You are on page 1of 16

Journal Pre-proofs

Pineapple (Ananás comosus) leaves ash as a solid base catalyst for biodiesel
synthesis

Silma de S. Barros, Wanison A.G. Pessoa Junior, Ingrity S.C. Sá, Mitsuo L.
Takeno, Francisco X. Nobre, William Pinheiro, Lizandro Manzato, Stefan
Iglauer, Flávio A. de Freitas

PII: S0960-8524(20)30841-5
DOI: https://doi.org/10.1016/j.biortech.2020.123569
Reference: BITE 123569

To appear in: Bioresource Technology

Received Date: 14 April 2020


Revised Date: 19 May 2020
Accepted Date: 20 May 2020

Please cite this article as: de S. Barros, S., Pessoa Junior, W.A.G., Sá, I.S.C., Takeno, M.L., Nobre, F.X.,
Pinheiro, W., Manzato, L., Iglauer, S., de Freitas, F.A., Pineapple (Ananás comosus) leaves ash as a solid
base
catalyst for biodiesel synthesis, Bioresource Technology (2020), doi: https://doi.org/10.1016/j.biortech.
2020.123569

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a
cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This
version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process,
errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.

Pineapple (Ananás comosus) leaves ash as a solid base catalyst for


biodiesel synthesis
Silma de S. Barros1, Wanison A. G. Pessoa Junior2, Ingrity S. C. Sá1, Mitsuo L.
Takeno2, Francisco X. Nobre3, William Pinheiro4, Lizandro Manzato2, Stefan Iglauer5,
Flávio A. de Freitas6,*

1Deparatamento de Engenharia de Materiais – PPGCEM/UFAM. Av. Octávio Hamilton


Botelho Mourão - Coroado, 69067-005, Manaus, Amazonas - Brazil.
2Instituto Federal de Educação, Ciência e Tecnologia do Amazonas – IFAM/CMDI. Av.
Gov. Danilo de Matos Areosa, 1672 - Distrito Industrial, 69075-351, Manaus,
Amazonas, Brazil.
3Instituto Federal de Educação, Ciência e Tecnologia do Amazonas – IFAM. Estr. Coari
Itapeua, s/n - Itamarati, 69460-000, Coari, Amazonas - Brazil.
4Instituto Nacional de Pesquisas da Amazônia – INPA. Av. André Araújo, 2936 -
Petrópolis, 69067-375, Manaus - Amazonas, Brazil.
5Petroleum Engineering Department, Edith Cowan University. 270 Joondalup Dr,
Joondalup WA 6027, Australia.
6Centro de Biotecnologia da Amazônia – CBA/SUFRAMA. Av. Gov. Danilo de Matos
Areosa, 690 - Distrito Industrial, 69075-351, Manaus, Amazonas - Brazil.

ABSTRACT

Homogeneous catalysts used for biodiesel synthesis have several limitations, including
non-recoverability/reusability, saponification, emulsification, equipment corrosion, and
environmental pollution. To overcome these limitations, we synthesized a novel catalyst
via calcination of pineapple leaves waste. This catalyst was characterized by X-ray
powder diffraction, x-ray fluorescence, Fourier transform infrared spectroscopy,
thermogravimetric analysis, scanning electron microscopy, and soluble alkalinity
measurements. The catalyst’s activity with regards to soybean oil transesterification was
analyzed, and multiple process parameters (temperature, catalyst amount, reaction time,
and methanol:oil molar ratio) were examined. A high catalytic activity, probably related
to the 85 wt% content of alkali/alkali metals (K, Ca and Mg), was observed after a 30
min reaction time, 60 °C, 4 wt% of catalyst, oil to methanol molar ratio of 1:40,
reaching an oil to biodiesel conversion above 98%. We conclude that the novel catalyst
presented here is efficient, cost-effective, and sustainable, while simultaneously
abundant waste is reduced.

Keywords: Pineapple waste problems; heterogeneous catalysis; transesterification;


biodiesel

*Corresponding author: Email: freitas.flavio@yahoo.com.br; Phone: +55 92 98189-6932

1. Introduction

Environmental management procedures have been adopted in several industrial


and agricultural sectors (Nilsson et al., 2007). One key aspect in environmental
management is the use of raw materials from renewable sources so that sustainability is
achieved. Fuel is a major component of most industrial processes, and it is therefore
imperative that alternatives are found to replace fossil fuels (Mendonça et al., 2019a,
2019b).

Biodiesel is such an alternative as it can be synthesized from renewable


agricultural resources (Bhatia et al., 2019; Nguyen et al., 2020; Park et al., 2016). In
addition, biodiesel is a biodegradable, and environmentally friendly fuel that can in
principle replace fossil fuels (Khan et al., 2014); furthermore, biodiesel emissions are
drastically lower than those of fossil fuels, while a variety of toxic pollutants are
eliminated (including carbon monoxide, hydrocarbons, highly carcinogenic
polyaromatics, particulate matter, sulfur oxides) (Mendonça et al., 2019a). Another
major advantage of biodiesel is that it can be used for full or partial replacement of
diesel in diesel engines without engine modifications (Ma et al., 2018).

This biofuel can be synthesized via transesterification of vegetable oils (Khan et


al., 2014). Triglycerides present in the vegetable oil react with added short-chain
alcohols (e.g. methanol) to form alkyl esters and glycerol (Baskar et al., 2018; Thushari
and Babel, 2018). To achieve high performance, alkaline homogeneous catalysts are
typically added as promotors (Rezania et al., 2019; as they have greater catalytic
activity than acidic catalysts; (Atadashi et al., 2013). However, a large amount of water
is needed to purify the biodiesel, and recycling such homogeneous catalysts is not
straightforward (Lee et al., 2020). Specifically, low-cost catalysts such as NaOH or
KOH require refined raw materials since oils or fats with significant acidity neutralize
and deactivate the catalysts (Atadashi et al., 2013).

Consequently, heterogeneous catalysts can significantly lower biodiesel


production costs as they do not require refined raw materials, and they can be easily
recycled (Gebremariam and Marchetti, 2018).

However, many solid alkaline catalysts were complicated and expensive to


synthesize, or they showed a high degree of detrimental leaching (Atadashi et al., 2013).
In contrast, bio-based catalysts can be easily obtained from agro-industrial wastes
through simple calcination at reduced. This process can bring economic and
environmental benefits through waste reduction, in addition to producing a low-cost
catalyst (Gebremariam and Marchetti, 2018). Thus, several such heterogeneous bio-
catalysts, based on renewable materials, have been identified, e.g. tucumã, banana or
colla peels, trees or peduncles of banana, woods and cupuaçu seed-based catalysts
(Balajii and Niju, 2019a; Betiku et al., 2016; Gohain et al., 2017; Mendonça et al.,
2019a, 2019b; Sharma et al., 2012). These catalysts showed good catalytic activities,
probably due to their high alkali content, including potassium (K), calcium (Ca), sodium
(Na) and magnesium (Mg) (Balajii and Niju, 2019a, 2019b; Betiku et al., 2016;
Mendonça et al., 2019a). However, high potassium carbonate concentrations led to
extensive leaching (Mendonça et al., 2019a; Sharma et al., 2012), soap formation,
which hinders efficient catalyst recycling. It is, therefore, necessary to further examine
additional prospective catalysts, to overcome these limitations.

In this context, an abundant and potentially effective resource to produce a bio-


based catalyst is pineapple (Ananás comosus), which is a native plant of America.
Pineapple is a monocotyledonous plant belonging to the bromeliad family and
bromelioideae subfamily (Crestani et al., 2010; Saito and Harborne, 1983); and it is one
of the most consumed fruits in the world, either directly or in processed foods (e.g. as a
base material for syrups, vinegar, wines, liqueurs, flour, and animal food, etc. (Dai and
Huang, 2017)).

In 2017, Brazil produced 2,253,897 tons of pineapple, generating twice as much


bio-waste (FAOSTAT, 2017). This waste consists of crown, bark, and leaves, which are
usually burned in the middle of the plantations at the end of the fruit harvest (Reinhardt
et al., 2018), polluting the environment.

Thus, this work examines the catalytic activity of calcined pineapple leaves with
regards to biodiesel production, to reduce waste and reliance on fossil fuels, and to aid
in the implementation of a sustainable agricultural and energy industry.

2. Materials and Methods

2.1 Materials

For this work methanol (> 99.5 vol%, -Vetec), n-hexane (99 vol%, Nuclear),
refined soybean oil (Concórdia), NaCl (99 wt%, Vetec), Na2SO4 (> 99 wt%, Vetec),
HCl (37 wt%, Vetec) and CDCl3 (99.8 vol%, Sigma-Aldrich) were used as received.

2.2 Catalyst Preparation

Mature pineapple (Ananas comosus – Vitoria variety) leaves were collected at


Nova Esperança site, municipality of Novo Aripuanã, Amazonas State, Brazil. After
manual separation of the leaves, they were cut with scissors to 1.5 - 2.5 cm length and
washed in running water to remove impurities. Subsequently, the leaves were dried in
an oven with air circulation (CIENLAB, model CE 220/81) at 65 ºC for 72 h. The
leaves were then ground in a knife mill (ACD LAB TE-650/1) with a 30 mesh sieve
(0.59 mm) and calcined at 600 °C for 2 h and 900 °C for 30 min in a muffle furnace; the
latter temperature is required to avoid CO2 chemisorption (which can be significant for
high calcium content materials (Technical Association of Pulp and Paper Industry,
2011). At this stage, ash content was measured via mass balance (the crucibles were
weighed with the material before and after calcination).

2.3 Catalyst Characterization

The catalyst obtained was characterized via X-ray fluorescence (FRX; using a
Panalytical Epsilon 3-XL instrument and a maximum voltage of 50 kV, a maximum
current of 3 mA and a helium gas pressure of 0.8 bar) and X-ray diffraction (XRD;
using a Bruker D2 Phaser diffractometer equipped with a copper anode and a nickel
filter; the powdered samples were inserted into an acrylic sample holder, with a

variation of 2θ, operating at a power of 40 kV, current of 30 mA and a sweeping range


of 10 – 100°). The greenish calcined pineapple leaves (CPL) ash was also examined by
Fourier transform infrared spectroscopy (with an attenuated reflectance
spectrophotometer FTIR-ATR Cary 630 - Agilent). FTIR spectra were acquired for the
range 4,000 to 650 cm-1, with a resolution of 8 cm-1 and 128 scans per sample.
Thermogravimetric and differential thermal analysis (TG-DTA) were used to evaluate
the influence of temperature on the CPL; here 4.174 mg of ash were placed in a
platinum crucible on a thermal analyzer (Shimadzu – TGA-50), applying a heating rate
of 10 °C/min (ambient temperature to 1000 °C in N2 atmosphere). Soluble alkalinity was
also analyzed, thus 20 mL of distilled water and 0.2 g of CPL were mixed in a beaker
while stirring for 48 h at room temperature. Thereafter, the mixture was centrifuged
(Cienlab CT-15000R) for 4 min at 13,500 rpm to separate the solid, which was then
titrated with 0.1 M aqueous HCl solution to determine the number of moles required to
neutralize 1 g of water-leached CPL. Catalyst morphology was imaged with a scanning
electron microscope (SEM; using a Tescan microscope, model VEGA 3 LMU), under
different magnifications. The CPL was macerated and deposited on carbon tapes for the
SEM analysis. The catalyst was sieved between 149 and 20 µm in order to obtain the
particle size distribution (Rafael García et al., 2020). The bulk density was obtained
according to the modified ASTM-C948 methodology (ASTM-C948, 2014), where 5 mL
volumetric flasks were filled with water (25 °C) and weighed to obtain the real volume
(VR). Knowing the VR, the dried flasks were filled with CPL catalyst and weighed,
obtaining its density (g/cm3).

2.3.2 Biodiesel synthesis

Biodiesel was synthesized via transesterification. Thus 1 g of refined soybean


oil, 5 wt% of CPL (relative to the oil mass), and 1.82 mL of methyl alcohol were added
into a 100 mL flask, maintaining an oil:methanol molar ratio of 1:40. The system was
coupled to a condenser to prevent alcohol loss and the reaction was conducted
isothermally (temperature was controlled by a heating mantle Mtops MS-DMSB) under
stirring; various reaction times were tested (15, 30, 60, 120 and 180 min). After the
reaction, the catalyst was separated by centrifugation (Cienlab CT-15000R, 4 min at

13,500 rpm) and washed with 20 % (w/v) NaCl solution and n-hexane. This procedure
separated the biodiesel from the byproduct glycerol; the biofuel was then purified by
evaporating the organic phase under reduced pressure using a rotary evaporator (Tecnal
TE-211). In order to evaluate the catalytic activity of CPL at different reaction
conditions, the catalyst dosage (from 1 to 5 wt%), temperature (50, 60, 70 and 80 °C),
and oil:methanol molar ratio (1:10, 1:20, 1:30 and 1:40) were varied. All reactions were
reproduced thrice, presenting standard deviations lower than 5% of conversion. Finally,
the resulting material was dissolved in CDCl3 and analyzed by 1H NMR (Bruker
Advance III HD 500 MHz/54 mm) to measure the conversion:

(Eq. 1) Conversion (%) =


(AOCH3) x 2

(ACH2C = O) x 3x 100

where AOCH3 and ACH2C=O are the integrals related to the methoxy protons (chemical
shift = 3.7 ppm) and acyl protons (chemical shift = 2.6 ppm), respectively (Masteri-
Farahani et al., 2020; Mendonça et al., 2019b).

2.3.3.Catalyst stability
CPL reusability was examined by repeating the transesterification reaction under

optimized conditions (oil:methanol molar ratio of 1:40, 4 wt% of CPL, 30 min reaction
time at 60 °C). After each reaction, the catalyst was centrifuged (4 min at 13,500 rpm),
washed 3 times with n-hexane, dried overnight at 80 ℃ and reused.

CPL stability was then tested via leaching tests applying optimized reaction
conditions. First, only the catalyst and methanol were added to a 100 mL flask, while
stirring constantly at 60 °C for 30 min. In order to observe possible CPL leaching into
the methanol and, consequently, a possible partial homogeneous catalysis, the catalyst
was separated by centrifugation and the methanol was transferred into another 100 mL
flask which contained soybean oil at the optimum oil:methanol molar ratio (1:40).
Hereafter, the reaction medium was treated, the organic phase evaporated under reduced
pressure, and the biodiesel conversion was analyzed by 1H NMR.

2.3.4.Activation energy

Since catalysts reduce the activation energy (Ea) and thereby increase the speed
of a specific reaction, Ea is an indicator of catalytic activity of the applied material.
Therefore, the activation energy was measured (Mendonça et al., 2019b).

3. Results and discussion


3.1. Catalyst Characterization

Ash content

An important factor for obtaining catalysts through the calcination of agro-


industrial wastes is the ash content. The material (CPL) obtained from calcining the
pineapple leaves had an ash content of 6 wt%, significantly higher than that obtained
from the calcination of cupuaçu seeds (2.6%) (Mendonça et al., 2019b).

Bulk density

The density obtained for CPL was 0.416 ± 0.084 g/cm3, similar to the
potassium-based catalyst supported on pumice (0.50 g/cm3; Cercado et al., 2018). It is
important to note that the analysis method used also considers the empty spaces
between the particles (ASTM-C948, 2014), which may result in a higher density using
other methods. Another factor that can influence this density is the formation of pores
during the calcination process.

X-ray fluorescence measurements

The CPL mainly consisted of potassium (K) and calcium (Ca), i.e. alkaline and
earth alkaline metals. This composition was similar to solid base catalysts obtained
through the calcination of agro-industrial wastes (Balajii and Niju, 2019; Betiku et al.,
2016; Mendonça et al., 2019a; Mendonça et al., 2019b), and was therefore deemed a
potential new catalyst for biodiesel production since high catalyst activity was
previously related to alkaline content (Betiku et al., 2016; Mendonça et al., 2019b;
Sharma et al., 2012).

X-ray diffraction measurements

Furthermore, the CPL catalyst consisted of multiple solid phases including


magnesium oxide (MgO), magnesium phosphate (Mg2P2O7), sodium-calcium
phosphate-silicate [Na2Ca4(PO4)2SiO4], aluminum sulfate (Al(SO4)3), manganese-
magnesium silicate (Mn0.97Mg0.03SiO3) and potassium sulfate (K2SO4). These XRD
results are consistent with the XRF analysis. Such a mixed composition is characteristic
of materials obtained through the calcination of lignocellulosic wastes (Betiku et al.,
2016; Mendonça et al., 2019a; 2019b). Importantly, oxide and carbonate phases were
not observed in large proportions; these phases produce soap during transesterification
and hinder catalyst recycling (Mendonça et al., 2019a; Sharma et al., 2012). Activated
carbon obtained from heat-treated pineapple leaves at 500 °C using phosphoric acid as
an activating agent also contained potassium, calcium, phosphorus, and silicon as major
inorganic elements (Mopoung et al., 2019). These diffraction patterns also showed a
characteristic profile of calcined lignocellulosic waste, suggesting the presence of
silicates and MnO2, SiO2, CaO, and K2O oxides. However, this study was conducted at
500 °C, which is significantly lower, and thus not directly comparable with the 600 and
900 °C for 2 h and 30 min, respectively, applied here.

Fourier transform infrared spectroscopy with attenuated reflectance

The vibrational modes active in the infrared region of the pineapple leaves
before and after heat treatment were identified by FTIR-ATR. The materials before and
after calcination exhibited quite distinct spectra; while the leaves showed characteristic
bands of cellulose (3429, 1372 and 1163 cm-1 related to ν(O-H), δ(CH) and νas(C-O-C)
vibrations, respectively) and lignin (1650, 1508, 1425 and 1270 cm-1 related to ν(C=O
conjugate), ν(C=C aromatic), δ(CH) and νas(C-O) vibrations, respectively) (Mendonça
et al., 2019a; Pastore et al., 2008), CPL presented a completely different profile. There
vibrational modes associated with water molecules and OH groups characteristic of
carbon chains were absent, confirming the elimination of biomass by thermal
decomposition, while vibrational bands were identified at wavenumbers 868, 1007,
1410, and 3026 cm-1. This corresponds to HSO4 groups (820 to 880 cm-1), in this case,
aluminum and potassium sulfates, and O-H bond deformations (Kloprogge et al., 2001;

Periasamy et al., 2009). Bands at 978 cm-1 were characteristic of P-O bonds in
magnesium phosphates, (Chitra et al., 2019); while bands at 1410 and 3026 cm-1 were
caused by asymmetric stretching (νas) of Si-O-Si bonds present in Ca2MgSi2O7 and
stretching of O-H bonds in silicates, consistent with observations made by Choudhary et
al., 2015.

Thermogravimetric analysis

A mass loss of 6.7 % was measured when heating the sample to 1000 ℃, Fig. 1.
The greatest loss was due to lost humidity (2.7%, at temperatures < 80 ℃) (de Freitas et
al., 2019). Another significant loss (1.7%) occurred between 550 and 800 ℃, which was
related to the degradation of the sulfate groups (Meng et al., 2017), corroborating the
presence of K2SO4 in the CPL as also observed by XRD analysis. Small intermediate
losses were also observed, probably resulting from the degradation of different phases
that form the material. Similar behavior was observed for tucumã peels (Mendonça et
al., 2019b). These small losses can be also from the desorption of the chemisorbed CO2

(between 360 and 886 °C), which is originated during the calcination process
(Mendonça et al., 2019a, 2019b).

FIGURE 1.

Scanning electron microscopy

CPL morphology was investigated via scanning electron microscopy in field


emission mode. The predominant occurrence of cylindrically shaped microstructures
composed of numerous nanometer plates arranged anisotropically over the entire
surface was observed. Fewer porous nanoplate structures were also detected. We
conclude that decomposition of the organic fraction resulted in the formation of
minerals which maintained the initial cylindrical profile of the original organic
composition.

Similar morphologies were also observed for catalysts obtained from (calcined)
elephant-ear tree pod husks (Falowo et al., 2019) and banana peels (Betiku et al., 2016),
where the micrograph showed that there was sintering of small mineral aggregates with
a high level of agglomeration of the nanoparticles.

SEM analysis also showed a particle size distribution is quite heterogeneous,


which was also confirmed by the sieving of the catalyst. Hence, based on these two
methodologies, a particle size distribution ranging from 15 to 32 µm was obtained.

Soluble alkalinity

A low concentration of soluble alkaline compounds (0.39 mmol g−1 of CPL) was
measured. This is in contrast to significantly higher alkaline solubilities observed for
cupuaçu seed ashes (1.05 mmol g−1 of catalyst, Mendonça et al., 2019a) and tucumã
peels (3.7 mmol g−1 of catalyst, calcined at 800 °C, Mendonça et al., 2019b).

This discrepancy was probably caused by the higher concentrations of alkali and
alkaline earth metals (in the form of carbonates and phosphates) in the cupuaçu seed and
tucumã peel ashes.

3.2. Catalytic activity

The catalytic activity of CPL was evaluated and the transesterification reaction
was optimized with regards to time, catalyst loading, temperature, and oil:methanol
molar ratio. The results are discussed in detail below.

Time

Time is a primary parameter in industrial biodiesel production, and it needs to be


minimized (Gebremariam and Marchetti, 2018). Here, reaction time was varied from 30
to 180 min, to achieve the highest conversion in the shortest possible time.

FIGURE 2.

Remarkably, 98.45 % conversion of soybean oil to biodiesel was achieved after


120 min reaction time, Fig. 2. Even after 30 min of reaction time, high conversion
(96.88 % %) was achieved. These results demonstrate that CPL has an excellent
catalytic activity. However, a reaction time around 15 min was insufficient to achieve
high conversions and only 25 % conversion was measured. This is somewhat faster than
typical reaction times reported in the literature, which are usually > 60 min, although
literature values vary widely. For example, tucumã peel ash (a solid base catalyst rich in
K) showed a maximum conversion of 95.7 % after 120 min, banana peel ash reached
94.97 % conversion after 65 min (in the transesterification of oleander oil; Betiku and
Ajala, 2014), and 91 - 99% conversion was achieved after 180 min using K2CO3/CaCO3
catalyst supported on wood ash (Sharma et al. 2012). Benchmark this with standard
homogeneous catalysts (e.g. NaOH and KOH), where reaction times are significantly
longer (240 - 480 min) (Atadashi et al., 2013).

Catalyst loading

The efficiency of a catalyst is also a function of the amount of catalyst required


to achieve high conversion. Thus, the amount of catalyst was varied from 1 to 5 wt%
relative to the mass of soybean oil to assess this parameter.

FIGURE 3.

Clearly, the amount of catalyst had a significant influence on the conversion; for
instance, when catalyst loading was increased from 1 to 4 wt%, conversion increased
from 18.87 to 98.92 % (Fig. 3). However, a further increase in catalyst loading to 5%
slightly decreased the conversion (95.11 %). This decrease at high loading was probably
caused by the increased viscosity of the reaction medium (viscosity of a slurry increases
with solid loading) (Mendonça et al., 2019b); note that reaction slurry viscosity is

inversely proportional to the transfer rate of the reagents from the reaction medium to
the catalyst surface (Mendonça et al., 2019a, 2019b). These results are in contrast to
those measured for cupuaçu seeds ash, where high conversions could only be achieved
with high catalyst loadings (10 wt%; Mendonça et al., 2019a). High loads of catalysts
can make biodiesel production costly. These loadings were higher than those required
for effective homogeneous catalysis, e.g. NaOH or KOH can be used at concentrations
less than 2 wt% (Atadashi et al., 2013; Rezania et al., 2019). Despite this advantage,
homogeneous catalysts have serious limitations, such as the formation of saponified
product, emulsion formation, high purification cost, high water, and energy
consumption, high wastewater discharges, which limits their usage (Atadashi et al.,
2013; Melero et al., 2009).
Temperature

For industrial biodiesel production, high temperatures require excessive energy


expenditure. Thus, operational temperatures need to be minimized; we, therefore,
studied here also the effect of temperature on the CPL-catalyzed transesterification.

FIGURE 4.

A drastic increase in conversion was observed when temperature increased from


50 to 60 ºC (from 20.30 to 97.55 %), Fig. 4. This strong increase is probably related to
the significantly lower reaction medium viscosity, increased alcohol solubility, and
accelerated reagent transfer rates at 60 ºC (Silva; and Oliveira, 2014). Further
temperature increase resulted in further slight improvements in conversion. Similar
behavior was also observed by Mendonça et al., 2019a, although they achieved a 97.3%
conversion applying 10 wt% of catalyst at relatively high temperatures (80 °C).
Atadashi et al., 2013 observed that for homogeneous base catalysts, when temperature
varied from 25 to 100 °C, conversions ranged from 55 to 99.6 %. Typically,
temperatures above the boiling point of methanol (64.7 °C) are used, hence, in this

study, the optimum temperature was 60 °C, at which already good conversions were
achieved.

Oil:methanol molar ratio

In the transesterification of vegetable oil with methanol, the stoichiometric


oil:alcohol molar ratio is 1:3 (Musa, 2016). Thus, based on Le Chatelier’s principle, by
increasing the alcohol concentration, the formation of biodiesel is favored (Sahani and
Sharma, 2018). These aspects were inspected further here experimentally.

FIGURE 5.

Conversion increased drastically when the oil:methanol was increased from 1:10
to 1:30, and then plateaued off at a ratio of 1:40, reaching 98.55 %, Fig. 5. CPL-
catalyzed soybean oil transesterification proved to be highly dependent on this
parameter. Note that excess methanol can be easily distilled and reused since it has a
low boiling point, even at ambient pressure. Furthermore, the optimum molar ratio
tested (1:40) was much higher than some literature values, although several studies also
reported significantly higher values (e.g. Xie and Zhao, 2014, and Nur Syazwani et al.,
2015 reported optimum molar ratios of 1:50 and 1:150, respectively). Compare these
results with optimum molar ratios less than 1:20 which are usually applied in
homogeneous catalysts (using KOH and NaOH; Atadashi et al. 2013). However, the
abovementioned disadvantages associated with homogenous catalysts remain.

Based on the performance presented during the parameter optimization


evaluation, we conclude that CPL has a significantly higher catalytic activity than most
homogeneous or heterogeneous catalysts reported in the literature.

3.3. Catalyst stability

One of the main advantages of a heterogeneous catalyst is its recyclability. Thus,


CPL recycling was examined with respect to successive transesterification reactions.
Note that the CPL was not re-calcined before each reuse to minimize energy
consumption. CPL catalytic activity was maintained during the first three reaction
cycles (1º - 98.92%, 2 º - 97.6% and 3º - 94.3%), but reduced by approximately 8% in
the fourth cycle. However, the conversion also remained high, above 85%, in the fourth
cycle. This reduced activity is likely due to the impregnation of organic compounds on
the catalyst surface/pores during the reaction, disabling active sites (Abukhadra and
Sayed, 2018), and could not be restored by washing with n-hexane. Falowo et al., 2019
observed similar behavior for calcined elephant-ear tree pod husks; they pointed out that
the decrease in biodiesel conversion could be due to the formation of potassium
methoxide in the reaction medium as a result of active site leaching. Mendonça et al.,
2019a and 2019b also observed similar results for cupuaçu seeds and tucumã peels
ashes, respectively; although CPL was generally more stable than these catalysts,
especially when compared with the third recycling cycle of the cupuaçu seeds, which
presented a very low conversion (22 %).

CPL leaching into methanol was also studied; note that for the CPL enriched
methanol phase a conversion of 4.5 % (homogeneous catalysis) was measured. Despite
this, no emulsion or soap formation was observed. This low conversion probably
stemmed from potassium leaching (forming K+ and CH3O-; Falowo et al., 2019).
Similarly, low homogeneous catalytic behavior was observed for tucumã peel ashes
(Mendonça et al., 2019b), while cupuaçu seeds ash had lower stability (only 60.4%
conversion was achieved after one reaction recycle; Mendonça et al., 2019b), allowing
it to be recycled only once. We conclude that CPL is very stable under the applied
reaction conditions. It would be possible, however, to improve CPL stability further, by
calcining CPL prior to each reaction, as recalcination prevents leaching and reduces the
amount of organic compounds that poison the catalyst (Betiku et al., 2016; Sharma et
al., 2012).

3.4. Activation energy (Ea)

The activation energy for the CPL catalyzed biodiesel synthesis was 86.84
kJ/mol. This is similar to the Ea measured for the transesterification of soybean oil with
0.5% sodium methoxide (Ea = 83.77 kJ/mol; Freedman et al., 1986), but much lower

than Ea for the transesterification of palm oil without catalyst under supercritical
conditions (Ea = 105 kJ/mol; Permsuwan et al., 2011) or catalysis by KF/Ca−Mg−Al
hydrotalcite solid base (Ea = 111.6 kJ/mol; Xiao et al., 2010). Ea measured for nano
CaO-catalyzed canola oil transesterification was also significantly higher (Ea = 136.48
kJ/mol; Zhao et al., 2013).

4. Conclusions

A low-cost eco-friendly biodiesel catalyst with high activity was developed from waste
pineapple leaves. The catalytic efficiency of the calcined pineapple leaves (CPL) was
high, and conversions greater than 98% were measured. Catalyst loading and
oil:methanol molar ratio had both a significant influence on CPL performance.
Furthermore, recycled CPL maintained high activity (until the fourth recycling cycle)
and presented a low transesterification activation energy (86.84 kJ/mol). We conclude
that CPL is a highly efficient biodiesel catalyst, while simultaneously reducing
biowaste.

E-supplementary data for this work can be found in e-version of this paper online.

Acknowledgment
The authors are thankful to Coordenação de Aperfeiçoamento de Pessoal de Nível
Superior for the scholarship for Silma de Sá Barros (CAPES - finance code 001), to
Laboratório de Materiais Amazônicos e Compósitos – LAMAC/UFAM for the XRF
analysis, to Laboratório Temático de Microscopia Ótica e Eletrônica – INPA, and to
NMR-LAB (UFAM) for all NMR analyses.

References

1 - Abukhadra, M.R., Sayed, M.A., 2018. K+ trapped kaolinite (Kaol/K+) as low cost
and eco-friendly basic heterogeneous catalyst in the transesterification of

commercial waste cooking oil into biodiesel. Energy Convers. Manag. 177, 468–
476. https://doi.org/10.1016/j.enconman.2018.09.083

2 - ASTM-C948, 2014. Standard Test Method for Dry and Wet Bulk Density, Water
Absorption, and Apparent Porosity of Thin Sections of Glass-Fiber Reinforced.
Am. Soc. Test. Mater. West Conshohocken, PA, USA.
https://doi.org/10.1520/D1895-96R10

3 - Atadashi, I.M., Aroua, M.K., Abdul Aziz, A.R., Sulaiman, N.M.N., 2013. The
effects of catalysts in biodiesel production: A review. J. Ind. Eng. Chem. 19, 14–
26. https://doi.org/10.1016/j.jiec.2012.07.009

4 - Balajii, M., Niju, S., 2019a. A novel biobased heterogeneous catalyst derived from
Musa acuminata peduncle for biodiesel production – Process optimization using
central composite design. Energy Convers. Manag.
https://doi.org/10.1016/j.enconman.2019.03.085

5 - Balajii, M., Niju, S., 2019b. Banana peduncle – A green and renewable
heterogeneous base catalyst for biodiesel production from Ceiba pentandra oil.
Renew. Energy. https://doi.org/10.1016/j.renene.2019.08.062

6 - Baskar, G., Aberna Ebenezer Selvakumari, I., Aiswarya, R., 2018. Biodiesel
production from castor oil using heterogeneous Ni doped ZnO nanocatalyst.
Bioresour. Technol. https://doi.org/10.1016/j.biortech.2017.12.010

7 - Betiku, E., Ajala, S.O., 2014. Modeling and optimization of Thevetia peruviana
(yellow oleander) oil biodiesel synthesis via Musa paradisiacal (plantain) peels
as heterogeneous base catalyst: A case of artificial neural network vs. response
surface methodology. Ind. Crops Prod. 53, 314–322.
https://doi.org/10.1016/j.indcrop.2013.12.046

8 - Betiku, E., Akintunde, A.M., Ojumu, T.V., 2016. Banana peels as a biobase catalyst
for fatty acid methyl esters production using Napoleon’s plume (Bauhinia
monandra) seed oil: A process parameters optimization study. Energy 103, 797–
806. https://doi.org/10.1016/j.energy.2016.02.138

9 - Bhatia, S.K., Gurav, R., Choi, T.R., Han, Y.H., Park, Y.L., Park, J.Y., Jung, H.R.,
Yang, S.Y., Song, H.S., Kim, S.H., Choi, K.Y., Yang, Y.H., 2019.
Bioconversion of barley straw lignin into biodiesel using Rhodococcus sp.
YHY01. Bioresour. Technol. https://doi.org/10.1016/j.biortech.2019.121704
10 - Cercado, A.P., Ballesteros, F., Capareda, S., 2018. Ultrasound assisted
transesterification of microalgae using synthesized novel catalyst. Sustain.
Environ. Res. https://doi.org/10.1016/j.serj.2018.03.002

11 - Chitra, S., Bargavi, P., Durgalakshmi, D., Rajashree, P., Balakumar, S., 2019. Role
of sintering temperature dependent crystallization of bioactive glasses on
erythrocyte and cytocompatibility. Process. Appl. Ceram. 13, 12–23.
https://doi.org/10.2298/PAC1901012C

12 - Choudhary, R., Koppala, S., Swamiappan, S., 2015. Bioactivity studies of calcium
magnesium silicate prepared from eggshell waste by sol-gel combustion
synthesis. J. Asian Ceram. Soc. 3, 173–177.
https://doi.org/10.1016/j.jascer.2015.01.002

13 - Crestani, M., Hawerroth, F.J., de Carvalho, F.I.F., de Oliveira, A.C., Barbieri, R.L.,
2010. From the Americas to the World - origin, domestication and dispersion of
pineapple. Cienc. Rural. https://doi.org/10.1590/S0103-84782010000600040

14 - Dai, H., Huang, H., 2017. Synthesis, characterization and properties of pineapple
peel cellulose-g-acrylic acid hydrogel loaded with kaolin and sepia ink.
Cellulose. https://doi.org/10.1007/s10570-016-1101-0

15 - Freitas, F.A., Keils, D., Lachter, E.R., Maia, C.E.B., Pais da Silva, M.I., Veiga
Nascimento, R.S., 2019. Synthesis and evaluation of the potential of nonionic
surfactants/mesoporous silica systems as nanocarriers for surfactant controlled
release in enhanced oil recovery. Fuel 1184–1194.
https://doi.org/10.1016/j.fuel.2018.12.059

16 - Falowo, O.A., Oloko-Oba, I.M., Betiku, E., 2019. Biodiesel production


intensification via microwave irradiation-assisted transesterification of oil blend
using nanoparticles from elephant-ear tree pod husk as a base heterogeneous
catalyst. Chem. Eng. Process. - Process Intensif.
https://doi.org/10.1016/j.cep.2019.04.010

17 - FAOSTAT, 2017. Food and Agriculture Organization of the United Nations


[WWW Document]. URL http://www.fao.org/faostat/en/?#data/QC

18 - Freedman, B., Butterfield, R.O., Pryde, E.H., 1986. Transesterification kinetics of


soybean oil 1. J. Am. Oil Chem. Soc. https://doi.org/10.1007/BF02679606

19 - Gebremariam, S.N., Marchetti, J.M., 2018. Economics of biodiesel production:


Review 168, 74–84. https://doi.org/10.1016/j.enconman.2018.05.002

20 - Gohain, M., Devi, A., Deka, D., 2017. Musa balbisiana Colla peel as highly
effective renewable heterogeneous base catalyst for biodiesel production. Ind.
Crops Prod. 109, 8–18. https://doi.org/10.1016/j.indcrop.2017.08.006

21 - Khan, T.M.Y., Atabani, A.E., Badruddin, I.A., Badarudin, A., Khayoon, M.S.,
Triwahyono, S., 2014. Recent scenario and technologies to utilize non-edible
oils for biodiesel production. Renew. Sustain. Energy Rev. 37, 840–851.
https://doi.org/10.1016/j.rser.2014.05.064
22 - Kloprogge, J.T., Ruan, H., Frost, R.L., 2001. Near-infrared spectroscopic study of
basic aluminum sulfate and nitrate. J. Mater. Sci. 36, 603–607.
https://doi.org/10.1023/A:1004860118470

23 - Lee, J.C., Lee, B., Ok, Y.S., Lim, H., 2020. Preliminary techno-economic analysis
of biodiesel production over solid-biochar. Bioresour. Technol.
https://doi.org/10.1016/j.biortech.2020.123086

24 - Ma, Y., Gao, Z., Wang, Q., Liu, Y., 2018. Biodiesels from microbial oils:
Opportunity and challenges. Bioresour. Technol.
https://doi.org/10.1016/j.biortech.2018.05.028

25 - Masteri-Farahani, M., Hosseini, M.S., Forouzeshfar, N., 2020. Propyl-SO3H


functionalized graphene oxide as multipurpose solid acid catalyst for biodiesel
synthesis and acid-catalyzed esterification and acetalization reactions. Renew.
Energy. https://doi.org/10.1016/j.renene.2019.11.108

26 - Melero, J.A., Iglesias, J., Morales, G., 2009. Heterogeneous acid catalysts for
biodiesel production: Current status and future challenges. Green Chem.
https://doi.org/10.1039/b902086a

27 - Mendonça, I.M., Machado, F.L., Silva, C.C., Duvoisin Jr., S., Takeno, M.L., Maia,
P.J.S., Manzato, L., de Freitas, F.A., 2019a. Application of calcined waste
cupuaçu (Theobroma grandiflorum ) seeds as a low-cost solid catalyst in
soybean oil ethanolysis : Statistical optimization. Energy Convers. Manag. 200,
112095. https://doi.org/10.1016/j.enconman.2019.112095

28 - Mendonça, I.M., Paes, O.A.R.L., Maia, P.J.S., Souza, M.P., Almeida, R.A., Silva,
C.C., Duvoisin Jr., S., de Freitas, F.A., 2019b. New heterogeneous catalyst for
biodiesel production from waste tucumã peels (Astrocaryum aculeatum Meyer):
Parameters optimization study. Renew. Energy 130, 103–110.
https://doi.org/10.1016/j.renene.2018.06.059

29 - Meng, C., Cao, G.P., Li, X.K., Yan, Y.Z., Zhao, E.Y., Hou, L.Y., Shi, H.Y., 2017.
Structure of the SO42−/TiO2 solid acid catalyst and its catalytic activity in
cellulose acetylation. React. Kinet. Mech. Catal. https://doi.org/10.1007/s11144-
017-1165-3

30 - Mopoung, S., Amornsakch, P., Mopoung, R., Thianngam, P., 2019. Potassium
Permanganate Loaded Activated Carbon Production from Pineapple Leaf at Low
Pyrolysis Temperature for Water Hardness Removal. Asian J. Sci. Res. 12, 126–
136. https://doi.org/10.3923/ajsr.2019.126.136

31 - Musa, I.A., 2016. The effects of alcohol to oil molar ratios and the type of alcohol
on biodiesel production using transesterification process. Egypt. J. Pet.
https://doi.org/10.1016/j.ejpe.2015.06.007

32 - Nguyen, H.C., Nguyen, M.L., Wang, F.M., Juan, H.Y., Su, C.H., 2020. Biodiesel
production by direct transesterification of wet spent coffee grounds using
switchable solvent as a catalyst and solvent. Bioresour. Technol.
https://doi.org/10.1016/j.biortech.2019.122334

33 - Nilsson, L., Olof Persson, P., Rydén, L., Darozhka, S., Zaliauskiene, A., 2007.
Cleaner Production: Technologies and Tools for Resource Efficient Production,
Encyclopedia of Global Environmental Change.

34 - Nur Syazwani, O., Rashid, U., Taufiq Yap, Y.H., 2015. Low-cost solid catalyst
derived from waste Cyrtopleura costata (Angel Wing Shell) for biodiesel
production using microalgae oil. Energy Convers. Manag.
https://doi.org/10.1016/j.enconman.2015.05.075

35 - Park, J., Kim, B., Lee, J.W., 2016. In-situ transesterification of wet spent coffee
grounds for sustainable biodiesel production. Bioresour. Technol.
https://doi.org/10.1016/j.biortech.2016.09.001

36 - Pastore, T.C.M., De Oliveira, C.C.K., Rubim, J.C., Santos, K.D.O., 2008. Effect of
artificial weathering ON TROPICAL WOODS monitored by infrared
spectroscopy (DRIFT). Quim. Nova 31, 2071–2075.
https://doi.org/10.1590/S0100-40422008000800030

37 - Periasamy, A., Muruganand, S., Palaniswamy, M., 2009. Vibrational studies of


Na2so4, K2SO4, NaHSO4 and KHSO4 crystals. Rasayan J. Chem. 2, 981–989.

38 - Permsuwan, A., Tippayawong, N., Kiatsiriroat, T., Thararux, C., Wangkarn, S.,
2011. Reaction Kinetics of Transesterification Between Palm Oil and Methanol
under Subcritical Conditions. Energy Sci. Technol.
https://doi.org/10.3968/j.est.1923847920110201.672

39 - Rafael García, J., María Bidabehere, C., Sedran, U., 2020. Non-uniform size of
catalyst particles. Impact on the effectiveness factor and the determination of
kinetic parameters. Chem. Eng. J. https://doi.org/10.1016/j.cej.2020.124994

40 - Reinhardt, D.H.R.C., Bartholomew, D.P., Souza, F.V.D., de Carvalho, A.C.P.P., de


Pádua, T.R.P., Junghans, D.T., de Matos, A.P., 2018. Advances in pineapple
plant propagation. Rev. Bras. Frutic. https://doi.org/10.1590/0100-29452018302

41 - Rezania, S., Oryani, B., Park, J., Hashemi, B., Yadav, K.K., Kwon, E.E., Hur, J.,
Cho, J., 2019. Review on transesterification of non-edible sources for biodiesel
production with a focus on economic aspects, fuel properties and by-product
applications. Energy Convers. Manag. 201, 112155.
https://doi.org/10.1016/j.enconman.2019.112155

42 - Sahani, S., Sharma, Y.C., 2018. Economically viable production of biodiesel using
a novel heterogeneous catalyst: Kinetic and thermodynamic investigations.
Energy Convers. Manag. 171, 969–983.
https://doi.org/10.1016/j.enconman.2018.06.059

43 - Saito, N., Harborne, J.B., 1983. A cyanidin glycoside giving scarlet coloration in
plants of the bromeliaceae. Phytochemistry 22, 1735–1740.
https://doi.org/10.1016/S0031-9422(00)80261-7

44 - Sharma, M., Khan, A.A., Puri, S.K., Tuli, D.K., 2012. Wood ash as a potential
heterogeneous catalyst for biodiesel synthesis. Biomass and Bioenergy 41, 94–
106. https://doi.org/10.1016/j.biombioe.2012.02.017
45 - Silva; C., Oliveira, J.V., 2014. Biodiesel production through non-catalytic
supercritical transesterification: current state and perspectives. Brazilian J.
Chem. Eng. 31, 271–285. https://doi.org/10.1590/0104-
6632.20140312s00002616

46 - Technical Association of Pulp and Paper Industry, 2011. T 413 om-11. Ash in
wood, pulp, paper and paperboard: combustion at 900°C. Talanta.

47 - Thushari, I., Babel, S., 2018. Sustainable utilization of waste palm oil and
sulfonated carbon catalyst derived from coconut meal residue for biodiesel
production. Bioresour. Technol. https://doi.org/10.1016/j.biortech.2017.06.106

48 - Xiao, Y., Gao, L., Xiao, G., Lv, J., 2010. Kinetics of the transesterification reaction
catalyzed by solid base in a fixed-bed reactor. Energy and Fuels.
https://doi.org/10.1021/ef100966t

49 - Xie WL, Z.L., 2014. Heterogeneous CaO-MoO3-SBA-15 catalysts for biodiesel


production from soybean oil. Energy Convers. Manag. 79, 34–42.
https://doi.org/10.1016/j.enconman.2013.11.041

50 - Zhao, L., Qiu, Z., Stagg-Williams, S.M., 2013. Transesterification of canola oil
catalyzed by nanopowder calcium oxide. Fuel Process. Technol.
https://doi.org/10.1016/j.fuproc.2013.03.027

LIST OF FIGURES

Figure 1. Thermogravimetric analysis of CPL.

Figure 2. Oil-biodiesel conversion (%) via transesterification for different reaction times
(oil:methanol molar ratio = 1:40, 80 °C and 5 wt% of catalyst).

Figure 3. Oil-biodiesel conversion (%) via transesterification versus catalyst loading.


(oil:methanol molar ratio = 1:40, 80 °C and 30 min).

Figure 4. Oil-biodiesel conversion (%) via transesterification versus reaction temperature.


(oil:methanol molar ratio = 1:40, 4 wt% of CPL, 30 min).

Figure 5. Oil-biodiesel conversion (%) via transesterification versus oil:methanol molar


ratios. (4 wt% CPL, 60 °C, 30 min).

SUPLEMENTARY INFORMATION
Application of pineapple (Ananás comosus) leaves ash as a solid base catalyst for
biodiesel synthesis
Silma de S. Barros1, Wanison A. G. Pessoa Junior2, Ingrity S. C. Sá1, Mitsuo L. Takeno2,
Francisco X. Nobre3, William Pinheiro4, Lizandro Manzato2, Stefan Iglauer5, Flávio A.
de Freitas6,*

1Deparatamento de Engenharia de Materiais – PPGCEM/UFAM. Av. Octávio Hamilton


Botelho Mourão - Coroado, 69067-005, Manaus, Amazonas - Brazil.
2Instituto Federal de Educação, Ciência e Tecnologia do Amazonas – IFAM/CMDI. Av.
Gov. Danilo de Matos Areosa, 1672 - Distrito Industrial, 69075-351, Manaus,
Amazonas, Brazil.
3Instituto Federal de Educação, Ciência e Tecnologia do Amazonas – IFAM. Estr. Coari
Itapeua, s/n - Itamarati, 69460-000, Coari, Amazonas - Brazil.
4Instituto Nacional de Pesquisas da Amazônia – INPA. Av. André Araújo, 2936 -
Petrópolis, 69067-375, Manaus - Amazonas, Brazil.
5Petroleum Engineering Department, Edith Cowan University. 270 Joondalup Dr,
Joondalup WA 6027, Australia.

6Centro de Biotecnologia da Amazônia – CBA/SUFRAMA. Av. Gov. Danilo de Matos


Areosa, 690 - Distrito Industrial, 69075-351, Manaus, Amazonas - Brazil.

Figure S1. XRD pattern of CPL obtained from the calcination of pineapple leaves.

Figure S2. FTIR spectra of CPL obtained before and after calcination of the milled
pineapple leaves.

Figure S3. FE-SEM images of CPL at different magnifications.

Figure S4. NMR spectrum showing the peaks of interest for obtaining the conversion of
oil into biodiesel.

Figure S5. Arrhenius plot of ln k vs 1/T for soybean methanolysis with 4 wt% of CLP at
50 – 80 °C.

HIGHLIGHTS
• A new catalyst from waste pineapple leaves was used in the biodiesel synthesis;
• 98.2% of conversion was obtained when applying 4 wt% of the new catalyst;
• High conversions were obtained in only 30 min of reaction time;
• Conversions higher than 85% were obtained after the fourth catalyst recycling cycle.

CREDIT AUTHOR STATEMENT

Silma de S. Barros: Investigation and Writing - Original Draft;


Wanison A. G. Pessoa Junior: Conceptualization and Resources;
Ingrity S. C. Sá: Validation;
Mitsuo L. Takeno: Methodology and Formal analysis;
Francisco X. Nobre: Formal analysis and Visualization;
William Pinheiro: Validation;

Lizandro Manzato: Funding acquisition;


Stefan Iglauer: Writing - Original Draft and Writing - Review & Editing;
Flávio A. de Freitas: Writing - Original Draft, Writing - Review & Editing and
Supervision.

You might also like