You are on page 1of 8

Journal of Molecular Liquids 279 (2019) 224–231

Contents lists available at ScienceDirect

Journal of Molecular Liquids

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

Diatomite supported by CaO/MgO nanocomposite as heterogeneous


catalyst for biodiesel production from waste cooking oil
Abdelrahman M. Rabie a,⁎, Mohamed Shaban b, Mostafa R. Abukhadra b,d, Rania Hosny b,c,
Sayed A. Ahmed e, Nabel A. Negm a,⁎
a
Petrochemical Department, Egyptian Petroleum Research Institute (EPRI), Cairo, 11727, Egypt
b
Nanophotonics and Applications Lab, Physics Department, Faculty of Science, Beni-Suef University, Beni-Suef 62514, Egypt
c
Chemistry Department, Faculty of Science, Beni-Suef University, Beni-Suef, Egypt
d
Geology Department, Faculty of Science, Beni-Suef University, Beni-Suef, Egypt
e
Chemistry Department (Organic Chemistry Division), Faculty of Science, Beni-Suef University, Beni-Suef, Egypt

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

Article history: Production of renewable, biodegradable, and non-toxic biodiesel by recycling waste cooking oil offers a profitable
Received 10 November 2018 way to reduce the raw materials of waste oil. Low-cost, efficient, reusable, and ecofriendly heterogeneous catalyst
Received in revised form 17 January 2019 is therefore obligatory for the massive and industrial production of biodiesel from used cooking oil. Here, a novel
Accepted 19 January 2019
basic heterogeneous catalyst was prepared from purified domestic diatomite supported by CaO/MgO composite.
Available online 23 January 2019
Different characterization techniques confirmed the loading of diatomite as catalyst support by CaO and MgO.
Keywords:
The catalyst casting by the porous structure of diatoms skeletons causing clear enhancing of the total porosity
Diatomite and surface area of the catalyst. The efficiency of diatomite@CaO/MgO catalyst for converting the waste cooking
Magnesium oxide oil into biodiesel was studied for different times, reaction temperatures, oil/methanol ratios, catalyst dosages, and
Calcium oxide reusability for several cycles of oil conversion. The maximum obtained biodiesel yield reaches 96.47% after
Nanocomposite 120 min at 90 °C using 6 wt% catalyst and 1:15 oil: methanol ratio. The produced biodiesel matched the standard
Transesterification biodiesel according to ASTM D-6571 and EN 14214 international biodiesel standards. Diatoms skeletons/CaO/
Biodiesel MgO composite is a promising basic heterogeneous catalyst for efficient, high throughput, and low-cost biodiesel
production.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction introducing renewable and environmental fuels as an alternative source


of energy is a critical demand for energy and environmental agencies.
Energy and environment are the most significant concerns of the Biodiesel was introduced as one of the best alternative energy
contemporary world and the greatest obsession of modern civilizations. sources for the traditional diesel fuel [9]. Chemically, it is of alkyl methyl
The conventional fossil fuels and their derivatives are the main sources esters with long chains of fatty acids that can be produced through es-
of energy for most of the industrial institutes and all the transportation terification of fatty acids or transesterification of vegetable oils, animal
techniques [1–3]. The enormous consumption of such fuel resources fats and cooking oil [1,10]. The reaction occurs in the existence of alco-
may cause enormous economic and environmental problems in the hol with a short chain and suitable homogenous or heterogeneous cat-
coming decades [4]. Economically, the estimated reserves for the com- alyst [8,11]. The biodiesel is clean, eco-friendly, non-toxic and
mon fossil fuel resources will be exhausted within less than ten decades, biodegradable fuel [12]. Also, it is characterized by its high flash point,
and the gradual increase in the fossil fuels demands continuously facing low viscosity, high octane number, and high lubricant properties. More-
the depletion of the related resources [3,5]. Environmentally, combus- over, it can be used in the present engines without modifications [13]. In
tion of the traditional fuels resulted in excessive emission of injurious the later period, transesterification production of biodiesel utilizing
gases (COx, SOx, and NOx) and other air pollutants which resulted in sev- suitable solid heterogeneous catalysts attracted the attention of numer-
eral problems for the environment, ecosystem and human health in ad- ous researchers in this field [7]. This ascribed to several advantages; low
dition to their major role in greenhouse effect [6–8]. Therefore, production cost, low corrosion properties, non-toxicity, high purity of
yielded biodiesel, easy separation and recovering of it from the reaction
system, high stability of such catalysts even at high temperatures and
⁎ Corresponding authors.
their suitability to be used for several transformation cycles [14,15]. Sev-
E-mail addresses: abdo3040@yahoo.com (A.M. Rabie), nabelnegm@hotmail.com eral natural and synthetic materials have been investigated as a hetero-
(N.A. Negm). geneous catalyst for efficient conversion of fatty acid and vegetable oils

https://doi.org/10.1016/j.molliq.2019.01.096
0167-7322/© 2019 Elsevier B.V. All rights reserved.
A.M. Rabie et al. / Journal of Molecular Liquids 279 (2019) 224–231 225

into biodiesel [16]. CaO and MgO received interest in the centrifuged and washed several times using deionized water to remove
transesterification reactions the production of biodiesel [7,17]. the excess sodium hydroxide and magnesium nitrate. The obtained
Supporting of the heterogeneous catalysts on suitable materials was product was dried at 120 °C for 24 h and powdered, then calcinated
studied to minimize the disadvantages of heterogeneous catalysts for 3 h at 800 °C to complete the conversion of magnesium hydroxide
[14]. However, little work focus on introducing cheap, efficient and into magnesium oxide.
available natural porous silica as an alternative catalyst supports in bio-
diesel production. The use of available porous natural materials will give 2.3. Characterization of the catalyst
a more realistic view of the applicability of using the final product on the
industrial scale. X-ray diffraction (XRD) charts of diatomite, Ca(OH)2/diatomite and
Diatomaceous earth or diatomite is a natural fine-grained siliceous diatomite@CaO/MgO were obtained by scanning the compounds from
rock composed of micronized amorphous silica particulates with amaz- 20 to 70o using PANalytical empyrean diffractometer with Cu (Kα) (λ
ing porous structure from the skeletons of single-celled algae known as = 0.15406 nm) at scanning increment of 0.02° and accelerating voltage
diatoms [18–20]. Diatomite exhibits unique physical properties such as of 40 kV and current intensity of 35 mA. High-resolution Scanning elec-
low bulk density, high porosity, high surface area, insulating ability, and tron microscope (SEM; Gemini, Zeiss-Ultra 55) was used to investigate
relative inertness. Such properties give the diatomite deposits great eco- the morphologies of the different compounds. Fourier Transform Infra-
nomic and industrial value [21,22]. From the previous literature survey, red spectrometer (FTIR; Bruker, Vertex 70 FTIR-FT Raman) was used to
the application of diatomite as catalyst support in biodiesel production identify the different function groups of the compounds. Brunauer–
was investigated [22]. Modiba et al., used diatomite impregnated in Emmett–Teller (BET) method was used to estimate the specific surface
KOH and achieved 90% biodiesel yield from vegetable oil after reaction area of the samples. Also, Pore-size distribution was measured from the
time 4 h at reaction temperature 75 °C using 30% methanol/oil ratio adsorption branches of the isotherms using the Barrett–Joyner–Halenda
and 5% catalyst/oil ratio [22]. Shan et al., applied diatomite supported (BJH) method.
by CaO in the production of biodiesel from vegetable oil achieving
92.4% conversion percentage [23]. 2.4. Procedures of transesterification reaction of waste cooking oil
Here, a hybrid system composed of purified diatomite supported by
CaO/MgO composite was synthesized as a basic heterogeneous catalyst. The transesterification reaction of the waste cooking oil using diato-
The synthetic catalyst was applied for the first time in the mite@CaO/MgO as an alkaline heterogeneous catalyst was performed
transesterification of waste cooking oil into biodiesel as a function of re- using three necked flask (250 mL) connected with a magnetic stirrer,
action time, catalyst dose, reaction temperature, catalyst/methanol digital heater, and a water-cooling condenser. Waste cooking oil
ratio, and reusability of the used catalyst for several conversion cycles. (50 g) was stirred with different amounts of the prepared catalyst
(1–8 wt%) for different reaction time intervals (0.5–5 h) at different
2. Experimental work temperatures (40 °C to 120 °C) and using different oil/methanol ratios
of 1:3, 1:6, 1:9, 1:12, 1:15 and 1:18. Each reaction condition, including
2.1. Materials reactants amounts, reaction time and temperature and time were re-
peated three times. After each experiment, the catalyst was removed
Waste oil samples from the surrounding environment (restaurants, from the reaction medium by centrifugation and the residual liquid
houses, farms) and methanol (99.8%, Sigma Aldrich) were used in the was retained in separating funnel overnight. Excess methanol was
transesterification reaction. Purified diatomite represented to El- stripped out under reduced pressure at 40 °C to obtain the
Fayoum area, Egypt was delivered from Central Metallurgical and De- transesterification product of biodiesel. The efficiency of conversion re-
velopment, Egypt as a catalyst support. The chemical composition of action (biodiesel yield %) was estimated from the amount of the re-
the purified diatomite sample is 97.87% SiO2, 1.79% Al2O3 and 0.34% sulted fatty acids methyl esters (FAMEs) which obtained from the Gas
L.O.I. Sodium hydroxide pellets (N98%, Sigma Aldrich), magnesium ni- chromatography (Agilent 7890A) analysis of the products [12,25]
trate hexahydrate (N98%, Sigma Aldrich) and calcium nitrate using Eq. (1):
tetrahydrate (N98%, Sigma Aldrich) were used in the synthesis of CaO
and MgO catalyst. All these reagents are of AR grade and used without ðWeight of biodiesel x%FAMEÞ
Biodiesel yield ð%Þ ¼  100: ð1Þ
further purification. Weight of oil

2.2. Preparation of the catalyst


3. Results and discussion
2.2.1. Synthesis of diatomite supported by Ca(OH)2 (CaOH/diatomite)
Synthesis of Ca(OH)2/diatomite was performed through a conven- 3.1. Catalyst characterization
tional impregnation method. In a typical procedure, 5 g of pure diato-
mite was dispersed in calcium nitrate solution (100 mL, 30%) and 3.1.1. Samples structures
stirred for 10 h at 50 °C. Then the solid was separated, washed by dis- The structural features of used diatomite support and the synthetic
tilled water and dried at 120 °C for 12 h. CaO and MgO catalysts were investigated based on XRD analysis
(Fig. 1). The XRD pattern of the purified diatomite sample exhibits a
2.2.2. Synthesis of diatomite supported by CaO/MgO composite (DCM) broad peak around 2θ of 22° which assigned to amorphous silica
Loading of MgO trough out the porous structure of pre-synthetic Ca (Fig. 1A). The XRD pattern of diatomite impregnated in calcium solution
(OH)2/diatomite was accomplished through chemical precipitation shows the presences of calcium carbonate (JCPDS card: No. 86-0174)
method [24]. Typically, 5 g of the synthesized Ca(OH)2/diatomite were and calcium hydroxide (JCPDS card No. 01-073-5492) as the main
dispersed in magnesium nitrate hexahydrate solution (100 mL, 0.5 M) formed crystalline calcium based phases in addition to the reduced
and stirred for 30 min at 800 rpm until homogeneous mixture of broad peak of amorphous silica related to the original diatomite support
suspended Ca(OH)2/diatomite particles was obtained. Then, sodium hy- (Fig. 1B). The presences of calcium carbonate assigned to the rapid at-
droxide solution (50 mL, 1 M) was added suddenly to the reaction me- mospheric carbonation of Ca(OH)2 by CO2 during washing and drying
dium under ultrasonic stirring 85 °C till a thick white gel is formed. The [25].
formed gel then transferred into a Teflon lined stainless steel autoclave The XRD pattern of the final prepared catalyst confirmed loading of
and heated at 150 °C for 12 h. Then, the obtained solid fractions were diatomite by of CaO and MgO as composite material; and the formation
226 A.M. Rabie et al. / Journal of Molecular Liquids 279 (2019) 224–231

the following absorption bands at: 3437, 1638, 1092, 799 and
465 cm−1 (Fig. 2A). The two broad bands at 3437 cm−1 and
1638 cm−1 were assigned to the characteristic surficial silanol group
(SiO\\H) of diatomite silica skeletons and the H\\O\\H vibration of
the adsorbed water molecules, respectively [27,28]. The presence of
asymmetric stretching of Si\\O\\Si siloxane group was confirmed by
the absorption bands at 1092 cm−1 and 465 cm−1, while the symmetric
stretching of Si\\O\\Si was confirmed by the absorption band at
799 cm−1 [29].
Loaded Ca(OH)2/diatomite and the associated carbonized product
(Calcite) was pointed in the FTIR spectrum (Fig. 2B). Loading of diato-
mite by Ca(OH)2 caused apparent reduction in the intensity of the char-
acteristic absorption bands of raw diatomite, e.g., absorption bands of
asymmetric vibration Si\\O\\Si group and SiO\\H group. Also, this re-
sulted in a disappearance of the absorption bands of symmetrical
Si\\O\\Si and H\\O\\H groups. The formation of the loaded phases
were confirmed by the appearance of new absorption band at
1465 cm−1 assigned for CO32– groups which produced from the carbon-
ation of calcium hydroxide to calcium carbonate [23].
Fig. 1. XRD patterns of purified diatomite (A), diatomite impregnated with calcium
solution (B) and the prepared diatomite supported by CaO/MgO (C). The prepared diatomite@CaO/MgO catalyst showed new absorption
bands related to the loaded oxides (Fig. 2C). Shifting the absorption
band located at 465 cm−1 to 443 cm−1 indicated the formation Ca\\O
of cristobalite (JCPDS card No. 89-3606 and 89-3434) as high- bond [3]. The observed spectrum band around 867 cm−1 was assigned
temperature polymorph of silica after calcination diatomite at 800 °C to the Mg\\O\\Mg interaction or C\\O bond [30]. The absorption
[26] (Fig. 1C). Additionally, small peaks were detected for the formation band of CO32– showed significant increases in the intensity and the
of calcium silicate phase represented by Ca2SiO4 at 2θ of 34.3° (JCPDS width of it which attributed to surface adsorption of CO2 from the atmo-
card No. 31-0302). MgO diffraction peaks were detected at 2θ of 43.2° sphere by the synthetic CaO or formation of magnesium carbonate [11].
and 74.7° which are corresponding to crystallographic planes (200)
and (311), respectively (JCPDS; Cubic No. 01-077-2364). This is related 3.1.3. Morphological properties
to cubic MgO with crystal lattice parameters of ?? = ?? = ?? = 4.213 Å The changes in the morphological features of diatomite with the
and space group Fm − 3 m (225). The crystallite size of the synthetic loading of it by CaO and MgO catalyst were studied through SEM images
MgO was determined based on the Scherrer equation and the obtained in Fig. 3. Raw diatomite sample appears with its characteristic shapes
value is about 81.6 nm. Several diffraction peaks were detected for CaO and its amazing nanoporous structure (Fig. 3A). Penates type diatom's
at 2θ of 32.23°, 37.6°, 55.4° and 67.14°, which were assigned to the crys- skeletons are the dominant type in the purified sample as shown in
tallographic planes of (111), (200), (202) and (222) of cubic CaO crystal the magnified SEM images. The average pores width is ~224 nm, and
with lattice parameters of ?? = ?? =?? = 4.8105 Å (JCPDS card No. 00- the average cell wall thickness is ~604 nm. After loading by calcium hy-
004-0777). The average crystallite size of CaO in CaO/MgO composite droxide and carbonization to calcium carbonate (calcite), diatom skele-
was found to be 63.3 nm. tons appear to be thoroughly coated with calcium hydroxide and
calcium carbonate particles (Fig. 3B). The porous morphology of diatom
3.1.2. Chemical functional groups skeletons was not destructed after the modification, but it was partially
The change in the chemical functional groups with loading of diato- closed by the deposited calcium-based materials. The coated particles
mite by CaO and MgO was followed using FTIR spectroscopy of raw di- present in two main features represented by nanofibrous grains distrib-
atomite, Ca(OH)2/diatomite, and diatomite@CaO/MgO nanocomposite uted throughout the porous structure of diatoms (Fig. 3C) and
(DCM) (Fig. 2). FTIR spectra of raw diatomite show the existence of nanoflakes or plated partially covered some parts of the diatomite
grains surfaces (Fig. 3D). The pores diameters were reduced to
b25 nm as shown in the insets magnified images of Fig. 3C. After MgO
coating, the diatomite surface was entirely covered by the studied
oxide (Fig. 3E). The diameter of the pores was reduced to b10 nm as
shown in the inset of Fig. 3E. The precipitated MgO has cast the porous
structure of diatomite, (inset of Fig. 3E), which enhanced the total sur-
face area and the overall porosity of the prepared diatomite@CaO/
MgO catalyst.

3.1.4. Textural properties


The textural properties of raw diatomite and the modified products
are described numerically in Table 1. The specific surface area of the raw
diatomite was increased significantly from 117 m2/g to 167 m2/g after
impregnation with calcium hydroxide to form Ca(OH)2/diatomite. This
can be attributed to the fibrous morphology of the loaded calcium hy-
droxide and its carbonated products. However, the loading of diatomite
by Ca(OH)2 cause blocking of some of the raw diatomite pores which
was associated with an increase in the pore volume from 0.09 cc/g to
0.168 cc/g. This was assigned to the formation of secondary pores due
to the interlocking between the fibrous particles of calcium hydroxide
Fig. 2. FTIR spectra of raw diatomite, Ca(OH)2/diatomite, and diatomite@CaO/MgO and calcium carbonate. In addition, the average pore diameter was de-
composite. creased from 16.57 nm of raw diatomite to 15.84 nm after loading of
A.M. Rabie et al. / Journal of Molecular Liquids 279 (2019) 224–231 227

Fig. 3. SEM images of: (A) raw diatomite, (B) diatomite loaded by Ca(OH)2 and CaCO3 particles (C) nanofibrous grains structures of diatomite loaded by CaCO3, (D) nanoflakes or plated
structures of diatomite loaded by CaCO3, (E) diatomite loaded by CaO/MgO.

it by calcium hydroxide. i.e., both samples can be defined as mesoporous 3.2.1. Effect of methanol to oil ratio
materials. Reducing the average pore diameter can be related to the par- The studying of transesterification reactions of waste cooking oil was
tial filling of raw diatomite pores by calcium hydroxide and calcium performed using different methanol to oil ratios (3:1, 6:1, 9:1, 12:1,
carbonate. 15:1, and 18:1) in the presence of 5% by weight of the synthesized diat-
The synthetic diatomite@CaO/MgO catalyst exhibits higher surface omite@CaO/MgO catalyst for 180 min at 60 °C. The yield % of the ob-
area of 134.85 m2/g than raw diatomite but lower than loaded Ca(OH) tained biodiesel was determined at each ratio, Fig. 4. It is clear from
2/diatomite. That can be commented to the blocking of raw diatomite data in Fig. 4 that a gradual increase in the biodiesel yield percentage
pores after modification by MgO. This also was resulted in an apparent
reduction in the average pore diameter and pore volume to be
0.137 cc/g and 11.57 m2/g, respectively.

3.2. Transesterification reaction

The transesterification reaction of waste cooking oil with methanol


to obtain the corresponding biodiesel was studied using the prepared
diatomite@CaO/MgO catalyst by studying the reaction parameters.
The studied parameters were: methanol to oil ratio, temperature, cata-
lyst amount, and the effect of time on the reaction.

Table 1
Textural properties of raw purified diatomite and its modified products.

Material BET surface area Average pore diameter Pore volume


(m2/g) (nm) (cc/g)

Raw diatomite 117.7 16.57 0.09


Ca(OH)2/diatomite 167.99 15.84 0.168
Fig. 4. Influence of oil to methanol ratio on the conversion percent of waste cooking oil into
Diatomite@CaO/MgO 134.85 11.57 0.137
biodiesel.
228 A.M. Rabie et al. / Journal of Molecular Liquids 279 (2019) 224–231

Fig. 5. Influence of reaction time on the conversion of waste cooking oil into biodiesel. Fig. 7. Influence of reaction time on the conversion of cooking oil into biodiesel.

was observed by increasing the methanol to oil ratio from 3:1 to 15:1. temperature range from 40 to 120 °C using 5% by weight of the synthe-
The yield percentage increased by 80.5%, 88.5%%, 90.9%, 92.3%, 92.7% sized catalyst for 180 min at 1:15 oil to methanol ratio (Fig. 5). This fig-
with increasing the ratio by 3, 6, 9, 12 and 15, respectively. Increasing ure showed a clear enhancement in the conversion reaction efficiency
the methanol to oil ratio has no positive effect on expanding the interfa- by increasing the reaction temperature from 40 to 90 °C. Then, the con-
cial area between the present immiscible liquid which might be resulted version efficiency decreased after increasing the temperature above 90
in higher frequency passage for the used heterogeneous catalyst °C. The biodiesel yield percentage through the transesterification pro-
through it. This phenomenon resulted in increasing the exposed surface cess using diatomite@Cao/MgO as a catalyst was achieved at the maxi-
area of the used catalyst and in turn a noticeable enhancing in the con- mum of 96.2% at 90 °C. This result was reported [35–37] and was
version rate of waste cooking oil into biodiesel [31]. related to the gradual increase in the activation energy of the reaction
The biodiesel yield percentage was started to decrease with increas- and the reaction rate by increasing the temperature of the
ing the ratio at 18:1. Increasing the methanol to oil ratio beyond 15 to transesterification reaction of vegetable oils. The increase of the reaction
18, the biodiesel yield percentage decreased to 91.7%. This behavior rate is resulted in more collisions between the used methanol and the
can be related to the deactivating role of excess methanol in dilution present triglyceride molecules which in turn promote the conversion
of the components in the transesterification system, which reduces reaction and become faster at high temperatures [36].
the conversion rate [32]. Additionally, this is commonly associated Increasing the reaction temperature to 100–120 °C resulted in de-
with difficulty in separation of the obtained biodiesel from the excess creasing the conversion efficiency by 96.1%, 95.88%, and 95.67%, respec-
methanol and increase in the solubility of glycerol by-product in the tively. This can be attributed to the evaporation of the methanol by
used system which shifts the conversion reactions backsides [33]. increasing the temperature which can reduce the methanol: oil molar
ratio during the conversion reaction and also due to the reverse behav-
3.2.2. Effect of reaction temperature ior of the transesterification reaction [34].
It was reported that increasing the transesterification reaction tem-
perature has a direct reflection on the transesterification reaction subse- 3.2.3. Effect of the catalyst amount
quently the conversion of the oil into biodiesel [34]. Consequently, the The effect of diatomite@CaO/MgO composite amount used on the
experiments using diatomite@CaO/MgO catalyst were carried out at a transesterification reaction of waste cooking oil was investigated using
catalyst amount of 1–6% by weight for 180 min at 90 °C and 1:15 oil
to methanol ratio (Fig. 6). The biodiesel yield percentage was increased
with increasing the catalyst weight from 1% to 6%. However, raising the

Fig. 6. Influence of amount of catalyst (%) by weight on the transesterification reaction of Fig. 8. Stability of diatomite@CaO/MgO for seven runs of transesterification of waste
waste cooking oil into biodiesel. cooking oil into biodiesel.
A.M. Rabie et al. / Journal of Molecular Liquids 279 (2019) 224–231 229

Table 2 biodiesel yield showed a gradual increase with increasing the


Physicochemical properties of the obtained biodiesel in comparison with international transesterification time from 30 to 210 min. Then, the conversion effi-
biodiesel standards.
ciency was slightly decreased with increasing the reaction time from
Contents Unit ASTM EN Prepared Literature 210 to 280 min. The biodiesel yield was increased by 79.7%, 83.3%,
D-6751 14214 biodiesel values 86.6%, 91.3%, 94.2%, 96.5% and 96.4% with increasing the
Viscosity mm2/s 1.9–6 3.5–5 3.12 3.20–3.22 transesterification time by 30, 60, 90, 120, 150, 180, and 210 min,
[40] respectively.
Moisture wt.(%) b0.05 b0.05 0.022 0.015 [40]
Enhancing the conversion efficiency of waste cooking oil with in-
content
Flash point °C N120 N120 133 129–131 [41] creasing the time can be assigned to the immiscibility of the oil and
Cloud point °C −3 to 15 – 3 3–4 [41] methanol at the initial stages of the reaction [38]. While reversing of
Pour point °C −5 to 10 – 3 0–2 [43] the behavior after a specific reaction time can be attributed to the re-
Cetane number – 47–65 51 to 120 51.4 46–49 [34] versible nature of the conversion reaction of oils into biodiesel [32].
Density g/cm3 0.82–0.9 0.86–0.9 0.876 0.869 [34]
Calorific value MJ/kg – 32.9 38.431 –
Thus, the reaction time of 210 min is the optimum time for the best con-
version of waste cooking oil into biodiesel using diatomite@CaO/MgO
composite catalyst.
catalyst amount N6% resulted in a gradual decrease in the obtained bio-
diesel yield %. The increment of the catalyst amount by 1–6% was re-
sulted in biodiesel yield of 88.7%, 93.1%, 95.3%, 96.2% and 96.47%, 3.2.5. Catalyst stability
respectively. This can be assigned to the increase in the present active The stability of the prepared catalyst during the transesterification
catalytic sites of the catalyst with increasing the amount [36]. reaction of waste cooking oil was determined at the optimum reaction
Further increase in the catalyst amount to 7–10% decreased the con- conditions for several runs of the reaction using the same catalyst. The
version efficiency to 96.3%, 95.8%, and 95.2%, respectively. This can be used catalyst after each run washed several times with a mixture of
referred to the increase in the viscosity of the reaction medium under methyl alcohol and n-hexane (1:1 V:V), then re-activated at 800 °C for
the formation of dense slurry from the catalyst powder and the used 3 h. The suitability of diatomite@CaO/MgO composite for several runs
oil, which resulted in a difficulty in the homogenous mixing of the reac- of transesterification of waste cooking oil into biodiesel was represented
tants and consequently decreases in the interaction between the cata- in Fig. 8, for seven runs at the pre-estimated optimum reaction condi-
lyst and reaction component [37]. tions (210 min, 1:15 oil to methanol ratio, 6% catalyst amount, 90 °C).
The catalytic activity of the synthetic catalyst decreased with increasing
the conversion runs. This is closely related to the decreasing of the active
3.2.4. Effect of conversion time catalytic sites under the sequential use of the catalyst due to the precip-
Fig. 7 showed the effect of the reaction time on the conversion of itation of the insoluble by-products on its surface, e.g., glycerol [39]. The
waste cooking oil into biodiesel. At fixed conditions of 1:15 oil to meth- biodiesel yield was decreased by 96.4%, 94.3%, 90.2%, 87.45%, 78.23%,
anol ratio, 90 °C reaction temperature, and 6% catalyst amount, the 65.45% and 50.4% with reusing the catalyst from 7 runs.

Fig. 9. Mechanism of alkaline transesterification reaction of waste cooking oil and methanol in the presence of the prepared catalyst.
230 A.M. Rabie et al. / Journal of Molecular Liquids 279 (2019) 224–231

3.3. Biodiesel properties Acknowledgement

The properties of the obtained biodiesel from the transesterification The Egyptian Academy for Scientific Research & Technology partially
reaction of waste cooking oil at the determined optimum reaction con- supports this work (ASRT/SNG/E/2014-5).
ditions was characterized based on the physicochemical specifications
of the international standards of EN-14214 and ASTM D-6751 References
(Table 2). It is clear from the data in Table 2 that the specification re-
[1] D.Y. Leung, X. Wu, M. Leung, A review on biodiesel production using catalyzed
quirements of the obtained biodiesel are in accordance with the ASTM transesterification, Appl. Energy 87 (2010) 1083–1095.
D-6571 and EN 14214 limits. The obtained values of the produced bio- [2] A.M. Rabie, M.A. Betiha, S.E. Park, Direct synthesis of acetic acid by simultaneous co-
diesel including viscosity, moisture content, flash point, cloud point, activation of methane and CO2 over Cu-exchanged ZSM-5 catalysts, Appl. Catal. B
Environ. 215 (2017) 50–59.
pour point, cetane number and density were compared by the recently [3] N.A. Negm, M.A. Shaalan, G.S. El Barouty, M.Y. Mohamed, Preparation and evalua-
published data [34,40,41,43]. Table 2 showed that the comparison com- tion of biodiesel from Egyptian castor oil from semi-treated industrial wastewater,
prised the comparable properties between the obtained biodiesel and J. Taiwan Inst. Chem. Eng. 63 (2016) 151–156.
[4] A. Elfadly, I. Zeid, F. Yehia, A. Rabie, S.E. Park, Highly selective BTX from catalytic fast
its validity for the use as alternative, environmentally and cheap fuel. pyrolysis of lignin over supported mesoporous silica, Int. J. Biol. Macromol. 91
(2016) 278–293.
[5] S. Lee, D. Posarac, N. Ellis, Process simulation and economic analysis of biodiesel pro-
duction processes using fresh and waste vegetable oil and supercritical methanol,
3.4. Mechanism of diatomite@CaO/MgO nanocomposite in
Chem. Eng. Res. Des. 89 (2011) 2626–2642.
transesterification reaction [6] A. Silitonga, H. Masjuki, H.C. Ong, T. Mahlia, F. Kusumo, Optimization of extraction of
lipid from Isochrysis galbana microalgae species for biodiesel synthesis, Energy
The transesterification reaction between vegetable oils and metha- Sources, Part A 39 (2017) 1167–1175.
[7] A. Silitonga, T. Mahlia, H.C. Ong, T. Riayatsyah, F. Kusumo, H. Ibrahim, S. Dharma, D.
nol are complicated reaction involves two main steps. The first is the Gumilang, A comparative study of biodiesel production methods for Reutealis
cracking of ester bonds in the triglyceride molecules and the second is trisperma biodiesel, Energy Sources, Part A 39 (2017) 2006–2014.
the substitution of the short chain alcohols in the molecules to produce [8] F. Kusumo, A. Silitonga, H.C. Ong, H. Masjuki, T. Mahlia, A comparative study of ultra-
sound and infrared transesterification of Sterculia foetida oil for biodiesel production,
the desired biodiesel [41]. The prepared catalyst is characterized by the Energy Sources, Part A 39 (2017) 1339–1346.
presence of highly alkaline catalytic sites which play an important role [9] T.H. Dang, B.H. Chen, D.J. Lee, Optimization of biodiesel production from
in the cleavage steps of the triglycerides and the alcohol molecules. transesterification of triolein using zeolite LTA catalysts synthesized from kaolin
clay, J. Taiwan Inst. Chem. Eng. 79 (2017) 14–22.
The alkaline catalytic sites in the catalyst are the metal oxide cores of [10] A. Endut, S.H.Y.S. Abdullah, N.H.M. Hanapi, S.H.A. Hamid, F. Lananan, M.K.A.
MgO and CaO. The presence of the Ca and Mg oxides in the chemical Kamarudin, R. Umar, H. Juahir, H. Khatoon, Optimization of biodiesel production
structure of the catalyst has a high influence on its alkalinity [42,43]. by solid acid catalyst derived from coconut shell via response surface methodology,
Int. Biodeterior. Biodegrad. 124 (2017) 250–257.
That is due to the presence of Ca and Mg metal ions as the first two el- [11] N.A. Negm, A.M. Rabie, E.A. Mohammed, Molecular interaction of heterogeneous
ements in the alkaline earth metal series increases the negative charges catalyst in catalytic cracking process of vegetable oils: chromatographic and biofuel
on the oxygen atoms. That increases their catalytic action in the alkaline performance investigation, Appl. Catal. B Environ. 239 (2018) 36–45.
[12] G. Chen, R. Shan, J. Shi, C. Liu, B. Yan, Biodiesel production from palm oil using active
hydrolytic mechanisms [44]. Also increases the stability of the polarized
and stable K doped hydroxyapatite catalysts, Energy Convers. Manag. 98 (2015)
molecules in the form of calcium or magnesium methoxide and fatty 463–469.
acid salts. The first catalytic step of the reaction involves polarization [13] M. Zabeti, W.M.A.W. Daud, M.K. Aroua, Activity of solid catalysts for biodiesel pro-
of the ester group of the triglycerides in the C\\O bond between the duction: a review, Fuel Process. Technol. 90 (2009) 770–777.
[14] Y. Ma, Q. Wang, X. Sun, C. Wu, Z. Gao, Kinetics studies of biodiesel production from
fatty acid moieties and the glycerol molecule [45]. Furthermore, a polar- waste cooking oil using FeCl3-modified resin as heterogeneous catalyst, Renew. En-
ization occurred for the methanol molecules due to the interaction by ergy 107 (2017) 522–530.
the different alkaline sites in the catalyst, Fig. 9. The second step of the [15] A. Silitonga, T. Mahlia, F. Kusumo, S. Dharma, A. Sebayang, R. Sembiring, A.
Shamsuddin, Intensification of reutealis trisperma biodiesel production using infra-
reaction is the cracking of the triglyceride and methanol molecules red radiation: simulation, optimization and validation, Renew. Energy 133 (2019)
which accompanied by combination of the formed methoxy group 520–527.
and the alkyl chains of the different fatty acids [46]. The interaction be- [16] J. Milano, H.C. Ong, H. Masjuki, A. Silitonga, W.H. Chen, F. Kusumo, S. Dharma, A.
Sebayang, Optimization of biodiesel production by microwave irradiation-assisted
tween the formed intermediates during the transeterification reaction transesterification for waste cooking oil-Calophyllum inophyllum oil via response
produces methyl esters of the different fatty acid moieties in the used surface methodology, Energy Convers. Manag. 158 (2018) 400–415.
oil (biodiesel) and the glycerol as a bi-product of the process. [17] J. Milano, H.C. Ong, H. Masjuki, A. Silitonga, F. Kusumo, S. Dharma, A. Sebayang, M.Y.
Cheah, C.T. Wang, Physicochemical property enhancement of biodiesel synthesis
from hybrid feedstocks of waste cooking vegetable oil and beauty leaf oil through
optimized alkaline-catalysed transesterification, Waste Manag. 80 (2018) 435–449.
4. Conclusion [18] J.A. Melero, J. Iglesias, G. Morales, Heterogeneous acid catalysts for biodiesel produc-
tion: current status and future challenges, Green Chem. 11 (2009) 1285–1308.
[19] S.S. Ibrahim, A.Q. Selim, Evaluation of Egyptian diatomite for filter aid applications,
Diatomite supported by CaO/MgO composite was synthesized and Physicochemical Problems of Mineral Processing, 47, 2011, pp. 113–122.
characterized as a novel basic heterogeneous catalyst for [20] M. Hassan, I. Ibrahim, I. Ismael, Diatomaceous deposits of Fayium, Egypt; character-
transesterification of waste cooking oil into biodiesel. CaO/MgO com- ization and evaluation for industrial application, Chin. J. Geochem. 18 (1999)
233–241.
posite was successfully loaded into the porous structure of diatoms skel- [21] H.A. Alyosef, S. Ibrahim, J. Welscher, A. Inayat, A. Eilert, R. Denecke, W. Schwieger, T.
eton. The suitability of diatomite@CaO/MgO as a heterogeneous catalyst Münster, G. Kloess, W.D. Einicke, Effect of acid treatment on the chemical composition
in the transesterification reactions was addressed according to the com- and the structure of Egyptian diatomite, Int. J. Miner. Process. 132 (2014) 17–25.
[22] N. Damanik, H.C. Ong, W. Chong, A. Silitonga, Biodiesel production from
monly used conversion parameters. The catalyst achieves high catalytic Calophyllum inophyllum−palm mixed oil, Energy Sources, Part A 39 (2017)
activity represented by maximum biodiesel yield percentage about 1283–1289.
96.5%. This result was obtained after conversion time 120 min using [[23] E. Modiba, C. Enweremadu, H. Rutto, Production of biodiesel from waste vegetable
oil using impregnated diatomite as heterogeneous catalyst, Chin. J. Chem. Eng. 23
6 wt% catalyst at 1:15 oil/methanol ratio and a reaction temperature (2015) 281–289.
of 90 °C. The qualifications of the final biodiesel product at the optimum [24] W. Jindapon, S. Jaiyen, C. Ngamcharussrivichai, Al2O3-supported mixed Ca and Zn
conditions match the specifications of the international standards bio- compounds prepared from waste seashells for synthesis of palm fatty acid methyl
esters, Chem. Eng. Commun. 202 (2015) 1591–1599.
diesel (ASTM D-6571 and most of EN 14214). Therefore, supporting of
[25] A.R. Ibrahim, J.B. Vuningoma, Y. Huang, H. Wang, J. Li, Rapid carbonation for calcite
diatoms skeletons by CaO/MgO composite is promising basic heteroge- from a solid-liquid-gas system with an imidazolium-based ionic liquid, Int. J. Mol.
neous catalyst in the transesterification reactions and can be used as an Sci. 15 (2014) 11350–11363.
[26] S. Zheng, C. Bai, R. Gao, Preparation and photocatalytic property of/diatomite-based
efficient, cheap and available catalyst support rather than the com-
porous ceramics composite materials, Int. J. Photoenergy 2012 (2012).
monly study synthetic silica porous materials.
A.M. Rabie et al. / Journal of Molecular Liquids 279 (2019) 224–231 231

[27] J.F. Mukerabigwi, S. Lei, L. Fan, H. Wang, S. Luo, X. Ma, J. Qin, X. Huang, Y. Cao, Eco- using novel sulphonated phenyl silane montmorillonite catalyst, J. Mol. Liq. 234
friendly nano-hybrid superabsorbent composite from hydroxyethyl cellulose and (2017) 157–163.
diatomite, RSC Adv. 6 (2016) 31607–31618. [37] N.A. Negm, G.H. Sayed, O.I. Habib, F.Z. Yehia, E.A. Mohamed, Heterogeneous catalytic
[28] J.F. Mukerabigwi, S. Lei, H. Wang, S. Luo, X. Ma, J. Qin, X. Huang, Y. Cao, Synthesis and transformation of vegetable oils into biodiesel in one-step reaction using super
properties of a novel ecofriendly superabsorbent hydrogel nanocomposite based on acidic sulfonated modified mica catalyst, J. Mol. Liq. 237 (2017) 38–45.
xyloglucan-graft-poly (acrylic acid)/diatomite, RSC Adv. 5 (2015) 83732–83742. [38] M. Feyzi, N. Hosseini, N. Yaghobi, R. Ezzati, Preparation, characterization, kinetic and
[29] S. Zhang, A new nano-sized calcium hydroxide photocatalytic material for the thermodynamic studies of MgO-La2O3 nanocatalysts for biodiesel production from
photodegradation of organic dyes, RSC Adv. 4 (2014) 15835–15840. sunflower oil, Chem. Phys. Lett. 677 (2017) 19–29.
[30] P. Devaraja, D. Avadhani, S. Prashantha, H. Nagabhushana, S. Sharma, B. [39] X. Niu, C. Xing, W. Jiang, Y. Dong, F. Yuan, Y. Zhu, Activity and stability of solid base
Nagabhushana, H. Nagaswarupa, Synthesis, structural and luminescence studies of KF/La2O3 catalysts for transesterification of tributyrin with methanol, React. Kinet.
magnesium oxide nanopowder, Spectrochim. Acta A Mol. Biomol. Spectrosc. 118 Mech. Catal. 109 (2013) 167–179.
(2014) 847–851. [40] A.M. Rabie, E.A. Mohammed, N.A. Negm, Feasibility of modified bentonite as acidic
[31] D.M. Marinković, J.M. Avramović, M.V. Stanković, O.S. Stamenković, D.M. Jovanović, heterogeneous catalyst in low temperature catalytic cracking process of biofuel pro-
V.B. Veljković, Synthesis and characterization of spherically-shaped CaO/γ-Al2O3 duction from nonedible vegetable oils, J. Mol. Liq. 254 (2018) 260–266.
catalyst and its application in biodiesel production, Energy Convers. Manag. 144 [41] A.W. Schwab, M.O. Baghy, B. Freedman, Preparation and properties of diesel fuels
(2017) 399–413. from vegetable oils, Fuel 66 (1987) 1372–1378.
[32] M.R. Abukhadra, F.M. Dardir, M. Shaban, E.A. Ahmed, M.F. Soliman, Spongy Ni/Fe [42] D.M. Marinkovic, J.M. Avramovic, M.V. Stankovic, O.S. Stamenkovic, D.M. Jovanovic,
carbonate-fluorapatite catalyst for efficient conversion of cooking oil waste into bio- V.B. Veljkovic, Synthesis and characterization of spherically-shaped CaO/γ-Al2O3
diesel, Environ. Chem. Lett. 16 (2018) 665–670. catalyst and its application in biodiesel production, Energy Convers. Manag. 144
[33] D. Zeng, L. Yang, T. Fang, Process optimization, kinetic and thermodynamic studies (2017) 399–413.
on biodiesel production by supercritical methanol transesterification with [43] J.P. Gomes, J.B. Puna, J.M. Bordado, M.N. Correia, A.S. Dias, Status of biodiesel produc-
CH3ONa catalyst, Fuel 203 (2017) 739–748. tion using heterogeneous alkaline catalysts, Int. J. Environ. Stud. 69 (2012) 635–653.
[34] N.A. Negm, M.K. Zahran, M.R. Abd Elshafy, I.A. Aiad, Transformation of Jatropha oil to [44] I.S. Resck, Polyaminophosphazenes as super bases nitrogenadas, Quim. Nova 17
biofuel using transition metal salts as heterogeneous catalysts, J. Mol. Liq. 256 (1994) 317–321.
(2018) 16–21. [45] P. Chand, C.V. Reddy, J.G. Verkade, T. Wang, D. Grewell, Thermogravimetric quanti-
[35] M. Feyzi, Z. Shahbazi, Preparation, kinetic and thermodynamic studies of Al–Sr fication of biodiesel produced via alkali catalyzed transesterification of soybean oil,
nanocatalysts for biodiesel production, J. Taiwan Inst. Chem. Eng. 71 (2017) Energy Fuel 23 (2009) 989–992.
145–155. [46] I.A.W. Mohamad, O.A. Ali, Evaluation of the transesterification of waste palm oil into
[36] N.A. Negm, G.H. Sayed, F.Z. Yehia, O.I. Habib, E.A. Mohamed, Biodiesel production biodiesel, Bioresour. Technol. 85 (2002) 251–256.
from one-step heterogeneous catalyzed process of Castor oil and Jatropha oil

You might also like