You are on page 1of 16

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

net/publication/335693992

Carica papaya stem: A source of versatile heterogeneous catalyst for biodiesel


production and C–C bond formation

Article  in  Renewable Energy · September 2019


DOI: 10.1016/j.renene.2019.09.016

CITATIONS READS

25 359

9 authors, including:

Khairujjaman Laskar Atanu Kumar Paul


Tezpur University Indian Institute of Technology Guwahati
13 PUBLICATIONS   78 CITATIONS    18 PUBLICATIONS   61 CITATIONS   

SEE PROFILE SEE PROFILE

Niran Daimary Mrutyunjay Maharana


Tezpur University Xi'an Jiaotong University
2 PUBLICATIONS   30 CITATIONS    27 PUBLICATIONS   87 CITATIONS   

SEE PROFILE SEE PROFILE

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

E-Nose Design and Evaluation View project

Biolubricant View project

All content following this page was uploaded by Atanu Kumar Paul on 23 September 2019.

The user has requested enhancement of the downloaded file.


Renewable Energy 147 (2020) 541e555

Contents lists available at ScienceDirect

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

Carica papaya stem: A source of versatile heterogeneous catalyst for


biodiesel production and CeC bond formation
Minakshi Gohain a, *, Khairujjaman Laskar b, Atanu Kumar Paul c, Niran Daimary a,
Mrutyunjay Maharana d, Imon Kalyan Goswami a, Anil Hazarika e, Utpal Bora b,
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
Bio-Energy Research Laboratory, Department of Chemical Engineering, Indian Institute of Technology, Guwahati, 781039, Assam, India
d
Centre for Energy, Indian Institute of Technology, Guwahati, 781039, Assam, India
e
Sophisticated Analytical Instrumentation Centre, Tezpur University, Napaam, 784028, Assam, India

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

Article history: Development of solid mixed oxide catalyst from waste biomass is a scarcely studied area. Thus, present
Received 30 January 2019 protocol aims to prepare an environmentally friendly, efficient, renewable and recyclable heterogeneous
Received in revised form base catalyst from Carica papaya stem. The chemical and structural properties of the catalyst were
25 July 2019
examined by Fourier-transform infrared spectroscopy (FTIR), X-ray diffractograms (XRD), Scanning
Accepted 6 September 2019
Available online 9 September 2019
electron microscopy (SEM), Energy Dispersive X-ray spectrometry (EDX), Transmission Electron Micro-
scopy (TEM) and Brunauer-Emmett-Teller (BET) analysis. The CO2-TPD and Hammett indicator test was
conducted to determine the basicity of the prepared catalyst. The study revealed the presence of alkali
Keywords:
Carica papaya stem
and alkaline earth metals that provide the basic sites to facilitate transesterification reaction for biodiesel
Waste production and formation of benzylidenemalononitrile (BMN). The conversion of the waste cooking oil
Heterogeneous catalyst (WO) and Scenedesmus obliquus (SO) lipid to biodiesel was confirmed by the NMR and Gas chroma-
Biodiesel tography Mass Spectroscopy (GC-MS) technique. Biodiesel conversions of 95.23% and 93.33% were
Knoevenagel condensation achieved using 2 wt % catalyst loading under optimized reaction conditions for WO and SO respectively.
Reusing the catalyst showed a slight drop in activity after 6 repeated uses. The reported catalyst has
shown its potential as an alternative and cheaper green solid catalyst for biodiesel production and
Knoevenagel reaction.
© 2019 Published by Elsevier Ltd.

1. Introduction fuel substitute [3]. Biodiesel is produced by transesterification of


triglycerides of oils or fats with alcohol (methyl or ethyl) in the
The world has witnessed much of the emphasis on energy and presence of a suitable catalyst [4,5].
related issues in the last few decades. The issue of continuous Researchers have reported biodiesel as an oxygenated, biode-
depletion of fossil fuel reserve along with the increase in fuel prices gradable and non-toxic fuel [6,7] Although biodiesel has been re-
has drawn global attention towards the development of alterna- ported as a potential substitute for petroleum diesel, much of the
tives that would be renewable, sustainable, efficient and cost- works involve homogeneous catalyst for its synthesis which is
effective [1,2]. Different countries have set goals to replace a part chemically synthesized and is toxic as well as corrosive in nature
of their energy requirement from renewable sources; biodiesel [8]. Use of homogeneous catalyst leads to additional cost in
being considered as a frontrunner to be taken up as an automobile wastewater disposal generated during its separation from the re-
action mixture [9]. On the other hand, the heterogeneous solid
catalyst has been reported to be advantageous due to its insolubility
in solvent or esters and hence can easily be recovered and reused
* Corresponding author. Department of Energy, Tezpur University, Napaam,
784028.
[10].
E-mail addresses: minakshiigohain@gmail.com, min_ene@tezu.ernet.in In recent years, the utilization of eco-friendly and reusable
(M. Gohain).

https://doi.org/10.1016/j.renene.2019.09.016
0960-1481/© 2019 Published by Elsevier Ltd.
542 M. Gohain et al. / Renewable Energy 147 (2020) 541e555

heterogeneous catalyst has emerged as a major constituent of green would be non-toxic, cost effective, renewable, recyclable and
chemistry [11]. Solid acids such as sulfated zirconia [12,13], sulfo- environmentally friendly in nature having widespread application
nated carbons [14e16], Amberlyst-15 [17] have attracted a lot of i.e. for transesterification as well as Knoevenagel condensation
attention as an effective catalyst for biodiesel production. But, in reaction. Thus, a newer approach for the preparation of green
the context of sustainability, economy and ecofriendly, the use of heterogeneous solid base catalyst under the protocol of bio-waste
renewable catalysts prepared from biomass have been increasingly utilization has effectively been investigated which meets the
targeted. Waste biomass has been fabricated to prepare acid and widespread scope of chemical transformation.
alkali catalysts using activated carbon [18,19] as catalyst support. A
number of works on the preparation of heterogeneous catalyst 2. Experimental
from waste biomass viz. coconut shell [20], palm shell [21], wood
ash [22], deoiled cake waste [23] etc., have been reported. Activated 2.1. Materials
carbon support based catalyst provides more specific surface area
and pores for active species which results in excellent catalytic Carica Papaya stem and WO were collected from the nearby
ability. But, the cost of production of such catalyst must be factored locality. The SO was obtained from Gauhati University, India.
in catalyst fabrication such as high carbonization temperature and Chemicals of analytical grades used in the present work were
wasteful chemical reaction during functionalization of the activated purchased from Sigma Aldrich and used without further
carbon. However, CaO based heterogeneous catalysts have been purification.
extensively synthesized without any chemical functionalization
from renewable sources like an eggshell [24], waste crab shell [25], 2.2. Extraction of lipid
etc. In this context, mixed oxide catalysts from waste biomass have
drawn little attention. The extraction of lipid from SO dried biomass was carried out
The main problem associated with transesterification reaction followed by Bligh and Dyer method [43]. Dried biomass sample was
using heterogeneous catalyst is the deactivation of the catalyst with weighed and homogenized in a mortar and pestle using 0.1e0.5 g of
time generally encountered when used oils are involved which may anhydrous Na2SO4 and 1e2 mL of 2% butylated hydroxytoluene
be due to poisoning, coking, sintering, and leaching [26,27]. Het- (BHT) (2.04 g of BHT in 100 mL CHCl3). Total lipid was extracted
erogeneous catalyst should possess some hydrophobic character to from the homogenized powder with 5e10 mL mixture of CHCl3:
support triglycerides adsorption and to avoid deactivation of cat- MeOH (2:1). The residue was extracted with chloroform until it
alytic sites by strong adsorption of polar byproducts such as glyc- became colorless. The extracts were mixed together, filtered and
erol and water [28,29]. Reported literature suggests that the then transferred to a separating funnel. Then 0.9% NaCl solution (1/
addition of metals onto the catalyst surface can improve its hy- 3 of the volume) and excess CHCl3 was added to the separating
drophobicity [30e33]. To enhance the catalytic efficacy biomass funnel, mixed thoroughly and kept undisturbed overnight at room
containing metals can be exploited as a forth runner to prepare temperature. A clear biphasic layer was observed. The lower layer
such catalyst without any chemical fabrication as a part of sus- (CHCl3) containing the lipid component was collected in a clean
tainable development. glass vial. The upper layer containing the methanol-water layer was
With the advancement in medicinal and pharmaceutical washed twice with chloroform and collected similarly. All the
chemistry, higher attention has also been paid towards the devel- collected chloroform layers were mixed together and were evap-
opment of environment-friendly processes that employ nontoxic orated in a water bath at 60  C and finally dried in desiccator in
reagents, solvents and catalysts [34]. The Knoevenagel condensa- presence of anhydrous Na2SO4.
tion is one of the most efficient and widely used methods for CeC
bond formation. Over the past years, various heterogeneous cata- 2.3. Catalyst preparation and its characterization
lysts have been developed for Knoevenagel condensation including
ion exchange resins, mesoporous zirconia [35], chitosan bio- Carica Papaya stem was cut into small pieces to speed up the
hydrogel [36] etc. Several studies [37] suggest the formation of drying process, washed thoroughly using double distilled water
BMN and widespread applications of such moiety inspire for the and then dried in an oven at 80  C for 48 h till constant weight. The
development of a newer protocol. dried Carica Papaya stem (CPS) pieces were ground and sieved to a
Papaya plant has been grown extensively in most of the coun- fine powder. The sieved powdered stem was calcined in a porcelain
tries including tropical and subtropical regions of the world. Papaya crucible placed in a muffle furnace for 4 h at a temperature of
fruit has high energy value as well as it contains a lot of vitamins 700  C. The calcinated Carica Papaya stem (CCPS) was stored
and minerals. Papaya plant has been extensively used for the properly for further use.
biochemical synthesis of different pharmaceutical compounds The FTIR spectra were measured using Nicolet (Impact 410) on
[38,39]. In 2017, the total world production of papaya was KBr pellets in 4000-500 cm1 wavenumber range. The XRD of CPS
13,016,281tonnes. India is leading in papaya production with and CCPS was analyzed by a Rigaku miniflex diffractometer in 2q
45.64% of the world total [40]. With such a large amount of papaya range 10e80 (CuKa radiation, l ¼ 1.5406 Å) at 2 scanning rate.
produced, and an estimated 30e50% cull rate, there is a large The BET sorption isotherm method (NOVA 1000E, NOVA WIN,
amount of agricultural waste produced [41]. The area under papaya QUANTACHROME) was used to determine the surface area of the
cultivation in India increased from 63% i.e. 45.2 thousand hectares catalyst. The SEM and EDX study was performed using Jeol, JSM-
in 1991e1992 to 73.7 thousand hectares in 2001e2002 [42]. There 6290 LV. The TEM study was analyzed using JEOL, JEM-2100 Plus
are around 1700 papaya plants per acre [42], where stems are Electron Microscope, Japan. The Hammett indicator method was
generally 5e10 m long. Cultivators discard the stem of papaya plant used to determine the basicity. The basicity was also measured by
after they find it of no use for productivity and then, the stem be- temperature programmed desorption of CO2 (Autochem 2920,
comes waste. micrometrics). To remove adsorbed water molecules and other
Owing to this, present work emphasizes utilization of waste impurities the sample was pre-treated with Helium gas at 350  C
(discarded papaya stem) as a source for the synthesis of the bio- for 1 h, followed by cooling to room temperature. The CO2
catalyst. To the best of our knowledge, it is the first study to explore desorption was monitored in the temperature range of
Carica papaya stem as a source of heterogeneous biocatalyst which 50  Ce670  C. The static contact angle of the catalyst pellet with
M. Gohain et al. / Renewable Energy 147 (2020) 541e555 543

Fig. 1. CO2 TPD profile of CCPS.

water was measured using a contact angle meter (Holmarc Opto- 2.4. Transesterification procedure
Mechatronics, HO-IAD-CAM-01B). The wettability of CCPS, deno-
ted by contact angle (q) of a water droplet, is given by Young's The WO was pre-treated using H2SO4 to reduce the acid value
equation [Eq. (1)]: which was found to be 9.18 mg KOH/g. The acid number of WO after
pre-treatment got reduced to 1.16 mg KOH/g. The acid value of SO
gSV  gSL lipid was found to be lower than 2 mg KOH/g i.e. 1.4 mg KOH/g
cosq ¼ (1)
gLV which is required for successful transesterification and hence it was
found ideal for effective base-catalyzed transesterification reaction
Where, [45]. In a typical process, 100 mL of WO was poured to a 500 mL
three-necked round bottom flask and heated (100  C) to eliminate
gSV ¼ Interfacial surface tension of solid (V) every trace of absorbed moisture. The WO was then cooled down to
gSL ¼ Interfacial surface tension of liquid (L) 60  C. Catalyst, methanol, WO was then mixed to the reactor under
gLV ¼ Interfacial surface tension of gas vapor (V) reflux and the reaction continued for desired reaction parameters
[46]. To ensure proper mixing of MeOH and oil the agitation speed
For reusability analysis, the reaction mixture was centrifuged was maintained at 500 rpm. The reaction mixture was allowed to
(3500 rpm) and washed with methanol (MeOH), to recover the cool and transferred to a separating funnel where three distinct
catalyst and to remove impurity. The catalyst was then dried in an layers were clearly visible. The upper layer was mainly composed of
oven at 120  C, for 3 h and experiments were repeated using the FAME (fatty acid methyl ester), the middle layer of glycerine and
recovered catalysts at required parameters [44].

Fig. 2a. . The N2 adsorption-desorption isotherm of CCPS. Fig. 2b. . Pore size distribution curve for CCPS.
544 M. Gohain et al. / Renewable Energy 147 (2020) 541e555

catalyst was the bottom layer. The unreacted MeOH from the upper Y ¼ Integration value of a-methylene (eCH2e) protons at
fraction was evaporated under reduced pressure. The mixture was 2.3 ppm
then washed with distilled water (40  C) to eliminate impurities
followed by drying over Na2SO4 (anhydrous) to obtain methyl es-
ters of waste cooking oil (WME).
The 1H and 13C NMR spectra of oil and biodiesel were taken in a 2.5. Knoevenagel condensation reaction
500 MHz NMR spectrometer (Oxford, AS400, China). The GC-MS
analysis was carried out in Agilent GC, 240 Ion Trap. The progress of the reactions was monitored by TLC (Thin layer
The efficient conversion of oil to FAME was determined in terms chromatography) which were carried out on Merck silica gel (60F-
of % FAME using nuclear magnetic resonance (NMR) technique [47] 254) plates and visualized under UV light (254 nm). To achieve the
using deuterated chloroform as solvent [Eq. (2)]. desired products, purification of the reaction mixture was done by
column chromatography using silica gel (60e120 mesh). The 1H
2X NMR spectra were recorded on a JEOL, JNM ECS NMR spectrometer
CMEster ð%Þ ¼  100 (2) operating at 400 MHz in CDCl3, using TMS as an internal standard
3Y
and 13C NMR spectra were recorded at 100 MHz in CDCl3 (d ¼ 77.00)
as standard. Chemical shifts (d) are quoted in ppm and coupling
constant (J) are expressed in Hz. The structures of the purified
X ¼ Integration value of methoxy protons (-OCH3) at 3.6 ppm products were confirmed by NMR spectra (See ESI, S1*).

Fig. 3. The SEM image of CPS (a) and CCPS (b).

Fig. 4. The TEM image of CPS (a, b) and CCPS (c, d).
M. Gohain et al. / Renewable Energy 147 (2020) 541e555 545

Fig. 5. FTIR spectra of CPS (a) and CCPS (b).

Fig. 6. XRD pattern of CPS (a) and CCPS (b).

2.5.1. Synthesis of BMN derivatives


An oven dried 10 mL RB flask was charged with a mixture of aryl
aldehydes (0.30 mmol), malononitrile (0.50 mmol) and catalyst
(8 wt %) in ethanol (2 mL). The reaction mixture was magnetically
stirred and refluxed at 55  C. On completion of the reaction
(monitored by TLC), the reaction mixture allowed to cool, the
catalyst was separated from the reaction mixture by centrifugation

Table 1
Elemental analysis by EDX.

Elements Mass fraction (%) (Fresh)

K 56.71
Ca 21.08
Na 14.78
Mg 4.41
Si 3.02
Fig. 7. EDX analysis of CCPS.
546 M. Gohain et al. / Renewable Energy 147 (2020) 541e555

and the filtrate was extracted with ethyl acetate (3  10 mL). The exhibit a highly porous structure, indicating its high porosity and
catalyst was washed thoroughly and then allowed to dry for further large specific area. The result is in accordance with the BET findings.
use. The resultant organic layer dried over anhydrous Na2SO4, For heterogeneous processes, highly porous catalyst achieves
evaporated under reduced pressure, and purified by column chro- higher conversion efficiency, similarly higher catalytic sites or
matography (2e4% v/v EtOAc: hexane) to afford the desired
product.
To explore the catalytic efficacy of prepared catalyst the recy-
cling potency of CCPS, for the Knoevenagel Condensation of aryl
aldehydes was tested using the model reaction of benzaldehyde
and malononitrile.

3. Results and discussion

3.1. Catalyst characterization

Hammett indicator test was conducted for basicity determina-


tion [48] and obtained according to color variation. After analysis,
the CPS and CCPS showed a basic strength of 9.8 < H  12.2 and of
9.8 < H  18.4 respectively. The increase in the basic strength after
calcination may be due to thermal activation of the metal carbonate
salts to active metal oxide [49]. Group I alkali metal hydroxide and
group II metal oxide infer higher transesterification activity due to
its higher basicity [50]. Strong bases which can remove protons Fig. 9a. Methanol/oil molar ratio effects on FAME conversion.
from MeOH and develop active methoxy species are responsible for
efficient transesterification [44].
Studies report that basicity of a catalyst is the main key in
improving the transesterification activity where the main param-
eter in higher FAME conversion is the number of strong basic sites
[51]. Fig. 1 depicts the TPD profile of CO2 adsorbed on CCPS. At a
higher temperature region i.e. above 400  C, desorption peak was
seen at 653.85  C, while two peaks were observed at 191.8  C and
309.2  C which could be ascribed to weak CO2 adsorption on OH
groups, moderate CO2 adsorption on metal-oxygen pairs and
strong adsorption on O2 anions [52]. These results strongly sug-
gest that both strong and moderately basic sites are present on the
CCPS [53].
As shown in Fig. 2 a, the N2 adsorption-desorption isotherm of
CCPS exhibited typical type IV and hysteresis loops type H3 ac-
cording to the Brunauer-Deming-Deming-Teller (BDDT) classifica-
tion, which indicates that CCPS is typical of a mesoporous in nature
[54]. The surface area of prepared catalyst (CCPS) was found to be
78.681 m2/g; with pore volume and pore radius of 0.349 cc/g and
1.6074 nm respectively. The surface area was found to be higher
Fig. 9b. Catalyst loading effects on FAME conversion.
than other reported works [22,55]. The corresponding pore size
distribution curve (Fig. 2 b) confirmed that the composite is mainly
mesoporous.
The SEM images (Fig. 3) of CPS and CCPS revealed both meso-
porous and microporous structures similar to the previously re-
ported literature [56]. From Fig. 3, it is clear that both CPS and CCPS

Fig. 8. Hydrophobicity test of CCPS. Fig. 9c. Reaction time effects on FAME conversion.
M. Gohain et al. / Renewable Energy 147 (2020) 541e555 547

higher surface areas are also essential [57]. Sintering of aggregates vibrations of O:H and KeO bonds respectively. The band at
was observed after the calcination of CPS which resulted in 1655 cm1, 1385 cm1, and 1129 cm1 are due to carbonate CO vi-
agglomeration (Fig. 3 b), and spongy nature was observed [58]. brations, which indicate the existence of carbonate (CO3). The band
According to previously reported literature, potassium leaching can at 2400-2000 cm1 are ascribed to N:O:K bonds (N¼ Mg, Si, etc).
be prevented by calcination, and hence may increase the catalyst The absorption band at 550 cm1 and 1000 cm1 is due to Ca:O,
reusability [22]. K:O stretching and Si:O:Si bonds respectively [58,59] (Fig. 5 a and
TEM images for CPS and CCPS showed a highly porous structure, 5 b). A strong characteristic peak at 1385 cm1 denotes the pres-
although they are not morphologically homogeneous (Fig. 4). The ence of K2CO3 which is prominent only in the CBPA due to car-
more microporous and less mesoporous structure were evident bonates produced by the adsorption of atmospheric CO2 onto metal
after calcination (Fig. 4 c and 4 d) which also supports the images oxides [22] (Fig. 5 b).
depicts in SEM analysis (Fig. 4). The microporous structure of the The occurrence of crystalline compounds in CPS and CCPS were
catalyst provides an adequate external surface which acts as active identified by XRD analysis (Fig. 6) wherein calcium and potassium
sites for rapid mass transfer for the transesterification process [55]. compounds in carbonates, oxides, sulfides, and silicates forms were
Fig. 5 shows the FTIR spectra of CPS and CCPS. Characteristic observed (Fig. 6 a and 6 b). Increase in potassium compounds can
bands at 3384 cm1 and 669 cm1 were attributed to the stretching be seen after calcination (Fig. 6 b) which is depicted by a change in

Fig. 10. Reusability of CCPS in biodiesel production.

Fig. 11. FTIR spectra of reused CCPS catalyst.


548 M. Gohain et al. / Renewable Energy 147 (2020) 541e555

intensity for potassium diffraction lines which had strong basic 3.3. Recycling and reusability of CCPS
sites [60] which is also confirmed by the K2CO3 dominance in FTIR
spectra of CCPS (Fig. 6 b). Strong characteristic diffraction lines Economical application of catalyst for biodiesel production de-
(Fig. 6) of K2O and K2CO3 were observed at 2q ¼ 25, 26, 31, 33, 41, mands recycling and non-leaching capacity of the catalyst [69].
44, 46 and the diffraction lines observed at 2q ¼ 30, 39, 43, 47 are MeOH/oil molar ratio of 9:1, CCPS of 2 wt % for 3 h maintaining the
due to CaO existence.35 Other small diffraction lines were also temperature of 60  C and 600 rpm was taken as the optimized re-
observed corresponding to KCl, K2SO4, K2MgSiO4 (See ESI, S14*). action parameters (Fig. 10). Reduced FAME conversion after the
The EDX analysis (Fig. 7) (Table 1) strongly supported the find- fourth cycle may be due to the leaching of active sites. Change in the
ings of XRD and FTIR which revealed a high distribution of K, Ca, Na, accessible basic sites due to active metal components lost present
Mg and Si oxides. in the catalyst or loss of the surface area of the base catalysts is
It was seen from the results that K is the major ingredient and basically the main reason for solid base catalysts deactivation. Here,
K2O being a strong base is mainly responsible for the basicity of the most important reason for the deactivation of the catalyst is due
CCPS. to the K leaching, as K2O is significantly soluble in glycerol and
The surface hydrophobicity was measured according to the methanol. The K and Ca have a tendency to get leached more
water contact angle (Fig. 8). The surface wettability can be divided readily than other elements [70]. Loss in catalyst quantity during
into four different categories: superhydrophilic (q < 10 ), hydro- transfer between each successive cycle might have resulted in
philic (10 < q < 90 ), hydrophobic (90 < q < 150 ), and super- reduced conversion. Another reason for reduced FAME conversion
hydrophobic (q > 150 ) [33]. The CCPS exhibited a hydrophobic may be that the inert species has absorbed the available active sites
nature (q ¼ 106.70 ). or has reacted to form other less active species. The trans-
esterification reaction is not only catalyzed by basic sites present on
the catalyst surface but also sometimes by soluble substances [71].
3.2. Optimization of parameters for biodiesel synthesis The recycled catalyst after the fifth cycle was tested to analyze
the leaching of K and Ca from CCPS during successive reaction cy-
One of the essential factors influencing the conversion of oil to cles (Fig. 11).
FAME is alcohol/oil molar ratio [61]. Various operating parameters The peak at 1389 cm1 and 1000 cm1 became very weak in the
such as quality of the oil used in the reaction and the ability of the reused CCPS which signifies the loss of K and Si resulting in the
catalyst to catalyze are also associated along with. Influence of availability of lower basic sites (Fig. 11). The bands at 1050 cm1 and
molar ratio of methanol/oil i.e. 3:1 to 12:1 on FAME conversion was
studied and found that the conversion % was high with higher
MeOH/oil molar ratio (Fig. 9a). Catalyst loading was kept at 2 wt %
and maintaining temperature (60  C, 3 h). A higher amount of
methanol promotes methoxide formation which helps in shifting
the reaction equilibrium to forward direction but, it may also result
in reversible transesterification [62] depending on the parameters
as seen in case of MeOH/oil molar ratio 15:1. The reason might be
that glycerol would have chiefly dissolved in too much methanol
and consequently inhibiting the transesterification reaction. The
optimum MeOH/oil molar ratio was found to be 9:1.
Fig. 9b shows the influence of different loading of catalyst on
biodiesel conversion. Various loading of CCPS was employed for the
conversion of WO to WME. In this study, the transesterification
reaction was performed with MeOH/oil molar ratio of 9:1 for 3 h.
The catalyst amount was varied from 1 to 4 wt % maintaining the
temperature at 60  C (Fig. 9b). At 1 wt % catalyst loading 85% FAME
conversion was achieved which may be due to the availability of a
Fig. 12. XRD spectra of reused catalyst.
less number of basic sites. As the catalyst loading increases the
FAME conversion increased and then remains constant. With
higher catalyst loading the number of active sites increases [45,63],
but beyond a certain optimum catalyst amount the biodiesel con-
version remains constant [64,65]. Higher concentrations of catalyst
might have made the reaction mixture more viscous which has
resisted the mass transfer [66] and since the reaction is base
catalyzed, saponification side reaction might have induced result-
ing in lower FAME conversion [67].
Influence of time on FAME conversion was examined using 9:1
(MeOH to oil molar ratio) and optimum catalyst amount of 2 wt %
(Fig. 9c). Sufficient contact time provides ample contact among the
reactants and catalyst resulting in the successful transesterification
reaction [64]. The reaction was carried out for 5 h and FAME con-
version was observed after every 1 h. The conversion speed is
usually slow at first during transesterification owing to slow
dispersion of triglycerides into the methanol phase. As the reaction
progresses the FAME i.e. generated acts as cosolvent and hence
enhance the miscibility of triglycerides and methanol resulting in
faster conversion until the reaction attains equilibrium [68]. Fig. 13. The N2 adsorption-desorption isotherm of reused CCPS.
M. Gohain et al. / Renewable Energy 147 (2020) 541e555 549

Fig. 14. 1H NMR spectra of WO (a) and WME (b).


550 M. Gohain et al. / Renewable Energy 147 (2020) 541e555

13
Fig. 15. C NMR spectra of WO (a) and WME (b).
M. Gohain et al. / Renewable Energy 147 (2020) 541e555 551

Fig. 16. GCMS chromatogram of WME.

at 2800 cm1 exhibits eC-O (primary alcohol) and CeH stretching catalyst which lowered the mass transfer during the reaction
indicating calcium methoxide was formed onto the surface of CCPS resulting in lower FAME conversion [72].
as CaO reacts with glycerine and methanol to produce calcium The XRD analysis of the recovered catalyst was investigated after
methoxide. The band at 1750 cm1 indicates esters adsorption on the sixth run (Fig. 12), and a sharp decrease in the intensity of
CCPS. The Ca(OCH3)2 and absorbed esters blocked the pores of the potassium oxide was observed. This clearly confirmed that

Table 2 Table 3
Fatty acid methyl esters composition of WME. Fuel properties of the feedstock.

Retention Time Component Property WO SO

21.392 min Propanoic acid, methyl ester Acid value (mg KOH/g) 1.18 1.4
27.517 min 8,11-Octadecadienoic acid, methyl ester Iodine value (g I2/100 g) 127 130
27.759 min Propionic acid, methyl ester Kinematic Viscosity (mm2/s, at 40  C) 31 30
27.901 min 9-Octadecenoic acid, methyl ester, Density (gm/cm3at 15  C) 0.9 0.9
28.239 min 6,9-Octadecadienoic acid, methyl ester Cloud point ( C) 7 5
28.846 min 9, 12, 15-Octadecatrieonic acid, methyl ester Pour point ( C) 2 1

Fig. 17. 1H NMR spectrum of SO biodiesel.


552 M. Gohain et al. / Renewable Energy 147 (2020) 541e555

potassium oxide played a major role in FAME conversion due to its Biodiesel synthesis form SO lipid was studied using the CCPS
high basicity. catalyst to test its versatile catalytic activity. Successful FAME
The spent catalyst was also analyzed for BET (Fig. 13), and a conversion of 93.33% was achieved with SO using 9:1 methanol/oil
decrease in the surface area was observed i.e. 61.973 m2/g. The molar ratio, catalyst loading of 2 wt % within 3 h. The 1H NMR of
decrease in the surface area and pore volume is related to the Scenedesmus obliquus biodiesel (SOB) is shown in Fig. 17.
partial blocking of the pores by Ca(OCH3)2 and absorbed esters as The physicochemical properties of the WO, SO and WME, SOB
seen in Fig. 11. The BET surface area data of the used CCPS, support are shown in Table 3 and Table 4 respectively. The physicochemical
the hypothesis of pore structure collapse for the used samples [73]. properties of WME meet the FAME standard specifications.

3.4. Biodiesel characterization 3.5. Optimization condition for Knoevenagel reaction

The 1H and 13C NMR spectroscopy confirms the WO and WME. To optimize the reaction condition for the titled compound, an
The presence of glyceridic protons in the range of 4.0e5.2 ppm in investigation has been done by changing the solvent system and
1
H NMR is the characteristics of triglycerides present in WO results summarized in Table 5. The reaction of benzaldehyde
(Fig. 14a) which is totally absent in the WME (Fig. 14b). The methyl (0.30 mmol) and malononitrile (0.50 mmol) at 55  C in presence of
ester moiety protons at 3.66 ppm and a-carbonyl methylene groups CCPS to form compound 3a (Table 6) was selected as a model re-
at 2.30 ppm are due to methyl ester formation i.e. WME (Fig. 14b). A action. Initially, the reaction was allowed to stir under reflux, the
chemical shift at 5.3 ppm is due to the alkene proton which signifies prolonged reaction time and lesser yield (Table 5, entries 1, and
that the main component of WO is unsaturated fatty acid and for 3e6) were observed. The yield of the product was good in protic
WME is methyl ester respectively. solvents like MeOH and iPrOH (Table 5, entries 2 and 7). However,
Fig. 15 (a) and Fig. 15 (b), shows the 13C NMR spectra of WO and maximum yield (93%) of the product with less reaction time was
WME respectively. Carboxyl carbons of the ester molecules in tri- obtained in EtOH (Table 5, entry 9).
glyceride are depicted at 173.28 ppm, 172.87 ppm and at In the present study, a series of benzylidenemalononitrile
68.90 ppm, 62.12 ppm were attributed to carbonyl methylene (BMN) derivatives (3a-k) (Table 6) have been synthesized using
groups (Fig. 15 a). The WME showed only the presence of carboxyl environmentally benign catalyst (CCPS). Table 6 shows the cat-
carbon at 174.30 ppm chemical shift (Fig. 15 b). The methoxy carbon alytic merit of formation of BMN derivatives in high yield, with a
of methyl esters (WME) can be clearly observed at the chemical short span of time (11e32 min) at 55  C. In view of high catalytic
shift of 51.42 ppm (Fig. 15 b) which confirms that WO was suc- efficacy of CCPS towards Knoevenagel condensation of aryl al-
cessfully transformed to biodiesel [57]. dehydes with malononitrile, substituted benzaldehydes were
The WME chromatogram has been presented in Fig. 16. Various studied under optimized reaction conditions. As nucleophilic
methyl esters peaks were observed at different retention time as addition is a rate-determining step in the Knoevenagel
shown in Table 2. condensation, electron-withdrawing groups are more reactive

Table 4
Fuel properties of FAME.

Fuel Property WME SOB ASTM D 6751 EN 14214

Acid value (mg KOH/g) 0.06 0.2 <0.80 0.5 max


Kinematic Viscosity (mm2/s, at 40  C) 3.10 4.1 1.9e6.0 3.5e5.0
Density (gm/cm3at 15  C) 0.87 0.86 0.86e0.90 0.85
Cloud point ( C) 6 8 e e
Pour point ( C) 10 12 e e
Flash point ( C) 150 115 93 min 120 min
Carbon residue (% wt) 0.016 0.021 0.050 max 0.3
Calorific value (MJ kg1) 40.30 40 e 35 min
Cetane Number 57 58 47 min 51 min

Table 5
Optimization conditions for Knoevenagel condensation of benzaldehyde with malononitrilea.

b
Entry Solvent Catalyst (mg) Time (min) Yield (%)c

1 H2O 2 120 Trace


2 MeOH 2 120 60
3 CH2Cl2 2 120 Trace
4 CH3CN 2 120 Trace
5 MeOH/H2O (2:1) 2 60 45
6 THF 2 60 40
7 i-PrOH 2 60 55
8 EtOH/H2O (1:1) 2 60 50
9 EtOH 2 12 93
10 EtOH e 180 Trace
a
Reaction condition: Benzaldehyde (0.30 mmol), malononitrile (0.50 mmol), catalyst (8 wt% of substrate) and 55  C.
b
Reaction progress monitored by TLC.
c
Isolated yield.
M. Gohain et al. / Renewable Energy 147 (2020) 541e555 553

Table 6
Knoevenagel condensation reaction of aryl aldehyde derivatives with malononitrile in the presence of CCPSa.

Entry Substrate Product Time (min)b Yield (%)c

1 12 93

2 12 94

3 11 95

4 11 95

5 15 90

6 17 88

7 18 82

8 13 85

9 14 93

10 15 91

11 32 76

a
Reaction condition: Aldehydes (0.30 mmol), malononitrile (0.50 mmol), catalyst (8 wt% of the substrate) at 55  C.
b
Reaction progress monitored by TLC.
c
Isolated yield.

than that of electron-donating groups [74] in BMN derivatives 4. Conclusion


formation. It is evident from Table 6, electron-withdrawing
groups (-Cl, -Br, eNO2) substituted benzaldehyde resulted in In conclusion, a nature benign, renewable, recyclable, hydro-
the higher condensed product (entries 2e4, and 8e10) than phobic solid base catalyst was synthesized from waste Carica
those of electron-donating benzaldehyde derivatives (-CH3, Papaya stem to produce biodiesel from waste cooking oil as well as
-OMe) (entries 5, 6 and 7). The successful formation of com- Scenedesmus obliquus lipid. The catalyst exhibited the highest effi-
pound 3k also observed with less yield of the product. Hence, ciency with maximum FAME conversion of 95.23% for waste
the reaction of aldehydes with malononitrile, substituents on cooking oil, 93.33% SO within 3 h. Furthermore, Knoevenagel
benzaldehyde derivatives has a significant role towards catalytic condensation reaction, the presented procedure with CCPS toler-
yield. ates a wide range of aryl aldehydes giving corresponding BMN
derivatives (3a-k) in appreciable yields. The reusability analysis
showed that the catalyst still exhibited a good catalytic activity
3.6. Catalyst recycling towards biodiesel production (i.e. conversion >85%) and for
Knoevenagel condensation even after five repeated cycles. The use
The catalyst was used five times for Knoevenagel condensation of reusable green heterogeneous catalyst, substrate tolerance, re-
and it was found that the catalyst maintained good activity for a action time and high yield make this protocol an energy sustainable
minimum of five consecutive cycles and loss in the yield can be choice to that of reported methods.
tolerated as exhibited in (Fig. 18).
554 M. Gohain et al. / Renewable Energy 147 (2020) 541e555

Fig. 18. Recycling data of CCPS in Knoevenagel condensation.

Conflicts of interest sustainable biodiesel production, Renew. Sustain. Energy Rev. 70 (2017)
1040e1051.
[10] B. Wang, S. Li, S. Tian, R. Feng, Y. Meng, A new solid base catalyst for the
There are no conflicts to declare. transesterification of rapeseed oil to biodiesel with methanol, Fuel 104 (2013)
698e703.
[11] R. Ballini, in: Eco-Friendly Synthesis of Fine Chemicals, vol. 3, R.S.C, 2009.
Acknowledgements
[12] J.J. Woodford, J.P. Dacquin, K. Wilson, A.F. Lee, Better by design: nano-
engineered macroporous hydrotalcites for enhanced catalytic biodiesel pro-
M. Gohain thanks DBT, New Delhi, India for providing financial duction, Energy Environ. Sci. 5 (3) (2012) 6145e6150.
support (Grant No- DBT/IC-2/Indo-Brazil/2016-19/04). K. Laskar is [13] S. Furuta, H. Matsuhashi, K. Arata, Biodiesel fuel production with solid
superacid catalysis in fixed bed reactor under atmospheric pressure, Catal.
grateful to the DST-SERB, India for providing National Post-Doctoral Ccommun. 5 (12) (2004) 721e723.
Fellowship (PDF/2017/001364). We acknowledge SAIC Tezpur [14] R. Luque, L. Herrero-Davila, J.M. Campelo, J.H. Clark, J.M. Hidalgo, D. Luna,
University, CSIR-NEIST Jorhat, CSIR-IIP Dehradun, IIT Guwahati and J.M. Marinas, A.A. Romero, Biofuels: a technological perspective, Energy En-
viron. Sci. 1 (5) (2008) 542e564.
Gauhati University for assisting in performing analyses. Authors [15] W. Thitsartarn, S. Kawi, An active and stable CaOeCeO2 catalyst for trans-
also thank Dr. Lakshi Saikia (Sr. Scientist, CSIR-NEIST Jorhat) for esterification of oil to biodiesel, Green Chem. 13 (12) (2011) 3423e3430.
providing the facility and necessary help to carry out the CO2-TPD [16] K. Nakajima, M. Hara, Amorphous carbon with SO3H groups as a solid
Brønsted acid catalyst, ACS Catal. 2 (7) (2012) 1296e1304.
analysis. [17] M.R. Talukder, J.C. Wu, S.K. Lau, L.C. Cui, G. Shimin, A. Lim, Comparison of
Novozym 435 and Amberlyst 15 as heterogeneous catalyst for production of
Appendix A. Supplementary data biodiesel from palm fatty acid distillate, Energy Fuels 23 (1) (2008) 1e4.
[18] T. Cordero-Lanzac, R. Palos, J. M Arandes, P. Castan ~ o, J. Rodríguez-Mirasol,
T. Cordero, J. Bilbao, Stability of an acid activated carbon based bifunctional
Supplementary data to this article can be found online at catalyst for the raw bio-oil hydrodeoxygenation, Appl. Catal., B 203 (2017)
https://doi.org/10.1016/j.renene.2019.09.016. 389e399.
[19] C. Cao, L. Wei, Q. Zhai, G. Wang, J. Shen, Biomass-derived nitrogen and boron
dual-doped hollow carbon tube as cost-effective and stable synergistic cata-
References lyst for oxygen electroreduction, Electrochim. Acta 249 (2017) 328e336.
[20] B.H. Hameed, C.S. Goh, L.H. Chin, Process optimization for methyl ester pro-
[1] P. Bora, J. Boro, L.J. Konwar, D. Deka, Formulation of microemulsion based duction from waste cooking oil using activated carbon supported potassium
hybrid biofuel from waste cooking oileA comparative study with biodiesel, fluoride, Fuel Process. Technol. 90 (12) (2009) 1532e1537.
J. Energy Inst. 89 (4) (2016) 560e568. [21] S. Baroutian, M.K. Aroua, A.A.A. Raman, N.M.N. Sulaiman, Potassium hydrox-
[2] M.A. Lacatus, L.C. Bencze, M.I. Tosa, C. Paizs, F.D. Irimie, Eco-friendly enzymatic ide catalyst supported on palm shell activated carbon for transesterification of
production of 2, 5-Bis (hydroxymethyl) furan fatty acid diesters, potential palm oil, Fuel Process. Technol. 91 (11) (2010) 1378e1385.
biodiesel additives, ACS Sustain. Chem. Eng. 6 (9) (2018) 11353e11359. [22] M. Sharma, A.A. Khan, S.K. Puri, D.K. Tuli, Wood ash as a potential heteroge-
[3] C. Zapata, P. Nieuwenhuis, Exploring innovation in the automotive industry: neous catalyst for biodiesel synthesis, Biomass Bioenergy 41 (2012) 94e106.
new technologies for cleaner cars, J. Clean. Prod. 18 (1) (2010) 14e20. [23] L.J. Konwar, R. Das, A.J. Thakur, E. Salminen, P. Ma €ki-Arvela, N. Kumar,
[4] B. Dhinesh, R. Niruban Bharathi, J. Isaac JoshuaRamesh Lalvani, J.P. Mikkola, D. Deka, Biodiesel production from acid oils using sulfonated
M. Parthasarathy, K. Annamalai, An experimental analysis on the influence of carbon catalyst derived from oil-cake waste, J. Mol. Catal. A Chem. 388 (2014)
fuel borne additives on the single cylinder diesel engine powered by cym- 167e176.
bopogon flexuosus biofuel, J. Energy Inst. 90 (4) (2017) 634e645. [24] N. Viriya-Empikul, P. Krasae, B. Puttasawat, B. Yoosuk, N. Chollacoop,
[5] N.S. Talha, S. Sulaiman, In situ transesterification of solid coconut waste in a K. Faungnawakij, Waste shells of mollusk and egg as biodiesel production
packed bed reactor with CaO/PVA catalyst, Waste Manag. 78 (2018) 929e937. catalysts, Bioresour. Technol. 101 (10) (2010) 3765e3767.
[6] C.C. Akoh, S.W. Chang, G.C. Lee, J.F. Shaw, Enzymatic approach to biodiesel [25] V. Shankar, R. Jambulingam, Waste crab shell derived CaO impregnated Na-
production, J. Agric. Food Chem. 55 (22) (2007) 8995e9005. ZSM-5 as a solid base catalyst for the transesterification of neem oil into
[7] S.H. Teo, A. Islam, T. Yusaf, Y.H. Taufiq-Yap, Transesterification of Nanno- biodiesel, Sustain. Environ. Res. 27 (6) (2017) 273e278.
chloropsis oculata microalga's oil to biodiesel using calcium methoxide cata- [26] A. Sivasamy, K.Y. Cheah, P. Fornasiero, F. Kemausuor, S. Zinoviev, S. Miertus,
lyst, Energy 78 (2014) 63e71. Catalytic applications in the production of biodiesel from vegetable oils,
[8] A. Zheng, S.J. Huang, S.B. Liu, F. Deng, Acid properties of solid acid catalysts ChemSusChem 2 (4) (2009) 278e300.
characterized by solid-state 31P NMR of adsorbed phosphorous probe mole- [27] M.K. Lam, K.T. Lee, A.R. Mohamed, Homogeneous, heterogeneous and enzy-
cules, Phys. Chem. Chem. Phys. 13 (33) (2011) 14889e14901. matic catalysis for transesterification of high free fatty acid oil (waste cooking
[9] S.H.Y.S. Abdullah, N.H.M. Hanapi, A. Azid, R. Umar, H. Juahir, H. Khatoon, oil) to biodiesel: a review, Biotechnol. Adv. 28 (4) (2010) 500e518.
A.A. Endut, A review of biomass-derived heterogeneous catalyst for a [28] A.A. Refaat, Biodiesel production using solid metal oxide catalysts, Int. J.
M. Gohain et al. / Renewable Energy 147 (2020) 541e555 555

Environ. Sci. Technol. 8 (1) (2011) 203e221. the effect of morphology on catalytic performance of porous CeO2 nano-
[29] E. Lotero, Y. Liu, D.E. Lopez, K. Suwannakarn, D.A. Bruce, J.G. Goodwin, Syn- crystals for H2S selective oxidation, Appl. Catal. B Environ. 252 (2019)
thesis of biodiesel via acid catalysis, Ind. Eng. Chem. Res. 44 (14) (2005) 98e110.
5353e5363. [53] S.T. Gadge, A. Mishra, A.L. Gajengi, N.V. Shahi, B.M. Bhanage, Magnesium oxide
[30] A.I. Osman, J.K. Abu-Dahrieh, D.W. Rooney, J. Thompson, S.A. Halawy, as a heterogeneous and recyclable base for the N-methylation of indole and O-
M.A. Mohamed, Surface hydrophobicity and acidity effect on alumina catalyst methylation of phenol using dimethyl carbonate as a green methylating
in catalytic methanol dehydration reaction, J. Chem. Technol. Biotechnol. 92 agent, RSC Adv. 4 (91) (2014) 50271e50276.
(12) (2017) 2952e2962. [54] M. Kruk, M. Jaroniec, Gas adsorption characterization of ordered
[31] J.J. Reinosa, J.J. Romero, P. Jaquotot, M.A. Bengochea, J.F. Fern
andez, J. F, Copper organicinorganic nanocomposite materials, Chem. Mater. 13 (10) (2001)
based hydrophobic ceramic nanocoating, J. Eur. Ceram. Soc. 32 (2) (2012) 3169e3183.
277e282. [55] S. Wang, C. Zhao, R. Shan, Y. Wang, H.A. Yuan, A novel peat biochar supported
[32] N. Zhao, F. Shi, Z. Wang, X. Zhang, Combining layer-by-layer assembly with catalyst for the transesterification reaction, Energy Convers. Manag. 139
electrodeposition of silver aggregates for fabricating superhydrophobic sur- (2017) 89e96.
faces, Langmuir 21 (10) (2005) 4713e4716. [56] W. Roschat, M. Kacha, B. Yoosuk, T. Sudyoadsuk, V. Promarak, Biodiesel pro-
[33] A.I. Osman, J.K. Abu-Dahrieh, A. Abdelkader, N.M. Hassan, F. Laffir, duction based on heterogeneous process catalyzed by solid waste coral
M. McLaren, D. Rooney, Silver-modified h-Al2O3 catalyst for DME production, fragment, Fuel 98 (2012) 194e202.
J. Phys. Chem. C 121 (45) (2017) 25018e25032. [57] E. Betiku, A.M. Akintunde, T.V. Ojumu, Banana peels as a biobase catalyst for
[34] A. Dewan, M. Sarmah, A.J. Thakur, P. Bharali, U. Bora, Greener biogenic fatty acid methyl esters production using Napoleon's plume (Bauhinia
approach for the synthesis of palladium nanoparticles using papaya peel: an monandra) seed oil: a process parameters optimization study, Energy 103
eco-friendly catalyst for CeC coupling reaction, ACS Omega 3 (5) (2018) (2016) 797e806.
5327e5335. [58] N. Mansir, S.H. Teo, U. Rashid, Y.H. Taufiq-Yap, Efficient waste Gallus
[35] K.M. Parida, S. Mallick, P.C. Sahoo, S.K. Rana, A facile method for synthesis of domesticus shell derived calcium-based catalyst for biodiesel production, Fuel
amine-functionalized mesoporous zirconia and its catalytic evaluation in 211 (2018) 67e75.
Knoevenagel condensation, Appl. Catal. A. 381 (1e2) (2010) 226e232. [59] O. Axelsson, J. Paul, M.D. Weisel, F.M. Hoffmann, Reactive evaporation of
[36] D. Kühbeck, G. Saidulu, K.R. Reddy, D.D. Díaz, Critical assessment of the effi- potassium in CO2 and the formation of bulk intermediates, J. Vac. Sci. Technol.
ciency of chitosan biohydrogel beads as recyclable and heterogeneous orga- 12 (1) (1994) 158e160.
nocatalyst for CeC bond formation, Green Chem. 14 (2) (2012) 378e392. [60] P. Qiu, B. Yang, C. Yi, S. Qi, Characterization of KF/g-Al 2 O 3 catalyst for the
[37] H. Goksu, E. Gultekin, Pd nanoparticles incarcerated in aluminium oxy-hy- synthesis of diethyl carbonate by transesterification of ethylene carbonate,
droxide: an efficient and recyclable heterogeneous catalyst for selective Catal. Lett. 137 (3e4) (2010) 232e238.
Knoevenagel condensation, ChemistrySelect 2 (1) (2017) 458e463. [61] L.C. Meher, D.V. Sagar, S.N. Naik, Technical aspects of biodiesel production by
[38] T. Kokila, P.S. Ramesh, D. Geetha, Biosynthesis of AgNPs using Carica Papaya transesterification-a review, J. Renew. Sustain. Energy 10 (2006) 248e268.
peel extract and evaluation of its antioxidant and antimicrobial activities, [62] F. Ma, M.A. Hanna, Biodiesel production: a review, Bioresour. Technol. 70 (1)
Ecotoxicol. Environ. Saf. 134 (2016) 467e473. (1999) 1e15.
[39] Y.C. Cheng, S.W. Tsai, Carica papaya lipase: an effective biocatalyst for ester- [63] P. Patil, S. Deng, J.I. Rhodes, P.J. Lammers, Conversion of waste cooking oil to
ification resolution of (RS)-2-(chlorophenoxy) propionic acid, Biochem. Eng. J. biodiesel using ferric sulfate and supercritical methanol processes, Fuel 89 (2)
35 (3) (2007) 318e324. (2010) 360e364.
[40] Food and agriculture organization corporate statistical database, United na- [64] D. Kumar, B. Singh, A. Banerjee, S. Chatterjee, Cement wastes as trans-
tions, http://www.fao.org/faostat/en/#data/QC/. (Accessed 24 July 2019). esterification catalysts for the production of biodiesel from Karanja oil,
[41] Z. Han, A. Park, W.W. Su, Valorization of papaya fruit waste through low-cost J. Clean. Prod. 183 (2018) 26e34.
fractionation and microbial conversion of both juice and seed lipids, RSC Adv. [65] V. Vadery, B.N. Narayanan, R.M. Ramakrishnan, S.K. Cherikkallinmel,
8 (49) (2018) 27963e27972. S. Sugunan, D.P. Narayanan, S. Sasidharan, Room temperature production of
[42] National Horticulture Board, India http://nhb.gov.in/Horticulture%20Crops/ jatropha biodiesel over coconut husk ash, Energy 70 (2014) 588e594.
Papaya/Papaya1.htm. (Accessed 24 July 2019). [66] C. Zhao, L. Yang, S. Xing, W. Luo, Z. Wang, P. Lv, Biodiesel production by a
[43] E.G. Bligh, W.J. Dyer, A rapid method of total lipid extraction and purification, highly effective renewable catalyst from pyrolytic rice husk, J. Clean. Prod. 199
JCP Biochem. Phys. 37 (8) (1959) 911e917. (2018) 772e780.
[44] M. Gohain, A. Devi, D. Deka, Musa balbisiana Colla peel as highly effective [67] H. Li, S. Niu, C. Lu, J. Li, Calcium oxide functionalized with strontium as het-
renewable heterogeneous base catalyst for biodiesel production, Ind. Crops erogeneous transesterification catalyst for biodiesel production, Fuel 176
Prod. 109 (2017) 8e18. (2016) 63e71.
[45] A.K. Tiwari, A. Kumar, H. Raheman, Biodiesel production from jatropha oil [68] S. Sirisomboonchai, M. Abuduwayiti, G. Guan, C. Samart, S. Abliz, X. Hao,
(Jatropha curcas) with high free fatty acids: an optimized process, Biomass K. Kusakabe, A. Abudula, Biodiesel production from waste cooking oil using
Bioenergy 31 (8) (2007) 569e575. calcined scallop shell as catalyst, Energy Convers. Manag. 95 (2015) 242e247.
[46] M.A. Olutoye, S.C. Lee, B.H. Hameed, Synthesis of fatty acid methyl ester from [69] M. Ferre, R. Pleixats, M.W.C. Man, X. Cattoen, Recyclable organocatalysts based
palm oil (Elaeis guineensis) with Ky (MgCa) 2xO3 as heterogeneous catalyst, on hybrid silicas, Green Chem. 18 (4) (2016) 881e922.
Bioresour. Technol. 102 (23) (2011) 10777e10783. [70] D.M. Alonso, R. Mariscal, R. Moreno-Tost, M.Z. Poves, M.L. Granados, Potas-
[47] G. Gelbard, O. Bres, R.M. Vargas, F. Vielfaure, U.F. Schuchardt, 1H nuclear sium leaching during triglyceride transesterification using K/g-Al2O3 cata-
magnetic resonance determination of the yield of the transesterification of lysts, Catal. Commun. 8 (12) (2007) 2074e2080.
rapeseed oil with methanol, J. Am. Oil Chem. Soc. 72 (10) (1995) 1239e1241. [71] M. Kouzu, S.Y. Yamanaka, J.S. Hidaka, M. Tsunomori, Heterogeneous catalysis
[48] K. Tanabe, T. Yamaguchi, Basicity and acidity of solid surfaces, J. Res. Inst. of calcium oxide used for transesterification of soybean oil with refluxing
Catal. Hokkaido Univ. 11 (1964) 179e184. methanol, Appl. Catal., A 355 (1e2) (2009) 94e99.
[49] A.F. Lee, J.A. Bennett, J.C. Manayil K. Wilson, Heterogeneous catalysis for [72] A. Kawashima, K. Matsubara, K. Honda, Acceleration of catalytic activity of
sustainable biodiesel production via esterification and transesterification, calcium oxide for biodiesel production, Bioresour. Technol. 100 (2) (2009)
Chem. Soc. Rev. 43 (22) (2014) 7887e7916. 696e700.
[50] M. Zabeti, W.M.A.W. Daud, M.K. Aroua, Activity of solid catalysts for biodiesel [73] H.R. Radfarnia, M.C. Iliuta, Development of zirconium-stabilized calcium oxide
production: a review, Fuel Process. Technol. 90 (6) (2009) 770e777. absorbent for cyclic high-temperature CO2 capture, Indus, Eng. Chem. Res. 51
n, The basicity of mixed
[51] J.M. Fraile, N. García, J.A. Mayoral, E. Pires, L. Rolda (31) (2012) 10390e10398.
oxides and the influence of alkaline metals: the case of transesterification [74] W. Jiang, J. Yang, Y.Y. Liu, S.Y. Song, J.F. Ma, A stable porphyrin-based porous
reactions, Appl. Catal. Gen. 387 (1e2) (2010) 67e74. mog metaleorganic framework as an efficient solvent-free catalyst for CeC
[52] X. Zheng, Y. Li, L. Zhang, L. Shen, Y. Xiao, Y. Zhang, C. Au, L. Jiang, Insight into bond formation, Inorg. Chem. 56 (5) (2017) 3036e3043.

View publication stats

You might also like