Industrial Crops and Products

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Industrial Crops and Products 89 (2016) 416–424

Contents lists available at ScienceDirect

Industrial Crops and Products


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

Solid acid as catalyst for biodiesel production via simultaneous


esterification and transesterification of macaw palm oil
Leyvison Rafael V. da Conceição, Livia M. Carneiro, J. Daniel Rivaldi, Heizir F. de Castro ∗
Engineering School of Lorena, University of São Paulo, Estrada Municipal do Campinho s/n, 12602-810 Lorena, São Paulo, Brazil

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

Article history: Heterogeneous catalysis applied to esterification and transesterification of non-edible oil offers a strat-
Received 20 January 2016 egy to the clean synthesis of the biodiesel and is driving research interested into the development of
Received in revised form 22 May 2016 acid catalysts for efficient conversion of low quality vegetable oils into fuels to meet future societal
Accepted 24 May 2016
demands. Thus, sulfated niobium oxide catalyst was synthesized by the impregnation method and used
as a heterogeneous catalyst aimed at biodiesel production via macaw palm oil through high free fatty acid
Keywords:
content transesterification with ethanol. The effect of two reaction parameters, molar ratio of ethanol to
Biodiesel
macaw palm oil and reaction temperature, on ester content and viscosity was studied by the response
Macaw palm oil (Acrocomia aculeata)
Sulfated niobium oxide
surface methodology (RSM). The ester content was determined by GC. The catalyst shows excellent
Optimization activity (99.2% ester content and 4.5 mm2 /s viscosity) towards biodiesel production. Its optimum reac-
Ethanolysis tion conditions were: 120:1 molar ratio of ethanol to macaw palm oil at 250 ◦ C reaction temperature.
The catalysts characterization was carried out by using the X-ray Diffraction (XRD), Scanning Electron
Microscopy (SEM), Fourier Transform Infrared Spectroscopy (FT-IR), and N2 Adsorption-desorption and
Surface Acidity Analyses.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction processing. However, edible oils for producing biodiesel is the sub-
ject of debate on its use as fuel or as food, thus these factors have
As traditional fossil fuels are not renewable and their depletion negatively affected its production from edible oils. Therefore, non-
is in the offing, probably in the next several decades, the search for edible vegetable oils have become more attractive for biodiesel
alternative fuels is becoming increasingly attractive (Fu et al., 2013). production (Ong et al., 2013). Hence, many resources have been
In this context, biodiesel stands out as an alternative fuel which analyzed in order to explore non-edible feedstocks for biodiesel
produces low aromatic gases and CO2 emissions. Moreover, it is also production, such as Jatropha oil (Jatropha curcas L.) (Nizah et al.,
renewable, biodegradable, non-flammable, non-toxic, and sulfur- 2014), Karanja oil (Pongamia pinnata L.) (Thiruvengadaravi et al.,
free (Syazwani et al., 2015). Biodiesel consists of a mixture of fatty 2012), Sea Mango oil (Cerbera odollam Gaertn.) (Kansedo and Lee,
acid alkyl esters, derived from renewable lipid feedstocks, such as 2013), Rubber seed oil (Hevea brasiliensis Muell. Arg) (Reshad et al.,
vegetable oils and animal fats, which are basically triacylglycerols 2015), Moringa oil (Moringa oleifera Lam.) (Fernandes et al., 2015),
(Banković-Ilić et al., 2014). Kusum oil (Schleichera oleosa L.) (Silitonga et al., 2015), Andiroba oil
Alcoholysis is the process that most frequently takes place, (Carapa guianensis Aubl.), Castanhola oil (Terminalia catappa L.) (Iha
which is an established method for the transformation of veg- et al., 2014), Mahua oil (Madhuca indica J.F. Gmel) and Simarouba
etable oils into biodiesel. The transesterification reaction can occur oil (Simarouba glauca DC) (Jena et al., 2010).
in the presence of homogeneous or heterogeneous catalysts (acids Macaw palm (Acrocomia aculeata (Jacq.) Lodd ex Mart.), com-
or bases) or enzyme (mostly lipases) with alcohol (methanol and monly known in Brazil as Macaúba, is a native oleaginous palm tree
ethanol) (Avhad and Marchetti, 2015; Banković-Ilić et al., 2014). which grows abundantly in the Brazilian cerrado, located mainly
Biodiesel production from non-edible oils has drawn the atten- in the center of the country, but adapted from cooler subtropical
tion of researchers due to its high biodiesel yield and easy scenarios to drier semiarid ecosystems, and widely spread across
Mexico, Antilles, Argentina, Uruguay and Paraguay (Moura et al.,
2009). This palm tree can reach 15–20 m in height, with 20–30 cm
of trunk diameter and covered by dark thorns that are usually 10 cm
∗ Corresponding author. in length (César et al., 2015). Its fruits are spherical, with 2.5–5.0 cm
E-mail address: heizir@dequi.eel.usp.br (H.F. de Castro).

http://dx.doi.org/10.1016/j.indcrop.2016.05.044
0926-6690/© 2016 Elsevier B.V. All rights reserved.
L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424 417

in diameter. A single plant has 2–6 clusters, with a total number of 2.3. Catalyst characterization
fruits ranging from 260 to 1270 per palm tree, producing estimated
yields of 4000 kg of oil/ha (Ciconini et al., 2013). 2.3.1. X-ray diffraction (XRD)
However, its oil presents high acidity and cannot be used X-ray powder diffractograms of the catalysts were collected
as feedstock for biodiesel production by a conventional alkaline with a PANalytical Model Empyrean X-ray diffractometer at 40 kV
route (Aguieiras et al., 2014). In this case, the biodiesel synthesis and 30 mA. Radiation of Cu K␣ (1.541874 Â) was employed with 2␪
should be carried out via simultaneous esterification and transes- of 8◦ –70◦ .
terification process using acid catalysts, which enables the use of
low-quality raw materials with high concentrations of free fatty 2.3.2. Scanning electron microscopy (SEM)
acids (FFA). The morphology of the catalyst was analyzed by high-resolution
The major focus of recent advances is on the rational devel- scanning electron microscopy in a LEO Model 1450 VP, operating
opment of recyclable solid acids catalysts. These catalysts have with accelerating voltage of 20 kV.
been established as being other alternatives to the heterogeneous
alkaline, unrecyclable-homogeneous acid and alkaline catalysts. 2.3.3. Fourier transform infrared spectroscopy (FT-IR)
Several encouraging results described in literature have highlighted The infrared spectra were recorded on a spectrum GX FT-IR sys-
the potentials of solid acid catalyzed-biodiesel production (Sani tem Perkin Elmer spectrometer. The sample was pelletized with
et al., 2014). Among the most studied solid acids are mixed oxides KBr, and the spectrum was obtained from the accumulation of a
(Liu et al., 2015; Amani et al., 2014), ion-exchange resin (Fu et al., total of 32 scans in the range of 4000–400 cm−1 with resolution of
2015; Shibasaki-Kitakawa et al., 2015), heteropolyacid (Narkhede 4 cm−1 .
et al., 2014; Gong et al., 2014), sulfated zirconia (Patel et al., 2013;
Saravanan et al., 2015) and zeolite (Vieira et al., 2015; Sun et al., 2.3.4. N2 adsorption-desorption analysis
2015). A sample of 0.2 g was added to a glass cell and heated at 200 ◦ C
The aim of the present work is to optimize the performance for two hours under vacuum to remove impurities adsorbed on
of sulfated niobium (Nb2 O5 /SO4 ) as a heterogeneous catalyst in the catalyst surface. Nitrogen adsorption isotherms were obtained
a simultaneous esterification and transesterification for biodiesel in a Nova 1000 mark Quantachrome. The surface area was calcu-
production using macaw palm oil with high free fatty acid con- lated with the BET equation, and the mean pore size diameter was
tent and ethanol as acylant agent. Ethanol has some advantages calculated by the BJH method.
when used in a process for biodiesel production. It has a supe-
rior dissolving power in vegetable oils and the ethyl esters (FAEE) 2.3.5. Surface acidity
show lower smoke opacity, lower exhaust temperature and lower The surface acidity of the catalyst was determined by using acid-
pour point (Yusoff et al., 2014). The evaluation of exhaust gas emis- base titration. An aqueous suspension of 0.1 g of catalyst in 20 mL of
sions (including nitrogen oxides, CO2 and smoke density) shows NaOH (0.1 M) was maintained under agitation for 3 h at room tem-
that FAEE has a less negative environmental effect in comparison to perature to allow ion exchange. The suspension was centrifuged
methyl esters (FAME) (Brunschwig et al., 2012; Stamenkovic et al., and the supernatant titrated with HCl (0.1 M) in the presence of
2011). Additionally, unlike methanol (which is generally derived phenolphthalein. The surface acidity of the catalyst was expressed
from fossil sources), ethanol is produced mainly from renewable in mmol H+ /g catalyst.
sources via fermentation processes, and because its large scale pro-
duction as a substitute fuel for gasoline already exists, the supply of 2.4. Catalytic reaction
bioethanol for the industrial production of biodiesel can be easily
achieved (Brunschwig et al., 2012). The Nb2 O5 /SO4 catalyst was previously dried at 120 ◦ C for 2 h
before utilization. Its catalytic performance was analyzed in a
simultaneous esterification and transesterification reaction. Exper-
2. Material and methods
iments were conducted using different ethanol-to-oil molar ratios
and reaction temperatures. All experiments were conducted at
2.1. Materials
300 rpm for a maximal period of 4 h in a pressurized stainless steel
reactor (Parr Series 5000 Multiple Reactor System) equipped with
Macaw palm oil with high free fatty acid content was sup-
an electrical heating jacket covering the reaction vessel, and indi-
plied by Association of Small Farmers D’Antas (Minas Gerais,
vidual temperature control.
Brazil). Hydrated niobium oxide HY340 (amorphous) with high
At the end of the transesterification, the catalyst was separated
surface area (BET ∼170 m2 /g) containing 80% Nb2 O5 was supplied
from the mixture containing products and reagents by centrifuga-
by Companhia Brasileira de Metalurgia e Mineração—CBMM and
tion at 1570x for 15 min. The supernatant containing ethyl esters
calcined at 500 ◦ C for 5 h before use. Sulfuric acid (98.0%) and anhy-
was transferred to a separating funnel and washed with the same
drous ethanol (98.0%) were purchased from VETEC® Sigma-Aldrich.
volume of hot water for 12 h, allowing the decantation and sep-
Anhydrous sodium sulfate, ethyl acetate (99.5%) and hexane (65.0%)
aration of glycerol and the ethyl ester phase (higher phase). This
were supplied by Cromoline. Methanol (99.95%) and acetonitrile
procedure was carried out three times to assure total glycerol
(99.9%) were purchased from J.T. Baker. Internal standard methyl
removal. The liquid phase containing ethyl esters was submitted to
heptadecanoate C17:0 (99.0%) was purchased from Sigma.
evaporation in a vacuum rotary evaporator to remove the remain-
ing ethanol. Finally, the ethyl esters were dried with anhydrous
2.2. Catalyst preparation sodium sulfate to remove traces of water.

The Nb2 O5 /SO4 catalyst was prepared with a mixture of 5 g 2.5. Experimental design
Nb2 O5 , 5 mL H2 SO4 solution (0.5 M) and 15 mL deionized water in a
glass reactor at 90 ◦ C under reflux and constant agitation (500 rpm) The effects of the process variables of ethanol-to-oil molar ratio
for 3 h. The solid was washed twice with deionized water, then (X1 ) and reaction temperature (X2 ) on ester content (Y1 ) and kine-
filtered and dried at 100 ◦ C for 12 h. The catalyst was calcined at matic viscosity (Y2 ) were verified using the experimental design.
500 ◦ C for 5 h. The experiments were carried out and optimized following the
418 L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424

Table 1
Macaw palm oil fatty acid composition and physico-chemical properties.

Property Value

Fatty acid (%)


Palmitic acid (C16:0) 18.7
Palmitoleic acid (C16:1) 4.0
Stearic acid (C18:0) 2.8
Oleic acid (C18:1) 53.4
Linoleic acid (C18:2) 17.7
Linolenic acid (C18:3) 1.5
Saturated total 21.5
Unsaturated total 78.5
Acid value (mg KOH/g) 39.0
Kinematic viscosity at 40 ◦ C (mm2 /s) 44.5
Density at 20 ◦ C (g/cm3 ) 0.902
Moisture content (%) 1.25

rate of 2.0 mL/min. The temperature setup was T1 = 150 ◦ C for 1 min,
heating up to T2 = 180 ◦ C at 7.5 ◦ C/min, holding for 2 min and heat-
Fig. 1. X-ray diffractograms (XRD) of a) Nb2 O5 , b) Nb2 O5 thermally treated at 500 ◦ C
and c) Nb2 O5 /SO4 catalyst.
ing up to T3 = 240 ◦ C for 3 min. The ester content, expressed as mass
fraction in percentage terms, is calculated by Eq. (2):

face centered central composite design (CCD) and the response sur-
A − API CPI
face methodology (RSM). The software Statistica (Statisoft, v.7) was EsterContent = × × 100 (2)
API Csample
used to analyze the results.

2.6. Statistical analysis Where,


A is the total peak area from the ethyl ester;
The RSM is a set of mathematical and statistical techniques API is the peak area corresponding to methyl heptadecanoate;
which are useful for modeling and analyzing problems in which CPI is the concentration of the methyl heptadecanoate solution;
a response of interest is influenced by several variables with the Csample is the concentration of the sample solution.
purpose of optimizing this response (Montgomery, 1997; Hameed The ethyl ester viscosity was determined according to the stan-
et al., 2009). The first step of the RSM was to obtain a mathe- dard method ASTM D 445 with a LVDVII Brookfield viscosimeter
matical model that describes the response variables (Y1 and Y2 ) (Brookfield Viscometers Ltd., England) equipped with a CP 42 cone
as a function of the independent variables (X1 and X2 ). The equa- at 40 ◦ C using 0.5 mL of sample. Its density value was determined
tion obtained from it only considers the significant coefficients, as according to the standard method ASTM D 4052 with a DMA 35N
shown in Eq. (1). Coefficient ˇ0 is the outcome (response) at cen- EX digital densimeter (Anton Paar) at 20 ◦ C using 2 mL of sample.
tral point, and the other coefficients measure the main effects and The remaining amount of glycerides after the transesterification
interactions of coded variables Xi on the response variable Y. was determined in an Agilent 1200 Series liquid chromatograph
(Agilent Technologies, USA) equipped with an Evaporative Light
k k k
Y = ˇ0 + ˇi Xi + ˇii X 2 i + ˇij Xi Xj +e (1) Scattering Detector and a Gemini C-18 (5 ␮m, 150 × 4.6 mm, 110 Å)
i=1 i=1 i<1 column at 40 ◦ C. The mobile phase contained a mixture of acetoni-
The model obtained from the regression analysis was used to trile (80%) and methanol (20%) at a flow rate of 1 mL/min for 6 min,
generate the response surface and contour plots. For selecting the 1.5 mL/min until 30 min, and 3.0 mL/min until 35 min. All samples
significant terms from the model, the significance level was set at were dissolved in ethyl acetate-hexane (1:1, v/v), filtered through
95% and p < 0.05. The second step was to determine the quality of 0.22 ␮m membrane filters (Millipore) and injected in a volume of
the model, which was assessed by the analysis of variance (ANOVA) 10 ␮L. All solvents were of an HPLC grade, and the assays were
using an F-test (Box et al., 1978). performed at least in duplicate.

2.7. Oil and biodiesel characterization


3. Results and discussion
2.7.1. Macaw palm oil
The macaw palm oil fatty acid composition was determined by 3.1. Properties of macaw palm oil
gas chromatography (GC) according to the AOCS Ce 1-62 method.
The physico-chemical characterization was conducted according to The fatty acid composition and physico-chemical characteriza-
the following standard methods: acid value (AOCS Cd 3d-63), vis- tion of the macaw palm oil are shown in Table 1.
cosity (ASTM D 445), density (ASTM D 4052) and moisture content Macaw palm oil contains a large amount of unsaturated fatty
(AOCS Ca 2b-38). acids (78.5%) primarily, oleic acid (53.4%), and a lower amount
of saturated acid (21.5%), typically palmitic (18.7%). Its acidity
2.7.2. Macaw palm biodiesel and moisture content was 39.0 mg KOH/g and 1.25%, respectively,
The ester content was determined by gas chromatography which was a determining factor in the choice of an acid catalyst
according to Silva et al. (2007). The samples were analyzed on a because the use of alkaline catalysts are recommended when the
Varian CP 3800 chromatograph equipped with an auto injector, oil has an free fatty acid (FFA) value of less than 3% and moisture
and a FID. A TR FAME capillary column was used with the follow- content lower than 1%, since the saponification reaction compro-
ing features: 30 m length, 0.25 mm internal diameter, and 0.25 ␮m mises the transesterification reaction’s efficiency and hinders the
film thickness. Nitrogen gas was used as the mobile phase at a flow separation of ester and glycerol phases (Meher et al., 2006).
L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424 419

Fig. 2. SEM images of Nb2 O5 /SO4 catalyst.

Table 2 lack of peaks is also associated with its structure, water and impu-
Textural and acidic properties of Nb2 O5 and Nb2 O5 /SO4 .
rities (Nowak and Ziolek, 1999).
Sample Surface area (m2 /g) Pore volume (cm3 /g) Surface acidity (mmol H+ /g) The Nb2 O5 /SO4 spectrum showed pronounced peaks regarding
Nb2 O5 169.1 0.093 0.041 the crystalline phase in the region of 23, 28, 36, 51 and 56 (2␪) of
Nb2 O5 /SO4 63.1 0.143 2.764 the spectrum (Fig. 1c). These peaks are similar to those observed for
niobium oxide submitted to a thermal treatment at 500 ◦ C without
impregnation (Fig. 1b). Calcined Nb2 O5 usually shows a polycrys-
Table 3 talline structure (polymorphism), depending on the temperature
Experimental design matrix and results of the synthesis of macaw palm oil biodiesel
and the presence of oxygen during the thermal treatment (Melo
via esterification and transesterification.
et al., 2012). The formation of crystals in this oxide starts at 500 ◦ C,
Run Variable Response promoting structures of monoclinic, hexagonal and orthorhombic
Molar ratio Temperature Ester content (%) Viscosity (mm2 /s) systems (Nowak and Ziolek, 1999). As the calcination temperature
(◦ C) increases, the surface area decreases due to the crystal formation
1 60 −1 150 −1 34.65 20.00 and compaction. However, this arrangement reduces the number of
2 60 −1 250 1 86.80 6.15 available active acid sites for catalysis, and promotes greater struc-
3 120 1 150 −1 32.10 19.00 tural stability. The impregnation with sulfuric acid did not affect
4 120 1 250 1 99.20 4.50
the crystalline structure of the niobium oxide after calcination, as
5 90 0 200 0 62.90 11.35
6 90 0 200 0 55.00 12.45
demonstrated by the same crystal spectrum observed for both the
7 90 0 200 0 54.85 11.00 catalyst and the calcined support (Fig. 1c).
8 60 −1 200 0 75.50 8.10 The catalyst SEM images are shown in Fig. 2(a) and (b). The
9 120 1 200 0 51.90 13.70 micrographs clearly show that the catalyst is not uniform and its
10 90 0 150 −1 20.50 24.60
particles do not present a single size.
11 90 0 250 1 92.30 5.85
The FT-IR spectra of Nb2 O5 and Nb2 O5 /SO4 are presented in
Fig. 3. The spectrum of the samples is predominated by a broad
Table 4 band at 675 cm−1 , which is typical for niobium oxides, and it is
Analysis of variance (ANOVA) of the response surface quadratic model. assigned to the vibrations of Nb O Nb bridges from slightly dis-
Sources Sum of squares Degree of freedom Mean squares F-value F0.05 torted octahedral NbO6 connected with sharing corners. The weak
band around 850 cm−1 can be assigned to the symmetric stretching
Ester model
Regression 5899.3 5 1179.86 13.05 5.05
mode of Nb O surface species, which is present in highly distorted
Residual 452.11 5 90.42 octahedral NbO6 structures (Stosic et al., 2014). At 1625 cm−1 , there
Lack of Fit 409.70 3 136.56 6.44 19.16 appears to be a peak which can be attributed to water molecules
Pure error 42.41 2 21.20 adsorbed on the surface of Nb2 O5 and Nb2 O5 /SO4 . There is also
Totals 6351.41 10
a shoulder at 3435 cm−1 , probably due to the hydroxyl moieties
Viscosity model (Nb OH) as well as from the water molecules themselves. In the
Regression 440.87 5 88.17 18.76 5.05 spectrum of Nb2 O5 /SO4 , the peak at 1125 cm−1 can be attributed
Residual 23.52 5 4.70
Lack of Fit 22.42 3 7.47 13.58 19.16
to the symmetric stretching of S O bonds (Ngee et al., 2014).
Pure error 1.10 2 0.55 Table 2 presents the values of surface area, pore volume and total
Totals 464.39 10 surface acidity of Nb2 O5 and Nb2 O5 /SO4 . The nitrogen adsorption-
desorption isotherms for Nb2 O5 and Nb2 O5 /SO4 are shown in Fig. 4.
The isotherms showed pronounced type IV characteristics with
3.2. Catalyst characterization hysteresis loop type H2, according to the IUPAC classification (Sing
et al., 1985). The hysteresis loop is associated with the capillary
The X-ray diffractograms of the Nb2 O5 support and Nb2 O5 /SO4 condensation taking place in the mesopores, which indicates the
catalyst were prepared by the impregnation method, which are existence of a mesoporous structure in the sulfated niobium. The
shown in Fig. 1. The spectrum of Nb2 O5 without any thermal treat- type H2 hysteresis loop is attributed to the pores shaped like ink
ment (Fig. 1a) showed a characteristic profile of a non-crystalline bottles (pores with narrow necks and wide bodies) (Deshmane and
material without peaks, corresponding to an amorphous solid. This Adewuyi, 2013). The sulfated niobium catalyst showed acidity of
420 L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424

Fig. 3. Infrared spectra of the Nb2 O5 and Nb2 O5 /SO4 catalyst.

Fig. 4. N2 adsorption-desorption isotherms of Nb2 O5 and Nb2 O5 /SO4 catalyst.


2.764 mmol H+ /g which is 67 times greater than that of the niobium
oxide used as support. This confirms the impregnation of sulfate
groups on the crystalline structure of the niobium oxide.

3.3. Optimization of the reaction conditions to produce biodiesel ethanol-to-oil molar ratio (X1 ) and reaction temperature (X2 ), were
from macaw palm oil analyzed by using the RSM, as shown in Table 3.
By applying a multiple regression analysis of Table 3 data, the
In the present work, the relations between ester content (Y1 ) experimental results of the centered face factorial design were fit-
and kinematic viscosity (Y2 ) responses and two reaction variables, ted to the polynomial Eq. (2). The models obtained for ester content

Fig. 5. Pareto charts for a) ester content and b) viscosity, and predicted vs. actual values for c) ester content and d) viscosity.
L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424 421

values. Notably, an increase in the reaction’s temperature from the


lowest to the highest level yielded an average increase of 61.6% in
ester content. In Fig. 5(b), variables X2 Temperature(L), X1 Molar
ratio(Q) and X2 Temperature(Q) were found to be significant for
the biodiesel’s viscosity. Their values were 27.64, −4.78, and 4.53,
respectively. A variation from the lowest to the highest level of
reaction temperature and molar ratio produced a 16.7% and 4.4%
decrease in the biodiesel’s viscosity, respectively.
Fig. 5(c) and (d) compares the observed experimental ester con-
tent and viscosity with the predicted values. The determination
coefficient (R2 ) value of ester content (Fig. 5c) is 0.9288, indicating
approximately 93% variability in the data accounted by the model.
Meanwhile, as regards its viscosity (Fig. 5d), the R2 value is 0.9494,
indicating approximately 95% of variation in the data explained by
the model.
Generally, the 3D response surface plots are graphical repre-
sentations of the regression equation for optimizing the reaction
variables. Fig. 6 shows the response surface and contour plots for
the effects of temperature and molar ratio on ester content (Fig. 6a)
and viscosity (Fig. 6b). Fig. 6(a) shows that any molar ratio value
(within the studied range) can be used to obtain samples with high
ester content (greater than 95%), provided that the reaction temper-
ature is higher than 230 ◦ C. A decrease in the reaction temperature
originated samples with low ester content, indicating once again
the statical significance of the variable temperature. Such behavior
can be confirmed in Fig. 6(b) which indicates that the biodiesel vis-
cosity was absolutely dependent on temperature, and values below
5.0 mm2 /s were obtained at any molar ratio for reaction tempera-
tures higher than 230 ◦ C.
As the fitted models presented in Eqs. (3) and (4) offered a good
estimation of experimental conditions, they were employed to pre-
dict the best conditions for the variables so as to attain samples
having maximum ester content with the lowest viscosity. The opti-
mal conditions for the selected transesterification variables were a
molar ratio of 120:1 and reaction temperature of 250 ◦ C, with an
actual macaw palm biodiesel ester content of 99.2% and viscosity
of 4.5 mm2 /s.

3.4. Effects of reaction time and catalyst concentration on the


response variables

In order to determine the effect of reaction time on ester content


and viscosity of macaw palm biodiesel, experiments were carried
out at various reaction times, ranging from 2 to 4 h, at molar ratio of
Fig. 6. Response surface and contour plots for (a) the ester content and (b) the 120:1, reaction temperature of 250 ◦ C and catalyst concentration of
viscosity vs. molar ratio and temperature.
30 wt.% Nb2 O5 /SO4 . The experimental results are shown in Fig. 7(a).
As it can be seen, the ester content increases as the reaction time
(Y1 ) and viscosity (Y2 ), as a function of the most significant inde- increases, and with a reaction time of 2 h, the biodiesel ester content
pendent variables, are shown in Eqs. (3) and (4). reaches 91.9%. When the reaction took 3 h to be complete, the ester
content reached the value of 98.8%, and a maximum of 99.2% ester
Y1 = 57.76 + 30.82X2 (3) content was achieved when it took 4 h.
Y2 = 12.22 − 2.22X1 2 − 8.3X2 + 2.11X2 2 (4) The reaction time also influences the viscosity of biodiesel. The
experimental results are shown in Fig. 7(b). By increasing the reac-
The fit of models Y1 and Y2 were verified by the coefficients of tion time of the process, the viscosity values tend to decrease. For
determination R2 (Y1) = 0.9288 and R2 (Y2) = 0.9493, indicating that biodiesel synthesized in a reaction time of 2 h, its viscosity was
approximately 93% and 95% of the responses variability could be 6.15 mm2 /s, while for the transesterification reaction that occurred
explained by these models, thus confirming the good correlations for 4 h, it was 4.5 mm2 /s. This fact is justified by the increased for-
between the independent variables. mation of ethyl esters along the reaction time, hence decreasing
The analysis of variance (ANOVA) in Table 4 demonstrates that the amounts of intermediates, such as monoacylglycerol and dia-
the empirical models can be considered statistically significant cylglycerol. Such compounds are partly responsible for increasing
according to the F-test at 95% confidence level. The F-values of Y1 the viscosity of biodiesel (Veljkovic et al., 2015).
and Y2 are 13.05 and 18.76, respectively, greater than F0.05 = 5.05. The catalyst concentration is an essential parameter for process
Fig. 5 shows the student’s t-distribution values in a Pareto chart optimization, thus making it feasible. In this part of the process, a
and plots of predicted versus actual values. According to Fig. 5(a), molar ratio of 120:1, reaction temperature of 250 ◦ C and reaction
variable X2 Temperature(L) has the greatest effect on the ester con- time of 4 h were maintained as constant by varying the catalyst
tent. The value of 16.4 was the greatest among the determined concentrations (5–30 wt.% of macaw palm oil) for the production of
422 L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424

Fig. 7. Effects of reaction time and catalyst concentration on the ester content and viscosity.

Fig. 8. Reusability of the Nb2 O5 /SO4 catalisty in simultaneous esterification and transesterification of macaw palm oil: a) Ester content and b) Viscosity.

ethyl ester. As shown in Fig. 7(c), increase in the catalyst concentra- tents higher than 98.0% were attained for catalyst concentrations
tion produces high ester content. The value of 92.8% ester content over 10%. The highest ester content was achieved from the biodiesel
was obtained using a concentration of 5% catalyst, and ester con- synthesized with catalyst concentration at 30%. The catalyst con-
L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424 423

centration in the transesterification reactions causes a decrease in Acknowledgements


the viscosity of biodiesel. As shown in Fig. 7(d), only the reaction
carried out with 30% catalyst concentration results in biodiesel with The authors gratefully acknowledge the financial support
viscosity level that is below 5.0 mm2 /s. of CNPq (Conselho Nacional de Desenvolvimento Científico e
Data reported from the optimization of reaction parameters pre- Tecnológico-Process Number 404812/2013-9). Leyvison Rafael V.
sented in this work suggest that the ester content and viscosity can da Conceição would especially like to thank the support of
be predicted by the models proposed herein. Coordenação de Aperfeiçoamento de Pessoal de Nível Superior
The physico-chemical properties of macaw palm biodiesel pro- (CAPES).
duced under optimum conditions were determined to assess its
quality. The ester content was 99.2%, which shows a high level References
of product purity. Other important properties that demonstrated
the quality of the synthesized fuel are its monoacylglicerol content Aguieiras, E.C.G., Cavalcanti-Oliveira, E.D., Castro, A.M., Langone, M.A.P., Freire,
D.M.G., 2014. Biodiesel production from Acrocomia aculeata acid oil by
(0.61%) and diacylglicerol content (0.44%). Its kinematic viscosity
(enzyme/enzyme) hydroesterification process: use of vegetable lipase and
(4.5 mm2 /s) and density (872.6 kg/m3 ) also demonstrate its quality. fermented solid as low-cost biocatalysts. Fuel 135, 315–321.
Amani, H., Ahmad, Z., Hameed, B.H., 2014. Highly active alumina-supported Cs-Zr
mixed oxide catalysts for low-temperature transesterification of waste
cooking oil. Appl. Cat. A 487, 16–25.
3.5. Catalyst recycle assays Avhad, M.R., Marchetti, J.M., 2015. A review on recent advancement in catalytic
materials for biodiesel production. Renew. Sust. Energy Rev. 50, 696–718.
Heterogeneous catalysts rather than homogeneous ones are Banković-Ilić, I.B., Stojković, I.J., Stamenković, O.S., Veljkovic, V.B., Hung, Y., 2014.
Waste animal fats as feedstocks for biodiesel production. Renew. Sust. Energy
recommended for industrial applications after considering their
Rev. 32, 238–254.
performance in activity, separation, and recycling. The recycling of Box, G.E.P., Hunter, W.G., Hunter, J.S., 1978. Statistic for Experimenters: An
catalysts is important from economic and environmental points of Introduction to Design, Data Analysis and Model Building. Wiley, New York.
Brunschwig, C., Moussavou, W., Blin, J., 2012. Use of bioethanol for biodiesel
view. When supported and bulk catalysts are used in liquid-phase
production. Prog. Energ. Combust. 38, 283–301.
reactions, there is a possibility that active species are leaching away César, A.S., Almeida, F.A., Souza, R.P., Silva, G.C., Atabani, A.E., 2015. The prospects
into the liquid phases (Zhao et al., 2000). of using Acrocomia aculeata (macauba) a non-edible biodiesel feedstock in
The stability of the Nb2 O5 /SO4 catalyst was assessed for Brazil. Renew. Sust. Energy Rev. 49, 1213–1220.
Ciconini, G., Favaro, S.P., Roscoe, R., Miranda, C.H.B., Tapeti, C.F., Miyahira, M.A.M.,
sequential recyclability in the simultaneous esterification and Bearari, L., Galvani, F., Borsato, A.V., Colnago, L.A., Naka, M.H., 2013. Biometry
transesterification of macaw palm oil to the corresponding ethyl and oil contents of Acrocomia aculeata fruits from the Cerrados and Pantanal
esters by recycling and reuse of the catalyst under the opti- biomes in Mato Grosso do Sul. Brazil. Ind. Crop. Prod. 45, 208–214.
Deshmane, V.G., Adewuyi, Y.G., 2013. Mesoporous nanocrystalline sulfated
mized reaction conditions. After the completion of the reaction and zirconia synthesis and its application for FFA esterification in oils. Appl. Cat. A
between each cycle, the catalysts was recovered from the reaction 462–463, 196–206.
mixture by simple filtration, washed with tert-butyl alcohol, dried Fernandes, D.M., Sousa, R.M.F., Oliveira, A., Morais, S.A.L., Richter, E.M., Muñoz,
R.A.A., 2015. Moringa oleifera: a potential source for production of biodiesel
at 120 ◦ C and subjected to the next cycle. Five recycle runs were and antioxidant additives. Fuel 146, 75–80.
carried out and the results attained at the end of each cycle are Fu, X., Li, D., Chen, J., Zhang, Y., Huang, W., Zhu, Y., Yang, J., Zhang, J., 2013. A
displayed in Fig. 8. microalgae residue based carbon solid acid catalyst for biodiesel production.
Bioresour. Technol. 146, 767–770.
The recycling experiments for simultaneous esterification and
Fu, J., Chen, L., Lv, P., Yang, L., Yuan, Z., 2015. Free fatty acids esterification for
transesterification reactions resulted in 98.65, 98.55, 96.05, 94.95 biodiesel production using self-synthesized macroporous cation exchange
and 93.70% successive ester content (Fig. 8a). For all the five runs, resin as solid acid catalyst. Fuel 154, 1–8.
Gong, S., Lu, J., Wang, H., Liu, L., Zhang, Q., 2014. Biodiesel production via
the rate of reaction was very high at the end of the 3 h reaction
esterification of oleic acid catalyzed by picolinic acid modified
period (with the esters content exceeding 90%). In all about 5.0% of 12-tungstophosphoric acid. Appl. Energy 134, 283–289.
total esters was reduced at the end of the fifth run, suggesting that Hameed, B.H., Lai, L.F., Chin, L.H., 2009. Production of biodiesel from palm oil
the catalyst was stable and no linking of the active specie observed. (Elaeis guineensis) using heterogeneous catalyst: an optimized process. Fuel
Process. Technol. 90, 606–610.
The viscosity results (5.25, 5.30, 5.65, 5.85 and 6.10 mm2 /s) (Fig. 8b) Iha, O.K., Alves, F.C.S.C., Suarez, P.A.Z., Silva, C.R.P., Meneghetti, M.R., Meneghetti,
revealed that the viscosity slowly increased with recycling exper- S.M.P., 2014. Potential application of Terminalia catappa L. and Carapa
iments, which are consistent with the corresponding to the ester guianensis oils for biofuel production: physical-chemical properties of neat
vegetable oils, their methyl-esters and bio-oils (hydrocarbons). Ind. Crop. Prod.
content decreased. 52, 95–98.
Jena, P.C., Raheman, H., Kumar, G.V.P., Machavaram, R., 2010. Biodiesel production
from mixture of mahua and simarouba oils with high free fatty acids. Biomass
Bioenerg. 34, 1108–1116.
4. Conclusion
Kansedo, J., Lee, K.T., 2013. Process optimization and kinetic study for biodiesel
production from non-edible sea mango (Cerbera odollam) oil using response
In this study, by optimizing the reaction parameters, macaw surface methodology. Chem. Eng. J. 214, 157–164.
Liu, L., Wen, Z., Cui, G., 2015. Preparation of Ca/Zr mixed oxide catalysts through a
palm biodiesel was produced with ester content of over 99%
birch-templating route for the synthesis of biodiesel via transesterification.
and viscosity level which is below 5.0 mm2 /s. These values were Fuel 158, 176–182.
obtained with the following reaction conditions: molar ratio of Meher, L.C., Sagar, D.V., Naik, S.N., 2006. Technical aspects of biodiesel production
120:1, reaction temperature of 250 ◦ C, reaction time of 4 h and 30% by transesterification-a review. Renew. Sust. Energy Rev. 10, 248–268.
Melo, R.S., Neto, I.S.S., Pinheiro, R.S., Moura, K.R.M., Silva, F.C., Maciel, A.P., 2012.
catalyst concentration. There are few reports in literature regard- Obtenção de catalisadores heterogêneos de Nióbio modificado para a
ing the use of Nb2 O5 /SO4 as catalyst for biodiesel production. Based conversão de sebo bovino em biodiesel. Cadernos de Pesquisa, São Luís, 19, n.
on the obtained experimental results, it can be concluded that the especial, 116–121.
Montgomery, D.C., 1997. Design and Analysis of Experiments, five ed. Wiley, New
application of Nb2 O5 /SO4 is promising and could be used as an York.
effective heterogeneous acid catalyst for a simultaneous esterifica- Moura, E.F., Motoike, S.Y., Ventrella, M.C., Sá Júnior, A.Q., Carvalho, M., 2009.
tion and transesterification in biodiesel production from reactions Somatic embryogenesis in macaw palm (Acrocomia aculeata) from zygotic
embryos. Sci. Hortic. Amsterdam 119, 447–454.
involving low-quality raw materials and ethanol as acylating agent. Narkhede, N., Brahmkhatri, V., Patel, A., 2014. Efficient synthesis of biodiesel from
The catalyst could be reused for five consecutive provided the excel- waste cooking oil using solid acid catalyst comprising 12-tungstosilicic acid
lent yield of the desired products. The lowering of the yield of the and SBA-15. Fuel 135, 253–261.
Ngee, E.L.S., Gao, Y., Chen, X., Lee, T.M., Hu, Z., Zhao, D., Yan, N., 2014. Sulfated
product after the fifthy recycle may be due to the handling loss of
mesoporous niobium oxide catalyzed hydroxymethylfurfural formation from
the catalyst in successive recycles. sugars. Ind. Eng. Chem. Res. 53, 14225–14233.
424 L.R.V. da Conceição et al. / Industrial Crops and Products 89 (2016) 416–424

Nizah, M.F.R., Taufiq-Yap, Y.H., Rashid, U., Teo, S.H., Nur, Z.A.S., Islam, A., 2014. Stamenkovic, O.S., Velickovic, A.V., Veljkovic, B.V., 2011. The production of
Production of biodiesel from non-edible Jatropha curcas oil via biodiesel from vegetable oils by ethanolysis: current state and perspectives.
transesterification using Bi2 O3 -La2 O3 catalyst. Energy Convers. Manage. 88, Fuel 90, 3141–3155.
1257–1262. Stosic, D., Bennici, S., Pavlovic, V., Rakic, V., Auroux, A., 2014. Tuning the acidity of
Nowak, I., Ziolek, M., 1999. Niobium compounds: preparation, characterization, niobia Characterization and catalytic activity of Nb2 O5 -MeO2 (Me = Ti, Zr, Ce)
and application in heterogeneous catalysis. Chem. Rev. 99, 3603–3624. mesoporous mixed oxides. Mater. Chem. Phys. 146, 337–345.
Ong, H.C., Silitonga, A.S., Masjuki, H.H., Mahlia, T.M.I., Chong, W.T., Boosroh, M.H., Sun, K., Lu, J., Ma, L., Han, Y., Fu, Z., Ding, J., 2015. A comparative study on the
2013. Production and comparative fuel properties of biodiesel from non-edible catalytic performance of different types of zeolites for biodiesel production.
oils: Jatropha curcas, Sterculia foetida and Ceiba pentandra. Energy Convers. Fuel 158, 848–854.
Manage. 73, 245–255. Syazwani, O.N., Rashid, U., Yap, Y.H.T., 2015. Low-cost solid catalyst derived from
Patel, A., Brahmkhatri, V., Singh, N., 2013. Biodiesel production by esterification of waste Cyrtopleura costata (Angel wing shell) for biodiesel production using
free fatty acid over sulfated zirconia. Renew. Energy 51, 227–233. microalgae oil. Energy Convers. Manage. 101, 749–756.
Reshad, A.S., Tiwari, P., Goud, V.V., 2015. Extraction of oil from rubber seeds for Thiruvengadaravi, K.V., Nandagopal, J., Baskaralingam, P., Bala, V.S.S., Sivanesan, S.,
biodiesel application: optimization of parameters. Fuel 150, 636–644. 2012. Acid-catalyzed esterification of karanja (Pongamia pinnata) oil with high
Sani, Y.M., Daud, W.M.A.W., Aziz, A.R.A., 2014. Activity of solid acid catalysts for free fatty acids for biodiesel production. Fuel 98, 1–4.
biodiesel production: a critical review. Appl. Cat. A 470, 140–161. Veljkovic, V.B., Banković-Ilic, I.B., Stamenkovic, O.S., 2015. Purification of crude
Saravanan, K., Tyagi, B., Shukla, R.S., Bajaj, H.C., 2015. Esterification of palmitic acid biodiesel obtained by heterogeneously-catalyzed transesterification. Renew.
with methanol over template-assisted mesoporous sulfated zirconia solid acid Sust. Energy Rev. 49, 500–516.
catalyst. Appl. Cat. B 172–173, 108–115. Vieira, S.S., Magriotis, Z.M., Ribeiro, M.F., Graça, I., Fernandes, A., Lopes, J.M.F.M.,
Shibasaki-Kitakawa, N., Hiromori, K., Ihara, T., Nakashima, K., Yonemoto, T., 2015. Coelho, S.M., Santos, N.Ap.V., Saczk, A.Ap., 2015. Use of HZSM-5 modified with
Production of high quality biodiesel from waste acid oil obtained during edible citric acid as acid heterogeneous catalyst for biodiesel production via
oil refining using ion-exchange resin catalysts. Fuel 139, 11–17. esterification of oleic acid. Micropor. Mesopor. Mat. 201, 160–168.
Silitonga, A.S., Masjuki, H.H., Mahlia, T.M.I., Ong, H.C., Kusumo, F., Aditiya, H.B., Yusoff, M.F.M., Xu, X., Gui, Z., 2014. Comparison of fatty acid methyl and ethyl
Ghazali, N.N.N., 2015. Schleichera oleosa L oil as feedstock for biodiesel esters as biodiesel base stock: a review on processing and production
production. Fuel 156, 63–70. requirements. J. Am. Oil Chem. Soc. 91, 525–531.
Silva, C., Weschenfelder, T.A., Rovani, S., Corazza, F.C., Corazza, M.L., Dariva, C., Zhao, F., Murakami, K., Shirai, M., Arai, M., 2000. Recyclable
Oliveira, J.V., 2007. Continuous production of fatty acid ethyl esters from homogeneous/heterogeneous catalytic systems for heck reaction through
soybean oil in compressed ethanol. Ind. Eng. Chem. Res. 46, 5304–5309. reversible transfer of palladium species between solvent and support. J. Catal.
Sing, K.S.W., Everett, D.H., Haul, R.A.W., Moscou, L., Pierotti, R.A., Rouquerol, J., 194, 479–483.
Siemieniewska, T., 1985. Reporting physisorption data for gas/solid systems
with special reference to the determination of surface area and porosity. Pure
Appl. Chem. 57, 603–619.

You might also like