Está en la página 1de 9

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/317947800

Optimization of biodiesel production by solid


acid catalyst derived from coconut shell via
response surface methodology

Article in International Biodeterioration & Biodegradation · June 2017


DOI: 10.1016/j.ibiod.2017.06.008

CITATIONS READS

0 154

9 authors, including:

Azizah Endut Fathurrahman Lananan


Universiti Sultan Zainal Abidin | UniSZA Universiti Sultan Zainal Abidin | UniSZA
47 PUBLICATIONS 347 CITATIONS 30 PUBLICATIONS 126 CITATIONS

SEE PROFILE SEE PROFILE

Hafizan Juahir Helena Khatoon


Universiti Sultan Zainal Abidin Universiti Malaysia Terengganu
202 PUBLICATIONS 1,275 CITATIONS 37 PUBLICATIONS 402 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Sustainable Aquaponics Recirculation System View project

Effect of Light Pollution View project

All content following this page was uploaded by Azizah Endut on 09 August 2017.

The user has requested enhancement of the downloaded file.


International Biodeterioration & Biodegradation xxx (2017) 1e8

Contents lists available at ScienceDirect

International Biodeterioration & Biodegradation


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

Optimization of biodiesel production by solid acid catalyst derived


from coconut shell via response surface methodology
Azizah Endut a, b, *, Sharifah Hanis Yasmin Sayid Abdullah a, Nur Hanis Mohamad Hanapi a,
Siti Hajar Abdul Hamid a, Fathurrahman Lananan a, Mohd Khairul Amri Kamarudin a,
Roslan Umar a, Hafizan Juahir a, c, Helena Khatoon d
a
East Coast Environmental Research Institute, Universiti Sultan Zainal Abidin, Gong Badak Campus, 21300 Kuala Terengganu, Terengganu, Malaysia
b
Faculty of Innovative Design and Technology, Universiti Sultan Zainal Abidin, Gong Badak Campus, 21300 Kuala Terengganu, Terengganu, Malaysia
c
Faculty of Bioresource and Food Industry, Universiti Sultan Zainal Abidin, Tembila Campus, 22000 Besut, Terengganu, Malaysia
d
School of Fishery Science and Aquaculture, Universiti Malaysia Terengganu, 21030 Kuala Terengganu, Terengganu, Malaysia

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

Article history: The application of solid acid catalyst in biodiesel production eliminates the tedious pretreatment of
Received 30 March 2017 feedstock and eradicates generation of high volume of wastewater. A novel acid catalyst was prepared via
Received in revised form sulfonation of incompletely carbonized coconut shell using concentrated sulfuric acid. The design of
12 June 2017
experiments were performed using four factor-three-level central composite design coupled with
Accepted 12 June 2017
Available online xxx
response surface methodology to evaluate the interaction between two factors in order to determine the
optimum process conditions. A quadratic model was suggested for the prediction of biodiesel yield. The
F-value and p-value of the model were 4.02 and 0.0129, respectively, indicate that the model was sta-
Keywords:
Biodiesel
tistically significant at 95 per cent confidence interval. In addition, R2 value of the model was 0.8364,
Biomass which indicates the acceptable accuracy of the model. From the optimization study, it was found that the
Solid acid catalyst carbonization temperature, carbonization time, sulfonation temperature and sulfonation time were
Transesterification 422  C, 4 h, 100  C and 15 h, respectively. At these optimum conditions the predicted and observed
Response surface methodology biodiesel yield were 88.15 per cent and 88.03 per cent, respectively, which experimentally verified the
accuracy of the model. The use of coconut shell-derived solid acid catalyst in biodiesel production
promotes the sustainable and green way of operations.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction the long branched triglyceride compound into diglyceride and


monoglyceride respectively. As a result, fatty acid methyl ester
Biodiesel is an environmental friendly solution for fuel and (FAME) can be obtained as the main transesterification product
power generation. Due to greater concern on environmental (Abdullah et al., 2017). Numerous studies conducted on biodiesel
pollution caused by over consumption of petroleum product and found that it is non-toxic, less harmful, produce cleaner emission
scarcity of the resources has turned the world's interest into bio- and did not contribute to global warming (Ahmad et al., 2011;
diesel production (Leung et al., 2010). Biodiesel is mainly derived Sumprasit et al., 2017). Hence, biodiesel is considered to be a
from vegetable oils of animal fats through chemical reaction of biodegradable and renewable source of fuels that is far more
methanol that result in the formation of fatty acid methyl ester and practical than fossil fuel (Ashnani et al., 2014).
glycerol as by-product (Borges and Diaz, 2012). The general rep- In a large scale application, biodiesel is commonly produced
resentation of transesterification reaction is illustrated in Fig. 1. using transesterification reaction catalyzed by base catalyst such as
Generally, 3 mol of methanol is required to convert and break down NaOH or KOH due to fast reaction time (Konwar et al., 2014).

* Corresponding author. Faculty of Innovative Design and Technology, Universiti Sultan Zainal Abidin, Gong Badak Campus, Terengganu 21300, Malaysia.
E-mail addresses: enazizah@unisza.edu.my (A. Endut), shanisyasmin@yahoo.com (S.H.Y.S. Abdullah), nhanismh@gmail.com (N.H.M. Hanapi), shah4488@gmail.com
(S.H.A. Hamid), fathur6@gmail.com (F. Lananan), mkhairulamri@unisza.edu.my (M.K.A. Kamarudin), roslan@unisza.edu.my (R. Umar), hafizanjuahir@unisza.edu.my
(H. Juahir), helena.khatoon@umt.edu.my (H. Khatoon).

http://dx.doi.org/10.1016/j.ibiod.2017.06.008
0964-8305/© 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
2 A. Endut et al. / International Biodeterioration & Biodegradation xxx (2017) 1e8

Fig. 1. General transesterification reaction scheme (Abdullah et al., 2017).

However, the process requires pre-treatment of feedstock to reduce prepared by the sulfonation of incompletely carbonized coconut
their free fatty acid (FFA) content to prevent from soap formation in shell. The effect of preparation variables including carbonization
the latter (Tariq et al., 2012). On top of that, generation of high temperature, sulfonation temperature and holding time on the
volume of wastewater as a result of washing and purification steps structure and activity of catalyst were systematically investigated.
present environmental threat to the receiving water (Helwani et al., The catalysts were employed in the transesterification of palm oil
2009). On the other hand, heterogeneous catalyst eliminate the with methanol. On top of that, the optimal conditions for catalyst
generation of wastewater and can be easily separated (Lam et al., preparation was also evaluated.
2010).
Application of biomass derived as solid acid catalyst in biodiesel 2. Materials and methods
production has been widely explored (Dehkhoda and Ellis, 2013;
Ayodele and Dawodu, 2014; Prabhavathi Devi et al., 2014). Solid 2.1. Materials
acid catalyst derived from biomass exhibit high catalytic activity
thus resulted in high biodiesel conversion. For example, Ezebor Coconut shell was locally obtained from daily market. Proximate
et al. (2014) studied the potential of sugarcane baggase and oil analysis was performed in accordance to the ASTM standard to
palm trunk as catalyst with 94 per cent and 93 per cent conversion determine its characteristic including moisture (D2867), ash
of biodiesel could be achieved in the transesterification of palm (D3174), volatile matter (D3175) and fixed carbon. Palm oil
olein. In another study, Li et al. (2014) reported that rice husk char (BURUH) was purchased from local grocery store. Concentrated
showed high catalytic activity (88 per cent) in the trans- sulfuric acid (98 per cent), methanol and sodium hydroxide (NaOH)
esterification of waste cooking oil. Since the solid acid catalyst is were purchased from Merck. Hydrochloric acid (HCl), sodium
mainly derived from waste material, it is basically low cost and chloride (NaCl) and potassium hydroxide (KOH) were purchased
abundance in supply (Emrani and Shahbazi, 2012). Apart from that, from Sigma-Aldrich.
it creates a potential market value for unwanted biomass (Sanjay,
2013). 2.2. Catalyst preparation
Response surface methodology (RSM) is a numerical tool that is
widely used for optimization study in large number of chemical The biomass was primarily washed with distilled water to
processes (Tabaraki et al., 2014; Jabeen et al., 2015). It is a set of remove impurities then sliced into thin pieces (<2 cm) and sun-
mathematical technique that describes the relationship of several dried for 24 h. The sliced biomass was then subjected to oven-
individual variables with one or more response (Witek-Krowiak drying at 100  C for 12 h to remove any moisture. About 20 g of
et al., 2014). It exterminates the one-factor-at-a-time method that the precursor carbon catalyst was incompletely carbonized through
produces large number of experimental run and rather time in a muffle furnace for a certain period of time. The carbonized
consuming. Furthermore, RSM is able to evaluate the interaction sample was then ground in a laboratory mortar to form fine powder
effect of two or more variables to the response that allows better (0.5e1 mm). Sulfonation of activated carbon was conducted by
understanding of the process. Apart from that, RSM allows the mixing of 10 g carbon catalyst with 100 ml H2SO4 in a 250 ml
determination of the best level of factors to optimize the desired conical flask for 15 min. Excess acid was removed. The residual wet
output (Choi et al., 2016). solids (carbon and acid) were then transferred to a ceramic crucible
Coconut shell (Cocos nucifera) belongs to the family of Arecaceae placed in a muffle furnace and heated at desired temperature for
is considered as one of the major agricultural waste in Malaysia. In specified time. After cooling to ambient temperature, the sulfo-
2009, it was estimated around 0.459 million tonnes coconut was nated carbon catalyst was washed with excessive distilled water
produced in Malaysia. This has led to production of significant until the wash water became neutral and dried at 105  C for 10 h in
amount of biomass. The major contributors are coconut shell (0.735 an air-drying oven. Sulfonated carbon catalyst was kept in an air-
million tonnes), coconut husk (0.166 million tonnes), coconut frond tight container to prevent contamination.
(0.103 million tonnes) and empty bunches (0.022 million tonnes)
throughout the year (Sulaiman et al., 2017). Biomass displays great 2.3. Catalyst characterization
potential to be a starting material for industrial process as it is less
expensive and does effect food supplies (Farah et al., 2016). Re-use The surface area and pore volume were determined using
of waste material is greatly desired towards developing environ- Brunauer-Emmett-Teller (BET) and Barrett-Joyner-Halenda (BJH)
mentally safe biodiesel processing. Coconut shell is abundantly equation with the aid of nitrogen gas adsorption/desorption tech-
available and comes from renewable resources (Prauchner and nique using Micromeritics V4.02. Samples were degassed prior to
Roriguez-reinoso, 2012; Sani et al., 2015). On top of that, coconut analysis. Surface morphology of the catalyst with Scanning Electron
shell is rich in carbon, low ash content, high in strength and Microscopy (SEM) was performed using SEM-JEOL unit with 10 kV
hardness make it suitable for catalyst development (Hidayu and accelerating voltage. The total acid density of the prepared catalyst
Muda, 2016). was measured using acid-base back titration method described by
In this study, a novel carbon based solid acid catalyst was Ezebor et al. (2014). Titration was performed three times and the

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
A. Endut et al. / International Biodeterioration & Biodegradation xxx (2017) 1e8 3

average number was reported. The acid density was calculated as Table 1
follows: Experimental range and levels of independent variables for catalyst preparation.

    Independent variables Range and levels


c Hþ ¼ c OH  ðA  BÞ=m (1) 1 0 þ1

Carbonization Temperature (degrees Celsius) (X1) 400 500 600


where c(Hþ) represents the acid density of the catalyst sample in Carbonization Time (h) (X2) 4 6 8
mmol g1; c(OH) represents the concentration of standard NaOH Sulfonation Temperature (degrees Celsius) (X3) 100 150 200
solution in mol L1; A represents the volume of NaOH consumed in Sulfonation Time (h) (X4) 15 20 25

the titration in mL; B represents the volume of NaOH consumed in


the blank titration in mL; and m represents the mass of catalyst
sample in g. Table 2
Central composite design matrix and FAME conversion.

Run Coded factor FAME Conversion (per cent)


2.4. Transesterification reaction X1 X2 X3 X4 Observed Predicted
values values
Transesterification reactions were carried out using method 1 1 1 1 1 78.84 80.00
described in Velasquez-Orta et al. (2013). Reaction was conducted 2 1 1 1 1 88.17 87.15
in a 50 ml centrifuge tubes filled with 2 g of palm oil. The reaction 3 0 0 1 0 83.50 83.36
4 1 1 1 1 83.50 83.07
conditions were kept constant at methanol to oil ratio of 30:1,
5 1 1 1 1 85.06 85.62
catalyst loading of 6 wt per cent, reaction time of 6 h. Centrifuge 6 0 1 0 0 88.95 87.16
tubes were kept in an incubator shaker at a constant temperature of 7 0 0 0 0 81.95 83.13
60  C and a stirring rate of 200 rpm as these provided non limiting 8 1 1 1 1 87.40 86.05
conditions. Once the reaction was concluded tubes were immedi- 9 1 1 1 1 86.50 87.94
10 1 1 1 1 83.50 83.51
ately cooled to room temperature. The biomass was separated from 11 1 1 1 1 85.06 86.19
the liquid by vacuum filtration. Excess methanol in the product is 12 0 0 0 0 81.95 83.13
being removed by fan drying for 12 h prior to determination of acid 13 1 1 1 1 88.17 86.84
value. The experiments were conducted in triplicate. 14 0 1 0 0 86.62 87.62
15 1 1 1 1 82.73 83.27
16 1 1 1 1 81.95 82.03
17 0 0 1 0 85.06 84.41
2.5. Analysis of biodiesel 18 1 1 1 1 86.62 86.17
19 1 0 0 0 78.72 79.81
20 1 0 0 0 80.84 78.97
Potentiometric titration was conducted to determine the acid
21 0 0 0 1 85.06 83.72
value of biodiesel (Wang et al., 2014). The acid value of biodiesel 22 1 1 1 1 81.39 82.81
was calculated as follows: 23 1 1 1 1 81.50 80.17
24 1 1 1 1 84.62 83.57
AV ¼ ðA  BÞ  C  M=W (2) 25 0 0 0 1 81.95 82.50
26 1 1 1 1 84.28 85.70
where AV represents the acid value of biodiesel sample in mg KOH/
g; A represents the volume of KOH standard solution consumed in
the titration in mL; B represents the volume of KOH standard so-
X
k XX X
k
lution consumed in the blank titration in mL; C represents the Y ¼ b0 þ bj Xj þ bij Xi Xj þ bjj Xj2 (4)
concentration of the standard KOH solution in mol L1; M repre- j¼1 i<j j¼1
sents the molecular weight of KOH (56.1 g mol1); and W repre-
sents the weight of biodiesel sample (g). The FFA conversion in where Y is the predicted response, X1, X2 and X3 are the coded
palm oil was determined to evaluate the catalytic activity of the factors, b0 is the intercept term, bi,bij andbii(i,j ¼ 1,2 … n) are the
prepared catalyst as follows: linear coefficients, interactive and quadratic coefficients respec-
tively. Design Expert version 7.0.0 (Stat-Ease Inc. Minneapolis, MN,
Conversion ð%Þ ¼ ðAV0  AV1 Þ=AV  100 (3) USA) software was applied to the model for regression analysis of
the experimental data obtained.
where AV0 and AV1 are the acid value of palm oil and biodiesel
respectively.
3. Results and discussion

3.1. Catalyst characterization


2.6. Experimental design
Table 3 summarizes the chemical compositions of coconut shell
A four-factor-three-level central composite design (CCD) was
employed in this study which generated a total of 26 experimental
runs in addition to two replications of center point. The indepen- Table 3
Proximate analysis of coconut shell as precursor carbon for solid
dent variables were chosen for the optimization study which in-
acid catalyst.
cludes carbonization temperature, carbonization time, sulfonation
temperature and sulfonation time. The coded and actual levels of Parameter Value (per cent)
the independent variables and the CCD arrangement for experi- Moisture content 6.72 ± 0.18
mental design are tabulated in Table 1 and Table 2, respectively. Volatile matter 64.24 ± 1.44
Regression procedure using second order polynomial equation Ash content 1.03 ± 0.09
Fixed Carbon 28.01 ± 1.64
Eq. (4) was used to analyze the experimental data obtained.

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
4 A. Endut et al. / International Biodeterioration & Biodegradation xxx (2017) 1e8

that were determined using proximate analysis. Table 4


Determination of biomass composition is essential as it might Catalyst yield, acid density and textural properties of solid acid catalyst derived from
coconut shell.
affect the structure and textural properties of a solid acid catalyst.
Coconut shell comprises of mainly volatile matter (64.24 per cent) Parameter AC CS SAC CS
and fixed carbon (28.01 per cent). Low ash content was observed in Catalyst Yield(per cent) ND 94.36
coconut shell with only 1.03 per cent. Ash content measures the Total Acid Density(mmol g-1) ND 3.213 ± 0.180
amount of calcium and potassium formed in the activated carbon Sulfonic Acid Density a(mmol g-1) ND 0.507 ± 0.144
BET Surface area b(m2 g-1) 144.6253 24.1703
(Kastner et al., 2012). Low ash content is preferable to prevent any
Average pore volume b(cm3 g-1) 0.0609 0.0115
interference with competitive carbon adsorption as well as adverse Pore size b(Å) 16.8484 19.0204
reaction (Mahamad et al., 2015). On the other hand, coconut shell
AC: Activated carbon, SAC: Solid acid catalyst, CS: Coconut shell, ND: not deter-
also contain small amount of water denoted by 6.72 per cent mined.
moisture content. Surface morphology of carbon catalyst was Carbonized at 500  C, 8 h, Sulfonation with conc. H2SO4, 15 h.
a
analyzed using SEM analysis. Fig. 2 shows the images of coconut Estimated by back titration method.
b
shell derived catalyst during carbonization and sulfonation process. Determined by nitrogen adsorption measurement.

Presence of pores on the carbon surface indicated that high


porosity nature of activated carbon upon carbonization at 500  C as
that correspond the response is summarized in terms of coded
shown in Fig. 2(a). After sulfonation with concentrated H2SO4, the
value.
pore distribution reduces upon contact with strong sulfonating
agent as illustrated in Fig. 2(b). Further analysis was conducted
using nitrogen adsorption measurement to determine BET surface Yð%Þ ¼ 83:13 þ 0:42X1 þ 0:23X2  0:53X3  0:61X4  6:8
area and porosity of the prepared catalyst. The results were sum-  103 X1 X2 þ 1:2X1 X3 þ 1:4X1 X4  0:86X2 X3
marized in Table 4. From the result, it was found that sulfonation
greatly affected the porosity of catalyst confirmed by the reduction þ 0:7X2 X4  0:15X3 X4  3:75X12 þ 4: 26X22 þ 0:75X32
in BET surface area to 24.1703 m2 g1 and 0.0115 cm3 g1 in pore  0:02X42
volume. This might be due to intercalation of -SO3H group in the
(5)
carbon sheet. It results in the reduction of the surface area in
derived catalyst. This is consistent to high sulfonic acid density of where Y is the FAME conversion, X1 is the carbonization tempera-
0.507 mmol g1observed in the catalyst. Similar observation was ture, X2 is the carbonization time, X3 is the sulfonation temperature
reported by Ayodele and Dawodu (2014) on the reduction in surface and X4 is the sulfonation time. The negative and positive signs of
area upon sulfonation that is due to the introduction of -SO3H the regression coefficients indicate the antagonistic and synergistic
group into the carbon framework. In another work by Babadi et al. effects of each variable towards the conversion.
(2016) reported that sulfonation of carbonized beet pulp has Analysis of variance (ANOVA) was conducted to evaluate the
resulted in the reduction in surface area and produced highly active statistical significance of the model equation and the result was
catalyst. The sulfonic acid density falls within the range reported in summarized in Table 5. The model F-value of 4.02 and p-value of
previous studies (Liu et al., 2013; Shah et al., 2015). 0.0129 indicated that the model was statistically significant at 95
per cent confidence interval (p < 0.05). Value of p-value (proba-
3.2. Experimental design and development of regression model bility of error value) less than 0.05 indicate that model terms are
significant, while values greater than 0.1 indicate that the model
In the present work, RSM model based on CCD with four factors terms are not significant. In this study X1X3, X1X4, X21 and X22 are
has been formulated to assess the performance of the developed significant model terms. In addition to that, the coefficient of
catalyst. All the 26 experimental runs were conducted and the re- determination, R2 was evaluated to test the model fitness. Value of
sults were analyzed through multiple regression analysis. The co- R2 closer to 1 is an indication of the stronger model and good
efficients were calculated using regression analysis and tested the prediction of the response. In this case, R2 value was found to be
significance based on p-value. The results of the experimental 0.8364, which indicated that the model could explain 83.64 per
design and FAME conversion in terms of actual and predicted cent of the variability. Apart from that, adequate precision value
values are presented in Table 4. greater than 4 is desirable to measure the signal-to-noise ratio. In
By applying the multiple linear regression analysis on the this instance, adequate value of 6.984 was obtained that indicate
experimental data, a polynomial quadratic equation was found to sufficient value which means this model can be used to navigate the
predict the conversion as presented in Eq. (5). The model equation design space.

Fig. 2. SEM images of coconut shell derived catalyst (a) carbonized at 500 Celcius for 8 h and (b) sulfonated using concentrated H2SO4 at 100 Celcius for 15 h.

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
A. Endut et al. / International Biodeterioration & Biodegradation xxx (2017) 1e8 5

Table 5 The interaction effect of carbonization temperature and sulfo-


Analysis of variance for second order polynomial model. nation temperature is presented in Fig. 4(a). It was found that the
Source Sum of Df Mean F-value P-value FAME conversion increase with increasing carbonization temper-
square Square Prob > F ature at a given sulfonation temperature. The FAME conversion is
Model 161 14 11.5 4.02 0.0129 more profound when higher carbonization temperature is utilized
X1 3.18 1 3.18 1.11 0.3134 until it reaches a maximum level. Then, further increase in the
X2 0.98 1 0.98 0.34 0.5703 carbonization temperature results in reduction in the conversion.
X3 4.96 1 4.96 1.73 0.2148
In other words, this result suggest that low carbonization temper-
X4 6.70 1 6.70 2.34 0.5144
X1X2 1 0.9873 ature improves the pore structure of solid acid catalyst. Hence,
X1 X3 23.11 1 23.11 8.07 0.0161 higher active sites available for reaction and improves overall
X1 X4 31.16 1 31.16 10.88 0.0071 conversion. Previous study stated that synthesis of carbon catalyst
X2X3 11.71 1 11.71 4.09 0.0681
at lower pyrolysis temperature (400e500  C) will result in the
X2X4 7.85 1 7.85 2.74 0.1259
X3X4 0.34 1 0.34 0.12 0.7372
generation of soft aggregated, cross-linked polymer. This carbon-
X12 35.93 1 35.93 12.55 0.0046 ized product is susceptible to be sulfonated at highest degrees thus
X22 46.47 1 46.47 16.23 0.0020 produce highly efficient solid acid catalyst (Kastner et al., 2012). In
X32 1.46 1 1.46 0.51 0.4903 regard to the interaction with sulfonation temperature, high FAME
X42 1 0.9849
conversion was observed at sulfonation temperature of 100  C
Residual
Lack of fit 31.50 10 3.15 compared to 200  C. Low sulfonation temperature favors the
Pure error 0.00 1 0.00 anchoring of -SO3H group onto the carbon framework. As a result,
Total correlation 192.5 24 higher numbers of catalytic sites are available for the reaction to
take place and improve the overall conversion. Previous study
found that prolonged sulfonation time at high temperature con-
The difference between observed and predicted values was dition will hinder the anchoring of -SO3H group on the catalyst
observed using parity plot in Fig. 3. The parity plot indicates a surface. This will reduce the number of acid sites on subsequently
satisfactory correlation between predicted and observed (actual) affected the catalytic performance (Kang et al., 2013).
value. The randomly scattering points close to diagonal line showed Fig. 4(b) shows the biodiesel conversion as function of carbon-
a good agreement between observed and predicted values. ization temperature and sulfonation time under experimental
conditions defined by the CCD designed in Table 2. Similar trend
3.3. Influence of the preparation variables on the FAME conversion was observed as in Fig. 4(a). Higher yield of biodiesel was obtained
when higher carbonization temperature was employed. Never-
Fig. 4 shows the response surface profile of FAME conversion. theless, the conversion is highly pronounced at low sulfonation
The plots were as a function of two variables while the third and time of 15 h. Prolonged sulfonation process resulted in significant
fourth were held at a center points. The graphical representation reduction in the FAME conversion. This is mainly due to the satu-
only show the significant interaction of factors determine by the ration of catalyst surface with acidic group. Additional sulfonation
presence of two non-parallel lines (figure not shown) in main time did not presented any positive effect on the incorporation of
interaction plot. -SO3H group. Thus, it negatively affected the FAME conversion
(Babadi et al., 2016).
Fig. 4(c) illustrates the response of the interactive factor of
carbonization time and sulfonation temperature. The surface plot
indicates that the FAME conversion decreases when prolonged
carbonization was employed at low sulfonation temperature.
Similar outcome was observed when the catalysts undergo sulfo-
nation at 100 and 200  C. On the other hand, Fig. 4(d) shows that
surface plots for the interaction of carbonization time and sulfo-
nation time. From the observation, high FAME conversion was ob-
tained when the catalyst was incompletely carbonized and
sulfonated at short time (4 and 15 h respectively). When higher
conditions were used the activity of the catalyst is negatively
affected presented by the abrupt reduction in FAME conversion.

3.4. Optimization of the catalyst preparation conditions

The optimal catalyst preparation conditions were predicted by


applying numerical optimization using Design-Expert software as
shown in Table 6.
The optimization criteria were set for both variables and
response. The goal was set for maximize the FAME conversion
(response) while minimize was set for all four factors involved. Five
solutions no were suggested by the software and solution with
highest desirability was chosen. The optimal conditions for palm
biodiesel production catalyzed by coconut shell derived solid acid
catalyst were found to be as follows: carbonization temperature
422  C, carbonization time 4 h, sulfonation temperature 100  C and
sulfonation time 15 h. Theoretical conversion to biodiesel predicted
Fig. 3. The parity plot of FAME conversion using coconut shell derived catalyst. under the above conditions was 88.15 per cent.

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
6 A. Endut et al. / International Biodeterioration & Biodegradation xxx (2017) 1e8

Fig. 4. Response surface plot on the FAME conversion for (a) Carbonization temperature:Sulfonation temperature, (b) Carbonization temperature:Sulfonation time, (c) Carbon-
ization time: Sulfonation temperature, and (d) Carbonization time:Sulfonation time. Other factors are held at zero level.

Table 6
Numerical optimization of the catalyst preparation conditions using response surface methodology.

No. Catalyst preparation conditions Conversion Desirability


(per cent)
Carbonization temperature Carbonization Sulfonation temperature Sulfonation
(degrees celsius) time (h) (degrees celsius) time (h)

1 422 4 100.00 15.00 88.15 0.956


2 428 4 100.00 15.03 88.33 0.956
3 425 4 100.55 15.00 88.24 0.955
4 438 4 100.00 15.00 88.63 0.955
5 427 4 100.00 15.08 88.31 0.954

For model verification, experiments were conducted with three values and the error between the two values were small (<1per
independent replicates and the result is summarized in Table 7. cent error for FAME conversion). Thus, it confirmed the efficacy of
The predicted value of 88.15 per cent was closed to the average the quadratic model to predict the optimal conditions for this
observed value of 88.03 ± 0.231 per cent. Therefore, the experi- process.
mental values were in acceptable agreement with the predicted

Table 7
Optimum conditions and validation test for biodiesel conversion using coconut shell derived solid acid catalyst.

No. Catalyst preparation conditions Observation Predicted Error


(per cent) (per cent) (per cent)
Carbonization temperature Carbonization Sulfonation temperature Sulfonation
(degrees celsius) time (h) (degrees celsius) time (h)

1 422 4 100 15 88.25 88.15 0.136


2 422 4 100 15 87.79 88.15 0.476
3 422 4 100 15 88.06 88.15 0.193

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
A. Endut et al. / International Biodeterioration & Biodegradation xxx (2017) 1e8 7

Table 8 biodiesel production by oil esterification and transesterification reactions: a


Properties of palm biodiesel catalyzed by coconut shell derived solid acid catalyst. review. Renew. Sustain. Energ. Rev. 16, 2839e2849.
Choi, J.M., Han, S.K., Kim, J.T., Lee, C.Y., 2016. Optimization of combined (acid-
Parameter Palm Biodiesel standard þthermal) pre-treatment for enhanced dark fermentative H2 production from
biodiesel Chlorella vulgaris using response surface methodology. Int. Biodeter. Biodegr
ASTM D6751 EN 14241
108, 191e197.
Density (g cm-3) 0.8953 0.82e0.90 0.86e0.90 Dawodu, F., Ayodele, O., Xin, J., Zhang, S., Yan, D., 2014. Effective conversion of non-
Kinematic viscosity (mm2 s-1) 5.726 1.9e6.0 3.5e5.0 edible oil with high free fatty acid into biodiesel by sulphonated carbon cata-
Moisture content (per cent) 0.08 0.05 0.075 lyst. Appl. Energ 114, 819e826.
Calorific value (MJ kg-1) 26.82 39.2 NS Dehkhoda, A.M., Ellis, N., 2013. Biochar-based catalyst for simultaneous reactions of
esterification and transesterification. Catal. Today 207, 86e92.
Cetane number 50.7 47 min 51 min
Emrani, J., Shahbazi, A., 2012. A single bio-based catalyst for bio-fuel and bio-diesel.
J. Biotechnol. Biomat 2 (1), 1e7.
Ezebor, F., Khairuddean, M., Abdullah, A.Z., Boey, P.L., 2014. Oil palm trunk and
sugarcane baggase derived solid acid catalysts for rapid esterification of fatty
3.5. Biodiesel properties acids and moisture-assisted transesterification of oils under pseudo-infinite
methanol. Bioresour. Technol. 157, 254e262.
Table 8 shows the properties of palm biodiesel in comparison to Farah, N.H., Salmah, H., Marliza, M., 2016. Effect of butyl methacrylate on properties
of regenerated cellulose coconut shell biocomposite films. P. Chem. 19,
both ASTM and European Union standard. Accordingly, the density 335e339.
of biodiesel is within the range required by both standards. Palm Helwani, Z., Othman, M.R., Aziz, N., Kim, J., Fernando, W.J.N., 2009. Solid hetero-
biodiesel shows viscosity value of 5.726 (mm2 s1) which is com- geneous catalysts for transesterification of triglycerides with methanol: a re-
view. Appl. Catal. A-Gen 363, 1e10.
parable to other reported values (Shah et al., 2015; Dawodu et al.,
Hidayu, A.R., Muda, N., 2016. Preparation and characterization of impregnated
2014). Low viscosity will provide smooth engine operation and activated carbon from palm kernel shell and coconut shell for CO2 capture.
increase engine life (Kumar and Sharma, 2016). Moisture content in P. Eng. 148, 106e113.
Jabeen, H., Iqbal, S., Anwar, S., Parales, R.E., 2015. Optimization of profenofos
palm biodiesel obtained was found to be slightly higher than those
degradation by a novel bacterial consortium PBAC using response surface
of the standards. This may be ascribed to the presence of traces of methodology. Int. Biodeter. Biodegr 100, 89e97.
water molecule that is not completely dried off prior to the analysis. Kang, S., Ye, J., Chang, J., 2013. Recent advances in carbon-based sulfonated catalyst:
The calorific value of the biodiesel showed lower value than the preparation and application. Int. Rev. Chem. Eng. 5 (2), 133e144.
Kastner, J.R., Miller, J., Geller, D.P., Locklin, J., Keith, L.H., Johnson, T., 2012. Catalytic
specification required by the standard. This can be explained by the esterification of fatty acids using solid acid catalysts generated from biochar and
presence of unreacted triglyceride and methanol in the biodiesel activated carbon. Catal. Today 190, 122e132.
product. This might affect the combustion of biodiesel. It is advis- Konwar, L.J., Boro, J., Deka, D., 2014. Review on latest developments in biodiesel
production using carbon-based catalysts. Renew. Sustain. Energ. Rev. 29,
able to refine the biodiesel via distillation prior to analysis to 546e564.
remove any unwanted compound and impurities. On the other Kumar, M., Sharma, M.P., 2016. Optimization of transesterification of Chlorella
hand, the cetane number of the derived biodiesel fulfill the protothecoides oil to biodiesel using box-Behnken design method. Waste
Biomass Valor 7 (5), 1105e1114.
requirement set by both standards. A high value of cetane number Lam, M.K., Lee, K.T., Mohamed, A.R., 2010. Homogeneous, heterogeneous and
indicates faster fuel ignition in engine (Oliveira and Da Silva, 2013). enzymatic catalysis for transesterification of high free fatty acid oil (waste
cooking oil) to biodiesel: a review. Biotechnol. Adv. 28, 500e518.
Leung, D.Y.C., Wu, X., Leung, M.K.H., 2010. A review on biodiesel production using
4. Conclusions catalyzed transesterification. Appl. Energ. 87, 1083e1095.
Li, M., Zheng, Y., Chen, Y., Zhu, X., 2014. Biodiesel production from waste cooking oil
The present study demonstrates successful application of co- using a heterogeneous catalyst from pyrolyzed rice husk. Bioresour. Technol.
154, 345e348.
conut shell-derived solid acid catalyst as an efficient catalyst for Liu, T., Li, Z., Li, W., Shi, C., Wang, Y., 2013. Preparation and characterization of
transesterification of palm oil. The effect of catalyst preparation biomass carbon-based solid acid catalyst for the esterification of oleic acid with
conditions on the activity of catalyst were extensively studied methanol. Bioresour. Technol. 133, 618e621.
Mahamad, M.N., Zaini, M.A.A., Zakaria, Z.A., 2015. Preparation and characterization
through RSM. It was found that carbonization temperature posi- of activated carbon from pineapple waste biomass for dye removal. Int. Bio-
tively influenced the yield and exhibited significant effect on the deter. Biodegr. 102, 274e280.
response. The catalyst can be prepared at carbonization tempera- Oliveira, L., Da Silva, M., 2013. Relationship between cetane number and calorific
pia visceral oil blends with mineral diesel. Renew.
ture (450  C) and low sulfonation temperature (100  C). This value of biodiesel from Tila
Energ. Power Qual. J. 687e690.
study uncovers another low-cost catalyst resource for catalyst Prabhavathi Devi, B.L.A., Vijai Kumar Reddy, T., Vijaya Lakshmi, K., Prasad, R.B.N.,
equipped with simpler production routes and mild process 2014. A green recyclable SO3H-carbon catalyst derived from glycerol for the
conditions. production of biodiesel from FFA-Containing karanja (Pongamia glabra) oil in a
single step. Bioresour. Technol. 153, 370e373.
Prauchner, M.J., Roriguez-reinoso, F., 2012. Chemical versus physical activation of
Acknowledgement coconut shell: a comparative study. Micropor. Mesopor. Mater 152, 163e171.
Sani, Y.M., Raji, A.O., Alaba, P.A., Abdul Aziz, A.R., Wan Daud, W.M.A., 2015. Palm
catalysts for biodiesel. Bioresources 10 (2), 3394e3408.
This work was financially supported by Ministry of Higher Ed- Sanjay, B., 2013. Heterogeneous catalyst derived natural resources for biodiesel
ucation of Malaysia [Project number RR067]. production: a review. Res. J. Chem. Sci. 3 (6), 95e101.
Shah, K., Parikh, J., Maheria, K., 2015. Use of sulfonic acid-functionalized silica as
catalyst for esterification of free fatty acids (FFA) in acid oil for biodiesel pro-
References duction: an optimization study. Res. Chem. Int. 41 (2), 1035e1051.
Sulaiman, S.A., Roslan, R., Inayat, M., Naz, M.Y., 2017. Effect of blending ratio and
Abdullah, S.H.Y.S., Hanapi, N.H.M., Azid, A., Umar, R., Juahir, H., Khatoon, H., catalyst loading on co-gasification of wood chips and coconut waste. J. Energ.
Endut, A., 2017. A review of biomass-derived heterogeneous catalyst for a Ins. http://dx.doi.org/10.1016/j.joei.2017.05.003 Available online on 13 May,
sustainable biodiesel production. Renew. Sustain. Energ. Rev. 70, 1040e1051. 2017 (in press).
Ahmad, A.L., Mat Yasin, N.H., Derek, C.J.C., Lim, J.K., 2011. Microalgae as a sustainable Sumprasit, N., Wagie, N., Glanpracha, N., Annachhatre, A.P., 2017. Biodiesel and
energy source for biodiesel production: a review. Renew. Sustain. Energ. Rev. 15, biogas recovery from spirulina plantesis. Int. Biodeter. Biodegr 119, 196e204.
584e593. Tabaraki, R., Nateghi, A., Ahmady-Asbchin, S., 2014. Biosorption of lead (II) ions
Ashnani, M.H.M., Johari, A., Hashim, H., Hasani, E., 2014. A source of renewable sargassum ilicifolium: application of response surface methodology. Int. Bio-
energy in Malaysia, why biodiesel? Renew. Sustain. Energ. Rev. 35, 244e257. deter. Biodegr 93, 145e152.
Ayodele, O.O., Dawodu, F.A., 2014. Production of biodiesel from Calophyllum ino- Tariq, M., Ali, S., Khalid, N., 2012. Activity of homogeneous and heterogeneous
phyllum oil using a cellulose-derived catalyst. Biomass Bioenerg 70, 239e248. catalysts, spectroscopic and chromatographic characterization of biodiesel: a
Babadi, F.E., Hosseini, S., Soltani, S.M., Aroua, M.K., Shamiri, A., Samadi, M., 2016. review. Renew. Sustain. Energy Rev. 16, 6303e6316.
Sulfonated beet pulp as solid acid catalyst in one-step esterification of industrial Velasquez-Orta, S.B., Lee, J.G.M., Harvey, A.P., 2013. Evaluation of FAME production
palm fatty acid distillate. J. Am. Oil Chem. Soc. 93, 319e327. from wet marine and freshwater microalgae by in situ transesterification.
Borges, M.E., Diaz, L., 2012. Recent developments on heterogeneous catalysts for Biochem. Eng. J. 76, 83e89.

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
8 A. Endut et al. / International Biodeterioration & Biodegradation xxx (2017) 1e8

Wang, L., Dong, X., Jiang, H., Li, G., Zhang, M., 2014. Preparation of a novel carbon Application of response surface methodology and artificial neural network in
based solid acid from cassava stillage residue and its use for the esterification of modelling and optimization of biosorption process. Bioresour. Technol. 160,
free fatty acids in waste cooking oil. Bioresour. Technol. 158, 392e395. 150e160.
Witek-Krowiak, A., Chojnacka, K., Podstawczyk, D., Dawiec, A., Pokomeda, K., 2014.

Please cite this article in press as: Endut, A., et al., Optimization of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology, International Biodeterioration & Biodegradation (2017), http://dx.doi.org/10.1016/j.ibiod.2017.06.008
View publication stats

También podría gustarte