You are on page 1of 11

Fuel 244 (2019) 569–579

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Multifunctionality of zinc carboxylate to produce acylglycerols, free fatty T


acids and fatty acids methyl esters
Alexis M. Escorsima, Fabiane Hamerskib, Luiz P. Ramosa, Marcos L. Corazzab,
Claudiney S. Cordeiroa,

a
Research Center in Applied Chemistry (CEPESQ), Department of Chemistry, Federal University of Paraná (UFPR), PO Box 19032, CEP 81531-980 Curitiba, PR, Brazil
b
Department of Chemical Engineering, Federal University of Paraná (UFPR), PO Box 19011, CEP 81531-990 Curitiba, PR, Brazil

GRAPHICAL ABSTRACT

FFA

TAG
MeOH
TAG

+
H2O

Hydrolyzed Oil
FAME

H2O

ZS ZS

ARTICLE INFO ABSTRACT

Keywords: Hydrolysis of acid oils is an important alternative for the production of non-ionic surfactants and fatty acids that
Palm oil can be converted to fatty acid methyl esters (biodiesel) by esterification. In this work, hydrolysis of palm oil was
Hydrolysis initially carried out in subcritical water using zinc stearate as catalyst. The experiments were performed in a
Simultaneous esterification and CSTR-type reactor where the effect of temperature (140–190 °C), palm oil-to-water MR (1:10–1:53) and catalyst
transesterification
amount (2–10 wt%) were evaluated. High contents of free fatty acids (up to 60 wt%) with 35 wt% diacylgly-
Fatty acid methyl esters
cerols and 20 wt% monoacylglycerols were obtained. Also, zinc stearate maintained its catalytic activity for palm
Glycerin
oil hydrolysis for five consecutive reuse cycles but changes in the catalyst chemical composition were observed,
mostly due to the conversion of zinc stearate to zinc palmitate. However, by increasing the reaction time from 4
to 8 h at 190 °C, there was a decrease in the accumulation of free fatty acids and an increase in the accumulation
of acylglycerols and this was a strong evidence for reaction reversion. After removal of water, palm oil hy-
drolysates still containing the reaction catalyst were subjected to simultaneous esterification and transester-
ification under similar reaction conditions, reaching a final monoester content of about 90 wt%. These values are
50 wt% higher than those obtained by direct transesterification using the same catalyst. Also, the resulting crude
glycerin phase was colorless in appearance, indicating little contamination with reaction products and inter-
mediates.


Corresponding author at: Department of Chemistry, Federal University of Paraná (UFPR), P. O. Box 19032, CEP 81531-980 Curitiba, PR, Brazil.
E-mail address: claudiney@quimica.ufpr.br (C.S. Cordeiro).

https://doi.org/10.1016/j.fuel.2019.01.178
Received 27 November 2018; Received in revised form 29 January 2019; Accepted 30 January 2019
Available online 19 February 2019
0016-2361/ © 2019 Elsevier Ltd. All rights reserved.
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

1. Introduction alkaline medium (KOH, NaOH or their corresponding alkoxides) [4]. In


addition to homogeneous catalysts, heterogeneous catalysts can also be
Oils and fats have important applications in the food and pharma- used for esterification and transesterification but not too many are able
ceutical industries as well as in the production of biofuels, especially to catalyze both reactions simultaneously. Some of the catalytic mate-
biodiesel. Diacylglycerols (DAG) and monoacylglycerols (MAG), often rials applied in such reaction systems are zeolites, oxides and inorganic
present as biodiesel contaminants, are used as emulsifiers in food, salts, coordination compounds, ion exchange resins, double layered
pharmaceuticals and cosmetic industries, and also as lubricants for hydroxides and metal carboxylates. The catalytic activity of these
plastics and additives for paints and perfumes [1,2]. In addition, several compounds is due to the presence of Brönsted–Lowry and/or Lewis acid
researchers have shown that DAG-rich oils are dietary efficient against or base sites in their chemical composition [17,18].
obesity and arteriosclerosis [1,3]. Metal carboxylates have been tested as catalysts for different pro-
Refined vegetable oils with low water and free fatty acid (FFA) cesses such as FFA esterification [15], TAG transesterification [19],
contents are the most widely used raw materials for biodiesel produc- TAG glycerolysis [20] and simultaneous esterification and transester-
tion by homogeneous alkaline transesterification. This catalytic route ification of acid oils [17,18]. Such variety of catalytic applications is
presents some advantages such as low energy consumption, low pres- mainly associated with their acid-base properties, high catalytic ac-
sure and high reaction rates. However, the downstream processing of tivity, ease of synthesis by precipitation in aqueous medium and pos-
fatty acid methyl esters and glycerin requires high capital costs while sible recovery and reuse in several consecutive reaction cycles [18].
the use of low grade oils and fats is prohibitive due to their high water One metal carboxylate that can be highlighted is zinc(II) carbox-
and FFA contents, which must be removed down to acceptable levels ylate, which can be obtained from saturated-chain carboxylic acids and
prior to transesterification [4,5]. are represented by the formula Zn(CnH2n+1COO)2. In these compounds,
One option to reduce biodiesel production costs is the use of low zinc occupies the center of a regular tetrahedron, where it is co-
value feedstocks such as rendered materials and waste oils and fats. ordinated with two carboxylate groups in a bidentate bridge whose
However, alternative technologies must be used to overcome barriers hydrocarbon chains interact by weak van der Waals forces. In addition,
that are usually attached their high FFA and free water contents. In this their all-trans (zig-zag) conformation provides an inclination angle of
case, hybrid processes such as hydroesterification can be used, where 27° among the hydrocarbon chains forming a bilayer structure [21].
hydrolysis of triacylglycerols (TAG) is followed by esterification of FFAs The melting point of layered carboxylates occurs at temperatures above
to produce fatty acid alkyl esters [5]. 135 °C; therefore, it presents a behavior similar to that of ionic liquids.
TAG hydrolysis occurs in three steps as shown in Equations 1, 2 and This causes a derangement of the organic chains and a partial change in
3, where one mol of TAG produces three mols of FFA. In these reactions, the coordination between zinc and carboxylate anions, moving it from a
water is used in excess to shift the reaction equilibrium towards the bidentate to a monodentate bridge in which the catalytic active sites are
products formation [6,7]. more exposed. Such features unveil a variety of different applications
for this catalytic material [17,18,22].
C3 H5 (OOCR)3 + H2 O C3 H5 (OH). (OOCR)2 + RCOOH
TAG Water DAG FFA (1) Although many works related to hydroesterification use catalysts
capable of performing esterification and transesterification, only few
C3 H5 (OH). (OOCR) 2 + H2 O C3 H5 (OH)2 . (OOCR) + RCOOH studies report on the use of the same catalytic system for TAG hydro-
DAG Water MAG FFA (2)
lysis. In general, TAG hydrolysis is performed without a catalyst in
C3 H5 (OH) 2 . (OOCR) + H2 O C3 H5 (OH)3 + RCOOH subcritical or supercritical water using relatively high temperatures
Water FFA (3)
compared to the use of catalysts such as oxides or enzymes [5,9,23].
MAG GLY

From techniques already described in the literature, the Colgate- Niobium oxide [5,24,25] and enzymes [23,26] performed well in these
Emery® [8] process applies subcritical water at 5 MPa and temperatures three reaction steps but these catalysts introduce high costs and lim-
above 250 °C without any added catalyst to obtain yields above 90 wt% itations to their practical application in industrial processes. On the
in FFA from oils and fats [2,9]. Similar results were obtained by other hand, catalytic systems for the conversion of low-cost feedstocks
Kusdiana and Saka [10] using rapeseed oil at temperatures within 255 into fatty acid methyl esters, with advanced properties such as multi-
and 350 °C, oil-to-water molar ratios from 1 to 217, and pressures functionality (active in hydrolysis, esterification and transesterifica-
ranging from 5 and 20 MPa and also by Almeida et al. [11] using re- tion), stability (including reduced leaching of catalytic species), low
sidual frying oils at temperatures between 250 and 275 °C, oil-to-water energy requirements, ease for recovery and reuse, facile chemical
mass ratios from 1 to 4, and pressures of 3.6 and 5.6 MPa. Regarding synthesis and good availability at low cost, are still in great industrial
TAG hydrolysis, catalysts can be efficiently used to improve yields as demand. Besides, advanced catalytic systems should also facilitate
demonstrated by Rocha et al. [5], who carried out hydrolysis of soy- glycerin recovery and produce a glycerin phase with less contamination
bean oil using niobium oxide as catalyst and reached conversions above and high glycerol contents.
80% in FFA after 1 h. Also, Cabral et al. [12] performed the enzymatic The main objective of this work was to study the hydrolysis, ester-
hydrolysis of olive oil using Lipozyme RM IN to obtain 45% of FFA after ification and transesterification of palm oil using zinc stearate as cat-
24 h at 55 °C. By contrast, Zanevicz et al. [13] used Lipozyme TL IN to alyst. The variables investigated for palm oil hydrolysis were: tem-
perform hydrolysis of soybean oil and used frying oil in an ultrasound perature, palm oil-to-distilled water molar ratio, and catalyst
system (132 W power) at temperatures ranging from 40 to 60 °C and concentration. The accumulation of MAG and DAG were also assessed,
obtained FFA yields around 60% after 2 h. as well as the characterization of the catalyst before and after use. In
Esterification consists of a single reaction step where FFA in the addition, the catalytic effect was demonstrated by comparing its per-
presence of a short-chain alcohol (methanol or ethanol) is converted to formance with hydrolysis in the absence of the catalyst.
monoesters and water in the presence or absence of a catalyst [14]. De
Paiva et al. [15] esterified long-chain FFAs with ethanol using zinc 2. Experimental
laurate as catalyst and obtained 92% in ethyl esters after 90 min at
165 °C. Dos Santos et al. [16] performed the esterification of different Palm oil and commercial zinc stearate (ZS) were supplied by
carboxylic acids (valeric, caprylic, lauric and oleic acids) in super- Agropalma S/A (Belém, PA, Brazil) and Sim Estearina (Araucária, PR,
critical ethanol and obtained yields of 90% in ethyl esters after 40 min. Brazil) as courtesy, respectively. Ethanol (99.8% purity) and hexane
Processing raw materials with high FFA content (above 4.0%) to (analytical grade) were purchased from Neon® (São Paulo, SP, Brazil)
alkyl monoesters can be carried out by esterification in a homogeneous and used without further purification.
acid medium (usually H2SO4), followed by transesterification in Palm oil was characterized in terms of its free fatty acid content

570
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

(AOCS method Ca 5a-40), saponification index (AOCS method Cd 400 °C, respectively. Helium was the carrier gas at 29.2 mL min−1 and
3–25), average molar mass (AOCS method Cd 3a-94) and fatty acid the injected volume was 1.0 μL. The column temperature began at 50 °C
profile by gas chromatography (see below for details). and was immediately heated to 180 °C at 15 °C min−1, then heated to
230 °C at 7 °C min−1, and finally to 380 °C at 10 °C min−1, where it
2.1. Zinc carboxylate characterization remained isothermally for 6 min. The quantification of reaction pro-
ducts (MAG, DAG and TAG) followed the EN 14105 method [30] using
X-ray diffraction (XRD) of ZS was carried out in a Shimadzu DRX- tricaprin (1,2,3-tricaprinoylglycerol) and lauric acid (dodecanoic acid)
7000 X-ray powder diffractometer using Cu Kα radiation (FFA) as the internal standards IS1 and IS2, respectively (Sigma-Al-
(λ = 1.5406 Å, 40 kV and 30 mA) in the range of 3 to 60° of 2θ and a drich).
step size of 0.02°. The basal spacings were obtained according to the For the fatty acid profiles of palm oil and ZS, an esterification
Bragg’s Law. procedure was carried out according to Menezes et al. [31]. Analyses of
Fourier transform infrared (FTIR) spectroscopy of ZS was obtained the resulting fatty acid methyl esters (FAME) were performed in a
from a Bomem MB 1000 spectrometer at room temperature in the range Shimadzu® 2010 Plus gas chromatograph using a Shimadzu SH-Rtx-
of 4000–400 cm−1, using a resolution of 4 cm−1 and 32 scans. The Wax (30 m × 0.32 mm; 0.25 μm) capillary column at a split ratio of
sample pellets were prepared by mixing the ground material with 1:10. The injector and detector temperatures were set at 250 and
spectroscopic grade KBr (Vetec) in 1:100 mass ratio. 260 °C, respectively. Helium was used as carried gas at a flow rate of
Thermogravimetry (TG) and differential scanning calorimetry (DSC) 50 mL min−1 and the injection volume was 1.0 μL. The column was
were carried out in a Mettler-Toledo TG/DTA model 851 under the programmed to start at 60 °C with a holding time of 2 min, followed by
oxygen flow of 50 mL min−1 and a heating rate of 10 °C min−1. Samples a heating rate of 10 °C min−1 to 180 °C, which was held for 1 min, and a
were placed in platinum crucibles and heated from 30 to 1000 °C. heating rate of 5 °C min−1 to 230 °C with a final holding time of 5 min.
Scanning electron microscopy (SEM; Tescan VEGA3 LMU, Czech The identification of each compound was performed by comparing their
Republic) of the catalyst was carried out before and after TAG hydro- retention times with those of FAME standards in the range of C4 to C24
lysis. Prior to analysis and to improve the material conductivity, sam- (Supelco™ 37 Component FAME Mix). Quantification was performed by
ples were gold plated in a SCD 030 flow exfoliant (Oerlikon Balzers, internal calibration according to the EN 14103 method [30] using
Liechtenstein). Different areas of the sample particles were analyzed. methyl nonadecanoate (C19:0, Sigma-Aldrich) as the reference stan-
Surface analyses were performed under vacuum using an acceleration dard.
voltage of 15.0 kV. Free glycerol (GLY) released after hydrolysis was quantified as de-
The catalytic efficiency was measured by the turnover number scribed by Bondioli et al. [32] using a UV 1100 spectrophotometer. The
(TON) (Equation 4), which represents the number of cycles that a cat- presence of free water (moisture) was determined according to the
alyst is able to perform before being deactivated, and by the turnover AOCS Ca 2e-84 method using Karl Fischer automatic coulometric ti-
frequency (TOF) (Equation 5), which represents how frequent the ob- trators (model MKC-610).
served catalytic activity was [18,27,28]. Zinc leaching to the reaction products was analyzed by inductively
coupled plasma optical emission spectrometry (ICP OES). Sample
TON = (mol of FFA/mol of catalyst) (4)
opening was performed according to the methodology described by
TOF = (TON /time) (5) Dugo et al. [33]. The analyzes were performed as described by Dos
Anjos et al. [34] using a Thermo Scientific ICAP 6500 spectrometer.
The effect of heating and cooling on the phase behavior of the re-
action system was visualized and registered using a variable volume 2.3. Reaction procedure
high pressure equilibrium cell coupled to a MicroZoom 10-200× USB
camera. TAG hydrolysis images were obtained at temperatures ranging Hydrolyses were performed in a 50 mL stainless-steel Parr batch
from 25 to 140 °C to demonstrate the processes of catalyst melting and reactor model 4597. The reagents used in ZS-catalyzed hydrolysis ex-
clustering. The reaction mixture was confined into the closed cell and periments were palm oil and distilled water. The volume occupied was
the camera was mounted next to the frontal inspection window. The always around 2/3 of the total volume of the reactor vessel. Palm oil,
system was continuously stirred using a magnetic stirrer. The tem- water and catalyst were fed into the reactor and the system was heated
perature set point (140 °C) was maintained constant using a resistive to the reaction temperature set point. The heating time was around
electric heater with a PID controller. Temperatures above 140 °C were 15 min and the reaction time was computed from the moment at which
not applied due to limitations of the equipment set up. More the temperature set point was reached. All reactions were carried out at
details about this experimental procedure can be obtained from 500 rpm.
Escorsim et al [29]. Initially, a central rotatable composite design based on a 23 factorial
experimental design with three replicates at the center point was used
2.2. Quantification of TAG, DAG, MAG, FFA, GLY, FAME and water to evaluate the effects of the following variables: palm oil-to-distilled
water molar ratio (MR), temperature (T) and catalyst concentration
Reactants and products were analyzed by gas chromatography after (CAT) in relation to the total amount of oil fed into the reactor chamber
extraction with hexane to remove water and free glycerol. Initially, (Table 1). After performing the assays and analyzing the reaction pro-
0.2 g of the reaction mixture was weighed in an Eppendorf tube and ducts, some additional trials were performed (experiments H9 to H17 in
1 mL of hexane was added. The Eppendorf tube was shaken vigorously Table 1) under more restricted experimental conditions. The kinetic
and centrifuged at 10,000 rpm for 10 min using an Eppendorf data of both auto-catalysed and ZS-catalyzed reactions were determined
Centrifuge (Marshal Scientific Model 5415C). The upper phase con- to investigate the effect of the catalyst on reaction performance. The
taining TAG, DAG, MAG and FFA was collected, the organic solvent was reaction time for all experiments was 240 min. The Statistica 7.0 soft-
evaporated, and the sample was saved for analysis. ware was used to evaluate the effect of the process variables (MR, T and
FFA and acylglycerols (MAG, DAG and TAG) were analyzed by gas CAT) on palm oil hydrolysis.
chromatography in a Shimadzu 2010 Plus chromatograph equipped
with an autosampler, a split/splitless injector at 1:10 split ratio and a 2.4. Simultaneous esterification and transesterification
flame ionization detector (FID). Analyses were performed in an Agilent
Select Biodiesel for Glycerides capillary column (15 m × 0.32 mm; The simultaneous esterification and transesterification of partially
0.10 μm). The injection and detector temperatures were 380 and hydrolyzed palm oil and its direct transesterification using ZS as

571
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

Table 1
Experimental conditions for zinc stearate-catalysed palm oil hydrolysis, ar-
A
ranged as a central rotatable composite design that was developed from a 23
factorial design with three replicates at the center point.
Experiment Conditionsa

T (°C) CAT (%) Palm oil-to-water MR B

H1 140 2 1:10
H2 190 2 1:10

Relative Intensity (u.a.)


H3 140 10 1:10
C
H4 190 10 1:10
H5 140 2 1:53
H6 190 2 1:53
H7 140 10 1:53
D
H8 190 10 1:53
H9b 131.2 6 1:31.5
H10 b 198.8 6 1:31.5
H11 b 165 6 2.4 E
H12 b 165 6 60.6
H13 b 165 0.6 1:31.5
H14 b 165 11.4 1:31.5 F
H15c 165 6 1:31.5
H16 c 165 6 1:31.5
H17c 165 6 1:31.5 2954
1730 1409 719
2916 2848 1538 1467
a
T, temperature; CAT, catalyst concentration; MR, molar ratio.
b
Central composite rotatable design. 4000 3500 3000 2000 1500 1000 500
c
Center point of the experimental design. -1
Wavenumber (cm )

catalyst were carried out for comparison. Reactions were performed in Fig. 1. FTIR spectra of zinc stearate before and after palm oil hydrolysis at
a 50 mL stainless-steel Parr batch reactor model 4597. Initially, palm oil different experimental condition: (A) ZS; (B) ZS in H25 after 8 h; (C) ZS in H26
hydrolysis was performed using a palm oil-to-water MR of 1:31.5 and after 8 h; (D) ZS in H27 after 8 h; (E) ZS in H28 after 8 h and (F) ZS in the 5th
cycle of H27 after 4 h. Conditions of each experiment are presented in Table 3.
6% catalyst at 190 °C for 4 h. After this period, the reactor was cooled,
water was removed by phase separation, methanol was added at a palm
oil hydrolysate-to-methanol MR of 1:10 and the reaction was carried 3.2. Catalyst characterization
out for another 2 h at 160 °C [18]. Palm oil hydrolysates and products
obtained after simultaneous esterification and transesterification were Fig. 1A shows the infrared spectra (FTIR) of the solid catalyst, which
collected for analysis after reaction completion. contained strong bands in the 2916–2848 cm−1 region and weak bands in
the 1467–719 cm−1 region that are typically attributed to methylene
2.5. Catalyst recovery from the reaction medium groups of fatty acid (carboxylate) hydrocarbon chains [36]. The band
centered at 2954 cm−1 is attributed to asymmetric axial deformation of
Removal of the catalyst was performed by washing the reaction terminal methyl groups. The absence of bands in the 3500–3300 cm−1
products with a 1:1 (v/v) ethanol:hexane mixture followed by cen- region (axial deformation of associated hydroxyl groups) confirms that the
trifugation and subsequent filtration on qualitative filter paper with a ZS catalyst was anhydrous. According to the literature [18,19,36], bands
grammature of 80 g m−2 [35]. This procedure was repeated at least five assigned to the carbonyl group are: ʋa(COO) at ∼1538–1531 cm−1
times. After filtration, the catalyst was dried in an air circulating oven (asymmetric stretching); ʋs(COO) at ∼1409–1394 cm−1 (symmetrical
(Solab, model SL100) at 60 °C for 12 h and reserved for chemical stretching); ʋd(C–C)COO at ∼949–957 cm−1 (axial deformation);
characterization and reuse. Also, the solvent was removed from the ʋδ(COO) at ∼ 745–743 cm−1 (angular deformation or bending); ʋτ(COO)
filtrate in a rotary evaporator (Ika, model RV 10 Digital) and the pro- at ∼580 cm−1 (out-of-plane twisting); and ʋρ(COO) at ∼550–547 cm−1
ducts were analyzed by gas chromatography for FFA, MAG, DAG and (chain oscillation).
TAG quantification using the methods mentioned above. Table 2 shows the fatty acid composition of the solid catalyst.
Stearic (50%) and palmitic (40%) acids (Table 2) corresponded to 90%
3. Results and discussion of its total fatty acid content, whose average molar mass was estimated
as 269.5 g mol−1. According to Barman and Vasudevan [22], saturated
3.1. Palm oil characterization fatty acids establish better van der Waals interactions among carbox-
ylate hydrocarbon chains resulting in solids with a better structural
The raw material (palm oil) used in this study presented a saponification stability. By contrast, when unsaturated fatty acids are present in levels
index of 195.39 ± 2.08 mg KOH g−1, an average molar masses (MM) of above 30%, the layered structure of metal carboxylates is weakened,
859.65 ± 2.62 g mol−1 for triacylglycerols and 273.87 ± 1.12 g mol−1 leading to less stable aggregates in which their original all-trans con-
for free fatty acids, and an acidity of 0.15 ± 0.02 g oleic acid 100 g−1 of formation is partially lost.
sample, which corresponds to 0.30 ± 0.03% in FFA. The FFA profile Fig. 2A shows the XRD for the ZS catalyst. The diffractogram was
contained 0.97 ± 0.01 wt% myristic acid (C14:0), 40.59 ± 0.03 wt% typical of layered carboxylates with basal peaks in the 3–18° region (in
palmitic acid (C16:0), 4.85 ± 0.03 wt% stearic acid (C18:0), 2θ) whose calculated basal spacing was 40.60 Å. The wide peaks ob-
43.90 ± 0.03 wt% oleic acid (C18:1), 0.20 ± 0.03 wt% linoleic acid served in the region of 18-25° of 2θ were attributed to the structural
(C18:2), and 9.11 ± 0.01 wt% arachidic acid (C20:0). Therefore, disorder of palmitate and stearate anions that were intercalated be-
palmitic and oleic acids corresponded to 84.5% of the total palm oil tween the lamellae.
fatty acid content. TAG, DAG, MAG and FFA contents in palm The thermal analyses (TG/DSC) of the catalyst revealed the pre-
oil were 98.16 ± 0.05 wt%, 1.07 ± 0.01 wt%, 0.47 ± 0.02 wt% and sence of an exothermic peak in the region of 135 °C (Fig. 3A), which
0.30 ± 0.01 wt%, respectively. was attributed to the melting of zinc stearate caused by rearrangement

572
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

Table 2
Zinc stearate (ZS) chemical composition before and after its use in palm oil hydrolysis at different reaction conditions (see Table 3 for details).
Fatty acid Cn:ma ZS ZS after use in:

H19 after 8 h H20 after 8 h H21 after 8 h H3 after 8 h H21 after 4 h, 5th cycle

Lauric C12:0 0.03 ± 0.01 – – – – –


Myristic C14:0 1.97 ± 0.02 1.00 ± 0.01 1.06 ± 0.01 1.08 ± 0.02 0.99 ± 0.01 1.42 ± 0.01
Pentadecanoic C15:0 0.41 ± 0.01 0.33 ± 0.02 0.42 ± 0.01 0.49 ± 0.01 0.35 ± 0.01 0.21 ± 0.01
Palmitic C16:0 39.86 ± 0.12 82.24 ± 0.13 80.74 ± 0.11 76.81 ± 0.10 58.53 ± 0.12 70.54 ± 0.14
Heptadecanoic C17:0 1.28 ± 0.03 – – – 0.27 ± 0.01 –
Stearic C18:0 50.18 ± 0.21 11.85 ± 0.05 10.52 ± 0.04 10.63 ± 0.04 18.29 ± 0.05 12.44 ± 0.03
Arachidic C20:0 – – 0.60 ± 0.01 0.88 ± 0.01 1.47 ± 0.01 1.78 ± 0.02
Myristoleic C14:1 0.15 ± 0.01 – – – 0.09 ± 0.01 –
Palmitoleic C16:1 0.11 ± 0.01 0.27 ± 0.01 – – 0.26 ± 0.01 0.21 ± 0.01
Heptadecanoic C17:1 0.08 ± 0.01 – – – – –
Oleic C18:1 2.77 ± 0.03 4.30 ± 0.03 6.40 ± 0.05 6.22 ± 0.04 18.32 ± 0.06 12.62 ± 0.04
Elaidic C18:1 1.89 ± 0.02 – – 0.20 ± 0.01 1.25 ± 0.02 0.34 ± 0.01
Linolenic C18:3 0.62 ± 0.01 – 0.25 ± 0.01 – 0.13 ± 0.01 0.40 ± 0.01

a
n is the number of carbon atoms in the hydrocarbon chain and m is the number of the existing double bonds.

Relative Intensity (u.a.)


B

B
C
C
Intensity (U.A.)

D
E
D
TG F

0 100 200 300 400 500 600 700 800 900 1000
E
Temperature (°C)

B
F
Relative Intensity (u.a.)

5 10 15 20 25 30 35 40 45 50 55 60
C
Angle 2θ

Fig. 2. X-ray diffraction analysis (XRD) of zinc stearate (ZS) before and after D
palm oil hydrolysis at different experimental condition: (A) ZS; (B) ZS in H25 E
after 8 h; (C) ZS in H26 after 8 h; (D) ZS in H27 after 8 h; (E) ZS in H28 after 8 h
F
and (F) ZS in the 5th cycle of H27 after 4 h. Conditions of each experiment are
presented in Table 3.
Exo
Endo
DSC
of molecules in the liquid phase, while the presence of endothermic
peaks was associated to the burning of organic matter (fatty acid car- 0 100 200 300 400 500 600 700 800 900 1000
boxylates). Thermal analysis gave also the amount of fixed residue Temperature (°C)
(ZnO) and organic matter that were present in the ZS catalyst and these
Fig. 3. TG and DSC profile of zinc stearate before and after palm oil hydrolysis:
were 17.2% and 82.8%, respectively (Fig. 3A). Then, the amount of (A) ZS; (B) ZS in H25 after 8 h; (C) ZS in H26 after 8 h; (D) ZS in H27 after 8 h;
fixed residue was used to calculate the zinc content of ZS, which cor- (E) ZS in H28 after 8 h and (F) ZS in the 5th cycle of H27 after 4 h. Conditions of
responded to 13.8% (138 g kg−1). This value was close to that found in each experiment are presented in Table 3.
the literature for freshly synthesized ZS (13.5%) [37].

3.3. Hydrolysis
of 2.60% was obtained at 190 °C using a palm oil-to-water MR of 1:53,
Hydrolysis of refined palm oil produced a blend of FFA, MAG, DAG, while the smallest was 0.53% at 140 °C using a palm oil-to-water MR of
TAG and GLY. Table 3 shows the results for reactions that were carried 1:10. This clearly demonstrates that ZS was catalytic active in palm oil
out at the experimental conditions described in Table 1. The results hydrolysis. The highest FFA content were observed in experiments H6
obtained for the thermal conversion of palm oil to its corresponding (45.8%), H10 (47.7%), H12 (59.2%) and H15–17 (47.7%) (Table 3).
FFA are observed in rows H18–H24 of Table 3. The highest FFA content The actual gains in relation to the highest thermal conversion (Table 3,

573
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

Table 3
Palm oil hydrolysis with and without zinc stearate (ZS) as catalyst and its corresponding catalytic efficiency (TON and TOF).
Exp. T CAT Palm oil-to-water MR Time P Component (wt%)a TONb TOFb
(°C) (%) (h) (bar) (h−1)
TAG DAG MAG FFA GLY

H1 140 2 1:10 4 1.2 64.89 21.23 4.78 8.23 0.87 39.6 9.9
H2 190 2 1:10 4 1.8 45.42 26.66 10.23 15.93 1.75 72.6 18.2
H3 140 10 1:10 4 1.2 33.35 21.41 15.37 26.94 2.93 24.5 6.1
H4 190 10 1:10 4 1.8 31.97 34.35 12.41 19.22 2.05 17.6 4.4
H5 140 2 1:53 4 3.5 37.05 30.88 15.94 14.54 1.60 64.3 16.1
H6 190 2 1:53 4 3.7 12.14 23.83 13.09 45.85 5.09 204.0 51.0
H7 140 10 1:53 4 3.5 24.03 18.88 17.06 35.92 4.11 33.5 8.4
H8 190 10 1:53 4 3.7 17.19 18.95 20.18 39.34 4.34 36.2 9.0
H9c 131.2 6 1:31.5 4 0.3 37.87 21.02 14.00 26.70 0.11 41.1 10.3
H10c 198.8 6 1:31.5 4 3.9 11.59 22.74 15.66 47.70 2.03 73.2 18.3
H11c 165 6 2.4 4 1.5 24.10 26.08 12.29 36.11 1.17 520.4 130.1
H12c 165 6 60.6 4 4.2 1.46 17.87 18.10 59.24 3.02 48.0 12.0
H13c 165 0.6 1:31.5 4 2.7 50.43 21.98 9.63 17.76 – 27.2 6.8
H14c 165 11.4 1:31.5 4 2.8 19.43 19.55 17.37 42.13 1.30 64.5 16.1
H15d 165 6 1:31.5 4 2.8 9.60 19.81 17.37 47.79 5.43 70.3 17.6
H16d 165 6 1:31.5 4 2.8 9.80 19.44 17.13 48.32 5.31 71.5 17.9
H17d 165 6 1:31.5 4 2.9 9.65 20.57 17.19 47.12 5.47 70.7 17.7
H18e 140 – 1:10 4 1.3 93.03 4.61 1.78 0.53 0.04 – –
H19e 190 – 1:10 4 1.9 91.02 5.41 1.99 1.42 0.16 – –
H20e 140 – 1:53 4 3.5 90.43 5.91 2.33 1.20 0.13 – –
H21e 190 – 1:53 4 3.8 83.63 9.24 4.20 2.60 0.32 – –
H22e 165 – 1:31.5 4 2.8 92.19 4.61 1.98 1.11 0.11 – –
H23e 165 – 1:31.5 4 2.7 92.43 4.30 2.00 1.12 0.15 – –
H24e 165 – 1:31.5 4 2.9 92.31 4.46 1.99 1.12 0.13 – –
H25f 140 6 1:31.5 8 1.5 14.83 17.80 15.70 46.48 2.57 – –
H26f 165 6 1:31.5 8 3.2 8.11 16.23 13.43 55.60 4.12 – –
H27f 190 6 1:31.5 8 4.1 2.93 29.15 16.57 46.15 1.92 – –
H28g 140 10 1:10 8 1.2 21.26 18.02 12.40 43.49 2.72 – –

a
TAG. triacylglycerol; DAG. diacylglycerol; MAG. monoacylglycerol; FFA. free fatty acids; GLY. Glycerol.
b
TON. turnover number; TOF. turnover frequency.
c
Central composite rotatable design.
d
Center point of the experimental design.
e
Experimental design without addition of the ZS catalyst.
f
Center point with temperature variation for a reaction time of 8 h.
g
H3 experiment performed by 8 h.

experiment H21) were 43.2%, 45.1%, 56.6% and 45.1% points percent, response function (FFA content) of the factorial design,
respectively.
FFA content at the center point of the experimental design had a FFA (%) = 454.0679 + 5.1629·T 0.0147· T 2 0.5432· MR
very low standard deviation of 1.12% and this demonstrated the good 0.0058·MR2 + 19.9667· CAT 0.7743· CAT 2
repeatability of the experimental procedures (Table 3). The FFA content (6)
+ 0.0079·T · MR 0.0562· T ·CAT
described in Table 3 were fitted to a quadratic model (Equation 6)
whose statistical analysis showed high linear regressions and small lack where FFA (%) is the content of free fatty acids, T is the reaction
of fit compared to Fisher F standard values at a 95% confidence level temperature (°C), MR the palm oil-to-water MR and CAT the catalyst
(Table 4). Maximum explained variances of 91% were obtained for both concentration (%) in relation to the amount of fatty material used for
palm oil hydrolysis and FFA contents and the regression factors were conversion. The linear effects of T, palm oil-to-water MR and CAT were
significant in relation to the residues. The lack of fit in relation to the statistically significant and had a positive influence on FFA content
pure error was also acceptable among the ANOVA data, with a good while the quadratic effects influenced it negatively (Eq. 6). The binary
dispersion of the residues in relation to the predicted answers, therefore interaction between T and MR had also a positive effect on FFA content,
demonstrating that the model was predictive and not biased but the interaction between T and CAT was negative. By contrast, no
(Supplementary Material – Fig. S2). significant effect was observed from the interaction between MR and
Eq. (6) describes the quadratic model that was developed to fit the CAT (Supplementary Material – Fig. S1).
The quality of the ANOVA data enabled the construction of the re-
Table 4 sponse surfaces (SDR) that are presented in Fig. 4. The non-linear aspect
Analysis of variance (ANOVA) and F-test for palm oil hydrolysis using zinc of the response surfaces (Fig. 4) confirmed the significance of the
stearate (ZS) as catalyst. second-order effects. With this, a broad region was defined to achieve
high FFA content of palm oil hydrolysis using a wide range of reaction
Source SS df MS F Ftab
conditions involving T, MR and CAT.
Regression 3555.0 8 444.4 90.4 3.44 Micic [9] studied the hydrolysis of sunflower oil under subcritical
Residual 39.4 8 4.9 conditions without the use of catalyst and observed a 1.8-fold increase
Lack of fit 38.6 6 6.4 17.8 19.33 in FFA content (35 for 63% at 200 °C) by increasing the palm oil-to-
Purr error 0.72 2 0.4
Total 3594.4 16
water MR from 1:5 to 1:25. Kusdiana and Saka [10] observed a 1.4-fold
increase in FFA content (62 to 85% at 270 °C) during hydrolysis of
SS, sum of squares; df, degree of freedom; MS, mean square; R-sq = 0.9557 and rapeseed oil under subcritical conditions without the presence of an
%Var = 0.9113. added catalyst when the rapeseed oil-to-water MR was increased from

574
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

14.54 to 45.85% as observed in experiments H5 and H6. Micic [9],


Pinto and Lanças [38] and Moquin and Temelli [2] observed also a
significant increase in TAG conversion to FFA when T was increased
from 200 to 250 °C, with FFA contents increasing from 65 to 95%.
A possible mechanism for the catalytic activity of zinc stearate in
palm oil hydrolysis is the initial interaction of the electron pair of the
carbonyl oxygen of the fatty acids with the metal present in the car-
boxylate, leading to the adsorption of the molecule. Thus, the metal
acting as a Lewis acid increases in the positive charge density on the
carbonyl carbon atom favoring the nucleophilic attack of the electron
pair of the water oxygen, with the consequent formation of a tetra-
hedral intermediate. Therefore, hydrogen exchanges between the two
oxygen atoms loses one DAG molecule, while the FFA formed remains
on the surface of the catalyst until desorption occurs. With this, the
catalyst active site is released for another reaction cycle (see a proposal
for the reaction mechanism in Supplementary Material – Fig. S3)
[18,19,39].
The best conditions for palm oil hydrolysis were obtained in the
center point of the experimental design (165 °C, 6% ZS and a palm oil-
to-water MR of 1:31.5), where a maximum FFA (50%) and a minimum
TAG (10%) contents were observed. This reaction condition was used to
evaluate the reaction kinetics with and without the presence of ZS as
catalyst. Furthermore, the three replicates that were performed at the
center point of the experimental design (H15, H16 and H17) indicated
the good repeatability of the experiments, since the resulting FFA
contents presented a standard deviation of only 0.60%.
Fig. 5A and 5B show the kinetic profile of ZS-catalyzed palm oil
hydrolysis at 165 and 190 °C, respectively, using a palm oil-to-water
MR of 1:31.5 and a CAT of 6%. After 8 h of hydrolysis, the FFA content

Fig. 4. Response surfaces relative to FFA content after palm oil hydrolysis. (A)
temperature (T) vs. molar ratio oil: water (MR), (B) temperature (T) vs. catalyst
content (%CAT) and (C) palm oil-to-water molar ratio (MR) vs. catalyst con-
centration (CAT).

1:13 to 1:54. Rocha [5] used castor oil and soybean oil and observed an
1.5-fold increase in FFA content (49 for 75% at 300 °C) with an increase
in the oil-to-water MR from 1:5 to 1:20 using 20% of niobium oxide as
catalyst. Hence, increased water concentrations in the reaction mixture
shift the reaction equilibrium towards products formation. In our study,
changes in the palm oil-to-water MR from 1:10 to 1:53 provided the
highest recovery of FFA (15.93 for 45.85% in experiments H2 and H6 of
Table 3) giving a 2.9-fold increase in FFA content. Fig. 5. Kinetics of palm oil hydrolysis catalyzed by zinc stearate under the
An increase in reaction temperature provides higher solubilities of following reaction conditions: (A) 165 °C, palm oil-to-water molar ratio of
both non-polar (oil) and polar (water) reactants due to the decrease of 1:31.5 and 6% catalyst, (B) 190 °C, palm oil-to-water molar ratio of 1:31.5 and
the dielectric constant of the latter, generating an increase in the re- 6% catalyst. Diamonds for free fatty acids (FFA), squares for monoacylglycerols
action rate [9]. Indeed, Table 3 shows that by increasing the tem- (MAG), asterisks for diacylglycerols (DAG), triangles for triacylglycerols (TAG)
perature from 140 to 190 °C, the FFA content was also increased from and circles for glycerol (GLY).

575
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

reached 55% at 165 °C and by increasing the reaction temperature from 3.5. Simultaneous esterification and transesterification
165 to 190 °C, the same FFA content was already reached at 4 h,
therefore reducing the reaction time by half. Also, in Fig. 5B, a decrease Table 6 shows the results for simultaneous esterification and
in the FFA content from 60 to 50% was observed between 4 and 8 h, transesterification of partially hydrolyzed palm oil and its direct
along with a slight increase in the acylglycerol contents (MAG, DAG and transesterification using ZS as catalyst. As observed, there was a 99%
TAG). According to Macierzanka and Szelag [20], this is due to the reduction in TAG, 92% in DAG, 85% in MAG and 89% in the FFA
esterification of glycerol (also demonstrated by the drop in GLY con- content of pre-hydrolyzed palm oil for a total ester content of 89%.
centration in the reaction media), which is favored at temperatures Direct transesterification of refined palm oil was also performed under
above 150 °C in the presence of the ZS catalyst. In addition, glycerol the same conditions used for its acid hydrolysis and the results in-
acylation may have occurred after 1 h (Fig. 5A and 5B) as suggested by dicated 85% TAG conversion for a product with a total ester content of
the observed decrease in GLY with the concomitant increase in DAG. 60%. Therefore, hydrolysis followed by simultaneous esterification and
Gomes and Cordeiro [40] demonstrated that ZS exhibits good catalytic transesterification improved the total ester content from palm oil by
activity in the breakdown of TAG by glycerol, however, this reaction 48% compared to its direct transesterification. Kusdiana and Saka [10]
was not observed at the conditions applied in this study for palm oil reported similar trends for the transesterification of rapeseed oil and the
hydrolysis. esterification of hydrolyzed rapeseed oil in supercritical methanol,
The hydrolysis kinetic data were used to validate the mathematical whose FAME yields were 60 and 98%, respectively. However, reactions
model presented in Equation 6. Two independent experimental proce- were carried out for 20 min at 270 °C and 17 MPa with an oil-to-me-
dures were carried out for this purpose. The model prediction for thanol MR of 1:42.
165 °C, palm oil-to-water MR of 1:31.5 and CAT of 6% after 4 h were The moisture content of the resulting FAMEs was 0.30 ± 0.03%
51.5 ± 2.9% while the FFA content obtained experimentally was after hydrolysis and 1.47 ± 0.05% after simultaneous esterification
48.6%. Also, the model predicted an FFA content of 47.4 ± 2.9% for and transesterification. Such increase in water content is associated
reactions carried out at 190 °C with a palm oil-to-water MR of 1:31.5 with the esterification of FFA, where water is obtained as co-product.
and 6% of CAT after 4 h, while the result obtained in a confirmatory The presence of water causes TAG hydrolysis, leading to the release of
experiment was 53.2 ± 1.3% (Fig. 5A and 5B). These data demon- FFAs that are rapidly converted to FAMEs. This observation was also
strated the validity of the proposed mathematical model for predicting confirmed by Kusdiana and Saka [41] as a result of the transester-
FFA contents derived from ZS-catalysed palm oil hydrolysis. ification of rapeseed oil in the presence of water using supercritical
The catalytic effect was assessed by performing palm oil hydrolysis methanol.
at the same conditions in the absence (thermal conversion) and pre- The crude glycerin phase from experiment H27 was analyzed for the
sence of the ZS catalyst. Low FFA, GLY, MAG and DAG contents where quantification of glycerol using the procedure described by Bondioli
obtained by thermal conversion (Supplementary Material – Fig. S4), et al. [32]. Samples were obtained from hydrolysis of palm oil followed
compared to the results of the catalytic reaction (Fig. 5). Also, the by simultaneous esterification and transesterification of palm oil hy-
significant decrease in TAG after 8 h of thermal conversion was asso- drolysate, as well as from the direct transesterification of palm oil. In
ciated to auto-hydrolysis because the first FFA molecules released in the both situations, analyses were carried out after removal of water and
medium were able to act as an acid catalyst [7]. methanol by evaporation under low pressure. The total mass recovery
Another observation from the experiments performed in this work of glycerol after hydrolysis followed by simultaneous esterification and
refers to the pressure reached in the reaction media, which remained transesterification was 77.0 ± 3.1% (47.4 and 29.6%, respectively),
low at values around the vapor pressure of water. Also, the efficiency of whereas for direct transesterification, this value reached 90.8 ± 4.5%.
the ZS catalyst was demonstrated by the poor reaction performance in Glycerol losses were observed in both situations because phase se-
its absence using otherwise identical reaction conditions. paration and washing stages were not optimized for its highest re-
covery. However, the FTIR spectra of these crude glycerin phases were
practically identical to the results obtained from pure glycerol, there-
3.4. Catalyst reuse assays fore evidencing the absence of symmetrical and asymmetric axial de-
formations of carboxylate anions (Supplementary Material – Fig. S5). A
Table 5 shows the catalyst recovery and reuse based on a single crude glycerin phase with high glycerol content (around 86%) was also
batch of ZS using reaction conditions in which the best FFA contents obtained by Ramos et al. [18] after transesterification of palm oil with
were achieved (190 °C, 1: 31.5 palm oil-to-water MR and 6% catalyst methanol using zinc carboxylates (zinc laurate) as a catalyst.
for 4 h). Despite the already mentioned conversion of zinc stearate to In general, zinc stearate showed good FFA contents in palm oil
zinc palmitate, the solids maintained the same catalytic efficiency in hydrolysis. Also, high FAME contents were obtained by hydrolysis fol-
five consecutive reaction cycles. Also, even with an increase in lamellar lowed by simultaneous esterification and transesterification. These re-
disorganization (characterized by XRD) and changes in its melting point sults demonstrate the potential and flexibility of zinc carboxylates for
(noted by DSC), no segregation of any crystalline impurity was ob- biodiesel production because, by applying this multifunctional catalytic
served. system, hydrolysis, esterification and transesterification would be

Table 5
Catalyst reuse at 190 °C for 4 h using 6% zinc stearate and a palm oil-to-water molar ratio of 1:31.5 (experiment H21).
Exp. Component (wt%)*

TAG DAG MAG FFA GLY

Initial reaction 2.05 ± 0.01 23.97 ± 0.13 18.82 ± 0.06 58.44 ± 0.48 6.63 ± 0.08
1st reuse 2.13 ± 0.06 23.97 ± 0.05 15.98 ± 0.28 57.69 ± 0.74 6.48 ± 0.07
2nd reuse 2.16 ± 0.10 24.43 ± 0.08 16.24 ± 0.49 57.46 ± 0.30 6.41 ± 0.10
3rd reuse 2.75 ± 0.19 23.73 ± 0.84 17.25 ± 0.28 56.73 ± 0.30 6.30 ± 0.02
4th reuse 2.12 ± 0.14 22.76 ± 0.21 18.06 ± 0.38 57.19 ± 0.37 6.23 ± 0.15
5th reuse 2.51 ± 0.20 23.13 ± 0.24 17.86 ± 0.35 58.02 ± 0.34 6.42 ± 0.02

* TAG, triacylglycerol; DAG, diacylglycerol; MAG, monoacylglycerol; FFA, free fatty acids; GLY, glycerol.

576
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

Table 6
Palm oil conversion by hydrolysis, hydrolysis followed by simultaneous esterification and transesterification, and direct transesterification of palm oil.
Experiment Components (wt%)1

TAG DAG MAG FFA GLY FAME

2
Hydrolysis 1.98 ± 0.06 22.12 ± 1.25 13.25 ± 0.28 56.01 ± 1.43 6.64 ± 0.09 –
Hydrolysis followed by esterification and transesterification3 0.01 ± 0.00 1.75 ± 0.17 1.93 ± 0.03 6.13 ± 0.48 3.38 ± 0.23 88.95 ± 0.70
Transesterification4 13.94 ± 0.27 9.23 ± 0.37 13.61 ± 0.27 0.49 ± 0.02 6.05 ± 0.20 57.04 ± 0.11

1
TAG, triacylglycerol; DAG, diacylglycerol; MAG, monoacylglycerol; FFA, free fatty acids; GLY, glycerol.
2
Hydrolysis at 190 °C for 4 h using 6% zinc stearate and an oil-to-water molar ratio of 1:31.5 (experiment H21).
3
Simultaneous esterification and transesterification of palm oil hydrolysates (experiment 21) at 160 °C for 4 h using 6% zinc stearate and an oil-to-methanol molar
ratio of 1:10.
4
Direct transesterification of palm oil at 160 °C for 4 h using 6% zinc stearate and an oil-to-methanol molar ratio of 1:10.

eventually feasible in three relatively simple reaction steps (hydrolysis, scattering. After hydrolysis, the catalyst assumed an intermediate basal
water withdrawal and simultaneous esterification and transesterifica- spacing between zinc palmitate (38.36 Å) and zinc stearate (42.67 Å)
tion). Also, the employment of zinc stearate allows the processing of and this variation was associated to its restructuring process [36].
much more complex feedstocks composed by different levels of FFAs, Hence, the slight migration to higher values of 2θ corresponded to a
MAGs, DAGs, TAGs and H2O, combining the use of low cost raw ma- decrease in basal spacing that is associated to the replacement of
terials with the good reuse capacity of a cheap and commercially stearate anions by shorter chain length anions such as palmitate. As
available catalytic system. In addition, the reactions occur at the vapor mentioned above and also demonstrated in the literature, palmitic acid
pressure of the solvents, enabling the use of hydroesterification reaction is the main component of palm oil and its acid hydrolysate [18,36].
systems without a significant change in the reactor design. After palm oil hydrolysis, the TG/DSC profiles of the recovered
solids (Fig. 3B to 3F) were also identical to the profile of the original
3.6. Characterization of the catalyst after the hydrolysis of palm oil structure (Fig. 3A). The TG/DSC data also provided the amount of fixed
residue (ZnO) and organic matter that were present in the recovered
The solids were characterized by FTIR and XRD after the hydrolysis solids and these values were 18.9 and 81.1% (Fig. 3B), 15.2 and 84.8%
of palm oil. Fig. 1B to 1F show the FTIR spectra of the solid catalysts (Fig. 3C), 16.1 and 83.9% (Fig. 3D). 12.8 and 87.2% (Fig. 3E) and 11.6
that were recovered after hydrolysis of refined palm oil. In general, and 88.4% (Fig. 3F), respectively. Then, the amount of fixed residue
these FTIR spectra were identical to that of the original structure was used to calculate the zinc content of these materials, which cor-
(Fig. 1A) except for the 1730 cm−1 band which, according to Taylor responded to 13.8% for ZS and 15.2%, 12.2%, 12.9%, 10.3% and 9.3%
and Ellis [36], are attributed to the presence of FFA (hydrolysis pro- for experiments H25, H26, H27, H28 and H27 (after five reuse cycles),
ducts) between the structural lamellae. respectively. These values were close to those found in the literature,
Table 2 shows the fatty acid composition of the ZS catalyst after which are 9.9% and 13.5% for zinc palmitate and zinc stearate, re-
hydrolysis. Changes in the chemical composition of the solid catalyst spectively [37,42]. The values found for recovered solids averaged
were observed mainly in relation to the presence of saturated FFA. The 12%, which are lower than the theoretical values reported by Gönen
concentration of stearic and palmitic acids, initially at 40:50 in ZS, et al. [37]. This decrease, according Taylor and Ellis [36], is due to
changed to 12:82, 11:81, 11:77 and 18:59 after experiments H25, H26, differences in the composition of the carboxylate hydrocarbon chains
H27 and H28, respectively, and to 12:71 after five complete reuse cy- (Table 2) and also to the presence of FFAs in-between the zinc car-
cles (experiment H27). These changes in chemical composition are re- boxylate lamellae, which was confirmed by the appearance of C]O
lated to the rearrangement of the zinc carboxylate anions during hy- axial deformations in the FTIR spectra of Fig. 1B to 1F. As mentioned
drolysis. This is explained by the ZS melting at 135 °C (see DSC profile above, this trend increased the organic matter and reduced the zinc
in Fig. 3A), leading to a structural derangement that led to a shift in the content of the recovered solids. In addition to the presence of FFA, mass
coordination between carboxylate anions and zinc atoms, moving from losses of solid catalyst may have occurred during water washing and
a bidentate to a monodentate bridge. With cooling, the solid catalyst some of the material may have remained adhered to the filter paper
reorganizes, and the bidentate bridge is restored while palmitic acid after filtration.
becomes the predominant fatty acid in the solid catalyst chemical The contribution of unsaturated fatty acids (mainly oleic and lino-
composition. leic acids) increased in the ZS composition after palm oil hydrolysis,
The reorganization and reconstitution of the zinc carboxylate reaching more than 15% of its dry mass (Table 2). This result led to a
layered structure does not follow a specific rule. However, saturated small reduction of the layered carboxylate melting point as observed by
fatty acids build more stable metal carboxylates due to their chain DSC in Fig. 3E and 3F. This reduction in melting point is caused by a
linearity and greater association by Van der Waals interactions. On the distortion of the all-trans (zig-zag) conformation of zinc carboxylates,
basis of their corresponding fatty acid profile, the average molar mass which is due to the stereochemical arrangement of the double bonds. In
of the recovered solid catalysts were in the range of 260–266 g mol−1 general, layered carboxylates melt when approximately 30% of their
(261.9 g mol−1 for H19, 260.9 g mol−1 for H20 and H21, all-trans conformation is lost [36]. Thus, the presence of unsaturated
266.0 g mol−1 for H22 and 262.4 g mol−1 for H21-4 h), compared to fatty acids during restructuration changes the linearity of the carbox-
269.5 g mol−1 for the starting material (ZS). ylate chains that are coordinated to the zinc atoms, resulting in a lower
Fig. 2B to F shows the XRD for the ZS catalyst after palm oil hy- crystallinity and a lower aggregate stability [22].
drolysis. The diffractograms were all typical of layered carboxylates Fig. 6 shows the beginning of palm oil hydrolysis at 25 °C, when
with basal peaks in the 3–18° region (in 2θ) and with the following water, palm oil and ZS phases are observed in the system. Melting of the
calculated basal spacing: 38.67, 38.62, 38.63, 39.12 and 38.77 Å for ZS catalyst started at 135 °C with its complete fusion at 140 °C. There-
experiments H19, H20, H21, H3 and H21, respectively. fore, zinc carboxylates behaved as a homogeneous catalyst during TAG
The observed peaks in the region of 18-25° of 2θ were attributed to hydrolysis; consequently, their catalytic activity is independent of the
the structural disorder of palmitate and stearate anions that were in- surface area of ZS solids originally added to the reaction mixture.
tercalated in-between the lamellae, in addition to some X-ray Cooling of the reaction mixture from 130 to 110 °C led to the formation

577
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

Fig. 6. Representative photographs of zinc stearate-catalyzed palm oil hydrolysis starting at 25 °C until the complete melting of the ZS catalyst at 140 °C, followed by
cooling after hydrolysis for 2 h to observe the gradual restructuration (clustering) of ZS.

of zinc carboxylate clusters from the border to the center (mist) of the observed, but not as much as that observed for an increase in the re-
equilibrium cell. With this, the layered structure of zinc carboxylates action temperature. The molar masses adopted for these calculations
was restored as already observed by other authors [17–19,35] during were 604.4 g mol−1 for ZS and 273.9 g mol−1 for palm oil FFAs.
the esterification of lauric acid using different metal laurates as cata- TOF values indicate the number of reaction cycles at one single
lysts. catalytically active center per unit of time (Table 3). TOF calculations
The crystal morphology of a ZS sample before and after palm oil for ZS indicated an excellent activity and fast regeneration of the cat-
hydrolysis (Supplementary Material – Fig. S6). ZS exhibited crystals in alytic site. In general, both TON and TOF values exhibited the same
the form of randomly arranged plates and/or sheets. After melting, reaction behavior, that is, more catalytic cycles were observed with an
there was a change in crystal morphology in relation to starting ma- increase in temperature and oil-to-water MR and a decrease in catalyst
terial. Larger aggregates in which the lamellae were better organized concentration.
and neatly packed. According to Gönem et al. [37], melting was better The TON and TOF values for the simultaneous esterification and
than precipitation for the synthesis of ZS solids. transesterification reactions of the palm oil hydrolysate were 85.1 and
Tests of zinc leaching were carried out after the ZS-catalyzed con- 42.5 h−1, respectively. The values of TON and TOF determined for the
version of palm oil to FAMEs. The zinc content in palm oil hydrolysates transesterification of palm oil were 60.1 and 30.3 h−1, respectively. The
varied from 38 to 60 mg kg−1, which corresponded to 0.04% of the molar masses adopted for these calculations were 604.4 g mol−1 for ZS,
total zinc content of the ZS catalyst. Lower zinc contents of 13 and 344.6 g mol−1 for palm oil hydrolysate and 859.7 g mol−1 for palm oil.
15 mg kg−1 were obtained after simultaneous esterification and trans- As noted, the ZS presented high TON values and such behavior de-
esterification of palm oil hydrolysates and after transesterification of monstrates the good regeneration capacity of these carboxylates in si-
refined palm oil, respectively, corresponding to less than 0.01% of the multaneous esterification and transesterification as well as in transes-
total zinc content of the ZS catalyst. Therefore, zinc leaching to the terification reactions. The high TOF values indicate excellent activity
reaction products was low, ranging from 0.01 to 0.04% in all cases. and rapid regeneration of the catalytic site.
There is no regulation for zinc content in biodiesel. However, the
maximum content of alkaline (sodium or potassium) and alkaline earth 4. Conclusions
metals (calcium and magnesium) is 5.0 mg kg−1 each. Hence, if these
values were extended to zinc, further washing steps would be required The potential of ZS as catalyst for producing MAG, DAG and FFA
to reach the current specifications for alkaline and alkaline earth metals from palm oil was demonstrated. Hydrolysis occurred under subcritical
of internationally accepted biodiesel standards such as EN 14214 and water, short reaction times and low pressures while its selectivity for
ASTM 6751. acylglycerols (DAG and MAG) or FFA could be adjusted by changes in
process variables such as temperature, palm oil-to-water MR and catalyst
3.7. Catalytic efficiency in palm oil hydrolysis concentration. The lamellar structure of the catalyst remained after its
use but its fatty acid composition changed by intercalation of a higher
The ZS catalyst presented high TON values (Table 3) and such be- percentage of saturated fatty acids. In addition, the high turnover fre-
havior demonstrates the good regeneration capacity of these carbox- quency of the system proved its great stability and capability to maintain
ylates, with no evidence of structural alterations after the first stage of its catalytic performance in several consecutive reaction cycles.
restructuration. Also, TON values increased with an increase in the Hydrolysis followed by simultaneous esterification and transester-
reaction temperature, therefore enhancing the ability of the catalyst to ification without removal of the catalyst afforded a 48% increase in
convert more TAG molecules to FFA. However, at the same tempera- FAME content compared to direct transesterification using the same
ture, TON values decreased by increasing the percentage of catalyst in catalytic system. Thus, zinc carboxylates had an effective performance
the reaction system because less interactions were expected to occur in hydrolysis as well as in both esterification and transesterification of
between reactants and a single catalytic site. By contrast, with an in- palm oil. In addition to the multiple applications demonstrated herein,
crease in the oil-to-water MR while keeping the other reaction para- zinc carboxylates are inert, relatively inexpensive, safe and commer-
meters unchanged, a slight positive influence in TON values was cially available solid catalysts.

578
A.M. Escorsim, et al. Fuel 244 (2019) 569–579

Acknowledgments transesterification of triglycerides catalyzed by zinc carboxylates. J Mol Catal A


Chem 2013;377:29–41.
[20] Macierzanka A, Szelag H. Esterification kinetics of glycerol with fatty acids in the
The authors are grateful to CNPq (Grants 306920/2013-1, 558836/ presence of zinc carboxylates: preparation of modified acylglycerol emulsifiers. Ind
2010-0 and 406737/2013-4) for the financial support to carry out this Eng Chem Res 2004;43:7744–53.
study. This work was also financed in part by the Coordenação de [21] Nielsen RB, Kongshaug KO, Fjellvåg H. Delamination, synthesis, crystal structure
and thermal properties of the layered metal-organic compound Zn(C12H14O4). J
Aperfeiçoamento de Pessoal de Nível Superior – Brasil (CAPES) – Mater Chem 2008;18:1002–7.
Finance Code 001. [22] Barman S, Vasudevan S. Mixed saturated-unsaturated alkyl-chain assemblies: Solid
solutions of zinc stearate and zinc oleate. J Phys Chem B 2007;111:5212–7.
[23] d’Avila Cavalcanti-Oliveira E, da Silva PR, Ramos AP, Aranda DAG, Freire DMG.
Appendix A. Supplementary data Study of soybean oil hydrolysis catalyzed by thermomyces lanuginosus lipase and
its application to biodiesel production via hydroesterification. Enzyme Res 2011,
Supplementary data to this article can be found online at https:// 2010,:618–92.
[24] Mohammed NI, Kabbashi NA, Alam MZ, Mirghani MES. Esterification of Jatropha
doi.org/10.1016/j.fuel.2019.01.178.
curcas hydrolysate using powdered niobic acid catalyst. J Taiwan Inst Chem Eng
2015;63:243–9.
References [25] Brandão RF, Quirino RL, Mello VM, Tavares AP, Peres AC, Guinhos F, et al.
Synthesis, characterization and use of Nb2O5 based catalysts in producing biofuels
by transesterification, esterification and pyrolysis. J Braz Chem Soc
[1] Hamerski F, Corazza ML. LDH-catalyzed esterification of lauric acid with glycerol in 2009;20:954–66.
solvent-free system. Appl Catal A Gen 2014;475:242–8. [26] José C, Austic GB, Bonetto RD, Burton RM, Briand LE. Investigation of the stability
[2] Moquin PHL, Temelli F. Kinetic modeling of hydrolysis of canola oil in supercritical of Novozym® 435 in the production of biodiesel. Catal. Today 2013;213:73–80.
media. J Supercrit Fluids 2008;45:94–101. [27] Michrowska A, Grela K. Quest for the ideal olefin metathesis catalyst. Pure Appl
[3] Voll FAP, Zanette AF, Cabral VF, Dariva C, De Souza ROMA, Cardozo Filho L, et al. Chem 2008;80:31–43.
Kinetic modeling of solvent-free lipase-catalyzed partial hydrolysis of palm oil. Appl [28] Maechling S, Zaja M, Blechert S. Unexpected results of a turnover number (TON)
Biochem Biotechnol 2012;168:1121–42. study utilising ruthenium-based olefin metathesis catalysts. Adv Synth Catal
[4] Kuss VV, Kuss AV, Da Rosa RG, Aranda DAG, Cruz YR. Potential of biodiesel pro- 2005;347:1413–22.
duction from palm oil at Brazilian Amazon. Renew Sustain Energy Rev [29] Escorsim AM, Cordeiro CS, Ramos LP, Ndiaye PM, Kanda LRSS, Corazza ML.
2015;50:1013–20. Assessment of biodiesel purification using CO2 at high pressures. J Supercrit Fluids
[5] Rocha LLL, Ramos ALD, Antoniosi NR, Furtado NC, Taft CA, Aranda DAG. 2015;96:68–76.
Production of biodiesel by a two-step niobium oxide catalyzed hydrolysis and es- [30] Centre EC for SM. European Committee for Standardization: Management Centre.
terification. Lett Org Chem 2010;7:571–8. 2003.
[6] Alenezi R, Leeke GA, Santos RCD, Khan AR. Hydrolysis kinetics of sunflower oil [31] Menezes RS, Leles MIG, Soares AT, Brandão PI, Franco M, Antoniosi Filho NR, et al.
under subcritical water conditions. Chem Eng Res Des 2009;87:867–73. Avaliação da potencialidade de microalgas dulcícolas como fonte de matéria-prima
[7] Milliren AL, Wissinger JC, Gottumukala V, Schall CA. Kinetics of soybean oil hy- graxa para a produção de biodiesel. Quim Nova 2013;36:10–5.
drolysis in subcritical water. Fuel 2013;108:277–81. [32] Bondioli P, Della Bella L. An alternative spectrophotometric method for the de-
[8] Barnebey HL, Brown AC. Continuous fat splitting plants using the colgate-emery termination of free glycerol in biodiesel. Eur J Lipid Sci Technol 2005;107:153–7.
process. J Am Oil Chem Soc 1948;25:95–9. [33] Dugo G, La Pera L, La Torre GL, Giuffrida D. Determination of Cd(II), Cu(II), Pb(II),
[9] Micic RD, Tomić MD, Kiss FE, Nikolić-Djorić EB, Simikić MD. Optimization of hy- and Zn(II) content in commercial vegetable oils using derivative potentiometric
drolysis in subcritical water as a pretreatment step for biodiesel production by es- stripping analysis. Food Chem 2004;87:639–45.
terification in supercritical methanol. J Supercrit Fluids 2015;103:90–100. [34] Dos Anjos VE, Abate G, Grassi MT. Determination of labile species of As(V), Ba, Cd
[10] Kusdiana D, Saka S. Two-step preparation for catalyst-free biodiesel fuel produc- Co, Cr(III), Cu, Mn, Ni, Pb, Sr, V(V), and Zn in natural waters using diffusive gra-
tion: hydrolysis and methyl esterification. Appl Biochem Biotechnol dients in thin-film (DGT) devices modified with montmorillonite. Anal Bioanal
2004;113–116:781–91. Chem 2017;409:1963–72.
[11] Almeida L, Corazza ML, Sassaki GL, Voll FAP. Experimental study and kinetic [35] Lisboa FDS, Da Silva FR, Cordeiro CS, Ramos LP, Wypych F. Metal glycerolates as
modeling of waste frying soybean oil hydrolysis in subcritical water. React Kinet catalysts in the transesterification of refined soybean oil with methanol under reflux
Mech Catal 2017:1–14. conditions. J Braz Chem Soc 2014;25:1592–600.
[12] Cabral PS, Zandoná Filho A, Voll FAP, Corazza ML. Kinetis of enzimatic hidrolysis [36] Taylor RA, Ellis HA. Room temperature molecular and lattice structures of a
of olive oil im batch and fed-batch system. Appl Biochem Biotechnol homologous series of anhydrous zinc(II) n-alkanoate. Spectrochim Acta – Part A
2014;173:1336–48. Mol Biomol Spectrosc 2007;68:99–107.
[13] Zenevicz MCP, Jacques A, Furigo AF, Oliveira JV, de Oliveira D. Enzymatic hy- [37] Gönen M, Balköse D, Inal F, Ülkü S. Zinc stearate production by precipitation and
drolysis of soybean and waste cooking oils under ultrasound system. Ind Crops Prod fusion processes. Ind Eng Chem Res 2005;44:1627–33.
2016;80:235–41. [38] Pinto JSS, Lanças FM. Hydrolysis of corn oil using subcritical water. J Braz Chem
[14] Jacobson K, Gopinath R, Meher LC, Dalai AK. Solid acid catalyzed biodiesel pro- Soc 2006;17:85–9.
duction from waste cooking oil. Appl Catal B Environ 2008;85:86–91. [39] Ramos LP, Silva FR, Mangrich AS, Cordeiro CS. Biodiesel Production Technologies.
[15] De Paiva EJM, Sterchele S, Corazza ML, Murzin DY, Wypych F, Salmi T. Rev Virtual Quim 2011;3:385–405.
Esterification of fatty acids with ethanol over layered zinc laurate and zinc stearate - [40] Gomes JH, Cordeiro CS. Estudos Preliminares da Glicerólise do Óleo de Palma
Kinetic modeling. Fuel 2015;153:445–54. Catalisada por Carboxilatos Lamelares Preliminary Studies of Glycerolysis of Palm
[16] Dos Santos PRS, Voll FAP, Ramos LP, Corazza ML. Esterification of fatty acids with Oil Catalyzed by Layered Resumo Estudos Preliminares da Glicerólise do Óleo de
supercritical ethanol in a continuous tubular reactor. J Supercrit Fluids Palma Catalisada por Carboxilatos. Rev Virtual Química 2017;9:3–10.
2017;126:25–36. [41] Kusdiana D, Saka S. Effects of water on biodiesel fuel production by supercritical
[17] Cordeiro CS, Da Silva FR, Wypych F, Ramos LP. Catalisadores heterogêneos para a methanol treatment. Bioresour Technol 2004;91:289–95.
produção de monoésteres graxos (biodiesel). Quim Nova 2011;34:477–86. [42] Kontopoulou I, Angelopoulou A, Bouropoulos N. ZnO spherical porous nanos-
[18] Ramos LP, Brugnago RJ, Da Silva FR, Cordeiro CS, Wypych F. Esterificação e tructures obtained by thermal decomposition of zinc palmitate. Mater Lett
transesterificação simultâneas de óleos ácidos utilizando carboxilatos lamelares de 2016;165:87–90.
zinco como catalisadores bifuncionais. Quim Nova 2015;38:46–54.
[19] Reinoso DM, Ferreira ML, Tonetto GM. Study of the reaction mechanism of the

579

You might also like