Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Waste Management 102 (2020) 212–221

Contents lists available at ScienceDirect

Waste Management
journal homepage: www.elsevier.com/locate/wasman

Towards sustainable biodiesel and chemical production: Multifunctional


use of heterogeneous catalyst from littered Tectona grandis leaves
Minakshi Gohain a,⇑, Khairujjaman Laskar b, Hridoyjit Phukon c,d, Utpal Bora b, Dipul Kalita c,d,
Dhanapati Deka a
a
Department of Energy, Tezpur University, Napaam 784028, Assam, India
b
Department of Chemical Sciences, Tezpur University, Napaam 784028, Assam, India
c
Cellulose Pulp and Paper Group (Material Sciences and Technology Division), North East Institute of Science and Technology, Jorhat 785006, Assam, India
d
Academy of Scientific and Innovative Research, Council of Scientific and Industrial Research - North East Institute of Science and Technology, Jorhat 785006, Assam, India

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

Article history: Waste biomass derived heterogeneous catalyst is an excellent alternative to chemically synthesized cat-
Received 9 September 2019 alysts. In this work, calcined Tectona grandis leaves were proposed as an eco-friendly, renewable and low
Revised 24 October 2019 cost heterogeneous base catalyst. The prepared catalyst was examined by FTIR, XRD, XPS, SEM, EDX, TEM,
Accepted 27 October 2019
TGA, BET and Hammett indicator test. The catalyst has an appealing nature towards various chemical
transformations due to its basic surface sites provided by alkali and alkaline earth metals. The efficiency
of the catalyst was successfully investigated by its application in biodiesel production. The products were
Keywords:
confirmed by 1H and 13C NMR. 100% FAME conversion was attained using a catalyst loading of 2.5 wt%
Waste biomass
Tectona grandis leaves
under optimized reaction parameters. The catalyst was further explored for Knoevenagel condensation
Renewable reaction, in which it showed its effectiveness and recyclability towards the formation of benzylidene-
Biodiesel malononitrile derivatives of aryl aldehydes. Thus, it is a potential ‘green catalyst’ derived from waste bio-
Knoevenagel condensation mass without any addition of chemicals that can replace the industrial base catalysts used for biodiesel
production and Knoevenagel reaction and makes the protocol environmentally benign.
Ó 2019 Elsevier Ltd. All rights reserved.

1. Introduction esterification reaction using homogeneous catalyst (NaOH, KOH,


CH3CH2ONa, H2SO4 etc), which are corrosive for the reactors
Depletion of fossil fuels, increasing concerns about environmen- (Sano et al., 2017), separation of the catalyst from the reaction mix-
tal pollution and greenhouse gases emission has led to the explo- tures is costly and chemically wasteful (Ma and Hanna, 1999;
ration of sustainable sources of energy (Li and Guo, 2017; Noureddini and Zhu, 1997) as well. Solid catalysts based on ZrO2
Sharma et al., 2009; Lardon et al., 2009). Among the alternatives, (Ajala et al., 2015), SrO (Liu et al., 2007), ZnO (Xie and Li, 2006,
renewable and sustainable biodiesel has the key advantage of Fe3O4 (Tang et al., 2012) etc., have been used in biodiesel synthesis
using the existing current technology for diesel fuels (Teo et al., however the use of biomass derived catalyst is highly desirable
2018). Biodiesel derived from edible or non-edible oils are a poten- from the perspective of sustainability and economy (Steen et al.,
tial alternative for diesel fuels which is biodegradable, non-toxic 2010).
and renewable (Lardon et al., 2009; Agarwal et al., 2018; Rafiei Activated carbon prepared from waste biomass has been exten-
et al., 2018). But the current conventional method used for biodie- sively fabricated by functionalization with acids or bases to pre-
sel synthesis is not sustainable due to use of edible vegetable oils pare various solid catalysts such as deoiled cake (Konwar et al.,
as feedstocks which accounts for more than 70% of the total pro- 2014; Toda et al., 2005), modified rice (Roschat et al., 2016) and
duction cost (Piker et al., 2016). Low cost and available feedstocks peat as support (Wang et al., 2017). Carbon based catalysts have
like waste cooking oil (Gohain et al., 2019; Tan et al., 2019; Xiong good thermal and mechanical stability and thus are ideal catalysts
et al., 2019), jatropha and palm oil should be used to replace the for various reactions. But it involves functionalization of the carbon
food grade vegetable oils for commercial biodiesel synthesis or activated carbons with harmful chemicals which increase the
(Janaun and Ellis, 2010). It is generally synthesized through trans- cost of production and is also time consuming. Hence, only a few
of them can be used on an industrial scale because of its high syn-
⇑ Corresponding author. thetic cost (Liu et al., 2016). These shortcomings can be addressed
E-mail address: minakshiigohain@gmail.com (M. Gohain). successfully by the use of heterogeneous catalyst derived from

https://doi.org/10.1016/j.wasman.2019.10.049
0956-053X/Ó 2019 Elsevier Ltd. All rights reserved.
M. Gohain et al. / Waste Management 102 (2020) 212–221 213

biomass. Therefore, researchers began to use waste materials for morphological characteristics of CTGL was studied by FTIR, XPS,
the preparation of heterogeneous catalyst from biomass: banana XRD, EDX, SEM, TEM, TGA, BET and Hammet Indicator test. The
peels (Lardon et al., 2009; Betiku et al., 2016), palm trunk proximate analysis of CTGL was carried out using ASTM standards.
(Ezebor et al., 2014), wood (Sharma et al., 2012), coconut husk The FTIR spectrum was recorded in Nicolet (Impact 410, USA) FTIR
(Vadery et al., 2014), egg shell (Tan et al., 2015), waste animal spectrophotometer. XPS was recorded in Thermo Fisher Scientific
bones (Nisar et al., 2017) etc. Calcined ash of some of these bio- (ESCALAB Xi+) instrument. The crystalline phase of the catalyst
mass derived catalysts can be used as a catalyst without any chem- was evaluated using Rigaku miniflex diffractometer in a 2h range
ical functionalization for biodiesel production. However, biomass of 20-80° with Cu Ka radiations (k = 1.5406 Å). The SEM was
wastes have been less explored for mixed oxide catalyst recorded on an electron microscope Jeol, JSM-6290 LV and TEM
preparation. study were recorded in JEOL, JEM-2100 Plus Electron Microscope,
Therefore, a simple methodology is highly required for the Japan. Elemental composition of the catalyst was determined in
development of heterogeneous catalyst from waste biomass which Carl ZEISS Microscopy, Sigma, Germany. The thermal stability
will be environmentally benign as well as cost effective. In this was tested in a Model TGA-50 and DSC-60, Shimadzu under an
work, we report the use of easily prepared catalyst from waste N2 flow rate of 1 mL/min and at 10 °C min1 ramping rate. The sur-
i.e. littered Tectona grandis leaves for transesterification reaction. face area and pore size distribution of the sample were analyzed
Native to South and Southeast Asia, Tectona grandis is grown exten- using NOVA 1000E, NOVA WIN, QUANTACHROME (degassing tem-
sively in India, Myanmar, Malaysia, Thailand, Sri Lanka and Bangla- perature of 150 °C, 16 h prior to adsorption/desorption). The basic-
desh as its wood is highly valued for its durability and water ity of CTGL was examined by the Hammett indicator test.
resistance property. It is a large, deciduous tree that occurs in
mixed hardwood forests and has large papery leaves that are often 2.3. FAME production
hairy on the lower surface (Wiersema et al., 2019). Different parts
of Tectona grandis have various applications in medicinal purpose The acid value of WO was found to be 3.73 mg KOH/g which
(Nalvothula et al., 2014). Extract obtained from its leaves has been was not ideal for effective base-catalyzed transesterification reac-
applied for nanoparticle synthesis (Devadiga et al., 2015). But, so tion. Hence, to minimize the acid value of WO to the desired acid
far, calcined ash of this leaf has not been evaluated for the applica- value of (2 mg KOH/g), the esterification step was carried out.
tion in organic transformations.
To evaluate the versatile nature of the catalyst, it was also 2.3.1. Pre-treatment of WO
tested for knoevenagel condensation reaction (KCR) of aryl aldehy- Esterification was carried out in a 3-necked round bottom flask
des containing both electron withdrawing and donating groups (500 mL) equipped with a reflux condenser. A 100 mL of WO was
with malononitrile, which are frequently applied for the produc- fed into the flask, heated to 60 °C and then cooled. To the cooled
tion of fine chemicals and biologically important compounds WO, H2SO4 (1 mL) and methanol (30 mL) were added and then
(Tran et al., 2011; Daneshvar et al., 2018). Over the past decades, continuously stirred at 800 rpm maintaining the temperature at
various solid-base catalysts have been applied to this reaction, 60 °C for 1 h. After reaction completion, it was allowed to settle
such as zeolites exchanged with alkylammonium cations for 2 h which resulted in two distinct liquid phases. The top layer
(Martins et al., 2008), chitosan biohydrogel (Kühbeck et al., of methanol-water was separated using a separating funnel. The
2012), mesoporous zirconia (Parida et al., 2010) and MOFs pre-treated WO layer was washed (double distilled water) and
(Kolmykov et al., 2016) etc. Thus, utilization of biowaste for the then residuals were evaporated using rotary evaporator. The acid
preparation of heterogeneous solid-base catalyst leads to a newer value of the pre-treated WO got reduced to 1.01 mg KOH/g.
dimension with opening the scope of chemical fabrication and
knoevenagel condensation of aryl aldehydes with malononitrile. 2.3.2. Transesterification reaction
The catalyst is renewable as well as cost effective as it is derived The pre-treated oil was heated to 100 °C and then cooled down
from bio-waste and is prepared without any functionalization. Fur- to remove any traces of absorbed water. The required reaction
thermore, the recycling potentiality of the catalyst was also con- parameters i.e. the molar ratio of methanol, pre-treated WO and
ducted. To our best knowledge, there has not been any report on CTGL loading were then mixed in a 3-necked glass flask reactor
the preparation of heterogeneous base catalyst from Tectona gran- (500 mL) and was equipped with a reflux condenser. The reaction
dis leaves for the production of biodiesel as well as KCR between continued for the desired time maintaining the agitation speed of
aldehydes and malononitrile (active methylene compound). 650 rpm for proper mixing of methanol and oil. After completion
of the reaction, the mixture was poured into a separating funnel
wherein three layers were observed: fatty acid methyl esters, glyc-
2. Materials and methods
erine and CTGL. The top layer i.e. FAME was washed with warm
distilled water (50 °C) and then evaporated using rotary evaporator
2.1. Materials
to remove any unreacted methanol to obtain WO methyl esters
(WOME).
Littered Tectona grandis leaves (TGL) were collected from the
local areas. Waste cooking oil (WO) was collected from Tezpur
2.4. Biodiesel analysis
University campus food court. All the chemicals were used without
further drying or purification and were purchased from Sigma
The biodiesel was analyzed using 1H NMR which was recorded
Aldrich.
in NMR spectrometer (500 MHz, Oxford, AS400, China).
The % FAME was determined using NMR technique (Gelbard
2.2. Catalyst preparation and characterization et al., 1995; Xie and Li, 2006) (CDCl3 solvent) (Eq. (1)).
2X
The TGL was washed thoroughly with double distilled water C MEster ð%Þ ¼  100 ð1Þ
and then dried under the sun. The leaves were then cut into pieces, 3Y
milled and sieved to obtain fine powder (250 lm). The sieved pow- X = Integration value of methoxy protons (-OCH3) at 3.6 ppm
der was then calcined in a muffle furnace at 700 °C for 4 h to pre- Y = Integration value of a-methylene (-CH2-) protons at
pare calcined Tectona grandis leaves (CTGL). The structural and 2.3 ppm
214 M. Gohain et al. / Waste Management 102 (2020) 212–221

The FAME was tested by the American Society for Testing and 3. Results and discussion
Material (ASTM) methods and compared with required biodiesel
standards. With the concept of environmental sustainability many atten-
tions have been given in utilization of biowaste for the synthesis
2.5. Reusability of the catalyst (Biodiesel) of fuels and chemicals in order to reduce the use of hazardous
chemicals. Thus, in present protocol we use CTGL an agro-waste
From an economic point of view, the reusability nature of the as a renewable catalyst for chemical transformation.
catalyst is an essential factor. The recovered catalyst was washed
with hot methanol (40 °C) to remove any traces of unwanted 3.1. CTGL characterization
impurities and dried in an oven (120 °C for 3 h) (Gohain et al.,
2017) and used for successive cycles. The XRD analysis of CTGL [Fig. 1(a)] revealed the presence of
metal oxides and carbonates in abundance. Strong characteristic
peaks of K2O and K2CO3 were seen at 2h = 25.28, 26.59, 28.41,
2.6. Knoevenagel condensation between aryl aldehydes and
31.4, 40.55, 44 (JCPDS reference code 00-026-1327, 00-16-820,
malononitrile
77-2176, 85-514, 87-0730). The peaks at 2h = 30.01, 39, 43, 47,
52.50 (JCPDS reference code 16-820, 87-1863, 82-1691, 01-085-
The product formation was monitored by TLC on pre-coated
0514) revealed the presence of CaO and CaCO3 in the CTGL. Previ-
Merck silica gel plates (60F-254) and visualized under UV light. .
ous literatures also reported the presence of CaO in their catalyst
The isolation of desired products was achieved by column chro-
comparable 2h values (Wilson et al., 2008; Pathak et al., 2018;
matography using silica gel (60–120 mesh). The 1H NMR spectra
Laskar et al., 2018; Zhao et al., 2018). The peaks at 2h = 34.13 con-
were recorded on a JEOL, JNM ECS NMR spectrometer operating
firmed the presence of SiO2 (JCPDS reference code 89-3609, 89-
at 400 MHz in CDCl3, using TMS as an internal standard. The 13C
7746) (Wang et al., 2017). It was found that CTGL contains oxides
NMR spectra were recorded at 100 MHz using CDCl3 (d = 77.00)
of K and Ca which have very strong basic sites. The group I alkali
as standard. The structures of the purified products were con-
metal hydroxide and group II metal oxide exhibits good transester-
firmed by NMR spectra (See ESI, S1-S15).
ification activity due to their higher basicity.
The functional groups present in CTGL were analyzed by FTIR
2.6.1. Synthesis of benzylidenemalononitrile derivatives [Fig. 1(b)]. The broad peak at 3400 cm1 [Fig. 1(b)] is associated
Aryl aldehydes (0.25 mmol), malononitrile (0.5 mmol) and with OAH stretching (Betiku et al., 2016), which is due to absorbed
CTGL (10 wt%) in ethanol (2 mL) were stirred at 60 °C for an appro- water molecules on the surface of CTGL. The bands at 2320 cm1
priate period time (Table 3). After completion of the reaction (mon- and 2175 cm1 are attributed to the SiAOAK, MgAOAK bonds
itored by TLC), the reaction mixture was centrifuged and the stretching vibrations. The SiAOASi bonds peak was observed at
catalyst was separated. The crude reaction mixture was extracted 1000 cm1 which further supports the XRD peak of SiO2 compo-
with ethyl acetate (3  10 mL). The catalyst was washed and then nent [Fig. 1(a)]. The characteristic peak at 560 cm1 is due to
allowed to dry for further use. The resultant organic phases were stretching vibrations of CaAO and KAO (Liu et al., 2016; Stork
dried using Na2SO4 (anhydrous), evaporated under reduced pres- and Pott, 1974). The peaks at 1654 cm1, 1389 cm1, 870 cm1
sure, and then purified by column chromatography using ethyl and 700 cm1 are attributed to CAO bond of metal carbonates that
acetate and hexane (1:5; v/v) as an eluent. may be formed due to CO2 absorption on to the surface of metal
oxides of CTGL which detects the presence of K, Ca and metal car-
2.7. Reusability of the catalyst (Knoevenagel Condensation) bonates (Sharma et al., 2012). Similar type of bands for carbonates
was also reported for some of the earlier reported ash catalysts
The recyclability of CTGL towards KCR of aldehydes was tested (Betiku et al., 2016; Gohain et al., 2017). Therefore, the predictions
using the model reaction of chlorobenzaldehyde and malononitrile concurred in the FT-IR study of CTGL confirms the presence of car-
under same optimal condition. After 1st cycle, the catalyst was fil- bonate and oxide of metals in the catalyst and is consistent with
tered and washed several times with ethyl acetate followed by the XRD analysis.
H2O. The recovered catalyst was dried in oven at 100 °C for over- EDX study supported the FTIR and XRD data and elucidates the
night and was further used for the consecutive runs. chemical nature of CTGL. The EDX (ESI, S16) analysis confirmed the

Fig. 1. (a) XRD pattern and (b) FTIR spectra of CTGL.


M. Gohain et al. / Waste Management 102 (2020) 212–221 215

Fig. 2. XPS spectra of CTGL (a) O 1s, (b) K 2p and (c) Ca 2p.

presence of K (53.25 wt%), Ca (30.28 wt%), Si (10.03 wt%), Mg surface area and high porosity. Higher surface area enhances inter-
(4.77 wt%) and Na (1.67 wt%). The proximate analysis of CTGL is action between active sites and reactants of the catalysts (Gohain
shown in ESI, S16. et al., 2017). The microporous nature of CTGL will provide an ade-
The XPS analysis supported the findings of EDX, and showed quate external surface which will act as active sites for rapid mass
peaks for K, Ca, Mg, Si and Na. The peak observed at 529.06 eV transfer for the transesterification process (Wang et al., 2017).
and 531.84 eV in O 1s spectra [Fig. 2(a)] can be attributed to metal BET study was performed to measure the surface area, pore vol-
oxides and metal carbonates respectively. Peaks appeared at bind- ume and pore size of CTGL which was found to be 116.833 m2 g1,
ing energy 292.5 eV and 295 eV in K 2p spectra [Fig. 2(b)] denotes 0.185 cm3 g1 and 112.210 Å respectively. The N2 physisorption of
the presence of potassium in the form of oxides or carbonates CTGL exhibits type-IV isotherm revealing its mesoporous charac-
(Pathak et al., 2018). The peaks at 346.48 eV and 350 eV in Ca 2p teristic (ESI, S17). Deka and Basumatary (2011), Wang et al.
spectra [Fig. 2(c)] denote the presence of Ca in oxides form (2017), Pathak et al. (2018) reported the surface area as 1.487 m2
(Nefedov et al., 1974). These results strongly suggest that both g1, 1.266 m2 g1, 4.442 m2 g1, 1.4546 m2 g1 respectively, which
strong and moderately basic sites are present on CTGL. are much lower compared to the present catalyst i.e. CTGL
From observed results, K was found to be the major ingredient (116.833 m2 g1). They showed lower catalytic ability which is
in CTGL and K2O being a strong base resulting strong basicity of probably due to their lower surface area as increase in the surface
CTGL. Strong bases can remove protons from methanol and area increases the catalytic activity.
develop active methoxy species which are responsible for efficient The basic strength of CTGL was found to be 9.8 < H  18.4 using
transesterification (Gohain et al., 2017). Hammett indicator test. After calcination TGL the metal carbonate
To assess the thermal stability and behavior of CTGL, TG analy- salts gets thermally activated to active metal oxide (Lee et al.,
sis (ESI, S16) was performed at elevated temperature upto 700 °C. 2014). According to reported literature, potassium leaching can
A total of 9% weight loss in CTGL was observed. The loss of surface be prevented by calcination, and hence may increase the catalyst
water present in CTGL was observed until 380.72 °C. Another slight reusability (Sharma et al., 2012).
loss from 380.72 °C to 450 °C is attributed to inorganic impurities
present in CTGL. At 450 °C oxidation of carbonaceous material 3.2. Biodiesel characterization
and release of CO, CO2 etc (Chouhan and Sarma, 2013) started
and continued upto 660 °C, beyond which no further weight loss Fig. 4(a) and (b) show the 1H NMR spectra of WO and WOME
was observed. respectively. A singlet appeared at d 3.6 due to the methoxy group
Both the SEM and TEM [Fig. 3(a) and (b)] analysis revealed and a triplet at d 2.3 corresponding to a-CH2 protons [Fig. 4(b)]
mesoporous and microporous nature of CTGL which indicates large which is the major differences in signals compared to WO [Fig. 4
216 M. Gohain et al. / Waste Management 102 (2020) 212–221

Fig. 3. (a) SEM image and (b) TEM image of CTGL.

(a)] confirms the formation of biodiesel. Other peaks appeared at d reverse reaction that reduces the FAME conversion (Ma and
0.85, d 1.2 and a multiplet at d 1.6 are assigned to terminal methyl Hanna, 1999; Xie and Huang, 2006).
protons, methylene protons and b-carbonyl methylene protons
respectively. The peak at d 5.3 is due to olefinic hydrogens (Tariq 3.3.3. Reaction time
et al., 2011; Diehl and Randel, 2007; Monteiro et al., 2009). The dis- Fig. 5(c) shows the influence of time on FAME conversion exam-
appearance of glycerol moiety signals at d 4.1–4.4 in Fig. 4(b), ined by varying the time from 1 h to 5 h, using 6:1 M ratio of
clearly showed the conversion of WO to FAME. methanol/oil and 2.5 wt% of CTGL loading. Conversion of FAME
Using Eq. (1), 100% FAME conversion was found with a 6:1 M was low at the first 1 h because of slow dispersion of triglycerides
ratio of methanol/oil, 2.5 wt% CTGL loading within 3 h. The physic- into the methanol phase. However, with further increase in reac-
ochemical properties of pre-treated WO are shown in ESI, S17. The tion time, FAME conversion got increased gradually until the reac-
WOME fuel properties (Table 1) meet the biodiesel standards tion attains the equilibrium (Vyas et al., 2009). Long reaction time
(ASTM D 6751 and EN 14214). may drive the reaction backward resulting in lower FAME conver-
sion (Wan et al., 2014) after 5 h.
The plausible mechanism for transesterification of triglyceride
3.3. Influence of reaction parameters on biodiesel production to biodiesel using CTGL as a heterogeneous base catalyst is shown
in Scheme 1. It is assumed that initially the basic sites present in
3.3.1. Catalyst loading CTGL abstract the proton from methanol to give methoxide. This
The influence of catalyst loading on biodiesel conversion was methoxide then react with carbonyl group present in triglyceride
analyzed by changing catalyst loadings (CTGL) from 0.5 to 4.5 wt to give biodiesel and at the same time catalyst is recovered to be
% [Fig. 5(a)] using a 6:1 M ratio of methanol/oil and performing used for next cycles.
the reaction at room temperature. The low conversion was
achieved at lower catalyst loading due to availability of lesser basic 3.3.4. Catalyst deactivation and reusability potency (Biodiesel)
sites which is required for conversion of oil to FAME. It was Recyclability of a catalyst is an important characteristic from
observed that FAME conversion increased with higher catalyst the point of green chemistry (Ferré et al., 2016). Methanol/oil
loading, due to an increase in the active sites which favors the con- molar ratio of 6:1, catalyst loading of 2.5 wt% and room tempera-
tact time of reactants and catalysts (Olutoye et al., 2011). But it was ture were selected as the optimized reaction parameters [Fig. 5
observed that with further increase in catalyst amount FAME con- (d)]. It was found that on reusing CTGL, biodiesel conversion got
version got reduced. This may be due to the lower diffusion of the reduced with each successive cycle. This lower FAME conversion
reaction mixtures (catalyst, methanol and oil) in higher catalyst may be attributed to the leaching of the catalyst which resulted
amount (Mansir et al., 2018). The three phase solution becomes in a reduction of active sites of the catalyst (Long et al., 2014).
more viscous with higher catalyst loading which could resist the The K and Ca have a greater tendency of getting leached into the
mass transfer among the reactants and saponification reaction solution more readily than other elements (Wang et al., 2017).
for base catalyzed reactions may get induce that decreased the Hence, K leaching may be one of the main reasons for catalyst
FAME conversion (Kaur and Ali, 2014; Meher et al., 2006). deactivation due to considerable solubility of K2O in methanol
and glycerol. Another reason may be the loss of catalyst amount
during washing and transfer between cycles.
3.3.2. Molar ratio of methanol/oil The FTIR analysis showed that upon repeated use of CTGL glyc-
The influence of methanol/oil molar ratio was also studied erol and ester got agglomerated on it (ESI, S17) which may result in
[Fig. 5(b)]. Transesterification reaction was performed with 2:1, blockage of active sites leading to lower FAME conversion. This was
4:1, 6:1, 8:1 and 10:1 methanol/oil molar ratio for 3 h using evident by the appearance of a band at 1750 cm1 which indicates
2.5 wt% CTGL loading. As there was less contact between the cata- adsorbed esters. The intensity of the peaks at 1389 cm1 and
lyst and the reactants, FAME conversion decreases at lower metha- 1000 cm1 got reduced in the reused CTGL which signifies the loss
nol/oil molar ratio. Transesterification being a reversible reaction, of K and Si resulting in the availability of lower basic sites. The
excess methanol is vital to shift the reaction in direction of the bands at 1050 cm1 and 2800 cm1 exhibits ACAO (primary alco-
products (Olutoye et al., 2011). With an increase in the molar ratio hol) and CAH stretching indicating Ca(OCH3)2 formation due to the
of methanol/oil FAME conversion got increased owing to the for- reaction of methanol and glycerine with CaO onto the surface of
mation of methoxy species on catalyst surface which shifted the CTGL. The Ca(OCH3)2 and absorbed esters resulted in pores block-
reaction towards the forward direction. But a decrease in the con- age which lowered the mass transfer during the reaction results in
version of FAME was also observed with further increase in metha- lowering FAME conversion (Kawashima et al., 2009). TEM and XRD
nol/oil molar ratio as higher methanol/oil molar ratio leads to analyses of recovered CTGL were also investigated. The TEM anal-
M. Gohain et al. / Waste Management 102 (2020) 212–221 217

1.0 W O.001.esp

0.9
(a)
0.8

0.7
Normalized Intensity

0.6

0.5

0.4

0.3

0.2

0.1

8 7 6 5 4 3 2 1 0 -1
Chemical Shif t (ppm)

1.0 W OME 1H.001.esp

0.9
(b)
0.8

0.7

0.6
Normalized Intensity

0.5

0.4

0.3

0.2

0.1

0
0.00 0.09 0.09 0.04 0.06 0.11 0.07 0.55 0.01 0.09 0.00

8 7 6 5 4 3 2 1 0 -1
Chemical Shif t (ppm)

Fig. 4. 1H NMR spectra of (a) WO and (b) WOME.

Table 1
Fuel properties of WOME.

Fuel Property Unit Biodiesel ASTM D 6751 EN 14214


Acid Number mg KOH per g 0.07 <0.80 0.5 max
Viscosity at 40 °C mm2 per sec 4.15 1.9–6.0 3.5–5.0
Density at 15 °C g per cc 0.85 0.86–0.90 0.85
Carbon residue % wt. 0.016 0.050 max 0.3
Calorific value MJ per kg 39.32 – 35 min
Cetane Number 58 47 min 51 min
Cloud point °C 5 – –
Pour point °C 9 – –
Flash point °C 148 93 min 120 min
218 M. Gohain et al. / Waste Management 102 (2020) 212–221

Fig. 5. Influence of (a) CTGL loading, (b) Methanol/oil molar ratio, (c) time and (d) reusability on FAME conversion.

ysis of the reused CTGL (ESI, S18) supported the FTIR findings and in trace amounts with no use of catalyst (Table 2, entry 10) signi-
depicted less porous nature of reused CTGL. The XRD analysis of fying the excellent catalytic activity of CTGL for KCR.
the recycled CTGL catalyst exhibited a sharp decrease in the inten- Moreover, Table 3 summarizes the successful conversion of all
sity of potassium oxide, the major basic site of CTGL (ESI, S18). aryl aldehyde derivatives to BMN derivatives in the presence of
Hence, pore blockage and reduction in the basic sites on reusing CTGL catalyst. The BMN derivatives were obtained in high yields
the catalyst might have resulted in lower FAME conversion. using minimum solvent for a shorter period (10–30 min) of time
at 60 °C. Considering the high catalytic efficiency of CTGL towards
3.4. Optimization of reaction condition for Knoevenagel condensation malononitrile and 4-Chlorobenzaldehyde, different benzaldehyde
derivatives were chosen as reaction substrates. For Knoevenagel
Table 2 well illustrated the optimization process for the amount condensation, nucleophilic addition is a rate-determining step,
of substrate, solvent effect, catalyst loading and period time. The thus electron-withdrawing groups are more reactive than those
efficiency CTGL was initially assessed in the reaction of 4- of electron-donating groups (Bhunia et al., 2016) in BMN deriva-
Chlorobenzaldehyde (0.25 mmol), and malononitrile (0.5 mmol) tives formation. Table 3, clearly showed the derivatives of ben-
using different solvents at 60 °C. The process aimed to provide zaldehyde having electron-withdrawing groups (ANO2, ACN, ACl
the absolute formation of benzylidenemalononitrile (BMN) deriva- etc) showed more conversion to catalytic products to those of
tives at 60 °C. Various reports (Göksu and Gültekin, 2017; Ding electron-donating benzaldehyde derivatives (ACH3, AOH, AOMe).
et al., 2015; Sakthivel and Dhakshinamoorthy, 2017) suggests the Compounds 21 and 29 were successfully synthesized but in low
production of BMN, but the longer period time and efficient pro- yields. 3,4,5-Trimethoxybenzaldehyde (entry 9), showed low yield
duct formation always a challenging task. To overcome this, the to the desired product is plausibly due to the sterical hindrance of
present protocol is based on the successful completion of KCR three methoxy groups along with the electron-donating effect.
remarkably with high yield in short reaction period and catalyst Thus, to obtain the catalytic yields, the substituents of benzalde-
reusability. Eventually, 0.5 mmol of 4-Chlorobenzaldehyde, hyde play a major role in the reaction of aldehydes with
1.0 mmol of malononitrile and 2.0 mg of catalyst provide sufficient malononitrile.
concert for knoevenagel condensation in 2 mL of ethanol (Table 2, A plausible mechanistic pathway for Knoevenagel condensation
entry 9). At the same time, the desired product formation took 3 h catalyzed by CTGL has been proposed in Scheme 1.
M. Gohain et al. / Waste Management 102 (2020) 212–221 219

Scheme 1. Plausible mechanism for transesterification of triglycerides to biodiesel and Knoevenagel condensation catalyzed by CTGL base catalyst.

Table 2 Table 3
Optimization conditions for KCR of 4-Cl-benzaldehyde with malononitrile.a KCR of aryl aldehyde derivatives with malononitrile in presence of CTGL catalyst.a

Sl. No. Solvent CTGL (mg) Time (min) Yield (%)b Sl. No. Substrate Product Time (min) Yield (%)b
1 H2O 2 120 Trace 1 12 94
2 MeOH 2 120 60 2 10 95
3 CH2Cl2 2 120 Trace 3 10 96
4 CH3CN 2 120 Trace 4 14 91
5 MeOH/H2O (2:1) 2 60 40 5 25 90
6 WEB 2 60 55 6 15 87
7 WEB/EtOH (2:1) 2 30 80 7 16 86
8 WEB/EtOH (2:2) 2 25 85 8 16 86
9 EtOH 2 10 95 9 22 84
10 EtOH – 180 Trace 10 20 86
a 11 30 81
Reaction condition: Chlorobenzaldehyde (0.25 mmol), malononitrile
12 14 93
(0.5 mmol), CTGL (10 wt% of substrate) at 60 °C.
b 13 11 97
Isolated yields.
14 10 96
15 17 86
16 12 92
17 14 91
3.4.1. Catalyst reusability potency (Knoevenagel condensation) 18 15 91
For KCR, the catalyst was used five times and it was found that 19 18 85
the catalyst maintained good activity for a minimum of five con- a
Reaction condition: Aldehydes (0.25 mmol), malononitrile (0.5 mmol), CTGL
secutive cycles demonstrating the catalyst can be reused without (10 wt% of substrate) at 60 °C.
b
a significant reduction in the yield (ESI, S18). Isolated yields.
220 M. Gohain et al. / Waste Management 102 (2020) 212–221

4. Conclusion Ferré, M., Pleixats, R., Man, M.W.C., Cattoën, X., 2016. Recyclable organocatalysts
based on hybrid silicas. Green Chem. 18, 881–922. https://doi.org/10.1039/
C5GC02579F.
We described a bio-refinery approach wherein the application Gelbard, G., Bres, O., Vargas, R.M., Vielfaure, F., Schuchardt, U.F., 1995. 1H nuclear
of CTGL having multifunctional use, capable of producing fuel as magnetic resonance determination of the yield of the transesterification of
rapeseed oil with methanol. J. Am. Oil. Chem. Soc. 72, 1239–1241. https://doi.
well as chemical as a renewable, reusable and low cost heteroge-
org/10.1007/BF02540998.
neous base catalyst. Considering the simplicity in catalyst prepara- Gohain, M., Devi, A., Deka, D., 2017. D, Musa balbisiana Colla peel as highly effective
tion which does not involve any chemical treatment or renewable heterogeneous base catalyst for biodiesel production. Ind. Crops
Prod. 109, 8–18. https://doi.org/10.1016/j.indcrop.2017.08.006.
functionalization makes our catalyst environmentally benign.
Gohain, M., Laskar, K., Paul, A.K., Daimary, N., Maharana, M., Goswami, I.K.,
FAME conversion of 100% was achieved at room temperature using Hazarika, A., Bora, U., Deka, D., 2019. Carica papaya stem: A source of versatile
2.5 wt% catalyst loading and 6: 1 methanol/oil molar ratio. For KCR, heterogeneous catalyst for biodiesel production and C-C bond formation.
CTGL catalyst showed a promising catalytic activity towards the Renew. Energy 147, 541–555. https://doi.org/10.1016/j.renene.2019.09.016.
Göksu, H., Gültekin, E., 2017. Pd nanoparticin aluminium oxy-hydroxide: an
formation of BMN derivatives of aryl aldehydes. The widespread efficient and recyclable heterogeneous catalyst for selective knoevenagel
scope, reusability of catalyst, improved reaction rate and yield of condensatles incarcerated ion. ChemistrySelect 2, 458–463. https://doi.org/
the product are the advantages of the present protocol. 10.1002/slct.201601721.
Janaun, J., Ellis, N., 2010. Perspectives on biodiesel as a sustainable fuel. Renew. Sust.
Energ. Rev. 14, 1312–1320. https://doi.org/10.1016/j.rser.2009.12.011.
Declaration of Competing Interest Kaur, N., Ali, A., 2014. Kinetics and reusability of Zr/CaO as heterogeneous catalyst
for the ethanolysis and methanolysis of Jatropha crucas oil. Fuel Process.
Technol. 119, 173–184. https://doi.org/10.1016/j.fuproc.2013.11.002.
There are no conflicts to declare. Kawashima, A., Matsubara, K., Honda, K., 2009. Acceleration of catalytic activity of
calcium oxide for biodiesel production. Bioresour. Technol. 100, 696–700.
Acknowledgements https://doi.org/10.1016/j.biortech.2008.06.049.
Kolmykov, O., Chebbat, N., Commenge, J.M., Medjahdi, G., Schneider, R., 2016. ZIF-8
nanoparticles as an efficient and reusable catalyst for the Knoevenagel
M. Gohain thanks DBT, New Delhi, India for providing financial synthesis of cyanoacrylates and 3-cyanocoumarins. Tetrahedron Lett. 57,
support (Grant No- DBT/IC-2/Indo-Brazil/2016-19/04). K. Laskar is 5885–5888. https://doi.org/10.1016/j.tetlet.2016.11.070.
Konwar, L.J., Das, R., Thakur, A.J., Salminen, E., Mäki-Arvela, P., Kumar, N., Mikkola, J.
grateful to the DST-SERB, India for providing National Post- P., Deka, D., 2014. Biodiesel production from acid oils using sulfonated carbon
Doctoral Fellowship (PDF/2017/001364). We acknowledge SAIC catalyst derived from oil-cake waste. J. Mol. Catal. A 388, 167–176. https://doi.
Tezpur University, CSIR-NEIST Jorhat and IIT Guwahati for assisting org/10.1016/j.molcata.2013.09.031.
Kühbeck, D., Saidulu, G., Reddy, K.R., Díaz, D.D., 2012. Critical assessment of the
in performing analyses. efficiency of chitosan biohydrogel beads as recyclable and heterogeneous
organocatalyst for C-C bond formation. Green Chem. 14, 378–392. https://doi.
org/10.1039/C1GC15925A.
Appendix A. Supplementary material Lardon, L., Helias, A., Sialve, B., Steyer, J.P., Bernard, O., 2009. Life-cycle assessment
of biodiesel production from microalgae. Environ. Sci. Technol. 43, 6475–6481.
Supplementary data to this article can be found online at https://doi.org/10.1021/es900705j.
Laskar, I.B., Rajkumari, K., Gupta, R., Chatterjee, S., Paul, B., Rokhum, L., 2018. Waste
https://doi.org/10.1016/j.wasman.2019.10.049.
snail shell derived heterogeneous catalyst for biodiesel production by the
transesterification of soybean oil. RSC Adv. 8, 20131–20142. https://doi.org/
References 10.1039/C8RA02397B.
Lee, A.F., Bennett, J.A., Manayil, J.C., Wilson, K., 2014. Heterogeneous catalysis for
sustainable biodiesel production via esterification and transesterification.
Agarwal, S., Singhal, S., Singh, M., Arora, S., Tanwer, M., 2018. Role of antioxidants in
Chem. Soc. Rev. 43, 7887–7916. https://doi.org/10.1039/C4CS00189C.
enhancing oxidation stability of biodiesels. ACS Sustain. Chem. Eng. 6, 11036–
Li, J., Guo, Z., 2017. Catalytic biodiesel production mediated by amino acid-based
11049. https://doi.org/10.1021/acssuschemeng.8b02523.
protic salts. ChemSusChem 10, 1792–1802. https://doi.org/10.1002/
Ajala, O.E., Aberuagba, F., Odetoye, T.E., Ajala, A.M., 2015. Biodiesel: sustainable
cssc.201700026.
energy replacement to petroleum-based diesel fuel–A review. ChemBioEng Rev.
Liu, H., Shuang Guo, H., Jing Wang, X., Zhong Jiang, J., Lin, H., Han, S., Peng Pei, S.,
2, 145–156. https://doi.org/10.1002/cben.201400024.
2016. Mixed and ground KBr-impregnated calcined snail shell and kaolin as
Betiku, E., Akintunde, A.M., Ojumu, T.V., 2016. Banana peels as a biobase catalyst for
solid base catalysts for biodiesel production. Renew. Energ. 93, 648–657.
fatty acid methyl esters production using Napoleon’s plume (Bauhinia
https://doi.org/10.1016/j.renene.2016.03.017.
monandra) seed oil: A process parameters optimization study. Energy 103,
Liu, X., He, H., Wang, Y., Zhu, S., 2007. Transesterification of soybean oil to biodiesel
797–806. https://doi.org/10.1016/j.energy.2016.02.138.
using SrO as a solid base catalyst. Catal. Commun. 8, 1107–1111. https://doi.
Bhunia, A., Dey, S., Moreno, J.M., Diaz, U., Concepcion, P., Van Hecke, K., Janiak, C.,
org/10.1016/j.catcom.2006.10.026.
Van Der Voort, P., 2016. A homochiral vanadium–salen based cadmium bpdc
Long, Y.D., Fang, Z., Su, T.C., Yang, Q., 2014. Co-production of biodiesel and hydrogen
MOF with permanent porosity as an asymmetric catalyst in solvent-free
from rapeseed and Jatropha oils with sodium silicate and Ni catalysts. Appl.
cyanosilylation. Chem. Commun. 52, 1401–1404. https://doi.org/10.1039/
Energy 113, 1819–1825. https://doi.org/10.1016/j.apenergy.2012.12.076.
C5CC09459C.
Ma, F., Hanna, M.A., 1999. Biodiesel production: a review. Bioresour. Technol. 70, 1–
Chouhan, A.P.S., Sarma, A.K., 2013. Biodiesel production from Jatropha curcas L. oil
15. https://doi.org/10.1016/S0960-8524(99)00025-5.
using Lemna perpusilla Torrey ash as heterogeneous catalyst. Biomass
Mansir, N., Teo, S.H., Rabiu, I., Taufiq-Yap, Y.H., 2018. Effective biodiesel synthesis
Bioenergy 55, 386–389. https://doi.org/10.1016/j.biombioe.2013.02.009.
from waste cooking oil and biomass residue solid green catalyst. Chem. Eng. J.
Daneshvar, N., Shirini, F., Langarudi, M.S.N., Karimi-Chayjani, R., 2018. Taurine as a
347, 137–144. https://doi.org/10.1016/j.cej.2018.04.034.
green bio-organic catalyst for the preparation of bio-active barbituric and
Martins, L., Vieira, K.M., Rios, L.M., Cardoso, D., 2008. Basic catalyzed Knoevenagel
thiobarbituric acid derivatives in water media. Bioorganic Chem. 77, 68–73.
condensation by FAU zeolites exchanged with alkylammonium cations. Catal.
https://doi.org/10.1016/j.bioorg.2017.12.021.
Today 133, 706–710. https://doi.org/10.1016/j.cattod.2007.12.043.
Deka, D.C., Basumatary, S., 2011. High quality biodiesel from yellow oleander
Meher, L.C., Sagar, D.V., Naik, S.N., 2006. Technical aspects of biodiesel production
(Thevetia peruviana) seed oil. Biomass Bioenergy 35, 1797–1803. https://doi.
by transesterification-a review. Renew. Sust. Energ. Rev. 10, 248–268. https://
org/10.1016/j.biombioe.2011.01.007.
doi.org/10.1016/j.rser.2004.09.002.
Devadiga, A., Shetty, K.V., Saidutta, M.B., 2015. Timber industry waste-teak (Tectona
Monteiro, M.R., Ambrozin, A.R.P., Liao, L.M., Ferreira, A.G., 2009. Determination of
grandis Linn.) leaf extract mediated synthesis of antibacterial silver
biodiesel blend levels in different diesel samples by 1H NMR. Fuel 88, 691–696.
nanoparticles. International. Nano Lett. 5, 205–214. https://doi.org/10.1007/
https://doi.org/10.1016/j.fuel.2008.10.010.
s40089-015-0157-4.
Nalvothula, R., Nagati, V.B., Koyyati, R., Merugu, R., Padigya, P.R.M., 2014. Biogenic
Diehl, B., Randel, G., 2007. Analysis of biodiesel, diesel and gasoline by NMR
synthesis of silver nanoparticles using Tectona grandis leaf extract and
spectroscopy–a quick and robust alternative to NIR and GC. Lipid Technol. 19,
evaluation of their antibacterial potential. Int. J. Chem. Tech. Res. 6, 293–298
258–260. https://doi.org/10.1002/lite.200700087.
http://sphinxsai.com/2014/ChemTech/JM14CT1_50/CT=34(293-298)JM14.pdf.
Ding, Y., Ni, X., Gu, M., Li, S., Huang, H., Hu, Y., 2015. Knoevenagel condensation of
Nefedov, V.I., Buslaev, Y.A., Kokunov, Y.V., 1974. X-ray electron spectroscopy of
aromatic aldehydes with active methylene compounds catalyzed by lipoprotein
alkali metal fluorides and alkaline earth metal fluorides. Zh. Neorg. Khim. 19,
lipase. Catal. Commun. 64, 101–104. https://doi.org/10.1016/
1166–1169 https://inis.iaea.org/search/search.aspx?orig_q=RN:6162099.
j.catcom.2015.02.007.
Nisar, J., Razaq, R., Farooq, M., Iqbal, M., Khan, R.A., Sayed, M., Shah, A., ur Rahman, I.,
Ezebor, F., Khairuddean, M., Abdullah, A.Z., Boey, P.L., 2014. Oil palm trunk and
2017. Enhanced biodiesel production from Jatropha oil using calcined waste
sugarcane bagasse derived heterogeneous acid catalysts for production of fatty
animal bones as catalyst. Renew. Energ. 101, 111–119. https://doi.org/10.1016/
acid methyl esters. Energy 70, 493–503. https://doi.org/10.1016/j.
j.renene.2016.08.048.
energy.2014.04.024.
M. Gohain et al. / Waste Management 102 (2020) 212–221 221

Noureddini, H., Zhu, D., 1997. Kinetics of transesterification of soybean oil. J. Tang, S., Wang, L., Zhang, Y., Li, S., Tian, S., Wang, S.B., 2012. Study on preparation of
Am. Oil. Chem. Soc. 74, 1457–1463. https://doi.org/10.1007/s11746-997- Ca/Al/Fe3O4 magnetic composite solid catalyst and its application in biodiesel
0254-2. transesterification. Fuel Process. Technol. 95, 84–89. https://doi.org/10.1016/
Olutoye, M.A., Lee, S.C., Hameed, B.H., 2011. Synthesis of fatty acid methyl ester j.fuproc.2011.11.022.
from palm oil (Elaeis guineensis) with Ky (MgCa)2xO3 as heterogeneous Tariq, M., Ali, S., Ahmad, F., Ahmad, M., Zafar, M., Khalid, N., Khan, M.A., 2011.
catalyst. Bioresour. Technol. 102, 10777–10783. https://doi.org/10.1016/j. Identification, FT-IR, NMR (1H and 13C) and GC/MS studies of fatty acid methyl
biortech.2011.09.033. esters in biodiesel from rocket seed oil. Fuel Process. Technol. 92, 336–341.
Parida, K.M., Mallick, S., Sahoo, P.C., Rana, S.K., 2010. A facile method for synthesis of https://doi.org/10.1016/j.fuproc.2010.09.025.
amine-functionalized mesoporous zirconia and its catalytic evaluation in Teo, S.H., Islam, A., Ng, C.H., Mansir, N., Ma, T., Thomas Choong, S.Y., Taufiq-Yap, Y.
Knoevenagel condensation. Appl. Catal. A. 381, 226–232. https://doi.org/ H., 2018. Methoxy-functionalized mesostructured stable carbon catalysts for
10.1016/j.apcata.2010.04.008. effective biodiesel production from non-edible feedstock. Chem. Eng. J. 334,
Pathak, G., Das, D., Rajkumari, K., Rokhum, L., 2018. Exploiting waste: towards a 1851–1868. https://doi.org/10.1016/j.cej.2017.11.110.
sustainable production of biodiesel using Musa acuminata peel ash as a Toda, M., Takagaki, A., Okamura, M., Kondo, J.N., Hayashi, S., Domen, K., Hara, M.,
heterogeneous catalyst. Green Chem. 20, 2365–2373. https://doi.org/10.1039/ 2005. Green chemistry: biodiesel made with sugar catalyst. Nature 438, 178.
C8GC00071A. https://doi.org/10.1038/438178a.
Piker, A., Tabah, B., Perkas, N., Gedanken, A., 2016. A green and low-cost room Tran, U.P.N., Le, K.K.A., Phan, N.T.S., 2011. Expanding applications of metal organic
temperature biodiesel production method from waste oil using egg shells as frameworks: zeolite imidazolate framework ZIF-8 as an efficient heterogeneous
catalyst. Fuel 182, 34–41. https://doi.org/10.1016/j.fuel.2016.05.078. catalyst for the knoevenagel reaction. ACS Catal. 1, 120–127. https://doi.org/
Rafiei, S., Tangestaninejad, S., Horcajada, P., Moghadam, M., Mirkhani, V., 10.1021/cs1000625.
Mohammadpoor-Baltork, I., Kardanpour, R., Zadehahmadi, F., 2018. Efficient Vadery, V., Narayanan, B.N., Ramakrishnan, R.M., Cherikkallinmel, S.K., Sugunan, S.,
biodiesel production using a lipase@ ZIF-67 nanobioreactor. Chem. Eng. J. 334, Narayanan, D.P., Sasidharan, S., 2014. Room temperature production of jatropha
1233–1241. https://doi.org/10.1016/j.cej.2017.10.094. biodiesel over coconut husk ash. Energy 70, 588–594. https://doi.org/10.1016/j.
Roschat, W., Siritanon, T., Yoosuk, B., Promarak, V., 2016. Rice husk-derived sodium energy.2014.04.045.
silicate as a highly efficient and low-cost basic heterogeneous catalyst for Vyas, A.P., Subrahmanyam, N., Patel, P.A., 2009. Production of biodiesel through
biodiesel production. Energy Convers. Manag. 119, 453–462. https://doi.org/ transesterification of Jatropha oil using KNO3/Al2O3 solid catalyst. Fuel 88, 625–
10.1016/j.enconman.2016.04.071. 628. https://doi.org/10.1016/j.fuel.2008.10.033.
Sakthivel, B., Dhakshinamoorthy, A., 2017. Chitosan as a reusable solid base catalyst Wan, L., Liu, H., Skala, D., 2014. Biodiesel production from soybean oil in subcritical
for Knoevenagel condensation reaction. J. Colloid. Interface. Sci. 485, 75–80. methanol using MnCO3/ZnO as catalyst. Appl. Catal. B. 152, 352–359. https://
https://doi.org/10.1016/j.jcis.2016.09.020. doi.org/10.1016/j.apcatb.2014.01.033.
Sano, N., Yamada, K., Tsunauchi, S., Tamon, H., 2017. A novel solid base catalyst for Wang, J., Xing, S., Huang, Y., Fan, P., Fu, J., Yang, G., Yang, L., Lv, P., 2017a. Highly
transesterification of triglycerides toward biodiesel production: Carbon stable gasified straw slag as a novel solid base catalyst for the effective
nanohorn dispersed with calcium ferrite. Chem. Eng. J. 307, 135–142. https:// synthesis of biodiesel: Characteristics and performance. Appl. Energy 190, 703–
doi.org/10.1016/j.cej.2016.08.010. 712. https://doi.org/10.1016/j.apenergy.2017.01.004.
Sharma, M., Khan, A.A., Puri, S.K., Tuli, D.K., 2012. Wood ash as a potential Wang, S., Zhao, C., Shan, R., Wang, Y., Yuan, H., 2017b. A novel peat biochar
heterogeneous catalyst for biodiesel synthesis. Biomass Bioenergy 41, 94–106. supported catalyst for the transesterification reaction. Energy Convers. Manag.
https://doi.org/10.1016/j.biombioe.2012.02.017. 139, 89–96. https://doi.org/10.1016/j.enconman.2017.02.039.
Sharma, Y.C., Singh, B., Korstad, J., 2009. High yield and conversion of biodiesel from Wiersema, J. H., GRIN Taxonomy. US National Plant Germplasm System. Checklist
a nonedible feedstock (Pongamia pinnata). J. Agric. Food Chem. 58, 242–247. dataset (2019) https://doi.org/10.15468/ao14pp accessed via GBIF.org on 2019-
https://doi.org/10.1021/jf903227e. 08-21.
Steen, E.J., Kang, Y., Bokinsky, G., Hu, Z., Schirmer, A., McClure, A., Cardayre, S.B., Wilson, K., Hardacre, C., Lee, A.F., Montero, J.M., Shellard, L., 2008. The application of
Keasling, J.D., 2010. Microbial production of fatty-acid-derived fuels and calcined natural dolomitic rock as a solid base catalyst in triglyceride
chemicals from plant biomass. Nature 463, 559. https://doi.org/ transesterification for biodiesel synthesis. Green Chem. 10, 654–659. https://
10.1038/nature08721. doi.org/10.1039/B800455B.
Stork, W.H.J., Pott, G.T., 1974. Studies of compound formation on alkali/gamma.- Xie, W., Huang, X., 2006. Synthesis of biodiesel from soybean oil using
aluminum oxide catalyst systems using chromium, iron, and manganese heterogeneous KF/ZnO catalyst. Catal. Lett. 107, 53–59. https://doi.org/
luminescence. J. Phys. Chem. 78, 2496–2506 https://pubs.acs.org/doi/pdf/10. 10.1007/s10562-005-9731-0.
1021/j100617a016. Xie, W., Li, H., 2006. Alumina-supported potassium iodide as a heterogeneous
Tan, Y.H., Abdullah, M.O., Nolasco-Hipolito, C., Taufiq-Yap, Y.H., 2015. Waste catalyst for biodiesel production from soybean oil. J. Mol. Catal. A. 255, 1–9.
ostrich-and chicken-eggshells as heterogeneous base catalyst for biodiesel https://doi.org/10.1016/j.molcata.2006.03.061.
production from used cooking oil: Catalyst characterization and biodiesel yield Xiong, Y., Miao, W.F., Wang, N.N., Chen, H.M., Wang, X.R., Wang, J.Y., Tan, Q.I., Chen,
performance. Appl. Energy 160, 58–70. https://doi.org/10.1016/j. S.P., 2019. Solid alcohol based on waste cooking oil: Synthesis, properties,
apenergy.2015.09.023. micromorphology and simultaneous synthesis of biodiesel. Waste Manag. 85,
Tan, Y.H., Abdullah, M.O., Kansedo, J., Mubarak, N.M., San Chan, Y., Nolasco-Hipolito, 295–303. https://doi.org/10.1016/j.wasman.2018.12.036.
C., 2019. Biodiesel production from used cooking oil using green solid catalyst Zhao, C., Yang, L., Xing, S., Luo, W., Wang, Z., Lv, P., 2018. Biodiesel production by a
derived from calcined fusion waste chicken and fish bones. Renew. Energy 139, highly effective renewable catalyst from pyrolytic rice husk. J. Clean. Prod. 199,
696–706. https://doi.org/10.1016/j.renene.2019.02.110. 772–780. https://doi.org/10.1016/j.jclepro.2018.07.242.

You might also like