You are on page 1of 10

Industrial Crops & Products xxx (xxxx) xxxx

Contents lists available at ScienceDirect

Industrial Crops & Products


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

Biodiesel production using a renewable mesoporous solid catalyst


Bishwajit Changmaia, Putla Sudarsanamb, Lalthazuala Rokhuma,*
a
Department of Chemistry, National Institute of Technology Silchar, Silchar, 788010, Assam, India
b
Center for Sustainable Catalysis and Engineering, Faculty of Bioscience Engineering, KU Leuven, Celestijnenlaan 200f, Heverlee, Belgium

A R T I C LE I N FO A B S T R A C T

Keywords: Heterogeneous solid catalysts have been largely developed for biodiesel production, because of their attractive
Fatty acid methyl ester acid-base properties, strong hydrothermal stability, and efficient recovery/reusability. In this framework, de-
Soybean oil veloping bio-waste derived heterogeneous catalysts has attracted immense attention for several catalytic ap-
Orange peel ash plications, owing to their inexpensive, high abundance, non-toxic, and adequate acid-base properties. In the
Reusability
present work, we investigated the catalytic performance of a biomass-derived orange peel ash (OPA), which
Biomass
contains a porous structure, as a raw heterogeneous catalyst for the transesterification of soybean oil to bio-
diesel. About 98 % conversion of soybean oil to biodiesel was obtained under the optimized reaction conditions
i.e., 6:1 methanol:oil ratio, 7 wt. % catalyst loading, 7 h reaction time at ambient reaction temperature, which
ascribed to the presence of abundant basic sites in the developed OPA catalyst. The catalyst can be reused for five
successive cycles and shows good stability towards the biodiesel production.

1. Introduction of triglycerides, and is mainly produced from vegetable oils, waste


cooking oils, non-edible oils, etc. (Agarwal and Das, 2001; Ma et al.,
Increasing rate of energy consumption and abnormal climate 2018).
change, due to the rapid population growth and the emission of en- In particular, biodiesel is largely produced by transesterification of
vironmental pollutants (e.g., CO2), respectively, have led to the search long chain fatty acids using homogeneous catalysts, like KOH, NaOH,
for new, sustainable, and renewable energy sources. Presently, non- etc. (Likozar et al., 2016) Although homogeneous catalysts showed
renewable fossil fuels are the major source for the production of energy, good activity in terms of conversions and yields, their toxic and cor-
fuels, and chemicals (Sahaym and Norton, 2008; Putla et al., 2019). rosive nature, problems associated with product purification and as
Concurrently, the resources of fossil fuels are declining gradually as the well as with catalyst recovery and its reusability limit their applications
consumption rate is increasing drastically to meet our energy needs. In in chemical industry. Thus, the overall cost of biodiesel production
addition, fossil fuels emit various greenhouse gases like CO2, S, NOx, increases hugely (Likozar and Levec, 2014; Mardhiah et al., 2017). In
etc., which are the key factors for global warming, a major threat facing recent years, heterogeneous catalysts have attracted much interest as a
humankind in the 21st century (Agarwal, 2007; Ambursa et al., 2016; potential substitute to homogenous catalysts because of their inter-
Caliskan, 2017). To minimize the emission of greenhouse gas and to esting properties, such as non-toxic, easy separation and recyclable, and
meet increasing energy demand, the use of an alternative renewable, non-corrosive nature (Sharma et al., 2011). Several heterogeneous
sustainable, and safer energy source is essential (Panwar et al., 2011). catalysts, such as Zn@CaO (Kumar and Ali, 2013), CaO@SnO2 (Xie and
Biofuels, like biodiesel and biogas are the excellent sources to meet Zhao, 2013), biguanide-functionalized hydroxyapatite encapsulated-γ-
the future energy demand (Kumar et al., 2018; Gilbert and Perl, 2008). Fe2O3 nanoparticles (Lee et al., 2015), CaO@La2O3 (Xie et al., 2017)
Currently, biodiesel, also known as fatty acid alkyl esters (FAMEs), has etc. were reported for the production of biodiesel. However, the che-
attracted much attention due to its biodegradability, low emission of mical synthesis of such catalysts is quite complicated and often involves
CO2 and CO, low sulphur content, renewable, and environmental drastic reaction conditions and multi steps, resulting in increased cat-
friendly nature. Moreover, biodiesel is compatible to normal diesel alyst synthesis costs (Correia et al., 2014). In this context, the use of
engine without any modification (Kulkarni et al., 2006; Kralova and waste biomass-derived material as a heterogeneous solid catalyst is an
Sjöblom, 2010; Maeda et al., 2011; Ong et al., 2011). It is an ester- excellent choice for a more sustainable biodiesel production
ification product of long chain fatty acids or transesterification product (Sudarsanam et al., 2018; Shan et al., 2018).


Corresponding author.
E-mail address: rokhum@che.nits.ac.in (L. Rokhum).

https://doi.org/10.1016/j.indcrop.2019.111911
Received 8 July 2019; Received in revised form 22 October 2019; Accepted 26 October 2019
0926-6690/ © 2019 Elsevier B.V. All rights reserved.

Please cite this article as: Bishwajit Changmai, Putla Sudarsanam and Lalthazuala Rokhum, Industrial Crops & Products,
https://doi.org/10.1016/j.indcrop.2019.111911
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

In spite of its wide availability and renewable nature, little attention electron microscope), XPS (X-ray photoelectron spectroscopy), EDX
has been directed towards modification of the waste biomass for im- (energy dispersive X-ray spectroscopy), BET (Brunauer-Emmett-Teller)
proving its catalytic performance. A few studies have reported on the and TGA (thermogravimetric analysis) techniques. The FTIR spectra
modification of waste biomass via surface functionalization using con- was taken in a Spectrum 100 spectrophotometer. HRTEM was taken in
ventional bases and acids. In this regard, functionalized coconut husk an electron microscope JEM-2100 instrument. SEM and element dis-
(Vadery et al., 2014), palm trunk (Ezebor et al., 2014), wood ash tribution of the catalyst i.e., EDX were taken in JSM-6360 (JEOL) in-
(Sharma et al., 2012) and rice husk (Zhao et al., 2018) were reported as strument. XRF analysis was performed on Bruker S4 Poinner instru-
heterogeneous catalysts for the production of biodiesel. ment. XPS analysis was performed using a ESCALAB Xi + instrument.
Although functionalized bio-waste derived carbon-based solid cat- XRD analysis was performed on an ULTIMA IV. The N2 ad-
alyst has its own merits, functionalization of waste biomass involves sorption–desorption analysis was carried out using a Micromeritics
some drawbacks, such as high cost, high temperature for carbonization ASAP 2010. The TGA analysis was performed using SII 6300 EXSTAR
and tedious preparation steps, which limits its applications. Hence, instrument under N2 atmosphere.
using a raw biomass waste as an effective and recyclable heterogeneous
catalyst is essential for sustainable biodiesel production in industrial 2.3. Transesterification of soybean oil
scale. In this line, limited literature reported the application of raw
biowaste (without functionalization) as a heterogeneous catalyst in the 14 g of soybean oil was taken in a round bottom flask, followed by
production of FAME (Betiku et al., 2016; Betiku and Ajala, 2014; the addition of 4 mL methanol and 1 g catalyst (7 wt. % w.r.t soybean
Gohain et al., 2017). Recently, our research group has successfully oil) and then, stirred at ambient temperature. The progress of the
utilized biowaste, banana (Musa acuminata species (available in transesterification reaction was observed by taking thin layer chroma-
Southeast Asia) peel ash as a heterogeneous catalyst for the production tography (TLC). The reaction was completed after 7 h, as confirmed by
of biodiesel (Pathak et al., 2018). The catalyst shows excellent activity TLC. The catalyst was recovered by centrifugation at 3500 rpm for
with a very high oil conversion of 98.95 %. Despite the high activity of 25 min. Biodiesel and glycerol were separated via separating funnel
the catalyst, the reusability of the catalyst limits its application in in- followed by evaporation of excess methanol via rotary evaporator.
dustrial scale.
Among all the citrus fruit produced annually, about 75 % is con- 2.4. Characterization of FAME
tributed by orange (Citrus sinensis) (Li et al., 2008). The global orange
production is about 49.3 million metric tons in the year 2017-2018. The synthesized FAME was characterized by 1H NMR, 13C NMR and
Therefore, the disposal of waste orange peel is a major concern. GC–MS (gas chromatography-mass spectrometry) analytic techniques.
1
Nonetheless, orange peel is considered as a valuable agricultural bio- H NMR and 13C NMR was carried out in Mercury Plus 300 MHz NMR
mass waste, composed of cellulose, carbohydrates, fats, etc. (Chutia spectrometer. GC–MS analysis was carried out in JEOL AccuTOF GCV
et al., 2009). Several reports had been already published regarding the instrument.
applications of orange peel in the field of pollution abatement, like
removal of heavy metals from wastewater and production of solid 2.5. Reusability of the catalyst
biofuels (Biswas et al., 2008; Santos et al., 2015). Until now, there are
no literature reports on the production of FAME by using orange peel Prior to catalyst reusability test, the catalyst was recovered using
ash. centrifugation, washed with methanol and kept in an oven at 100 °C for
In this present work, in continuation of our interest in the devel- 5 h to dry out the catalyst. Reusability test was investigated under the
opment of economic heterogeneous catalysts, green synthetic meth- optimized reaction conditions i.e., 6:1 methanol:oil ratio, 7 wt.% cata-
odologies (Rajkumari et al., 2017; Das et al., 2016) and renewable lyst, 7 h and room temperature.
energy (Pathak et al., 2018; Laskar et al., 2018a; Chatterjee et al., 2017;
Laskar et al., 2018b), we used waste biomass (orange peel ash) as a 3. Results and discussion
heterogeneous catalyst for the conversion of soybean oil to biodiesel at
room temperature (RT). The investigated orange peel ash catalyst is 3.1. Characterization of orange peel (OP) and orange peel ash (OPA)
highly abundant, renewable, and cost-effective. In addition, it is easily catalyst
prepared by burning the dried orange peel and can be used for biodiesel
production without further modification. EDX was performed to investigate the elemental composition of the
orange peel (SI, Fig. S1) and the OPA catalyst (Fig. 1A). EDX data shows
2. Materials and methods the presence of K, Ca, O, C, P, Mg, and Cl in both samples. The ele-
mental distribution along with their atomic wt.% are shown in Table 1.
2.1. Chemicals used The wt.% of K and Ca increased from 0.18 % and 0.03 % (in the dried
orange peel) to 14.67 % and 7.34 % (in the OPA), respectively, due to
Soybean oil was purchased from the local market in Silchar, Assam, the burning of organic matters present in the dried orange peel to
India. Methanol (analytical grade) was purchased from Merck, India. produce OPA. Due to the presence of basic sites (e.g., K and Ca) OPA
The chemicals were used without further purification. could show good catalytic activity in the base-catalyzed reactions like
transesterification reaction of soybean oil to biodiesel.
2.2. Preparation and characterization of the catalysts TGA analysis was performed under N2 atmosphere to verify the
stability of the OPA catalyst with temperature. TGA thermogram dis-
The fresh orange were collected from Silchar, Assam, India. The plays an initial weight loss of 6 % up to 181 °C, due to the removal of
peels were separated, washed properly with distilled water, cut into adsorbed moisture from the catalyst. The further weight loss of the
small pieces and then dried under sun for 3 days. Afterwards, about catalyst under high temperature might be due to the oxidation of the
50 g of dried orange peels were burnt in open air for 30 min and ground carboneous materials in the form of CO and CO2 (Chouhan and Sarma,
to produce orange peel ash (OPA) weighing 4.94 g. The structure, 2013) (Fig. 1B).
morphology, crystallinity and composition of the resultant catalyst To investigate the pore volume, pore diameter and BET surface area
were characterized by FTIR (Fourier transform infrared spectroscopy), of the prepared catalyst, N2 adsorption-desorption analysis was carried
XRF (X-ray fluorescence), XRD (powder X-ray diffraction), SEM (scan- out. Fig. 1C shows N2 adsorption–desorption isotherm of type IV, which
ning electron microscope), HRTEM (high resolution transmission indicates that the catalyst is mesoporous in nature. BJH model (Fig. 1D)

2
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

Fig. 1. EDX spectrum of OPA catalyst (A), TGA thermogram (B), N2 adsorption-desorption isotherm (C), and BJH pore size distribution (D).

Table 1 Table 2
EDX data for elemental composition of OP and OPA catalyst. Composition of OP and OPA as determined by XRF analysis.
Sl. No Element OP (Atomic %) OPA (Atomic %) Sl. No Compound formula OPA (wt. %)

1 O 42.00 40.86 K2 O 51.64


2 C 57.73 32.50 2 CaO 25.67
3 K 0.18 14.67 3 SiO2 13.24
4 Ca 0.03 7.34 4 MgO 4.76
5 Mg 0.04 2.02 5 P2O5 2.95
6 P 0.01 1.57 6 Na2O 1.81
7 Fe2O3 0.070
8 MnO 0.033
9 TiO2 0.114
depicted high pore density in the mesoporous region centered at
2.85 nm. The pore volume and BET surface area of the catalyst were
found to be 0.428 cc/g and 605.60 m2/g, respectively. Due to the high
FTIR analysis was performed to detect the functional groups present
BET surface area and mesoporous nature of the catalyst, it may show
in the OPA (Fig. 2). The band at 3251 cm-1 represents the stretching
good catalytic activity towards the production of FAME from soybean
frequency of OeH bond, due to the absorption of water molecules from
oil.
the atmosphere. The characteristic peaks at 1447 cm-1, 1064 cm-1 and
To investigate the composition of OPA catalyst, XRF analysis was
876 cm-1 were attributed to carbonate stretching and bending vibration.
performed. The composition of the catalyst listed in Table 2, shows that
The absorption peaks at 708 cm-1 and 608 cm-1 were attributed to KeO
alkaline K2O and CaO are the major components present in the catalyst
and CaOe stretching frequency, respectively.
(OPA). Thus, K2O and CaO could play a major catalytic role in the
XRD analysis was performed to investigate the crystalline compo-
production of biodiesel from soybean oil.
nents present in the catalyst. XRD data (Fig. 3) show the presence of
To investigate the basicity of the catalyst, Hammett indicator
K2O, CaO, K2CO3, CaCO3, MgO, SiO2 etc. The peaks for K2O and K2CO3
method was performed and it was found to be 9.8 < H_ < 12.2 for OPA
were observed at 2θ = 30.106, 41.11, 33.72, and 42.160° (JCPDS file
catalyst. The high basicity of catalyst can be ascribed to the presence of
no. 77-2178 and JCPDS file no. 71-1466). The peaks for CaO and CaCO3
K2O, CaO, MgO, etc., as observed from XRF data (Table 2). Hence, the
were observed at 2θ = 26.007, 39.83, 28.41, and 48.75° (JCPDS file no.
presence of high basic sites can lead to improved catalytic activity of
28-0775 and JCPDS file no. 72-1214). The existence of SiO2 and MgO in
OPA catalyst towards biodiesel synthesis via transesterification reac-
the catalyst was confirmed by the peaks at 2θ = 33.72, 54.37, 43.60,
tion.
and 57.83° (JCPDS file no. 89-3609 and JCPDS file no. 57.83). Thus,

3
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

Fig. 2. FTIR spectrum of fresh OPA.

Table 3, which supports the results obtained from both XRF (Table 2)
and EDX data (Table 1).

3.2. Characterization of synthesized biodiesel

Once the transesterification reaction to produce FAME was com-


pleted, the resultant product was confirmed by performing 1H NMR and
13
C NMR analysis. Fig. 6 shows the 1H NMR spectrum of soybean oil
(Fig. 6A) and FAME (Fig. 6B). The appearance of peaks at 4.31 ppm and
5.36 ppm confirms the presence of glyceridic protons and olefinic
protons (Fig. 6A) in soybean oil. While the appearance of a peak at
3.68 ppm and disappearance of peak at 4.31 ppm confirms the forma-
tion of FAME (Fig. 6B). The singlet peak at 3.68 ppm represents the
methoxy protons of FAME and the triplet peak at 2.32 ppm can be at-
tributed for α−CH2 protons. From 13C NMR, the appearance of peak at
51.41 ppm for −OCH3 carbon (SI, Fig. S2B) and disappearance of
glyceridic carbon peaks at 68.89 ppm and 62.08 ppm (SI, Fig. S2A)
confirms the formation of FAME. The percentage yield of soybean oil to
Fig. 3. XRD spectrum of OPA catalyst. FAME can be determined by using the integration value for methoxy
protons and integration value for α−CH2 protons in the equation de-
oxides and carbonates of K and Ca are the dominating components in rived by Knothe and Kenar (Eq. 1). The percentage yield of FAME was
the catalyst, which supports the XRF data (Table 2). calculated and it is found to be 98 % at optimized reaction conditions as
TEM and SEM were performed to investigate the external surface discussed in the following sections.
morphology and structure of OPA (Fig. 4). The TEM and SEM images 2AMe
showed well-arranged microporous and mesoporous structure with C = 100 ×
3ACH2 (1)
some spongy nature of the OPA catalyst.
`To further investigate the elemental composition of the catalyst, Here, C is the percent conversion of soybean oil, AMe is the integration
XPS analysis was performed. Fig. 5 shows the wide range XPS spectrum area of the methoxy protons and ACH2 is the integration area of the
along with deconvoluted spectra for C 1s, O 1s, K 2p, Ca 2p and Mg 1s. α−CH2 protons.
The catalyst comprises of C, O, K, Ca, Mg, Na, P, etc., as shown in The chemical composition of the synthesized FAME was obtained by
Fig. 5A. The deconvoluted spectrum for C 1s (Fig. 5B) shows two peaks GC–MS analysis (Fig. 7). Furthermore, the quantitative measurement of
at 283.92 eV and 288.53 eV, which can be ascribed to CeC and CO] FAME components was calculated by using the peak areas of FAME and
bonds, respectively. Fig. 5C shows two peaks for K 2p with binding internal standard peak area (methyl heptadecanoate, C17:0). The ob-
energies of 291.73 eV and 294.45 eV, which are ascribed to K in the tained data was presented in Table 4.
form of metal oxides and metal carbonates. Fig. 5D shows two peaks for The properties of soybean oil and synthesized FAME were measured
Ca 2p with binding energies of 345.94 eV and 349.48 eV, which con- by using standard methods depicted in the Table 5. The properties of
firms the presence of CaO and CaCO3. Mg 1 s spectrum contains one the synthesized biodiesel meets the standard of ASTM D 6751.
peak with binding energy of 1303.72 eV, confirming the presence of
MgO (Fig. 5E). The O 1s spectrum shows two peaks at 530.12 eV and 3.3. Influence of reaction parameters on soybean oil to FAME conversion
532.07 eV, which can be ascribed to the presence of oxides and car-
bonates of metals (Fig. 5F). The atomic wt. % of the major elements, 3.3.1. Influence of catalyst loading
obtained from XPS analysis, present in the OPA catalyst are listed in The effect of catalyst loading in the conversion of soybean oil to

4
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

Fig. 4. TEM images (A, B) and SEM images (C and D) of OPA catalyst.

Fig. 5. (A) Wide range XPS spectrum and (B), (C), (D), (E), (F) are the deconvoluted peaks of C1s, K 2p, Ca 2p and Mg 1s and O 1s of OPA catalyst, respectively.

5
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

Table 3 run. In contrast, orange peel ash catalyst shows very high activity and
XPS data for elemental composition and their corresponding amount (%) of the stability towards the FAME production at room temperature and can be
catalyst. reused for 5 catalytic cycles with high percentage conversion.
Sl. No Elements Atomic wt%
3.5. Reusability test of OPA catalyst
1 O1 s 39.45
2 C1 s 33.07
Catalyst reusability is a very important factor to use the catalyst in
3 K2p 13.53
4 Ca2p 7.56 commercial scale and industrial processes. Reusability test was per-
5 Mg1 s 2.11 formed for 5 successive cycles under the optimized reaction conditions
6 P2p 1.23 i.e., 6:1 methanol:oil ratio with 7 wt.% at RT for 7 h (Fig. 11). The
conversion of FAME after each cycle was recorded and it was observed
that conversion decreases in each cycle and after the 5th cycle the
FAME (Fig. 8) was performed by varying the catalyst amount from 2 wt. conversion of FAME was only 85 %. This might be due to the slight
% to 13 wt. % with 1:6 oil:methanol ratio at RT. When catalyst loading leaching of active sites of the catalyst during the reaction and/or weight
was 2 wt. % the conversion was 76.5 %. Increasing the catalyst loading loss of the catalyst during recovery/purification steps.
to 7 wt. %, about 98 % conversion of oil to FAME was found. Further The XPS analysis of the recovered catalyst after 5th cycle was per-
increase in catalyst loading, a decrease in the conversion of oil was formed. The wide range XPS spectrum and deconvoluted spectra of K
observed. This might be due to the initiation of saponification reaction 2p, O 1s and Ca 2p (SI, Fig. S3), show a significant drop of K, O and Ca
at higher catalyst loadings, which also makes the reaction mixture more atomic wt. % from 13.53 %, 39.45 % and 7.56 % to 11.32 %, 34 % and
viscous, resulting in severe mass transfer limitations (Pathak et al., 5.95 %, respectively. The slight loss of K and Ca leads to the slight
2018; Hsiao et al., 2011). Thus, the conversion of oil decreases at higher decrease of catalytic activity towards the transesterification reaction
catalyst loadings and therefore, 7 wt. % is the optimal catalyst loading and hence, the yield of FAME was considerably decreased to 85 % after
for achieving high conversion of soybean oil to FAME. the 5th catalytic cycle. Thus, we can conclude that K and Ca play a
major catalytic role in the production of FAME.
3.3.2. Influence of methanol:oil ratio To investigate the morphology and structure of the recovered cat-
The significant effect of oil:methanol ranging from 1:3 to 1:12 on alyst after the 5th cycle, SEM and TEM analyses were performed (SI,
the production of FAME (Fig. 9) was investigated with catalyst loading Fig. S4), which show an irregular surface, morphology and structure of
of 7 wt.%, 7 h reaction time and at ambient temperature. 1:3 of oil:- the catalyst after the 5th cycle.
methanol ratio gives only 74 % conversion of FAME. While, 1:6 of To investigate the functional groups present in the recovered cata-
oil:methanol ratio drastically increases the conversion of FAME to 98 lyst after the 5th cycle, FTIR analysis was performed (SI, Fig. S5 A). An
%. Further increase in the oil:methanol ratio upto 1:12; it is expected to absorption peak was observed at 3328.31 cm-1, which is due to the
exponentially increase the conversion of FAME. However, the conver- presence of moisture absorbed by the catalyst. The peak at 2924.92 cm-
sion decreased to 96 % for 1:12 oil:methanol molar ratio since increase 1
could be attributed for C–H stretching of Ca(OCH3)2, which is formed
in the amount of methanol will drive the reaction to backward direction in little amount on the catalyst surface, due to the reaction of CaO with
to form monoglyceride and diglyceride as the transesterification is a methanol and glycerin (Gohain et al., 2017). The peaks at 1430.78 cm-1,
reversible reaction (Laskar et al., 2018a). Thus, 1:6 oil:methanol molar 1069.53 cm-1 and 874.86 cm-1 could be attributed for carbonate
ratio is the optimized condition for the maximum conversion of soy- stretching and bending vibration. The spectra also show the peaks at
bean oil to FAME. 718 cm-1 and 576.84 cm-1, which are characteristics of KeO and CaOe
stretching vibrations.
3.3.3. Influence of reaction time on conversion of soybean oil to FAME The chemical composition of the recovered OPA catalyst after 5th
The influence of reaction time on biodiesel production was also cycle, investigated via EDX analysis (SI, Fig. S5B), shows a significant
investigated by performing the reaction for different time-period at loss of K, Ca and O wt. %, from 14.67 %, 40.86 % and 7.34 % to 11.95
optimized conditions (Fig. 10). It was observed that after 7 h it gives 98 %, 34.64 % and 6.10 %, respectively. Due to the slight loss of major
% conversion of biodiesel. catalytic components, such as K, Ca and O, the conversion of soybean
oil decreased to 85 % after 5th cycle.
3.4. Comparison of other reported heterogeneous catalysts with the present
catalyst 4. Conclusions

Several literatures were reported on the production of FAME from In conclusion, we for the first time reported the application of or-
soybean oil using heterogeneous catalysts. Comparison of reported ange peel ash (OPA) as a raw heterogeneous catalyst for sustainable
heterogeneous catalysts with the present catalyst is shown in Table 6. production of biodiesel via transesterification of soybean oil at room
CaO–MoO3–SBA-15 (Xie and Zhao, 2014), CaO/Al2O3 (Pasupulety temperature. The mesoporous structure and the high BET surface area
et al., 2013), WO3/SnO2 (Xie and Wang, 2013) etc., showed very good of the OPA (605.60 m2/g) enhanced its catalytic activity towards
stability and high conversion of soybean oil to FAME, but they have transesterification reaction. Potassium and calcium in the form of their
some drawbacks such as preparation of these catalyst require toxic oxides are the major components present in the catalyst, which makes
chemicals, complex chemical process and high cost. However, several the catalyst highly basic. Owing to heterogeneity and basic nature, OPA
heterogeneous catalysts, such as chicken manure (Maneerung et al., catalyst can potentially replace conventionally used homogeneous basic
2016), egg shell (Wei et al., 2009), waste carbide slag (Li et al., 2015), catalysts for biodiesel production. Moreover, the catalyst was used for
snail shell (Laskar et al., 2018a) etc., can be prepared from natural biodiesel production without any functionalization or post-modifica-
waste which can reduce the total cost of the FAME production. Un- tion. Our catalyst has the advantages of being renewable, non-toxic,
fortunately, preparation of these catalyst requires high calcination biodegradable, safe to handle and carbon neutral. Although the burning
temperature and longer time, which limits their wide applications in of orange peel produces carbon dioxide, it is considered ‘carbon neutral’
industrial scale. Recently, (Pathak et al. (2018) reported the use of a as orange trees absorb carbon dioxide during its whole life cycle.
banana species Musa acuminata peel ash catalyst for the preparation of Hence, our catalyst has a huge impact on conserving the environment
FAME from soybean oil at room temperature. Despite the high activity as compared to chemically synthesized catalysts which are often highly
of the catalyst, the catalyst shows only 52 % conversion after the 4th toxic, polluting, unstable and largely non-biodegradable. In addition,

6
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

Fig. 6. 1H NMR of soybean oil (A) and FAME (B).

owing to its wide availability at low cost, the application of bio-waste scale.
materials as a catalyst source not only complements the potential re- The authors declare the following financial interests/personal re-
cycling the natural waste resources but also reduce the total cost and lationships which may be considered as potential competing interests:
negative environmental impact of biodiesel production for industrial

7
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

Fig. 7. GC–MS spectrum of FAME.

Table 4
Components present in the resultant FAME along with their percentage weight
and retention time (R.T.).
Entry R.T. Name of components Corresponding acids Amount (%)

1 18.05 Methyl palmitate C16:0 11.63


2 20.13 Methyl oleate C18:1 25.32
3 21.44 Methyl linoleate C18:2 54.34
4 24 Methyl cis-11-eicosenoate C20:1 1.27
5 24.36 Methyl eicosenoate C20:0 2.56
6 27.05 Methyl docosanoate C22:0 4.88

Table 5
Physico-chemical properties of soybean oil and soybean oil biodiesel.
Properties ASTM test ASTM limits Soybean oil Soybean oil
method for biodiesel biodiesel

Kinematic viscosity D 445 1.90-6 32.6 5.88


(cst at 40 °C) Fig. 9. Influence of oil:methanol molar ratio for the transesterification of soy-
Flash point (°C) D 93 < 93 317 146 bean oil to biodiesel. Reaction conditions: catalyst loading of 7 wt. %, reaction
Cetane number D 6890 ≥47 37.9 51 time of 7 h and room temperature.
Cloud point (°C) D 2500 −3 to 12 −9 2
Pour point(°C) D 97 −15 to 6 12 −0.2
Density (g cm−3) D 1448–1972 0.86-0.900 0.93 0.86
Calorific Value D240 37.27 39.6 38.2
(MJ/kg)

Fig. 10. Influence of reaction time for the transesterification of soybean oil to
biodiesel. Reaction conditions: 6:1 of methanol:oil ratio, catalyst loading of
7 wt. % and RT.
Fig. 8. Influence of catalyst loading for the transesterification of soybean oil to
biodiesel. Reaction conditions: 6:1 of methanol:oil ratio, Catalyst loading of Declaration of Competing Interest
7 wt. % and RT.
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influ-
ence the work reported in this paper.

8
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

Table 6
Comparison of other reported heterogeneous catalysts with the present catalyst.
Entry Catalyst Methanol/oil molar ratio Catalyst amount (wt. %) Temperature (°C) Time (h) Conversion (%) Reference

1 CaO–MoO3–SBA-15 50:1 6 63 50 83.2 (Xie and Zhao, 2014)


2 CaO/Al2O3 9:1 3 150 3 90 (Pasupulety et al., 2013)
3 WO3/SnO2 30:1 5 110 5 79.2 (Xie and Wang, 2013)
4 Chicken manure 15:1 7.5 65 3 90 (Maneerung et al., 2016)
5 Egg shell 9:1 3 65 3 95 (Wei et al., 2009)
6 Waste carbide slag 9:1 1 65 0.5 91.3 (Li et al., 2015)
7 Duck egg shell 10:1 10 60 1.2 94.6 (Yin et al., 2016)
8 Snail shells 6:1 3 RT 7 100 (Laskar et al., 2018a)
9 Banana peel ash 6:1 7 RT 4 98.95 (Pathak et al., 2018)
10 Orange peel ash 6:1 7 RT 7 98 Present work

8, 355–357. https://doi.org/10.1016/j.wasman.2018.07.022.
Kulkarni, M.G., Dalai, A.K., Bakshi, N.N., 2006. Utilization of green seed canola oil for
biodiesel production. J. Chem. Technol. Biotechnol. 81, 1886–1893. https://doi.org/
10.1002/jctb.1621.
Kralova, I., Sjöblom, J., 2010. Biofuels–renewable energy sources: a review. J. Dispersion
Sci. Technol. 31, 409–425. https://doi.org/10.1080/01932690903119674.
Maeda, Y., Thanh, L.T., Imamura, K., Izutani, K., Okitsu, K., Van Boi, L., Lan, P.N., Tuan,
N.C., Yoo, Y.E., Takenaka, N., 2011. New technology for the production of biodiesel
fuel. Green Chem. 13, 1124–1128. https://doi.org/10.1039/C1GC15049A.
Ong, H.C., Mahlia, T.M.I., Masjuki, H.H., Norhasyima, R.S., 2011. Comparison of palm
oil, jatropha curcas and calophyllum inophyllum for biodiesel: a review. Renew. Sust.
Energy Rev. 15, 3501–3515. https://doi.org/10.1016/j.rser.2011.05.005.
Agarwal, A.K., Das, L.M., 2001. Biodiesel development and characterization for use as a
fuel in compression ignition engines. J. Eng Gas Turb Power. 123, 440–447. https://
doi.org/10.1115/1.1364522.
Ma, Y., Gao, Z., Wang, Q., Liu, Y., 2018. Biodiesels from microbial oils: opportunity and
challenges. Bioresour. Technol. Rep. 263, 631–641. https://doi.org/10.1016/j.
biortech.2018.05.028.
Likozar, B., Pohar, A., Levec, J., 2016. Transesterification of oil to biodiesel in a con-
Fig. 11. Reusability of the catalyst for the transesterification of soybean oil to
tinuous tubular reactor with static mixers: modelling reaction minetics, mass transfer,
biodiesel. Reaction conditions: 6:1 of methanol:oil ratio, catalyst loading of scale-up and optimization considering fatty acid composition. Fuel Process. Technol.
7 wt.%, reaction time of 7 h and RT. 142, 326–336. https://doi.org/10.1016/j.fuproc.2015.10.035.
Likozar, B., Levec, J., 2014. Effect of process conditions on equilibrium, reaction kinetics
and mass transfer for triglyceride transesterification to biodiesel: experimental and
Acknowledgements modeling based on fatty acid composition. Fuel Process. Technol. 122, 30–41.
https://doi.org/10.1016/j.fuproc.2014.01.017.
Mardhiah, H.H., Ong, H.C., Masjuki, H.H., Lim, S., Lee, H.V., 2017. A review on latest
Financial support from the SERB (Grant No. SB/FT/CS-103/2013 developments and future prospects of heterogeneous catalyst in biodiesel production
and SB/EMEQ-076/2014), New Delhi is acknowledged. We acknowl- from non-edible oils. Renew. Sust. Energy Rev. 67, 1225–1236. https://doi.org/10.
edge MHRD andDST, India for providing fund to BC. We also thank 1016/j.rser.2016.09.036.
Sharma, Y.C., Singh, B., Korstad, J., 2011. latest developments on application of het-
SAIF-NEHU, CSIR-NEIST, IIT-Roorkee, IIT-Bombay, IASST Guwahati
erogenous basic catalysts for an efficient and eco friendly synthesis of biodiesel: A
and Assam University for analysis. Review. Fuel 90, 1309–1324. https://doi.org/10.1016/j.fuel.2010.10.015.
Kumar, D., Ali, A., 2013. Transesterification of low-quality triglycerides over a zn/cao
Appendix A. Supplementary data heterogeneous catalyst: kinetics and reusability studies. Energy Fuels 27, 3758–3768.
https://doi.org/10.1021/ef400594t.
Xie, W., Zhao, L., 2013. Production of biodiesel by transesterification of soybean oil using
Supplementary material related to this article can be found, in the calcium supported tin oxides as heterogeneous catalysts. Energy Convers. Manage.
online version, at doi:https://doi.org/10.1016/j.indcrop.2019.111911. 76, 55–62. https://doi.org/10.1016/j.enconman.2013.07.027.
Lee, H.V., Juan, J.C., Taufiq-Yap, Y.H., 2015. Preparation and application of binary
acid–base cao–la2o3 catalyst for biodiesel production. Renew Energy 74, 124–132.
References https://doi.org/10.1016/j.renene.2014.07.017.
Xie, W., Han, Y., Tai, S., 2017. Biodiesel production using biguanide-functionalized hy-
droxyapatite-encapsulated-γ-fe2o3 nanoparticles. Fuel 210, 83–90. https://doi.org/
Sahaym, U., Norton, M.G., 2008. Advances in the application of nanotechnology in en-
10.1016/j.fuel.2017.08.054.
abling a hydrogen economy. J. Mater. Sci. 43, 5395–5429. https://doi.org/10.1007/
Correia, L.M., Saboya, R.M.A., Campelo, N.S., Cecilia, J.A., Rodrguez-Castelln, E.,
s10853-008-2749-0.
Cavalcante, C.L., Vieira, R.S., 2014. Characterization of calcium oxide catalysts from
Putla, S., Peeters, E., Makshina, E., Parvulescu, V., Sels, B., 2019. Advances in porous and
natural sources and their application in the transesterification of sunflower oil.
nanoscale catalysts for viable biomass conversion. Chem. Soc. Rev. 48, 2366–2421.
Bioresour. Technol. 151, 207–213. https://doi.org/10.1016/j.biortech.2013.10.046.
https://doi.org/10.1039/c8cs00452h.
Sudarsanam, P., Zhong, R., den Bosch, S.V., Coman, S.M., Parvulescu, V.I., Sels, B.F.,
Agarwal, A.K., 2007. Biofuels (alcohols and biodiesel) applications as fuels for internal
2018. Functionalised heterogeneous catalysts for sustainable biomass valorisation.
combustion engines. Prog. Energy Combust. Sci. 33, 233271–233275. https://doi.
Chem. Soc. Rev. 47, 8349–8402. https://doi.org/10.1039/C8CS00410B.
org/10.1016/j.pecs.2006.08.003.
Shan, R., Lu, L., Shi, Y., Yuan, H., Shi, J., 2018. Catalysts from renewable resources for
Ambursa, M.M., Ali, T.H., Voon, L.H., Sudarsanam, P., Bhargava, S.K., Hamid, S.B.A.,
biodiesel production. Energ convers. Manage. 178, 277–289. https://doi.org/10.
2016. Hydrodeoxygenation of dibenzofuran to bicyclic hydrocarbons using bimetallic
1016/j.enconman.2018.10.032.
cu–ni catalysts supported on metal oxides. Fuel 180, 767–776. https://doi.org/10.
Vadery, V., Narayanan, B.N., Ramakrishnan, R.M., Cherikkallinmel, S.K., Sugunan, S.,
1016/j.fuel.2016.04.045.
Narayanan, D.P., Sasidharan, S., 2014. Room temperature production of jatropha
Caliskan, H., 2017. Environmental and enviroeconomic researches on diesel engines with
biodiesel over coconut husk ash. Energy 70, 588–594. https://doi.org/10.1016/j.
diesel and biodiesel. Fuels. J. Clean. Prod. 154, 125–129. https://doi.org/10.1016/j.
energy.2014.04.045.
jclepro.2017.03.168.
Ezebor, F., Khairuddean, M., Abdullah, A.Z., Boey, P.L., 2014. Oil palm trunk and su-
Panwar, N.L., Kaushik, S.C., Kothari, S., 2011. Role of renewable energy sources in en-
garcane bagasse derived heterogeneous acid catalysts for production of fatty acid
vironmental protection: a review. Renew. Sust. Energy Rev. 15, 1513–1524. https://
methyl esters. Energy 70, 493–503. https://doi.org/10.1016/j.energy.2014.04.024.
doi.org/10.1016/j.rser.2010.11.037.
Sharma, M., Khan, A.A., Puri, S.K., Tuli, D.K., 2012. Wood ash as a potential hetero-
Kumar, A., Gupta, R., Rathod, V.K., 2018. Waste cooking oil and waste chicken eggshells
geneous catalyst for biodiesel synthesis. Biomass Bioenergy 41, 94–106. https://doi.
derived solid base catalyst for the biodiesel production: optimization and kinetics.
org/10.1016/j.biombioe.2012.02.017.
Waste Manag. 79, 169–178. https://doi.org/10.1016/j.wasman.2018.07.022.
Zhao, C., Yang, L., Xing, S., Luo, W., Wang, Z., Lv, P., 2018. Biodiesel production by a
Gilbert, R., Perl, A., 2008. Book review: rob konings, hugo priemus and peter nijkamp
highly effective renewable catalyst from pyrolytic rice husk. J. Clean. Prod. 199,
(2008) the future of intermodal freight transport operations, design and policy. EJTIR

9
B. Changmai, et al. Industrial Crops & Products xxx (xxxx) xxxx

772–780. https://doi.org/10.1016/j.jclepro.2018.07.242. Laskar, I.B., Rajkumari, K., Gupta, R., Chatterjee, S., Paul, B., Rokhum, L., 2018a. Waste
Betiku, E., Akintunde, A.M., Ojumu, T.V., 2016. Banana peels as a biobase catalyst for Snail shell derived heterogeneous catalyst for biodiesel production by the transes-
fatty acid methyl esters production using Napoleon’s plume (bauhinia monandra) terification of soybean oil. RSC Adv. 8, 20131–20142. https://doi.org/10.1039/
seed oil: a process Parameters Optimization Study. Energy 103, 797–806. https://doi. C8RA02397B.
org/10.1016/j.energy.2016.02.138. Chatterjee, S., Dhanurdhar, Rokhum, L., 2017. Extraction of a cardanol based liquid bio-
Betiku, E., Ajala, S.O., 2014. Modeling and optimization of thevetia peruviana (yellow fuel from waste natural resource and decarboxylation using a silver-based catalyst.
oleander) oil biodiesel synthesis via musa paradisiacal (plantain) peels as hetero- Renew. Sust. Energy Rev. 72, 560–564. https://doi.org/10.1016/j.rser.2017.01.035.
geneous base catalyst: a case of artificial neural network vs. Response surface Laskar, I.B., Rajkumari, K., Gupta, R., Rokhum, L., 2018b. Acid-functionalized meso-
methodology. Ind. Crops Prod. 53, 314–322. https://doi.org/10.1016/j.indcrop. porous polymer-catalyzed acetalization of glycerol to solketal, a potential fuel ad-
2013.12.046. ditive under solvent-free conditions. Energy Fuel. https://doi.org/10.1021/acs.
Gohain, M., Devi, A., Deka, D., 2017. Musa Balbisiana colla peel as highly effective re- energyfuels.8b02948.
newable heterogeneous base catalyst for biodiesel production. Ind. Crops Prod. 109, Chouhan, A.P.S., Sarma, A.K., 2013. Biodiesel production from jatropha curcas l. Oil
8–18. https://doi.org/10.1016/j.indcrop.2017.08.006. using lemna perpusilla torrey ash as heterogeneous catalyst. Biomass Bioenergy 55,
Pathak, G., Das, D., Rajkumari, K., Rokhum, L., 2018. Exploiting waste: towards a sus- 386–389. https://doi.org/10.1016/j.biombioe.2013.02.009.
tainable production of biodiesel using Musa acuminata peel ash as a heterogeneous Hsiao, M.C., Lin, C.C., Chang, Y.H., 2011. Microwave irradiation-assisted transester-
catalyst. Green Chem. 20, 2365–2373. https://doi.org/10.1039/C8GC00071A. ification of soybean oil to biodiesel catalyzed by nanopowder calcium oxide. Fuel 90,
Li, X., Tang, Y., Caoa, X., Lua, D., Luoa, F., Shao, W., 2008. Preparation and evaluation of 1963–1967. https://doi.org/10.1016/j.fuel.2011.01.004.
Orange peel cellulose adsorbents for effective removal of cadmium, zinc, cobalt and Xie, W., Zhao, L., 2014. Heterogeneous CaO–MoO3–SBA-15 catalysts for biodiesel pro-
nickel. Colloid Surf.A 317, 512–521. https://doi.org/10.1016/j.colsurfa.2007.11. duction from soybean oil. Energy Conv. Manag. 79, 34–42. https://doi.org/10.1016/
031. j.enconman.2013.11.041.
Chutia, M., Bhuyan, P.D., Pathak, M.G., Sarma, T.C., Boruah, P., 2009. Antifungal activity Pasupulety, N., Gunda, K., Liu, Y., Rempel, G.L., Ng, F.T.T., 2013. Production of biodiesel
and chemical composition of citrus reticulata blanco essential oil against phyto- from soybean oil on CaO/Al2O3 solid base catalysts. Appl. Catal. A Gen. 452,
pathogens from north east india. LWT - Food Sci. Technol. 42, 777–780. https://doi. 189–202. https://doi.org/10.1016/j.apcata.2012.10.006.
org/10.1016/j.lwt.2008.09.015. Xie, W., Wang, Tao, 2013. Biodiesel production from soybean oil transesterification using
Biswas, B.K., Inoue, J., Inoue, K., Ghimire, K.N., Harada, H., Ohto, K., Kawakita, H., 2008. tin oxide-supported WO3 catalysts. Fuel Process Tech. 109, 150–155. https://doi.org/
Adsorptive removal of As (V) and As (III) from water by a Zr (IV)-loaded Orange 10.1016/j.fuproc.2012.09.053.
waste gel. J. Hazard. Mater. 154, 1066–1074. https://doi.org/10.1016/j.jhazmat. Maneerung, T., Kawi, S., Dai, Y., Wang, C.H., 2016. Sustainable biodiesel production via
2007.11.030. transesterification of waste cooking oil by using CaO catalysts prepared from chicken
Santos, C.M., Dweck, J., Viotto, R.S., Rosa, A.H., de Morais, L.C., 2015. Application of manure. Energy Convers. Manage. 123, 487–497. https://doi.org/10.1016/j.
Orange peel waste in the production of solid biofuels and biosorbents. Bioresour. enconman.2016.06.071.
Technol. 196, 469–479. https://doi.org/10.1016/j.biortech.2015.07.114. Wei, Z., Xu, C., Li, B., 2009. Application of waste eggshell as low-cost solid catalyst for
Rajkumari, K., Kalita, J., Das, D., Rokhum, L., 2017. Magnetic Fe3O4@silica sulfuric acid biodiesel production. Bioresour. Technol. 100, 2883–2885.
nanoparticles promoted regioselective protection/deprotection of alcohols with di- Li, F.J., Li, H.Q., Wang, L.G., Cao, Y., 2015. Waste carbide slag as a solid base catalyst for
hydropyran under solvent-free conditions. RSC Adv. 7, 56559–56565. https://doi. effective synthesis of biodiesel via transesterification of soybean oil with methanol.
org/10.1039/C7RA12458A. Fuel Process Tech. 131, 421–429. https://doi.org/10.1016/j.fuproc.2014.12.018.
Das, D., Pathak, G., Rokhum, L., 2016. Polymer supported DMAP: an easily recyclable Yin, X., Duan, X., You, Q., Dai, C., Tan, Z., Zhu, X., 2016. Biodiesel production from
organocatalyst for highly atom-economical Henry reaction under solvent-free con- soybean oil deodorizer distillate using calcined duck eggshell as catalyst. Energy Con
ditions. RSC Adv. 6, 104154–104163. https://doi.org/10.1039/C6RA23696K. Manag. 112, 199–207. https://doi.org/10.1016/j.enconman.2016.01.026.

10

You might also like