You are on page 1of 7

Catalysis Today xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Biphasic reaction of glycerol and oleic acid: Byproducts formation and phase
transfer autocatalytic effect
Mariana de S. Gomesa, Mariana R.D. Santosa, Adriana B. Salvianoa, Fernanda G. Mendonçaa,
Izadora R.S. Menezesa, Marina Jurischa, Guilherme D. Rodriguesa, Rodinei Augustia,

Paulo S. Martinsb, Rochel M. Lagoa,
a
Chemistry Department, Universidade Federal de Minas Gerais, MG, Belo Horizonte, CEP31270-901, Brazil
b
Fiat Chrysler Automobiles, MG, Betim, CEP 32669-185, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: This work investigates two new aspects of the biphasic esterification reaction of glycerol (G) and oleic acid (A),
Glycerol i.e. the formation of several oxidation/dehydration co-products and the effect of the ester products as phase
Oleic acid transfer agents to accelerate the reaction. Potentiometric titration, FTIR, TG and especially paper electrospray
Oleate ester mass spectrometry (PS-MS) were employed to analyze the products. Reactions with different G:A molar ratios
Biphasic
(1:0.5, 1:1 and 1:2) without catalyst at 140, 160 and 180 °C showed oleic acid conversions of 50–80% to form the
Phase transfer
mono, di and triesters with several other by co-products due to dehydrogenation/oxidation of the oleic chain and
dehydration of the glycerol moiety identified by PS-MS. The phase transfer autocatalytic effect was investigated
by the addition of the oleate esters (2, 4 and 10 wt%) to the reaction. It was observed a strong acceleration of at
the initial stage of the process indicating an autocatalysis likely due to a phase transfer effect. The reaction is
discussed in terms of three stages. In the initial biphasic Stage 1, oleic acid suffers some dehydrogenation/
oxidation and the esterification reaction takes place slowly at the glycerol-oleic acid interface to form mainly the
monooleate ester which is an amphiphilic/tensoactive molecule. In the Stage 2, the amphiphilic ester products
act as phase transfer agents taking glycerol molecules to the oleic acid phase for the formation of mono and
mainly dioleate ester. At the Stage 3 in the presence of relatively high concentration of the tensoactive ester
products the reaction becomes homogeneous/monophasic.

1. Introduction considerable interest. Several works investigated the reaction of gly-


cerol with fatty acids using solvents, e.g., t-butanol, in the presence of
Conversion of glycerol byproduct of biodiesel to different products, tin-organic framework and montmorillonite clays as catalysts [25,26].
such as hydrogen [1,2], ethanol [3], methanol [4], allylic alcohol [5], On the other hand, the majority of the works have focused on solvent-
dust suppressant [6], 1,2-propanediol [7,8], trehalose and propanoic free esterification in the presence of different catalysts such as ionic
acid [9], has been intensively investigated in the last years [10–12]. liquids [27,28], methane sulfonic acid under microwave-heating [29],
Catalytic glycerol esterification with different acids such as acetic modified natural zeolites [30], layered double hydroxide [18] and
[13,14], lauric [15,16], and valeric [17] has been proposed as a pro- immobilized lipases [31,32]. It is interesting to observe that although
mising route to obtain value-added products. Glycerol esterification the reaction of glycerol and fatty acid, e.g. oleic acid, is biphasic, most
with fatty acids in the presence of different catalysts has been reported of the published work did not mention nor considered diffusion lim-
to form mixtures of mono, di and triacyl glycerides [11,18]. These itations in the process. The reaction of glycerol with oleic acid in the
products show promising applications as additives in polymeric films presence of modified clinoptilolite indicated the importance of an am-
[19], formulations with antibacterial activity [20,21], food emulsifiers phiphilic catalyst to interact with the polar glycerol phase and the
[22,23], and as well as in pharmaceutical and cosmetic formulations hydrophobic oleic acid [30]. The location of the lipase immobilized on
[24]. Therefore, these glycerol esterification reactions have attracted poly(methylmethacrylate-co-divinylbenzene) [31] at the interface


Corresponding author.
E-mail addresses: mariana.ufmg@hotmail.com (M. de S. Gomes), marianards@ufmg.br (M.R.D. Santos), adrianasalviano@colab.ead.ufmg.br (A.B. Salviano),
nandagm@ufmg.br (F.G. Mendonça), izadora.rhaynna.s.m@gmail.com (I.R.S. Menezes), marina.jurisch@gmail.com (M. Jurisch),
guilhermedr@ufmg.br (G.D. Rodrigues), augusti@ufmg.br (R. Augusti), paulosergio.martins@fcagroup.com (P.S. Martins), rochel@ufmg.br (R.M. Lago).

https://doi.org/10.1016/j.cattod.2019.02.011
Received 1 May 2018; Received in revised form 16 December 2018; Accepted 10 February 2019
0920-5861/ © 2019 Published by Elsevier B.V.

Please cite this article as: Mariana de S. Gomes, et al., Catalysis Today, https://doi.org/10.1016/j.cattod.2019.02.011
M. de S. Gomes, et al. Catalysis Today xxx (xxxx) xxx–xxx

glycerol-oleic acid was also mentioned as an important factor for the


esterification. Up to date, no detailed study on the reaction dynamics
and diffusion limitations of the biphasic system glycerol-oleic acid has
been published in the literature. Moreover, despite the relatively high
temperatures used for the esterifications, no attention has been paid to
collateral reactions, such as oxidation and dehydration commonly ob-
served for fatty acids and glycerol.
In this work, it was studied the reaction between glycerol and oleic
acid looking at the potential effect of the tensoactive reaction products,
i.e. oleate esters of glycerol, as phase transfer agents in an autocatalytic
reaction. Although several examples of different reactions are reported
to be accelerated by phase transfer catalysts [33–35], molecules such as
glycerol oleate esters have never been reported to have this effect.
Moreover, the paper spray mass spectrometry (PS-MS) was used to in-
vestigate the formation of coproduct by products related to oxidation
and dehydration processes.

2. Materials and methods

The reaction of glycerol (G) with oleic acid (A) was studied in dif- Fig. 1. Conversion of oleic acid during the reaction with glycerol at different
ferent molar proportions (1:0.5, 1:1 and 1:2) at 140, 160 and 180 °C. temperatures 140, 160 and 180 °C.
These reactions are named hereon according to the molar ratio and
temperature, e.g. G1:A1160 (glycerol : oleic acid molar ratio 1:1 at oleic acid takes place in the glycerol phase. The titrations of the oleic
160 °C). acid phase were used to calculate the acid conversion. The effect of
The reaction was carried out in a beaker containing glycerol and temperature on the oleic acid conversion is shown in Fig. 1.
oleic acid in an oil bath under stirring for 5 h. The reaction was mon- It can be observed at 140 °C no significant oleic acid conversion up
itored by titration with sodium hydroxide (0.5 mol L−1). to 1 h of reaction. During the second hour, however, reaction a strong
Infrared spectra (FTIR) were obtained on a Perkin Elmer conversion was observed. After 3 h the oleic acid conversion increased
Spectrophotometer Frontier Single Range, by the Attenuated Total linearly reaching 49% at 6 h. A similar behavior was observed for the
Reflectance (ATR) technique and registered in the region of 4000 - reaction carried out at 160 °C which reached oleic acid conversions ca.
550 cm−1, resolution of 4 cm−1 and 16 scans per sample. 66%. It was interesting to observe that the biphasic reactions gradually
The thermal stability of the obtained product was studied by TG/ become yellow/brown in color and at near 2.5 h changed to mono-
DTG (TG DTG Shimatzu-60 H) instrument under N2 dynamic atmo- phasic.
sphere, with a heating rate of 10 °C min−1 at 900 °C. On the other hand, the reaction at 180 °C showed a more important
The reactions were monitored using the paper spray mass spectro- oleic acid conversion of 21% already in the first hour and reached ca.
metry (PS-MS) technique. Under the optimized conditions, the ester- 85% after 6 h. Moreover, it was observed a mixing and homogenization
ification reaction reached equilibrium after approximately 5 h. All PS- of the biphasic mixture at near 2 h reaction.
MS experiments were performed using a LCQ Fleet (Thermo Scientific, The reaction at 160 °C at different times was analyzed by FTIR and
San Jose, CA, USA) mass spectrometer operating in the positive ion TG. The FTIR spectra showed a gradual disappearance of the band near
mode. The instrumental conditions were as follows: voltage applied to 1710 cm−1, related to COOH (C]O stretching) of oleic acid [36,37],
the paper, 4 kV; capillary temperature, 275 °C; capillary voltage, 35 V; and the appearance of the band near 1737 cm−1, characteristic of the
tube lens voltage, 65 V. Full scan mass spectra were recorded for all ester C]O stretching [38,39] (Fig. 2).
analyzes over a 100–1000 m/z range. For the PS-MS experiments, small
pieces of a chromatographic paper (grade 1, purchased from Whatman
International Ltd., Maidstone, Kent, England) were cut in a triangular
format (10 mm height and 5 mm base width). The paper was then po-
sitioned approximately 5 mm away from the mass spectrometer inlet,
using a metal clip fixed to a platform with three-dimensional move-
ment. An optimized high voltage (4.0 kV) was applied to the paper base
through the metal clip whereas sample (5 μL) and HPLC grade methanol
(5 μL) were added at the central portion of the triangle. To avoid
memory effects, each analysis was performed on a new paper.
For the autocatalysis study, the reactions were carried out with the
addition of the product of t = 5 h of the reaction G1: A1 - 160 °C at
t = 0 h of the reactions in the concentrations of 2, 4 and 10 wt% in the
reaction system G1:A1 – 160 °C at t = 0 h.

3. Results and discussion

The reactions of glycerol with oleic acid were performed with gly-
cerol:oleic acid molar ratio 1:1 (G1:A1) at 140, 160 and 180 °C without
catalyst.
The reaction was initially biphasic and the oleic acid consumption
was monitored by FTIR and NaOH titration of both phases, i.e. glycerol
and oleic acid. The analyses showed no significant chemical modifica- Fig. 2. FTIR spectra of the reaction G1:A1160 at different times and glycerol
tion of the glycerol phase. The results also showed that no dissolution of conversion into mono, di and trioleate ester.

2
M. de S. Gomes, et al. Catalysis Today xxx (xxxx) xxx–xxx

RCOOCH2CH(OH)CH2OH → RCOOCH2C = OCH3/RCOOCH = CHCHO


+ [H2O] (2)

Similar dehydration processes have been observed before during


oligomerization of glycerol [48].
The ions at m/z 603, 617 and 637 are related to the beginning of the
diesters [D] formation. The ion at m/z 617 is related to the diester after
the loss of H2, [(D)-H2]H +. The ion at m/z 637 refers to the diester with
an oxygen atom (epoxy), [(D)+O]H+. The m/z 603 corresponds to the
diester that suffered the loss of H2O, [(D)-H2O]H+. No ion related to
trioleate was observed at this stage. After 5 h reaction, howevwe, the
relative intensities of the monoester-related ions decreased and lower-
intensity ions at m/z 901 and 917 likely due to the trioleate [T]
emerged. The first ion refers to trioleate with the loss of 2 molecules of
hydrogen [(T)-2H2]H + whereas the second ion one is related to oxi-
dized trioleate, i.e. [(T)+O]H+, both resulting from oxidation of the
main reaction product (trioleate) [49,50].
Based on the MS data, a plot with the relative abundances of oleic
acid and esters was built, as shown in Fig. 5.
Fig. 3. DTG analyses of the reactions G1: A1140, G1: A1160, G1: A1180, glycerol, It can be observed that the relative intensity of the oleic acid signal
oleic acid and soybean oil. rapidly decreased up to 2 h reaction. In this period, the relative in-
tensities of the signals for the mono and specially diester increased.
TG analyses of the reactions G1:A1140, G1:A1160 and G1:A1160 after After 2 h the monoester relative signal intensity decreased whereas the
6 h of reaction showed broad weight losses centered at ca. 250, 320, diester went on increasing. This result seems to suggest that both mono
380 °C. DTG analyses of pure oleic acid, glycerol and the tri-oleic ester but specially diester are formed at the beginning of the reaction. After
of glycerol (see Fig. 3) indicated that the peak at 250 °C could be related 2 h, part of the monoester is apparently converted to the diester. The
to oleic acid, whereas the weight losses centered at 320 and 380 °C can relative signal intensity concentration of triester remains low
be related to the mono and the diester, respectively. throughout the reaction.
The reaction at 160 °C at different times was also monitored by The Paper ESI-MSPS-MS relative intensity data for the mono, di and
Paper Spray ionization mass spectrometry (PS-MS). PS-MS is a recently- triester was also used to investigate the effect of temperature on the
developed ambient ionization method. In PS-MS, the tip of a triangular reactions G1:A1140, G1:A1160 and G1:A1180 (Fig. 6).
paper is placed in front of the mass spectrometer inlet whereafter a high It can be observed that at 140 °C the main product is the diester with
voltage is applied through a metal clip connected to its base. The a relative abundance of ca. 54% and the monoester with 40%. As the
sample is then added to the central part of the triangle while a solvent is temperature increased to 160 °C the diester increased to 65% and the
added at its base. A spray emerges from the tip and charged analytes are monoester decreased to 28%. At 180 °C the diester reached ca. 70% and
attracted into the mass spectrometer to be analyzed [40–45]. Due to its the monoester further decreased to 22%.
peculiar characteristics, PS-MS has shown to be a convenient approach It is interesting to consider based on the PS-MS and TG results that
to monitor reactions taking place in the condensed phase. Fast re- no free glycerol was observed after 5 h reactions. As the amount of
sponses can be achieved, thus allowing the attainment of real-time and glycerol used in the reaction was enough to produce the monoester,
quite useful information regarding a given reactional process. The PS- these results suggest that glycerol is lost during the reaction by eva-
MS obtained are shown in Fig. 4. poration.
The results obtained for the reaction at t = 0 h (after ca. 1 h heating The effect of the glycerol:oleic acid molar ratio was also in-
to reach 160 °C) showed no significant formation of esters. On the other vestigated at 160 °C by the reactions G1:A0,5160, G1:A1160 and G1:
hand, the PS-MS indicated some modification of the oleic acid. The A2160. The ESI-MS spectra PS-MS data were used to estimate the
mass spectrum showed no ion at m/z 283 related to the protonated oleic abundance ratio of the mono, di and triesteres (Fig. 7).
acid molecule. However, strong signals are observed at m/z 561 and It can be observed for G1:A0,5160 (a high glycerol:oleic acid molar
563 likely related to the dehydrogenated oleic acid dimers, i.e. [2(OA) - ratio) that the main product is the diester with 66% relative abundance.
H2]H+ and [2(OA) - 2H2]H+, respectively. These results seem to in- As the concentration of oleic acid increased the relative formation of
dicate that during the heating phase to reach 160 °C, the oleic acid the diester increased as expected, reaching a relative concentration of
molecule lost hydrogen probably by an oxidation process according to ca. 70%.
the simplified reaction (Eq. (1)): It was also investigated a possible phase transfer catalysis effect by
adding the dioleate glycerol ester to the biphasic reaction at propor-
C17H33COOH → C17H32COOH/C17H30COOH + [H2O] (1) tions equal to 2, 4 and 10 wt%. It was observed that upon the addition,
the diester was located at the glycerol-oleic acid interface. After vig-
It can also be observed two ions at signals m/z 577 and 579 which
orous stirring, the diester was mixed to the oleic acid phase. No sig-
are likely related to oxidized molecules, e.g. epoxides. Similar dehy-
nificant diester concentration was detected by FTIR in the glycerol
drogenation and oxidation processes have been observed in previous
phase.
works on the ozonization of oleic acid [46] and thermal stress of soy
Upon heating of the biphasic mixtures glycerol-oleic acid
bean oil [47,48].
(G1:A1160) containing 2, 4 and 10 wt% of the diester, the conversion of
After 2 h, the esterification product is observed by the new ion at m/
oleic acid was monitored as shown in the curves presented in Fig. 8.
z 357 related to the protonated form of the oleate glycerol monoester
It can be observed for the reaction G1:A1160 that the oleic acid
[M+]H+]. It is interesting to observe an intense ion signal at m/z 339
conversion increases slowly only after 1.5 h reaction. In the presence of
likely related to the monoester dehydrated [(M)–H2O]H+, suggesting a
2 and 4 wt% diester, oleic acid conversion is already observed at 1 h
possible dehydration process. Although this dehydration process is not
reaction. On the other hand, with 10 wt% diester the conversion
clear, it likely takes place in the glycerol moiety forming groups such as
reached ca. 45% at near 1 h reaction. These results clearly indicate that
ketone and aldehyde (Eq. 2).

3
M. de S. Gomes, et al. Catalysis Today xxx (xxxx) xxx–xxx

Fig. 4. PS(+)-MS of the aliquots of the reaction 1G:A1160 collected after 0, 2 and 5 h.

the ester products accelerated the reaction. The ester products have a the interface.
tensoactive nature, where the glycerol moiety is more polar/hydro- At the second stage (Fig. 10), the amphiphilic ester product acts as a
philic whereas the oleate moiety is strongly hydrophobic and interacts phase transfer catalyst and carry glycerol molecules to the oleic acid
with the oleic acid phase. Therefore, the mono and diester products are phase and the esterification is accelerated. As the oleic acid is present in
likely acting as phase transfer catalyst carrying glycerol molecules to high concentration, the second esterification takes place to form pre-
the oleic acid phase. ferentially the diester. Likely, the esterification to form the triester does
Considering these results, the reaction can be discussed in terms of not occur significantly due to sterical hindrance.
three different stages. At Stage 1 (Fig. 9) glycerol and oleic acid do not In Stage 3, the amphiphilic molecules are formed in quantities en-
mix and the reaction takes place at the interface. The monoester formed ough to promote the homogeneous mixture of the two phases, i.e.
is an amphiphilic molecule, with the polar glycerol moiety and the glycerol and oleic acid.
hydrophobic oleate ramification and it will likely be located mainly at

4
M. de S. Gomes, et al. Catalysis Today xxx (xxxx) xxx–xxx

Fig. 5. Relative abundances of oleic acid, mono, di and trioleate esters of gly- Fig. 7. Relative abundance of mono, di and triglycerides after 5 h reactions in
cerol for the reaction G1:A1160 obtained acquired from the PS-MS data. the molar ratios of G1:A0,5160, G1:A1160 and G1:A2160.

Fig. 8. Autocatalysis study of the esterification reaction with the addition of 2,


Fig. 6. Selectivity of mono, di and triesters after 5 h G1:A1 reactions at 140 and 4 and 10 wt% of the product G1: A1160 at the beginning of the reaction, com-
160 °C. pared to the reaction without the addition of the diester.

4. Conclusions

The biphasic esterification reaction of glycerol (G) and oleic acid (A)
has been reported in many recent works but several aspects of this re-
action have been neglected. The use of paper spray MS mass spectro-
metry showed that at temperatures 140–180 °C, besides the esterifica-
tion reaction, several collateral processes take place, such as the oleic
acid chain oxidation/dehydrogenation and glycerol moiety dehydra-
tion. Moreover, the esterification reaction is limited by diffusion pro-
cesses. As the reaction initiates, the amphiphilic/tensoactive oleate
ester derivatives are likely acting as phase transfer agents and carry
glycerol molecules to the oleic acid phase where the esterification re-
action takes place to form the mono but mainly the dioleate ester. As Fig. 9. Schematic representation of ester formation at the glycerol-oleic acid
the concentration of the oleate esters increase the reaction become interface (Stage 1).
homogeneous/monophasic.

Acknowledgements

The authors gratefully acknowledge the financial support of Fiat


Chrysler Automobiles group for the financial assistance of the fluid

5
M. de S. Gomes, et al. Catalysis Today xxx (xxxx) xxx–xxx

Fig. 10. Schematic representation of the phase transfer of glycerol to oleic acid followed by the esterifications to form the mono and diester (Stage 2).

cutting project. Also the support of INCT Midas, CNPq, CAPES, as co-surfactants, Eur. J. Pharm. Biopharm. 108 (2016) 100–110.
FAPEMIG. [22] F.C. Wang, A.G. Marangoni, Advances in the application of food emulsifier α-gel
phases: saturated monoglycerides, polyglycerol fatty acid esters, and their deriva-
tives, J. Colloid Interface Sci. 483 (2016) 394–403.
References [23] N.C. Acevedo, J.M. Block, A.G. Marangoni, Unsaturated emulsifier-mediated mod-
ification of the mechanical strength and oil binding capacity of a model edible fat
crystallized under shear, Langmuir 28 (2012) 16207–16217.
[1] R. Mangayil, M. Karp, V. Santala, Bioconversion of crude glycerol from biodiesel
[24] A. Akinshina, C. Das, M.G. Noro, Effect of monoglycerides and fatty acids on a
production to hydrogen, Int. J. Hydrog. Energy 37 (2012) 12198–12204.
ceramide bilayer, Phys. Chem. Chem. Phys. 18 (2016) 17446–17460.
[2] F.G.F. de Paula, M.G. Rosmaninho, A.Pd.C. Teixeira, P.P. de Souza, R.M. Lago,
[25] L.H. Wee, T. Lescouet, J. Fritsch, F. Bonino, M. Rose, Z. Sui, E. Garrier, D. Packet,
Alcoxycle: a novel route for glycerol reform into H2 and COx in separate stages,
S. Bordiga, S. Kaskel, M. Herskowitz, D. Farrusseng, J.A. Martens, Synthesis of
Catal. Today 289 (2017) 127–132.
monoglycerides by esterification of oleic acid with glycerol in heterogeneous cat-
[3] T. Ito, Y. Nakashimada, K. Senba, T. Matsui, N. Nishio, Hydrogen and ethanol
alytic process using tin-organic framework catalyst, Catal. Lett. 143 (2013)
production from glycerol-containing wastes discharged after biodiesel manu-
356–363.
facturing process, J. Biosci. Bioeng. 100 (2005) 260–265.
[26] A. Chaari, S.B. Neji, M.H. Frikha, Fatty acid esterification with polyols over acidic
[4] M.H. Haider, N.F. Dummer, D.W. Knight, R.L. Jenkins, M. Howard, J. Moulijn,
montmorillonite, J. Oleo Sci. 66 (2017) 455–461.
S.H. Taylor, G.J. Hutchings, Efficient green methanol synthesis from glycerol, Nat.
[27] P. Lozano, C. Gomez, S. Nieto, G. Sanchez-Gomez, E. Garcia-Verdugo, S.V. Luis,
Chem. 7 (2015) 1028–1032.
Highly selective biocatalytic synthesis of monoacylglycerides in sponge-like ionic
[5] A. Konaka, T. Tago, T. Yoshikawa, A. Nakamura, T. Masuda, Conversion of glycerol
liquids, Green Chem. 19 (2017) 390–396.
into allyl alcohol over potassium-supported zirconia–iron oxide catalyst, Appl.
[28] W.N.R.W. Isahak, Z.A.C. Ramli, M. Ismail, M.A. Yarmo, Highly selective glycerol
Catal. B Environ. 146 (2014) 267–273.
esterification over silicotungstic acid nanoparticles on ionic liquid catalyst, Ind.
[6] M.A. Medeiros, C.M.M. Leite, R.M. Lago, Use of glycerol by-product of biodiesel to
Eng. Chem. Res. 53 (2014) 10285–10293.
produce an efficient dust suppressant, Chem. Eng. J. 180 (2012) 364–369.
[29] P.S. Kong, M.K. Aroua, W.M.A.W. Daud, P. Cognet, Y. Pérès, Enhanced microwave
[7] Z. Yuan, J. Wang, L. Wang, W. Xie, P. Chen, Z. Hou, X. Zheng, Biodiesel derived
catalytic-esterification of industrial grade glycerol over Brønsted-based methane
glycerol hydrogenolysis to 1,2-propanediol on Cu/MgO catalysts, Bioresour.
sulfonic acid in production of biolubricant, Process. Saf. Environ. Prot. 104 (Part A)
Technol. 101 (2010) 7088–7092.
(2016) 323–333.
[8] A. Marinas, P. Bruijnincx, J. Ftouni, F.J. Urbano, C. Pinel, Sustainability metrics for
[30] M. Akgül, A. Karabakan, Selective synthesis of monoolein with clinoptilolite,
a fossil- and renewable-based route for 1,2-propanediol production: a comparison,
Microporous Mesoporous Mater. 131 (2010) 238–244.
Catal. Today 239 (2015) 31–37.
[31] X. Meng, G. Xu, Q.-L. Zhou, J.-P. Wu, L.-R. Yang, Improvements of lipase perfor-
[9] R. Ruhal, S. Aggarwal, B. Choudhury, Suitability of crude glycerol obtained from
mance in high-viscosity system by immobilization onto a novel kind of poly(me-
biodiesel waste for the production of trehalose and propionic acid, Green Chem. 13
thylmethacrylate-co-divinylbenzene) encapsulated porous magnetic microsphere
(2011) 3492–3498.
carrier, J. Mol. Catal. B Enzym. 89 (2013) 86–92.
[10] J.Q. Albarelli, D.T. Santos, M.R. Holanda, Energetic and economic evaluation of
[32] Y. Zhao, J. Liu, L. Deng, F. Wang, T. Tan, Optimization of Candida sp. 99-125 lipase
waste glycerol cogeneration in Brazil, Braz. J. Chem. Eng. 28 (2011) 691–698.
catalyzed esterification for synthesis of monoglyceride and diglyceride in solvent-
[11] P.S. Kong, M.K. Aroua, W. Daud, Catalytic esterification of bioglycerol to value-
free system, J. Mol. Catal. B Enzym. 72 (2011) 157–162.
added products, Rev. Chem. Eng. 31 (2015) 437–451.
[33] L. Zong, C.-H. Tan, Phase-transfer and ion-pairing catalysis of Pentanidiums and
[12] M.V. Sivaiah, S. Robles-Manuel, S. Valange, J. Barrault, Recent developments in
bisguanidiniums, Acc. Chem. Res. 50 (2017) 842–856.
acid and base-catalyzed etherification of glycerol to polyglycerols, Catal. Today 198
[34] S. Liu, Y. Kumatabara, S. Shirakawa, Chiral quaternary phosphonium salts as phase-
(2012) 305–313.
transfer catalysts for environmentally benign asymmetric transformations, Green
[13] P.U. Okoye, A.Z. Abdullah, B.H. Hameed, Synthesis of oxygenated fuel additives via
Chem. 18 (2016) 331–341.
glycerol esterification with acetic acid over bio-derived carbon catalyst, Fuel 209
[35] M. Ikunaka, PTC in OPRD: an illustrative overview, Org. Process Res. Dev. 12
(2017) 538–544.
(2008) 698–709.
[14] P. Arun, S.M. Pudi, P. Biswas, Acetylation of glycerol over sulfated alumina: reac-
[36] N. Shukla, C. Liu, P.M. Jones, D. Weller, FTIR study of surfactant bonding to FePt
tion parameter study and optimization using response surface methodology, Energy
nanoparticles, J. Magn. Magn. Mater. 266 (2003) 178–184.
Fuels 30 (2016) 584–593.
[37] D.M. Perígolo, F.G.F. de Paula, M.G. Rosmaninho, P.P. de Souza, R.M. Lago,
[15] F. Hamerski, M.L. Corazza, LDH-catalyzed esterification of lauric acid with glycerol
M.H. Araujo, Conversion of fatty acids into hydrocarbon fuels based on a sodium
in solvent-free system, Appl. Catal. A Gen. 475 (2014) 242–248.
carboxylate intermediate, Catal. Today 279 (2017) 260–266.
[16] X.X. Han, X.F. Zhang, G.Q. Zhu, J.J. Liang, X.H. Cao, R.J. Kan, C.T. Hung, L.L. Liu,
[38] A. Rohman, Y.B.C. Man, Fourier transform infrared (FTIR) spectroscopy for analysis
S.B. Liu, Ionic liquid-silicotungstic acid composites as efficient and recyclable cat-
of extra virgin olive oil adulterated with palm oil, Food Res. Int. 43 (2010)
alysts for the selective esterification of glycerol with lauric acid to monolaurin,
886–892.
Chemcatchem 9 (2017) 2727–2738.
[39] N. Salih, J. Salimon, E. Yousif, Synthesis of oleic acid based esters as potential
[17] K. Kaur, R.K. Wanchoo, A.P. Toor, Elementary transformation of glycerol to triva-
basestock for biolubricant production, Turk. J. Eng. Environ. Sci. 35 (2011)
lerin: design of an experimental approach, ACS Sustain. Chem. Eng. 5 (2017)
115–123.
802–808.
[40] H. Wang, J. Liu, R.G. Cooks, Z. Ouyang, Paper spray for direct analysis of complex
[18] F. Hamerski, M.A. Prado, V.R. da Silva, F.A.P. Voll, M.L. Corazza, Kinetics of
mixtures using mass spectrometry, Angew. Chemie Int. Ed. 49 (2010) 877–880.
layered double hydroxide catalyzed esterification of fatty acids with glycerol, React.
[41] J. Ji, L. Nie, L. Liao, R. Du, B. Liu, P. Yang, Ambient ionization based on mesoporous
Kinet. Mech. Catal. 117 (2016) 253–268.
graphene coated paper for therapeutic drug monitoring, J. Chromatogr. B 1015-
[19] P.G. Boakye, K.C. Jones, N.P. Latona, C.K. Liu, S.A. Besong, S.E. Lumor, V.T. Wyatt,
1016 (2016) 142–149.
Modification of absorbent poly(glycerol‐glutaric acid) films by the addition of
[42] C.W. Klampfl, M. Himmelsbach, Direct ionization methods in mass spectrometry: an
monoglycerides, J. Appl. Polym. Sci. 134 (2017).
overview, Anal. Chim. Acta 890 (2015) 44–59.
[20] J.A. Jackman, B.K. Yoon, D. Li, N.-J. Cho, Nanotechnology formulations for anti-
[43] N.E. Manicke, Q. Yang, H. Wang, S. Oradu, Z. Ouyang, R.G. Cooks, Assessment of
bacterial free fatty acids and monoglycerides, Molecules (Basel, Switzerland) 21
paper spray ionization for quantitation of pharmaceuticals in blood spots, Int. J.
(2016) 305-305.
Mass Spectrom. 300 (2011) 123–129.
[21] A. Umerska, V. Cassisa, N. Matougui, M.-L. Joly-Guillou, M. Eveillard, P. Saulnier,
[44] H. Wang, N.E. Manicke, Q. Yang, L. Zheng, R. Shi, R.G. Cooks, Z. Ouyang, Direct
Antibacterial action of lipid nanocapsules containing fatty acids or monoglycerides
analysis of biological tissue by paper spray mass spectrometry, Anal. Chem. 83

6
M. de S. Gomes, et al. Catalysis Today xxx (xxxx) xxx–xxx

(2011) 1197–1201. catalyzed oligomerization of glycerol investigated by electrospray ionization mass


[45] R.D. Espy, A.R. Muliadi, Z. Ouyang, R.G. Cooks, Spray mechanism in paper spray spectrometry, J. Braz. Chem. Soc. 20 (2009) 1667–1673.
ionization, Int. J. Mass Spectrom. 325-327 (2012) 167–171. [49] W.C. Byrdwell, W.E. Neff, Dual parallel electrospray ionization and atmospheric
[46] R.L. Grimm, R. Hodyss, J.L. Beauchamp, Probing interfacial chemistry of single pressure chemical ionization mass spectrometry (MS), MS/MS and MS/MS/MS for
droplets with field-induced droplet ionization mass spectrometry: physical ad- the analysis of triacylglycerols and triacylglycerol oxidation products, Rapid
sorption of polycyclic aromatic hydrocarbons and ozonolysis of oleic acid and re- Commun. Mass Spectrom. 16 (2002) 300–319.
lated compounds, Anal. Chem. 78 (2006) 3800–3806. [50] R.R. Catharino, H.M.S. Milagre, S.A. Saraiva, C.M. Garcia, U. Schuchardt,
[47] R.R. Catharino, R. Haddad, L.G. Cabrini, I.B.S. Cunha, A.C.H.F. Sawaya, M.N. Eberlin, R. Augusti, R.C.L. Pereira, M.J.R. Guimarães, G.F. de Sá,
M.N. Eberlin, Characterization of vegetable oils by electrospray ionization mass J.M.R. Caixeiro, V. de Souza, Biodiesel typification and quality control by direct
spectrometry fingerprinting: classification, quality, adulteration, and aging, Anal. infusion electrospray ionization mass spectrometry fingerprinting, Energy Fuels 21
Chem. 77 (2005) 7429–7433. (2007) 3698–3701.
[48] M.A. Medeiros, M.H. Araujo, R. Augusti, L.C.Ad. Oliveira, R.M. Lago, Acid-

You might also like