase (called as crude biodiesel) for
the purication experiments.
The crude ester phase was washed with hot distilled water
(50  C) at a 1:1 mass ratio under mild agitation, as suggested by He
et al. [11], until that the pH of the water washing was neutral. This
condition was reached with three washing steps.
2.2. Filtration process to purify crude biodiesel
Besides water washing purication method, the crude biodiesel
was also puried by membrane ltration using at polymeric
membranes. Mixed cellulose ester microltration membranes of
0.22 and 0.30 mm pore size (Millipore, Ireland) and poly(ethersulfone) ultraltration membranes of 10 and 30 kDa nominal

Fig. 1. Scheme for biodiesel (FAME) purication by the membrane separation process.

M.J. Alves et al. / Renewable Energy 58 (2013) 15e20

Manometer

Filtration
module

Permeate
collection
Magnetic stirrer

N 2 gas tank

Digital balance

measured according to AOCS Ofcial Method Cd 3e25 [21]. The


saponication value represents milligrams of KOH required to
saponify 1 g of fat or oil. Water content was measured according to
the EN ISO 12937 [22] by Karl Fischer coulometric titration
(Metrohm).
Free glycerol content was measured by the volumetric method
described by AOCS methodology for the analysis of free glycerol in
oils and fats (Ca 14e56 [23]). Free glycerol refers to the amount of
glycerol that is left in the nished biodiesel. Glycerol is insoluble in
biodiesel so almost all of the glycerol is easily removed by settling.
Some glycerol may remain either as suspended droplets or a very
small amount that is dissolved in the biodiesel, and it is known as
free glycerol.
Although the volumetric method for free glycerol content is
accurate for this determination [24], the chromatography method
described in ASTM-D6584 [25] was employed to verify free glycerol
percent in some biodiesel samples. An Agilent 7890A gas chromatography (GC) with a ame ionization detector (FID) and with a
DB-5HT capillary column (J&W Scientic), capillary column of 15 m
length, 0.32 mm ID, 0.1 mm lm thickness with (5% phenyl)methylpolysiloxane was used.
The yield of biodiesel was calculated as presented in Equation
(1) [26].


Yield%
Fig. 2. Schematic diagram of ltration module.

MWCO (GE Osmonics, USA) were evaluated in this work. Only new
membranes were used in the experiments.
The microltration experiments were carried out at two
different transmembrane pressures, 1 and 2 bar. The ultraltration
tests were performed at 4 bar of transmembrane pressure for both
evaluated ultraltration membranes. These pressures were chosen
based on preliminary tests in which suitable biodiesel uxes were
observed.
The ltrations were carried out in a semi-batch module (Fig. 2)
using membranes with ltration radius of 45 mm. A volume of
500 mL of biodiesel was fed into the membrane module for each
experiment. The pressure was adjusted with N2 gas and monitored
in a manometer. The permeate ux was monitored through the
ltration time. Micro and ultraltrations were carried out during 10
and 60 min, respectively. These times were xed based on the
volume of crude biodiesel fed in the ltration module.
Besides crude biodiesel ltrations, additional tests were carried
out by the adding of low quantities of water to the crude biodiesel
before the ltration. Deionized water was added to the biodiesel
sample at concentrations of 0.1 and 0.2 wt%, as suggested by Saleh
et al. [16]. Deionized water was mixed with the crude biodiesel
using a magnetic stirrer for 1 h at ambient temperature prior to the
ltration run.
2.3. Chemical analysis
Soybean oil and crude, washed and ltrated biodiesel samples
were analyzed for free glycerol content, density at 20  C, kinematic
viscosity at 40  C, acid value, saponication value, amount of soap
and water content.
Kinematic viscosity, density, and acid value, were determined
according to ASTM D445 [17], D1298 [18], and D664 [19], respectively. The acid number is a direct measure of free fatty acids in
biodiesel. It is measured in terms of the quantity of KOH required to
neutralize the sample.
Amount of soap was measured following a modied version of
AOCS method Cc 17e79 [20], soap in oil. Saponication value was

17


mbiodiesel
 100
mraw oil

(1)

where mbiodiesel is the quantity of crude biodiesel obtained after the


transesterication reaction, and mraw oil is the quantity of soybean
oil used as feedstock.
3. Results and discussion
3.1. Transesterication reaction and water washing purication
The used soybean oil had a density of 915 kg m3 at 25  C, kinematic viscosity of 29.3 mm2 s1 at 40  C, acid value of
0.17 mgKOH g1, and saponication value of 162.12 mgKOH g1. These
results show that this feedstock is appropriate to obtain high ester
yields. According to Freedman et al. [27] the feedstock must present
an acid value less than 1 to obtain maximum ester formation by
transesterication of vegetable oils. Sharma and Singh [28]
reviewed the literature and stated that the acid value of the feedstock has to be reduced to less than 2 mgKOH g1 for alkaline
transesterication.
According to Equation (1), the product yield was 93%, similar to
the result reported by Sharma and Singh [28].
After the transesterication reaction and the settling process to
separate ester and glycerin phases, the obtained crude biodiesel
was characterized (Table 1). Crude biodiesel was washed with
water and the characteristics of the washed biodiesel are also
presented in Table 1.
Table 1
Characterization of crude and washed biodiesel.
Parameter

Biodiesel sample
Crude

Washed

Density (kg m3)


Kinematic viscosity (mm2 s1)
Water content (ppm)
Acid value (mgKOH g1)
Saponication value (mgKOH g1)
Amount of soap (gsoap g1sample)a
Free glycerol (%, wt/wt)

880
5.0
525.4
0.08
191.5
2.6  103
0.029

880
5.0
1026.2
0.14
179.2
n. d.
0.007

n. d. not detected.
a
As potassium oleate.

18

M.J. Alves et al. / Renewable Energy 58 (2013) 15e20

Transesterication reaction reduced the feedstock oil density


and viscosity, as expected. Water washing process did not change
density and viscosity values of crude biodiesel.
According to EN 14214, the limit for water content in fatty acid
methyl esters is 500 ppm. As shown in Table 1, the water washing
process increased the water content and the puried biodiesel has
more water than the limit allowed by the legislation. He et al. [11]
observed that water washing can increase the water content if water is added at 20  C. Gonzalo et al. [29] reported the values of water
content in puried biodiesel samples ranging from 800 to 1000 ppm.
Moreover, Gonzalo et al. [29] also observed that this water content
increases after the water washing process. This behavior can be
associated to the interaction between mono and diglycerides and
water, since water solubility in ester is very small. Mono and diglycerides molecules, left from an incomplete reaction, can act as an
emulsier, allowing the water to be mixed with the biodiesel.
The acid values of crude and washed biodiesel are smaller than
the acid value of soybean oil. Usually, for a base catalyzed process,
the acid value after production will be low since the base catalyst
will strip the available free fatty acids. The alkali catalyst also serves
as a caustic stripper and removes the free fat acids by converting
them into soap that is removed during washing. The acid value of
crude biodiesel is smaller than the acid value of washed biodiesel.

Fig. 3. Permeate ux of biodiesel at 1 and 2 bar of transmembrane pressures


throughout microltration membranes of 0.22 mm (a) and 0.30 mm (b) pore size.

In this case, the presence of alkali catalyst in crude biodiesel


diminished the quantity of base (KOH) that was necessary to
neutralize the available free fatty acids. The catalyst was removed
by the washing process and then the acid value increased. Soap was
removed by the washing process and the amount of soap decreased
after the washing process.
The water washing process was able to achieve the limit
imposed by ASTM D6751 (USA) [4] and EN 14214 (Europe) [5] for
free glycerol content. Except for the water content, all the other
analyzed parameters of the washed biodiesel are in accordance to
the limits imposed by the legislation.
3.2. Membrane separation process
Crude biodiesel was ltered throughout micro and ultraltration
membranes. Fig. 3 presents the obtained ux of biodiesel through
the microltration membranes at 1 and 2 bar. A ux decline is
observed in the rst 2 min of operation for the membrane of 0.22 mm
(Fig. 3a). Fig. 3b shows that for the 0.30 mm membrane steady declines are observed for 1 and 5 min at 1 and 2 bar of transmembrane
pressures, respectively. The stabilized ux is greater at 2 bar than at
1 bar, showing that greater transmembrane pressure enables
greater uxes within the analyzed pressure range. Moreover, the
ux obtained is greater with the more open membrane.
Fig. 4 presents the permeate ux throughout the ultraltration
membranes (10 and 20 kDa MWCO) at the transmembrane pressure of 4 bar. Besides the higher transmembrane pressure, the ux
with ultraltration membranes was smaller than with the microltration ones and a less pronounced ux decline was observed in
this case. The membrane with higher cutoff presented higher
uxes.
Table 2 presents the observed nal ux for each evaluated
membrane. Othman [30] evaluated nanoltration polymeric
membranes for biodiesel ltration and the higher observed ux
was approximately 51 kg h1 m2. Saleh et al. [31] proposed the
application of ceramic membranes for biodiesel purication and
observed stabilized uxes around 40 and 30 kg h1 m2 for
micro and ultraltration processes, respectively. Wang et al. [13]
used a ceramic microltration membrane of 0.1 mm pore size at
1.5 bar of transmembrane pressure for biodiesel purication and
observed a stabilized ux of about 300 kg h1 m2. Mah et al. [32]
also used at polymeric membrane of 30 kDa for dead-end

Fig. 4. Permeate ux of biodiesel throughout ultraltration membranes.

M.J. Alves et al. / Renewable Energy 58 (2013) 15e20

19

Table 2
Final ux throughout micro and ultraltration membranes.
Membrane pore size

0.22 mm

Pressure (bar)
Final ux (kg h1 m2)

1
109

0.30 mm
2
253

1
536

2
923

10 kDa

30 kDa

4
55

4
120

ltration of synthetic mixtures containing glycerol, oleic acid and


palm oil. Mah et al. [32] observed nal uxes ranging from 349 to
82.8 kg h1 m2, depending on the composition of the feed mixtures. Several aspects can inuence the observed ux throughout
the membrane, as the membrane material and its pore size, the
operation mode and the feed composition. According to the
observed uxes, the membranes proposed in this work are a
promising alternative for biodiesel purication.
The physico-chemical characterization of permeate samples
showed that the ltration process did not affect the density and the
viscosity of crude biodiesel. These results show that the membrane
was not able to remove catalyst excess and/or free acid content of
crude biodiesel, since the acid value of crude biodiesel did not
change after the ltration. Table 3 presents the amount of soap and
the free glycerol content measured in ltrate biodiesel samples.
Fig. 5 presents the percentage reductions related to the parameter
values of crude biodiesel.
Only the ultraltration membrane of 30 kDa (4 bar) was not able
to reduce the amount of soap and free glycerol content detected in
crude biodiesel sample. And additional test was carried out with
this membrane at a smaller transmembrane pressure (3 bar). This
reduction in the transmembrane pressure value increases the
performance of the 30 kDa membrane for biodiesel purication.
However, this membrane did not show promising values for biodiesel purication, probably related to the more open pore size (in
relation to the 10 kDa membrane) and to the higher applied
transmembrane pressure.
Fig. 5 shows how the removal of soap is related to the free
glycerol removal. Higher soap removals lead to higher free glycerol
removals. Wang et al. [13] stated that the size of reversed micelle
formed by the soap and free glycerol with the mean size of 2.21 mm
(analyzed by zeta potential analyzer) was larger than that of biodiesel, and was easier to be removed the membrane separation.
Saleh et al. [14] showed that the presence of water increases glycerin removal by the membrane separation process.
The transmembrane pressure applied in the microltration experiments did not affect the biodiesel quality. Regarding to free glycerol content, as presented in Table 3, only the ultraltration
membrane of 10 kDa was able to reduce the glycerol content according
to the international legislation limit for biodiesel (less than 0.02 wt%).
This result was conrmed by chromatography analysis. Table 4
presents the results obtained by chromatography analyzes of
crude, washed and permeate samples. This result shows that the
ultraltration process with the 10 kDa membrane was able to
reduce the glycerol content to the desired level (less than 0.02 wt%).
Table 3
Characterization of ltrate biodiesel samples.
Membrane and pressure

0.22 mm (1 bar)
0.22 mm (2 bar)
0.3 mm (1 bar)
0.3 mm (2 bar)
10 kDa (4 bar)
30 kDa (4 bar)
30 kDa (3 bar)
a

Parameter
Amount of soap
(103 gsoap g1sample)a

Free glycerol
(%, wt/wt)

1.3
1.3
1.6
1.6
1.0
2.7
2.4

0.022
0.025
0.026
0.026
0.020
0.031
0.029

Fig. 5. Reductions for amount of soap and free glycerol content in puried biodiesel
samples.

3.3. Water addition


Further ltrations were carried out using the membrane of
10 kDa and adding water to the crude biodiesel sample prior to the
ltration. Fig. 6 presents ux results for ltrations of the samples
with and without water addition throughout the 10 kDa ultraltration membrane. According to these results (Fig. 6) the water
addition decreases the observed stabilized ux. Gomes et al. [15]
evaluated the addition of acidied water to crude biodiesel sample at 20 wt%. These authors observed a sharp drop in the ux and a
smaller stabilized ux for the sample with water addition.
The decrease in the ux is probably associated with higher removals. Table 5 presents free glycerol and water concentrations in
the permeate at some ltration times. For these samples, free
glycerol content was measured by gas chromatography.
Water addition improved the glycerol removal since glycerol and
water (completely soluble) formed an immiscible phase with the
FAME phase [14]. The molecules of water joined to glycerol and these
larger molecules were unable to pass through the membrane pores.
After some ltration time the glycerol concentration in the permeate
increased as the water concentration decreased. This increase in the
glycerol content in the permeate as a function of the ltration time is
probably related to the dead-end operation mode. The accumulation
of glycerol particles near to the membrane surface caused their
permeation at longer ltration times. Mah et al. [32] reported the
cake formation in dead-end ltrations of synthetic mixtures of palm
oil with glycerin.
Addition of water at higher concentration (0.2 wt%) improved the
glycerol removal by the membrane at the end of the ltration time.
Lower water concentration was measured in the permeate for the
sample with water addition at 0.2 wt%, showing that this added water
was effectively used to be joined to the glycerol molecules. Saleh et al.
[16] showed that a polyacrylonitrile (PAN) membrane with a molecular
weight cutoff of 100 kDa was able to reduce the glycerol content in
biodiesel obtained from canola oil. The authors applied a
Table 4
Free glycerol content in biodiesel samples determined by gas chromatography
analysis.
Parameter

Free glycerol (%, wt/wt)


As potassium oleate.

Biodiesel sample
Crude

Washed

Filtrated (10 kPa; 4 bar)

0.049

0.011

0.019

20

M.J. Alves et al. / Renewable Energy 58 (2013) 15e20

References

Fig. 6. Permeate uxes throughout the ultraltration membrane of 10 kDa with water
addition to the crude biodiesel.
Table 5
Free glycerol and water contents in biodiesel permeate samples with water addition
for the 10 kDa membrane at 4 bar.
Filtration
time (min)

10
40
60

Sample: crude biodiesel


0.1 wt% of water

Sample: crude biodiesel


0.2 wt% of water

Free glycerol
(%, wt/wt)

Water
(ppm)

Free glycerol
(%, wt/wt)

Water
(ppm)

0.002
0.006
0.010

1577.17
1362.36
987.90

0.007
0.007
0.009

1001.08
937.76
910.28

transmembrane pressure of 552 kPa. However, according to the results


showed by Saleh et al. [16], the glycerol content in the permeated
achieved the limit of less than 0.2 wt% only when pure water was added
to the feed solution. Saheh et al. [14] showed that small quantities of
water have a great effect in removing glycerol from biodiesel.
4. Conclusions
This work evaluated the application of micro and ultraltration
membranes for biodiesel purication. Microltration membranes
presented higher uxes, but the permeate characteristics did not
attempt the international legislation regarding to free glycerol
content. The ultraltration membrane of 30 kDa was not able to
produce a puried biodiesel according to the international legislation parameters. Between the analyzed membranes, the glycerol
content level (less than 0.02 wt%) was achieved only with the
membrane of 10 kDa. Water addition in the biodiesel sample
improved the glycerol removal. This membrane (10 kDa) also presented a suitable permeate ux, showing that the membrane separation process is a suitable alternative for biodiesel purication.

Acknowledgments
The research work was funded by VALE S.A. and Conselho
Nacional de Desenvolvimento Cientco e Tecnolgico (CNPQ).

[1] Basha SA, Gopal KR, Jebaraj S. A review on biodiesel production, combustion,
emissions and performance. Renew Sustain Energy Rev 2009;13:1628e34.
[2] Nogueira LAH. Does biodiesel make sense? Energy 2011;36:3659e66.
[3] Cheng JJ, Timilsina GR. Status and barriers of advanced biofuel technologies: a
review. Renew Energy 2011;36:3541e9.
[4] ASTM D6751. Standard specication for biodiesel fuel blend stock (B100) for
middle distillate fuels. 2012.
[5] EN 14214. The pure biodiesel standard. 2008.
[6] Atadashi IM, Aroua MK, Aziz AA. Biodiesel separation and purication: a review. Renew Energy 2011;36:437e43.
[7] Cao PG, Dube MA, Tremblay AY. High-purity fatty acid methyl ester production from canola, soybean, palm, and yellow grease lipids by means of a
membrane reactor. Biomass Bioenergy 2008;32:1028e36.
[8] Cao PG, Tremblay AY, Dube MA, Morse K. Effect of membrane pore size on the
performance of a membrane reactor for biodiesel production. Ind Eng Chem
Res 2007;46:52e8.
[9] Baroutian S, Aroua MK, Aziz ARA, Sulaiman NMN. TiO2/Al2O3 membrane
reactor equipped with a methanol recovery unit to produce palm oil biodiesel.
Int J Energy Res 2012;36:120e9.
[10] Cheng LH, Yen SY, Chen ZS, Chen JH. Modeling and simulation of biodiesel
production using a membrane reactor integrated with a prereactor. Chem Eng
Sci 2012;69:81e92.
[11] He HY, Guo X, Zhu SL. Comparison of membrane extraction with traditional
extraction methods for biodiesel production. J Am Oil Chem Soc 2006;83:
457e60.
[12] Cheng LH, Cheng YF, Yen SY, Chen JH. Ultraltration of triglyceride from
biodiesel using the phase diagram of oil-FAME-MeOH. J Membr Sci 2009;330:
156e65.
[13] Wang Y, Wang XG, Liu YF, Ou SY, Tan YL, Tang SZ. Rening of biodiesel by
ceramic membrane separation. Fuel Process Technol 2009;90:422e7.
[14] Saleh J, Dube MA, Tremblay AY. Effect of soap, methanol, and water on
glycerol particle size in biodiesel purication. Energy Fuel 2010;24:
6179e86.
[15] Gomes MCS, Arroyo PA, Pereira NC. Biodiesel production from degummed
soybean oil and glycerol removal using ceramic membrane. J Membr Sci
2011;378:453e61.
[16] Saleh J, Tremblay AY, Dube MA. Glycerol removal from biodiesel using
membrane separation technology. Fuel 2010;89:2260e6.
[17] ASTM D445. Standard test method for kinematic viscosity of transparent and
opaque liquids (and calculation of dynamic viscosity); 2006.
[18] ASTM D1298. Standard test method for density, relative density (specic
gravity), or API gravity of crude petroleum and liquid petroleum products by
hydrometer method. 2003.
[19] ASTM D664. Standard test method for acid number of petroleum products by
potentiometric titration. 2006.
[20] Knothe G. Analytical methods used in the production and fuel quality
assessment of biodiesel. Trans ASAE 2001;44:193e200.
[21] AOCS. Ofcial method Cd 3-25 saponication value. 2003.
[22] EN ISO 12937. Petroleum products, determination of water, coulometric Karl
Fisher titration method. 2000.
[23] AOCS. Ofcial method Ca14-56 total, free and combined glycerol iodometricperiodic acid method. 1997.
[24] Monteiro MR, Ambrozin ARP, Liao LM, Ferreira AG. Critical review on
analytical methods for biodiesel characterization. Talanta 2008;77:593e
605.
[25] ASTM D6584. Test method for determination of free and total glycerin in B100 biddies methyl esters by gas chromatography. 2000.
[26] Leung DYC, Guo Y. Transesterication of neat and used frying oil: optimization
for biodiesel production. Fuel Process Technol 2006;87:883e90.
[27] Freedman B, Pryde EH, Mounts TL. Variables affecting the yields of fatty esters
from transesteried vegetable-oils. J Am Oil Chem Soc 1984;61:1638e43.
[28] Sharma YC, Singh B. Development of biodiesel: current scenario. Renew
Sustain Energy Rev 2009;13:1646e51.
[29] Gonzalo A, Garcia M, Sanchez JL, Arauzo J, Pena JA. Water cleaning of biodiesel. Effect of catalyst concentration, water amount, and washing temperature on biodiesel obtained from rapeseed oil and used oil. Ind Eng Chem Res
2010;49:4436e43.
[30] Othman R, Mohammad AW, Ismail M, Salimon J. Application of polymeric
solvent resistant nanoltration membranes for biodiesel production. J Membr
Sci 2010;348:287e97.
[31] Saleh J, Dube MA, Tremblay AY. Separation of glycerol from FAME using
ceramic membranes. Fuel Process Technol 2011;92:1305e10.
[32] Mah SK, Leo CP, Wu TY, Chai SP. A feasibility investigation on ultraltration of
palm oil and oleic acid removal from glycerin solutions: ux decline, fouling
pattern, rejection and membrane characterizations. J Membr Sci 2012;389:
245e56.

You might also like