Uploaded by Dương Linh

Bioinspired Enzymatic Synthesis

advertisement
pubs.acs.org/journal/ascecg
Research Article
Bioinspired Enzymatic Synthesis of Terpenoid-Based (Meth)acrylic
Monomers: A Solvent‑, Metal‑, Amino‑, and Halogen-Free Approach
Thibault Castagnet, Garbine Aguirre, Jose ́ M. Asua,* and Laurent Billon*
Cite This: ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
Downloaded via LG CHEM LTD on January 3, 2022 at 10:14:42 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
ACCESS
Metrics & More
Read Online
Article Recommendations
sı Supporting Information
*
ABSTRACT: In an attempt to pave the way toward the substitution of
petroleum-based monomers for monomers from sustainable and renewable
resources, this work investigates the synthesis of a portfolio of new
(meth)acrylic monomers using terpenoids as raw materials that can be
conveniently extracted from wood waste and citrus fruits. The synthetic
process is based on the enzymatic catalysis of the esterification of terpenoids
and is solvent-, metal-, amino-, and halogen-free, overcoming the limitations
of the common esterification techniques. The effect of employing microwaves
in combination with the enzyme and the effect of the acyl donor choice
[(meth)acrylic acids vs (meth)acrylic anhydrides] were studied. The process
yielded complete conversion and then high-purity monomers with a variety of
structures that would allow them to substitute many of the most commonly used petrochemical (meth)acrylic monomers.
KEYWORDS: terpenoids, esterification, microwave, enzyme, (meth)acrylic monomers
■
reagents,41 metal or amino-based catalysts,42 and/or harsh
conditions (acidic medium and high temperatures).43,44 In
addition, under these conditions, it is difficult to avoid the
polymerization of unsaturated bonds, which are largely present
on the molecular structure of terpenes. Therefore, there is an
urgent need of greener and more efficient methods. This work
focuses on the environmentally friendly synthesis of renewable
(meth)acrylic monomers from terpenoids. The current
technology is based on the esterification of terpenoids
employing (meth)acryloyl chloride in solvents such as
dichloromethane. In this reaction, hydrochloric acid is
produced, which is trapped with triethylamine (TEA),
followed by washing with brine and water to remove TEA
and HCl.41 All this makes the current technology environmentally unfriendly and time-/energy-consuming.
This study reports on a green synthetic method that allows
the synthesis of terpenoid-based (meth)acrylate monomers by
avoiding the use of solvents, metal or amino-based catalysts,
halogen, and high temperature. This method is based on the
combination of enzymatic catalysis and microwave irradiation.
Enzymatic biocatalysis is an important tool for the green
production of chemicals because of the mild conditions used.
The wide range of enzymes available includes lipases, esterases,
INTRODUCTION
Polymers are ubiquitous in our society because their properties
can be adjusted to widely different applications. The
development of such a versatile material has been possible
because of the availability of a range of monomers that,
through clever combinations, yield the performance polymers.
Most of the monomers currently used to produce synthetic
polymers are petroleum-based, and biosourced alternatives
should be found for long-term sustainability. Biobased
compounds do not have readily polymerizable double bonds,
and therefore, they need to be modified to incorporate a
polymerizable double bond in their structure. Numerous
examples of this type of modification can be found in the
literature.1−12 In particular, the incorporation of (meth)acrylate moieties has been often reported using saccharides13−17 (glucose as the main example18−21), animal and
vegetable fatty acids, and resin acids.22−31
Terpenes and terpenoids are a category of biomass, which is
extracted from trees and especially from conifers and pine
trees. Terpenes are the main components of turpentine, which
is the volatile fraction of resin. Terpenoids are modified
terpenes that contain oxygen in various functional groups.
They are industrially available at the industrial scale from the
paper industry. A wide variety of terpene and terpenoid
molecules have the potential to be the raw material for the
production of a wide range of monomers, which in turn can
form polymers having a broad spectrum of properties.5,32−39
A serious drawback of the synthetic methods used to prepare
the monomers mentioned above is that they are not
environmentally friendly. They use solvents,40 alkyl chloride
© 2020 American Chemical Society
Received: March 24, 2020
Revised: April 10, 2020
Published: April 24, 2020
7503
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
Research Article
Scheme 1. Synthetic Methodologies for Obtaining Renewable (Meth)acrylic Monomers
monomers was investigated using the industrially available
terpenoids (Table 1).
cutinases, and proteases. These enzymes have attractive
characteristics such as substrate selectivity, enantioselectivity,
and versatility. Among them, lipases are widely used at
industrial scale because of their high performance.45−53
In this work, enzymes known to catalyze esterification are
used. Candida Antarctica lipases are very efficient in the
esterification of primary alcohols and even show some activity
with secondary alcohols.54 Novozym 435 (a supported
Candida Antarctica lipase) is widely used for its high
performance in the direct esterification of primary alcohols54
or in their transesterification.55 Candida Rugosa lipases are
commonly used for esterification of secondary alcohols. Both
lipases are industrial supported catalysts to be easily separated
and removed from the reactive media.
A drawback of the enzymes is that they often need long
process times, as recently illustrated in the study of Droesbeke
and Du Prez,55 where Novozym 435 was employed to
transesterify geraniol, 3,7-dimethyl-1-octanol, citronellol, and
menthol with methyl (meth)acrylates. Microwave irradiation is
intensively used in organic synthesis because of the low
reaction times needed to complete the reaction in comparison
with standard heating systems.56 A first difference between
standard heating and microwave heating is the shorter time
needed to reach the desired temperature when microwaves are
used. In addition and more importantly, for systems, which
present a polar transition state, the microwaves are able to
reduce the activation energy, enabling new reaction pathways
for lower energy.57 The efficiency of microwaves has already
been demonstrated in different systems involving enzymes.57−64 Therefore, as esterification has polar transition
states,57 microwave irradiation may greatly decrease the energy
of the transition state, thus leading to quicker reactions and
higher conversions.
The article is organized as follows (Scheme 1). Using
tetrahydrogeraniol as a representative terpenoid, first the
solvent-free synthesis of (meth)acrylate monomers by direct
esterification of tetrahydrogeraniol with (meth)acrylic acids
catalyzed by Novozym 435 under mild conditions was studied.
As methacrylic acid is not soluble in THG, methacrylic
anhydride was used, finding that the process is much faster
than using the acid. The effect of carrying out the reaction
under microwave irradiation was also studied. Afterward, the
versatility of the enzyme-catalyzed method under microwave
irradiation to synthesize a portfolio of terpenoid-based
Table 1. Chemical Structures of Terpenoids and Their
(Meth)acrylic Homologs (R = H or CH3)
The main concern is to develop a new bioinspired
sustainable approach based on industrially renewable resources
without other chemicals than the reactants, that is, free from
metals, halogens, amines, and solvent-free synthesis.
■
EXPERIMENTAL SECTION
Materials. Citronellol (95% min), Nopol (95% min), cyclademol
(95% min), and tetrahydrogeraniol (95% min) were kindly provided
by the DRT company (France). Novozym 435 (Strem Chemicals),
Candida Rugosa Lipase Type VII CRL (Sigma-Aldrich, ≥700 unit/mg
solid) immobilized on Celite 545 from Merck, acrylic/methacrylic
acids (Sigma-Aldrich, 99%), and acrylic/methacrylic anhydrides
(Sigma-Aldrich, 99%) were used without further purification.
Methods. The different experiments carried out in this study are
described in Table 2. Magnetic stirring (150 rpm) was used. The %
w/w of enzyme was calculated based on the total mass of the
reactants.
Synthesis of Monomers under Conventional Heating. The
terpenoid (20 mM) and either (meth)acrylic acid or anhydride at
different molar ratios (3:1, 2:1, and 1:1) were introduced in a 10 mL
7504
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
Research Article
Table 2. Summary of All Experimental Conditions Used in This Study
alcohol
exp
no
1
2
terpenoid
THG
acid or anhydride
acrylic
acid
acrylic
anhydride
methacrylic
acid
X
X
3
X
4
X
5
6
X
X
X
8
X
9
X
10
X
11
X
12
X
13
X
Nopol
X
15
16
X
Cyclademol
X
17
X
18
X
19
X
20
X
21
X
22
X
23
24
25
26
27
X
THG
Nopol
Cyclademol
Citronellol
X
X
X
30
31
X
X
32
33
X
X
34
35
36
enzyme
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Candida
Rugosa
Lipase
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
X
X
X
28
29
solvent
2-methyl
butan-2-ol
(mL)
Novozyme
435
Novozyme
435
Novozyme
435
7
14
methacrylic
anhydride
catalyst
X
X
X
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
7505
5
5
5
experimental conditions
molar ratio
alcohol/acid or
anhydride
enzymes
(% w/w)
drying agent
or molecular
sieves
MW
power
(watt)
T (°C)
3:1
3:1
5
3:1
5
3:1
5
3:1
3:1
5
1:1
5
40
2:1
5
40
1:1
5
40
2:1
5
2:1
5
40
1:1
5
40
2:1
5
300
40
2:1
5
300
40
2:1
5
300
40
2:1
5
40
2:1
5
40
2:1
5
50
2:1
5
60
2:1
10
40
2:1
10
60
2:1
5
2:1
5
2:1
2:1
2:1
2:1
40
40
drying agent
(10% w/w)
molecular
sieves (1 g)
40
40
300
300
300
300
40
40
40
40
40
5
300
300
300
300
40
40
40
40
2:1
5
300
40
2:1
5
200
40
2:1
5
200
40
2:1
5
150
40
2:1
5
150
40
2:1
5
50
40
2:1
5
50
40
2:1
5
40
2:1
5
40
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
Research Article
Table 2. continued
alcohol
exp
no
terpenoid
37
acid or anhydride
acrylic
acid
acrylic
anhydride
THG
solvent
enzyme
2-methyl
butan-2-ol
(mL)
experimental conditions
molar ratio
alcohol/acid or
anhydride
enzymes
(% w/w)
Novozyme
435
2:1
5
X
Novozyme
435
2:1
5
X
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
Novozyme
435
2:1
methacrylic
anhydride
X
38
39
methacrylic
acid
catalyst
40
X
41
X
42
Cyclademol
X
43
Nopol
X
44
THG
X
45
Cyclademol
X
46
Nopol
X
round-bottom flask. Novozym 435 or Candida Rugosa Lipase was then
added to the reactants and stirred to start the reaction at the selected
temperature.
Synthesis of Monomers under Microwaves Heating. The
terpenoids (20 mM) and either (meth)acrylic acid or anhydride at
different molar ratios (3:1, 2:1, and 1:1) were introduced in a 10 mL
round-bottom flask. Novozym 435 was then added to the reactants
and stirred to start the reaction at 40 °C. The reactions involving
microwaves were performed with a CEM Discover SP oven in pulse
mode (SPS) consisting a set of controlled temperature (±2 °C) by IR
detection.
Characterizations. Aliquots (50 μL) were removed from the flask
at different times. Each aliquot was analyzed by 1H NMR, 13C NMR,
and DOSY. NMR spectra were recorded at 25 °C in CDCl3 at a
concentration of 350 g L−1 on a Bruker AVANCE DPX-400 or a
DPX-500 Spectrometer. 1D 1H spectra were acquired using 32 K data
points, which were zero-filled to 64 K data points prior to Fourier
transformation. 1D 13C spectra were recorded at a 13C Larmor
frequency of 125.77 MHz using 20,000 transients. All NMR DOSY
experiments were performed using the bipolar longitudinal whirl
current delay pulse sequence (BPLED). The spoil gradients were also
applied at the diffusion period and the whirl current delay. Typically,
the gradient duration (δ) used was 2 ms, the diffusion time (Δ) was
150 ms, and the gradient strength (g) varied from 1.67 to 31.88 G
cm−1 in 32 steps. Each parameter was elected to get 95% signal
weakening for the slowest diffusion components at the last step
experiment. The pulse repetition delay (including acquisition time)
between two scans was greater than 2 s. Data acquisition and analysis
were fulfilled using the Bruker Topspin software (version 2.1). The
calculation of the diffusion coefficients and the creation of twodimensional spectra with NMR chemical shifts along one dimension
and the calculated diffusion coefficients along the other were executed
via the T1/T2 analysis module of Topspin. All final products were
characterized by 1H, DOSY, and 13C NMR and are displayed in the
Supporting Information.
drying agent
or molecular
sieves
MW
power
(watt)
T (°C)
5
200
60 for
15 min
then 40
60 for
15 min
then 40
40
2:1
5
150
40
2:1
5
50
40
2:1
5
150
40
2:1
5
150
40
2:1
5
150
40
2:1
5
150
40
2:1
5
150
40
approaches. These monomers were then used for controlled
radical polymerization in bulk and solution65 as well as in
dispersed aqueous media.32,33,62,63
Acrylic Acid. No solvent was needed for the esterification
reaction, as acrylic acid is soluble in tetrahydrogeraniol THG,
which is liquid at room temperature. Experiments 1−6 aimed
at finding the most efficient conditions to get complete
conversion in the shortest time (Figure 1). First, the effect of
Figure 1. Comparison of the different experimental conditions for the
esterification of THG with acrylic acid: control reaction, that is,
without enzyme (X; experiment 1), 5% w/w of Novozym 435 (○;
experiment 2), 5% w/w of Novozym 435 and 10% w/w of drying
agent (△; experiment 3), 5% w/w of Novozym 435 and 1 g of
molecular sieves (□; experiment 4), only 300 W microwave (+;
experiment 5), and 5% w/w and 300 W microwave (◇; experiment
6).
■
the enzyme Novozym 435 is visible by comparing the control
reaction (experiment 1) and the reaction with enzyme
(experiment 2). As water is produced in the medium during
the reaction, it can shift the equilibrium to the reverse reaction,
that is, hydrolysis of the terpene acrylate. Therefore, some
attempts to reduce the water content were tested with both a
drying agent CaSO466 (experiment 3) and molecular sieves
RESULTS AND DISCUSSION
Esterification with (Meth)acrylic Acids. The comparison
between acrylic and methacrylic acids was first studied using
tetrahydrogeraniol THG, as a model terpenoid, because the
reactivity of THG has been already studied using less “green”
7506
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
(experiment 4). Interestingly, no effect of these attempts on
the kinetics was observed. A possible explanation of these
results is that the equilibrium constant is high, and therefore,
the effect of water was less in the range of conversion studied.
This agrees with the data reported for the esterification of
butyric acid with methanol (equilibrium conversion greater
than 90%).64
Experiment 5 shows that the use of microwaves alone, that
is, without enzyme, was equivalent to the use of the enzyme
only (66% after 6 h in experiment 5 vs 69% after 6 h in
experiment 2). On the other hand, the combination of
enzymes and microwaves allowed reaching complete conversion in only 5 h (experiment 6). This experiment clearly
demonstrates that the combination of enzyme and microwaves
is definitively a good solution to improve the esterification
kinetics. This agrees with what has been reported for the
synthesis of cosmetic esters.67 In addition, experiment 6
supports the hypothesis that the equilibrium is shifted toward
the ester as full conversion was reached. Achieving complete
conversion indicates that water was removed from the reaction
medium. This might be due to the combined effect of trapping
of some water by the enzyme, and water evaporation due to
hot spots in this heterogeneous system present under
microwave irradiation might have an effect. Xu et al.68 showed
that a certain amount of water is required around the enzyme
to make the enzymatic activity to its best. A quick calculation
shows that the amount of water produced during the reaction
is of the same magnitude as the one for which Xu et al. found
the best enzyme activity. Moreover, it has been proven that
under microwave irradiation in a heterogeneous system, hot
spots where the temperature could increase up to 20% would
lead to water evaporation in the reaction medium.57
Methacrylic Acid. Methacrylic acid is not soluble in
tetrahydrogeraniol; therefore, a solvent was added into the
medium. In order to stay in a green approach, an environmentfriendly solvent, if possible biobased, would be preferred. 2Methyl-2-butanol, which can be obtained from biomass, was
chosen as a solvent69 (experiment 7). This tertiary alcohol was
expected to be less reactive than THG regarding the enzyme.
As shown in Figure 2, this was the case, although still a
noticeable amount of 2-methyl-2-butanol methacrylate (2.5%)
was formed. Even though this is a biosourced methacrylate
that, in some cases, may be used together with the
tetrahydrogeraniol methacrylate, for fine-tuning the properties
of the polymers, it is desirable to have pure monomers.
Although nonreactive solvents could be used, the unavoidable
removal of the solvent at the end of the process will increase
the environmental impact of the process. Therefore, alternative
routes using (meth)acrylic anhydrides instead of (meth)acrylic
acids for the synthesis of tetrahydrogeraniol (meth)acrylate
were explored.
Esterification with (Meth)acrylic Anhydrides. Experiment 8 was a Novozym 435 catalyzed reaction of methacrylic
anhydride and THG in stoichiometric amounts. No solvent
was needed because the methacrylic anhydride is soluble in
THG. Experiment 9 was a comparative reaction using
methacrylic acid instead of methacrylic anhydride (Figure 3).
Research Article
Figure 3. Comparison between the use of methacrylic anhydride (□;
experiment 8) and methacrylic acid (◇; experiment 9) in the
presence of Novozym435 at 40 °C.
It can be seen that using methacrylic anhydride, almost full
conversion was achieved in about 1 h in a solvent-free process,
whereas with methacrylic acid, a solvent was needed and the
conversion was only 60% in 6 h. Therefore, the use of
methacrylic anhydride leads to a fast esterification. In addition,
a pure THG methacrylate was obtained as shown by the 1H
and 13C NMR spectra presented in Figure 4. This figure also
includes the DOSY spectrum that demonstrates the presence
of only one diffusion coefficient which is a fingerprint of a
unique molecular structure of the obtained (meth)acrylic
monomers. Notice that in the absence of side products, yield =
conversion. A combination of enzymatic catalysis and microwave irradiation was used in experiment 10, finding that a
complete reaction was achieved in only 5 min.
Encouraged by the results obtained with the methacrylates,
the possibility of using acrylic anhydride instead of acrylic acid
was explored. Experiments 11 and 12 were Novozym 435
catalyzed reactions of stoichiometric amounts of THG and
acrylic anhydride and acrylic acid, respectively. No solvent was
used in these reactions. Complete conversion was reached after
24 h for the anhydride and in 48 h for the acid. This shows that
also for the acrylates, the reactivity of the anhydride is higher
than that of the acid. The use of microwaves in combination
with Novozym 435 (experiment 13) was also beneficial, and
100% conversion and then yield were obtained in only 90 min.
Figure S1 presents the 1H, 13C, and DOSY NMR spectra of the
product of the reaction. It can be seen that a pure THG
acrylate was obtained. As the reaction was performed with the
stoichiometric ratio between the terpenoids and the acyl group
and a pure THG acrylate was obtained, the full conversion
indicates 100% of yield. This is a tremendous practical
Figure 2. Evolution of THG methacrylate (◇; experiment 9) and 2methyl-2-butanol methacrylate (□; experiment 7) conversion during
the esterification reaction between THG and methacrylic acid at 40
°C with 5% w/w of Novozym 435.
7507
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
Research Article
adhesives,33 biofuels,69 or cosmetics.62 These applications
require widely different properties that are obtained by
copolymerization of a relatively small group of (meth)acrylic
monomers. The main components are monomers that yield
polymers with high glass-transition temperatures, Tg. In
practice, most of the formulations are based on methyl
methacrylate (Tg = 105 °C) and soft monomers that yield low
Tg polymers. In practice, the most often used acrylates are
butyl acrylate (Tg = −54 °C) and 2-ethylhexyl acrylate (Tg =
−65 °C). In addition, small fractions of functional monomers
(e.g., carboxyl-containing monomers and cross-linkers) are
used. For example, by only varying the monomer ratio, both
adhesives and coatings can be obtained with the hard methyl
methacrylate and the soft butyl acrylate. Therefore, in order to
substitute the oil-based (meth)acrylic monomers by biomassbased ones, a portfolio of these biomass-based monomers
should be produced. The first two components of this portfolio
are the THG acrylates and methacrylates synthesized above.
Because of the long-branched alkyl chain, they are soft
monomers, the acrylate being softer (Tg acrylate = −46 °C65 and
Tg methacrylate = −30 °C).32 Therefore, harder biobased
monomers are needed. Fortunately, the paper industry
provides a broad variety of renewable C10 terpenoids with
different chemical structures.
Among the possible alternatives, in this work, we focused on
three additional terpenoids that are complementary to THG
(Table 1). Nopol and cyclademol33 have a cyclic structure and
lead to hard (high Tg) (meth)acrylates, which could be
compared to the hard benzyl(meth)acrylate and isobornyl(meth)acrylate monomers, respectively. Moreover, Nopol also
has intracyclic unsaturation that could be used for cross-linking
by postmodification. Citronellol will yield (meth)acrylic
monomers with a branched ester chain that from the point
of view of the Tg will be similar to THG. However, its alkene
moiety opens many possibilities such as acting as a cross-linker
during polymerization as well as allowing the postmodification
of the polymer after the polymerization if the double bond can
be preserved.
Nopol. The combination of microwaves and enzyme was
employed for the synthesis of (meth)acrylic Nopol (experiments 14 and 15). A complete reaction was obtained after 110
min for both acrylic and methacrylic Nopol while preserving
intracyclic unsaturation (1H, 13C, and DOSY NMR spectrum
in Figure S2).
Cyclademol. Cyclademol is a secondary alcohol, which is
expected to be less reactive than the primary alcohols of THG
and Nopol. Therefore, Candida Rugosa Lipase, an enzyme that
is claimed to be appropriate for secondary alcohols,21 was used
in experiment 16. However, the reaction was very slow and
only about 8% conversion was achieved in 7 h (esterification
conversion vs time shown in Figure S3).
An attempt with Novozym 435 (experiment 17) was made
under similar conditions, and the reaction rate was twice faster
(42% of conversion after 24 h) even if Novozym 435 is not
considered suitable for secondary alcohols. The reaction was
performed at higher temperatures (50 and 60 °C) (experiments 18 and 19), leading to a conversion of 52 and 58%,
respectively, after 24 h. This shows that conversion was not
significantly influenced by temperature. Higher temperatures
were not used to avoid denaturation of the enzyme. An
experiment using twice the concentration of enzyme (10% w/
w) was carried out at 40 °C (experiment 20) and also yielded a
very modest increase of the conversion (up to 55% after 24 h).
Figure 4. (a) 1H NMR spectrum, (b) 13C NMR spectrum, and (c)
DOSY spectrum of THG methacrylate (from experiment 8) done in
CDCl3 without purification step.
advantage as no complex purification process to remove the
solvent, coreactants, or side products is needed; just filtration
to remove the supported enzyme is enough.
This leads to the conclusion that anhydrides are more
convenient than acids for the synthesis of THG (meth)acrylate
using enzyme-catalyzed esterification. The difference could be
explained by the reaction mechanism, which is based on the
formation of the acyl-enzyme intermediate formed between the
anhydride and the enzyme.68 This unstable intermediate reacts
with the alcohol to form the ester and with water to produce
the acid. As the anhydride moiety is very reactive, esterification
with the anhydride is the preferential route. The water
produced in situ by esterification with methacrylic acid
seems to be high enough to allow the full activity of the
enzyme (Scheme S1).
In conclusion, the study carried out with THG shows that
the most efficient way of synthesizing the terpenoid-based
(meth)acrylates is by a combination of enzymatic catalysis and
microwave irradiation using anhydrides as acyl donors instead
of acids. This combination will be used in the following
experiments.
Application to Other Terpenoids. (Meth)acrylic polymers are used in a variety of applications including coatings,68
7508
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
An increase of the temperature to 60 °C with 10% w/w of
enzyme (experiment 21) led to 71% after 24 h, but the kinetics
still remained slower than for primary alcohols. A combination
of Novozym 435 and microwaves at 40 °C (experiment 22)
was found to be more efficient as 83% of cyclademol was
converted into methacrylate in 60 min and resulted in a
complete reaction after 2 h. However, comparison with
experiment 10 shows that the reaction was still slow compared
to primary alcohols. Microwaves were applied to the Novozym
435 catalyzed reaction of cyclademol and acrylic anhydride
(experiment 23). Complete reaction was obtained in 150 min
with the use of a microwave at 300 W, and the NMR spectra
show that pure cyclademol (meth)acrylate was obtained (1H,
13
C, and DOSY NMR spectrum shown in Figure S4).
A comparison of the reactivity of different terpenoids (THG,
Nopol, and cyclademol) with methacrylic anhydride using only
microwave irradiation (without enzyme) was made in
experiments 24−26. Total conversions were reached after 60,
110, and 240 min for THG, Nopol, and cyclademol,
respectively (Figure 5). However, the differences between
conversion was obtained (Figure 6). No gel was observed,
and the NMR spectra clearly show that the double bond was
Research Article
Figure 6. Evolution of the esterification conversion vs time between
citronellol and methacrylic anhydride with 5% w/w of Novozym 435
with temperature program (◇, experiment 38) and without
temperature program (◇, experiment 36).
totally preserved (1H NMR spectra shown in Figure S2). In
order to achieve complete conversion, a temperature program
was used in which the first 15 min were conducted at 60 °C
and then the temperature was reduced to 40 °C (experiments
37 and 38). A full conversion was achieved after 120 min for
citronellol methacrylate (Figure 6). In addition, no gel was
formed, and the NMR spectra show the presence of
unsaturation (1H, 13C, and DOSY NMR spectrum shown in
Figure S5). The reaction involving citronellol acrylate leads to
the same type of kinetics.
These results show that microwave-assisted enzymatic
esterification of terpenoids is an efficient method to prepare
a portfolio of monomers that can lead to polymers with Tgs,
covering the range of Tgs of the most often used oil-based
(meth)acrylates.
Influence of Microwave Power. As the microwave power
employed in the previous experiments was quite high (300 W),
an attempt to reduce the power was made using a model
system, tetrahydrogeraniol and methacrylic anhydride. For this
purpose, 300, 200, 150, and 50 W were applied (experiments
10 and 39−41). No significant difference was observed
between 300, 200, and 150 W (Figure 7), whereas reaction
was slower at 50 W but still reached full conversion and 100%
yield. This means that there is room to make the synthesis
even less energy-consuming. A 150 W irradiation was applied
to other terpenoids (cyclademol, experiment 42; and Nopol,
experiment 43), and no change in kinetics with respect to 300
W was observed (complete reaction observed after 120 and
110 min, respectively). This decrease of the microwave power
was also tested on systems using acrylic anhydride as acyl
donor (THG, experiment 44, cyclademol, experiment 45; and
Nopol, experiment 46). Similarly, to the methacrylic anhydride
systems, no change in the kinetics with respect to 300 W was
observed. Total reaction was achieved in 30, 150, and 110 min,
respectively.
Polymerization of Monomers. The ability of the
synthesized acrylic monomers to polymerize was checked in
free radical polymerizations carried out using AIBN as an
initiator (monomer/AIBN = 516) during 2 h at 80 °C. The
results are summarized in Table 3. It can be seen that all of the
Figure 5. Evolution of the esterification conversion vs time between
methacrylic anhydride and (◇) THG, (□) Nopol, and (△)
cyclademol at 40 °C under 300 W microwave radiation.
primary and secondary alcohols were maintained; that is,
cyclademol esterification is still the slowest, although the gap
between primary and secondary alcohol is far smaller with the
use of microwaves. This illustrates that Novozym 435 is more
suitable to catalyze esterification of primary alcohols instead of
secondary ones even under microwaves.
Citronellol. The synthesis of (meth)acrylic citronellol was
performed using the enzyme/microwave method described
above (citronellol acrylate, experiment 27; and citronellol
methacrylate, experiment 28). Both reactions led to gel
formation. Citronellol is similar to tetrahydrogeraniol except
for the presence of unsaturation. Gel formation clearly points
out that the alkene moiety reacted. Therefore, different
microwave powers were applied to the system (experiments
29−34). However, gel was formed in all cases. This means that
the microwave irradiation was too energetic for the proper
esterification of citronellol. The differences with Nopol were
due to the easier activation of the citronellol double bond as
compared with Nopol whose alkene group is intracyclic.
Consequently, reactions without microwaves were carried
out to preserve the double CC bond. Experiments 35 and 36
were carried out at a constant temperature of 40 °C for acrylic
and methacrylic anhydrides, respectively. The reaction was
faster for the methacrylic anhydride, but no complete
7509
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
However, microwaves were not useful for citronellol as they
activated its double bond, yielding a cross-linked polymer.
Therefore, for the synthesis of citronellol (meth)acrylates, a
temperature program was used instead of microwaves. In all
cases, high-purity monomers and full conversion were
obtained. The monomers can be easily polymerized by free
radical polymerization.
The monomers of this work include both soft [tetrahydrogeraniol (meth)acrylates]4 and hard [cyclademol32,33 and
Nopol (meth)acrylates] monomers as well as cross-linking
[citronellol (meth)acrylates] monomers. This green portfolio
of monomers would give polymers with Tgs in the same range
as the currently used petroleum-based poly(meth)acrylics,
paving the way toward the substitution of petroleum-based
monomers for a new class of (meth)acrylic monomers from
sustainable reactions and renewable resources.
Figure 7. Evolution of the esterification conversion vs time between
THG and methacrylic anhydride at 40 °C in the presence of
Novozym 435 under (◇) 300, (◇) 200, (△) 150, and (○) 50 W
microwaves.
■
ASSOCIATED CONTENT
sı Supporting Information
*
monomers showed a good activity for polymerization and that
high molecular weights were obtained.
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acssuschemeng.0c02307.
1
H, 13C, and DOSY NMR spectrum of THG (meth)acrylate, Nopol (meth)acrylate, cyclademol (meth)acrylate, citronellol (meth)acrylate, conversion versus
time of the cyclademol/methacrylic anhydride esterification with lipase, and hypothetic enzyme-catalyzed
esterification mechanism (PDF)
Table 3. Free Radical Polymerizations Carried Out at 80 °C
with the Raw Acrylic Monomers Synthesized in This Work
(as Synthesized without Any Purification Step)
monomer
conversion (%)
molar mass (g mol−1)
Dispersity Đ
THG
Citronellol
Nopol
Cyclademol
80
78
75
77
215,000
205,000
250,000
211,000
2.6
2.4
2.4
2.5
Research Article
■
AUTHOR INFORMATION
Corresponding Authors
José M. Asua − Kimika Aplikatua Saila, Kimika Zientzien
Fakultatea, Joxe Mari Korta Zentroa, POLYMAT, University of
the Basque Country UPV/EHU, 20018 Donostia-San
Sebastián, Spain; orcid.org/0000-0002-7771-1543;
Email: jm.asua@ehu.eus
Laurent Billon − Universite de Pau & des Pays de l’Adour, E2S
UPPA, CNRS, IPREM-UMR 5254, 64000 Pau, France; BioInspired Materials Group: Functionalities & Self-Assembly,
Université de Pau & des Pays de l’Adour, E2S UPPA, 64000
Pau, France; orcid.org/0000-0003-0999-899X;
Email: laurent.billon@univ-pau.fr
■
CONCLUSIONS
This study reports on an environmentally friendly method
based on bioinspired enzymatic catalysis to synthesize a new
set of (meth)acrylic monomers from terpenoids extracted from
wood biomass. The method is solvent-, metal-, amino-, and
halogen-free. The monomers were synthesized by esterification
of terpenoids (tetrahydrogeraniol, citronellol, cyclademol, and
Nopol) using both (meth)acrylic acids and anhydrides. It was
found that anhydrides were more reactive.
Microwave irradiation strongly accelerated the esterification
and was beneficial for tetrahydrogeraniol, cyclademol, and
Nopol, achieving pure final products with complete conversion
and high yield in a short period, from 5 to 150 min depending
on the alcohol and acyl nature. Primary alcohols and
methacrylates required less time, and the best conditions
found in this work are summarized in Table 4.
Authors
Thibault Castagnet − Universite de Pau & des Pays de l’Adour,
E2S UPPA, CNRS, IPREM-UMR 5254, 64000 Pau, France;
Bio-Inspired Materials Group: Functionalities & Self-Assembly,
Université de Pau & des Pays de l’Adour, E2S UPPA, 64000
Pau, France; Kimika Aplikatua Saila, Kimika Zientzien
Fakultatea, Joxe Mari Korta Zentroa, POLYMAT, University of
the Basque Country UPV/EHU, 20018 Donostia-San
Sebastián, Spain
Garbine Aguirre − Universite de Pau & des Pays de l’Adour,
E2S UPPA, CNRS, IPREM-UMR 5254, 64000 Pau, France;
Bio-Inspired Materials Group: Functionalities & Self-Assembly,
Université de Pau & des Pays de l’Adour, E2S UPPA, 64000
Pau, France
Table 4. Best Conditions for Monomer Syntheses with
Enzyme Novozym 435 and Anhydride
time (min)
terpenoid
microwave power
(watts)
temperature
(°C)
acrylate
methacrylate
THG
Citronellol
Nopol
Cyclademol
150
n.a.
150
150
40
60a/40
40
40
30
120
110
150
5
120
110
120
Complete contact information is available at:
https://pubs.acs.org/10.1021/acssuschemeng.0c02307
Notes
The authors declare no competing financial interest.
a
15 min.
7510
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
ACKNOWLEDGMENTS
The PhD fellowship of T.C. was supported by the UPPA/UPV
funding.
ization Using a Well-Defined Block Glycopolymer Stabilizer. Soft
Matter 2011, 7, 2493−2499.
(19) León, O.; Bordegé, V.; Muñoz-Bonilla, A.; Sánchez-Chaves, M.;
Fernández-García, M. Well-Controlled Amphiphilic Block Glycopolymers and Their Molecular Recognition with Lectins. J. Polym. Sci.,
Part A: Polym. Chem. 2010, 48, 3623−3631.
(20) Nekrasova, T. N.; Andreeva, L. N.; Nazarova, O. V.; Bezrukova,
M. A.; Zolotova, Y. I.; Imanbaev, R. T.; Skorbunova, O. V.; Pautov, V.
D.; Panarin, E. F. Structural and Dynamic Characteristics of Thermoand PH-Sensitive Copolymers of 2-(Diethylamino)Ethyl Methacrylate and 2-Deoxy-2-Methacrylamido-d-Glucose. Polymer 2015, 77,
246−253.
(21) Miyata, T.; Jikihara, A.; Nakamae, K.; Hoffman, A. S.
Preparation of Poly(2-Glucosyloxyethyl Methacrylate)-Concanavalin
A Complex Hydrogel and Its Glucose-Sensitivity. Macromol. Chem.
Phys. 1996, 197, 1135−1146.
(22) Galià, M.; de Espinosa, L. M.; Ronda, J. C.; Lligadas, G.; Cádiz,
V. Vegetable Oil-Based Thermosetting Polymers. Eur. J. Lipid Sci.
Technol. 2010, 112, 87−96.
(23) Lligadas, G.; Ronda, J. C.; Galià, M.; Cádiz, V. Renewable
Polymeric Materials from Vegetable Oils: A Perspective. Mater. Today
2013, 16, 337−343.
(24) Decostanzi, M.; Lomège, J.; Ecochard, Y.; Mora, A.-S.; Negrell,
C.; Caillol, S. Fatty Acid-Based Cross-Linkable Polymethacrylate
Coatings. Prog. Org. Coat. 2018, 124, 147−157.
(25) Wang, S.; Ding, W.; Yang, G.; Robertson, M. L. Biorenewable
Thermoplastic Elastomeric Triblock Copolymers Containing Salicylic
Acid-Derived End-Blocks and a Fatty Acid-Derived Midblock.
Macromol. Chem. Phys. 2016, 217, 292−303.
(26) de Espinosa, L. M.; Ronda, J. C.; Galià, M.; Cádiz, V. A New
Route to Acrylate Oils: Crosslinking and Properties of Acrylate
Triglycerides from High Oleic Sunflower Oil. J. Polym. Sci., Part A:
Polym. Chem. 2009, 47, 1159−1167.
(27) De Espinosa, L. M.; Ronda, J. C.; GaliÀ , M.; Cádiz, V. A
Straightforward Strategy for the Efficient Synthesis of Acrylate and
Phosphine Oxide-Containing Vegetable Oils and Their Crosslinked
Materials. J. Polym. Sci., Part A: Polym. Chem. 2009, 47, 4051−4063.
(28) Müller, R.; Wilke, G. Synthesis and Radiation Curing of
Acrylated Castor Oil Glycerides. J. Coatings Technol. Res. 2014, 11,
873−882.
(29) Pelletier, H.; Belgacem, N.; Gandini, A. Acrylated Vegetable
Oils as Photocrosslinkable Materials. J. Appl. Polym. Sci. 2006, 99,
3218−3221.
(30) Moreno, M.; Goikoetxea, M.; Barandiaran, M. J. BiobasedWaterborne Homopolymers from Oleic Acid Derivatives. J. Polym.
Sci., Part A: Polym. Chem. 2012, 50, 4628−4637.
(31) Black, M.; Messman, J.; Rawlins, J. W. Chain Transfer of
Vegetable Oil Macromonomers in Acrylic Solution Copolymerization.
J. Appl. Polym. Sci. 2011, 120, 1390−1396.
(32) Noppalit, S.; Simula, A.; Billon, L.; Asua, J. M. On the Nitroxide
Mediated Polymerization of Methacrylates Derived from Bio-Sourced
Terpenes in Miniemulsion, a Step towards Sustainable Products.
Polym. Chem. 2020, 11, 1151.
(33) Noppalit, S.; Simula, A.; Billon, L.; Asua, J. M. Paving the Way
to Sustainable Waterborne Pressure-Sensitive Adhesives Using
Terpene-Based Triblock Copolymers. ACS Sustainable Chem. Eng.
2019, 7, 17990−17998.
(34) Yao, K.; Tang, C. Controlled Polymerization of NextGeneration Renewable Monomers and Beyond. Macromolecules
2013, 46, 1689−1712.
(35) Zhao, J.; Schlaad, H. Synthesis of Terpene-Based Polymers. In
Bio-Synthetic Polymer Conjugates; Schlaad, H., Ed.; Springer Berlin
Heidelberg: Berlin, Heidelberg, 2013; pp 151−190.
(36) Bolton, J. M.; Hillmyer, T. R.; Hoye, R. T. Sustainable
Thermoplastic Elastomers from Terpene-Derived Monomers. ACS
Macro Lett. 2014, 3, 717−720.
(37) Winnacker, M.; Rieger, B. Recent Progress in Sustainable
Polymers Obtained from Cyclic Terpenes: Synthesis, Properties, and
Application Potential. ChemSusChem 2015, 8, 2455−2471.
■
■
REFERENCES
(1) Wilbon, P. A.; Chu, F.; Tang, C. Progress in Renewable
Polymers from Natural Terpenes, Terpenoids, and Rosin. Macromol.
Rapid Commun. 2013, 34, 8−37.
(2) Thomsett, M. R.; Storr, T. E.; Monaghan, O. R.; Stockman, R.
A.; Howdle, S. M. Progress in the Synthesis of Sustainable Polymers
from Terpenes and Terpenoids. Green Mater. 2016, 4, 115−134.
(3) Ando, S.; Someya, Y.; Takahashi, T.; Shibata, M. Thermal and
Mechanical Properties of Photocured Organic−Inorganic Hybrid
Nanocomposites of Terpene-Based Acrylate Resin and MethacrylateSubstituted Polysilsesquioxane. J. Appl. Polym. Sci. 2010, 115, 3326−
3331.
(4) Stößer, T.; Li, C.; Unruangsri, J.; Saini, P. K.; Sablong, R. J.;
Meier, M. A. R.; Williams, C. K.; Koning, C. Bio-Derived Polymers for
Coating Applications: Comparing Poly(Limonene Carbonate) and
Poly(Cyclohexadiene Carbonate). Polym. Chem. 2017, 8, 6099−6105.
(5) Firdaus, M.; Montero De Espinosa, L.; Meier, M. A. R. TerpeneBased Renewable Monomers and Polymers via Thiol-Ene Additions.
Macromolecules 2011, 44, 7253−7262.
(6) Min, E. H.; Wong, K. H.; Setijadi, E.; Ladouceur, F.; Straton, M.;
Argyros, A. Menthol-Based Chiral Copolymers for Polymer Optical
Fibres (POF). Polym. Chem. 2011, 2, 2045−2051.
(7) Alvès, M.-H.; Sfeir, H.; Tranchant, J.-F.; Gombart, E.; Sagorin,
G.; Caillol, S.; Billon, L.; Save, M. Terpene and Dextran Renewable
Resources for the Synthesis of Amphiphilic Biopolymers. Biomacromolecules 2014, 15, 242−251.
(8) Wilbon, P. A.; Zheng, Y.; Yao, K.; Tang, C. Renewable Rosin
Acid-Degradable Caprolactone Block Copolymers by Atom Transfer
Radical Polymerization and Ring-Opening Polymerization. Macromolecules 2010, 43, 8747−8754.
(9) Athawale, V.; Manjrekar, N.; Athawale, M. Lipase-Catalyzed
Synthesis of Geranyl Methacrylate by Transesterification : Study of
Reaction Parameters. Tetrahedron Lett. 2002, 43, 4797−4800.
(10) Athawale, V.; Manjrekar, N.; Athawale, M. Effect of Reaction
Parameters on Synthesis of Citronellyl Methacrylate by LipaseCatalyzed Transesterification. Biotechnol. Prog. 2003, 19, 298−302.
(11) Sainz, M. F.; Souto, J. A.; Regentova, D.; Johansson, M. K. G.;
Timhagen, S. T.; Irvine, D. J.; Buijsen, P.; Koning, C. E.; Stockman, R.
A.; Howdle, S. M. A Facile and Green Route to Terpene Derived
Acrylate and Methacrylate Monomers and Simple Free Radical
Polymerisation to Yield New Renewable Polymers and Coatings.
Polym. Chem. 2016, 7, 2882−2887.
(12) Noppalit, S.; Simula, A.; Ballard, N.; Callies, X.; Asua, J. M.;
Billon, L. Renewable Terpene Derivative as a Biosourced Elastomeric
Building Block in the Design of Functional Acrylic Copolymers.
Biomacromolecules 2019, 20, 2241−2251.
(13) Kloosterman, W. M. J.; Jovanovic, D.; Brouwer, S. G. M.; Loos,
K. Amylase Catalyzed Synthesis of Glycosyl Acrylates and Their
Polymerization. Green Chem. 2014, 16, 203−210.
(14) Kloosterman, W. M. J.; Brouwer, S. G. M.; Loos, K. EnzymeCatalyzed Synthesis of Saccharide Acrylate Monomers from Nonedible Biomass. Chem.Asian J. 2014, 9, 2156−2161.
(15) Kloosterman, W. M. J.; Roest, S.; Priatna, S. R.; Stavila, E.;
Loos, K. Chemo-Enzymatic Synthesis Route to Poly(GlucosylAcrylates) Using Glucosidase from Almonds. Green Chem. 2014, 16,
1837−1846.
(16) Akashi, M.; Sakamoto, N.; Suzuki, K.; Kishida, A. Synthesis and
Anticoagulant Activity of Sulfated Glucoside-Bearing Polymer.
Bioconjugate Chem. 1996, 7, 393−395.
(17) Raynor, J. E.; Petrie, T. A.; Fears, K. P.; Latour, R. A.; García, A.
J.; Collard, D. M. Saccharide Polymer Brushes to Control Protein and
Cell Adhesion to Titanium. Biomacromolecules 2009, 10, 748−755.
(18) Muñoz-Bonilla, A.; Heuts, J. P. A.; Fernández-García, M.
Glycoparticles and Bioactive Films Prepared by Emulsion Polymer7511
Research Article
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
ACS Sustainable Chemistry & Engineering
pubs.acs.org/journal/ascecg
(38) Llevot, A.; Dannecker, P.-K.; von Czapiewski, M.; Over, L. C.;
Söyler, Z.; Meier, M. A. R. Renewability Is Not Enough: Recent
Advances in the Sustainable Synthesis of Biomass-Derived Monomers
and Polymers. Chem.Eur. J. 2016, 22, 11510−11521.
(39) Weems, A. C.; Delle Chiaie, K. R.; Worch, J. C.; Stubbs, C. J.;
Dove, A. P. Terpene- and Terpenoid-Based Polymeric Resins for
Stereolithography 3D Printing. Polym. Chem. 2019, 10, 5959−5966.
(40) Black, M.; Rawlins, J. W. Thiol-Ene UV-Curable Coatings
Using Vegetable Oil Macromonomers. Eur. Polym. J. 2009, 45, 1433−
1441.
(41) Okoye, P. U.; Abdullah, A. Z.; Hameed, B. H. Synthesis of
Oxygenated Fuel Additives via Glycerol Esterification with Acetic
Acid over Bio-Derived Carbon Catalyst. Fuel 2017, 209, 538−544.
(42) Esfandmaz, S.; Chaibakhsh, N.; Moradi-Shoeili, Z.;
Mohammadi, A. Eco-Friendly Synthesis of Maleate Ester: A
Comparison between Solid Acid and Enzyme-Catalyzed Esterification. Sustainable Chem. Pharm. 2018, 8, 82−87.
(43) Natalino, R.; Varejão, E. V. V.; Da Silva, M. J.; Cardoso, A. L.;
Fernandes, S. A. P-Sulfonic Acid Calix[n]Arenes: The Most Active
and Water Tolerant Organocatalysts in Esterification Reactions. Catal.
Sci. Technol. 2014, 4, 1369−1375.
(44) Kaur, K.; Jain, P.; Sobti, A.; Toor, A. P. Sulfated Metal Oxides:
Eco-Friendly Green Catalysts for Esterification of Nonanoic Acid with
Methanol. Green Process. Synth. 2016, 5, 93−100.
(45) Gupta, P.; Paul, S. Solid Acids: Green Alternatives for Acid
Catalysis. Catal. Today 2014, 236, 153−170.
(46) Bauwelinck, J.; Cornet, I.; Wijnants, M.; Dams, R.; Tavernier, S.
Investigation of the Enzyme-Catalysed Transesterification of Methyl
Acrylate and Sterically Hindered Alcohol Substrates. ChemistrySelect
2018, 3, 5169−5175.
(47) Pereira, G. N.; Holz, J. P.; Giovannini, P. P.; Oliveira, J. V.; de
Oliveira, D.; Lerin, L. A. Enzymatic Esterification for the Synthesis of
Butyl Stearate and Ethyl Stearate. Biocatal. Agric. Biotechnol. 2018, 16,
373−377.
(48) Perin, G. B.; Felisberti, M. I. Enzymatic Synthesis and
Structural Characterization of Methacryloyl-D-Fructose- and Methacryloyl-D-Glucose-Based Monomers and Poly(Methacryloyl-D-Fructose)-Based Hydrogels. J. Chem. Technol. Biotechnol. 2018, 93, 1694−
1704.
(49) Karaki, N.; Aljawish, A.; Humeau, C.; Muniglia, L.; Jasniewski,
J. Enzymatic Modification of Polysaccharides: Mechanisms, Properties, And Potential Applications: A Review. Enzyme Microb. Technol.
2016, 90, 1−18.
(50) Lopresto, C. G.; Calabrò, V.; Woodley, J. M.; Tufvesson, P.
Kinetic Study on the Enzymatic Esterification of Octanoic Acid and
Hexanol by Immobilized Candida Antarctica Lipase B. J. Mol. Catal.
B: Enzym. 2014, 110, 64−71.
(51) Molina-Gutiérrez, M.; Hakalin, N. L. S.; Rodríguez-Sanchez, L.;
Prieto, A.; Martínez, M. J. Green Synthesis of β-Sitostanol Esters
Catalyzed by the Versatile Lipase/Sterol Esterase from Ophiostoma
Piceae. Food Chem. 2017, 221, 1458−1465.
(52) Chojnacka, A.; Obara, R.; Wawrzeńczyk, C. Kinetic Resolution
of Racemic Secondary Aliphatic Allylic Alcohols in Lipase-Catalyzed
Transesterification. Tetrahedron: Asymmetry 2007, 18, 101−107.
(53) Aghababaie, M.; Beheshti, M.; Razmjou, A.; Bordbar, A.-K.
Enzymatic Biodiesel Production from Crude Eruca Sativa Oil Using
Candida Rugosa Lipase in a Solvent-Free System Using Response
Surface Methodology. Biofuels 2017, 11, 93−99.
(54) Gofferjé, G.; Stäbler, A.; Herfellner, T.; Schweiggert-Weisz, U.;
Flöter, E. Kinetics of Enzymatic Esterification of Glycerol and Free
Fatty Acids in Crude Jatropha Oil by Immobilized Lipase from
Rhizomucor Miehei. J. Mol. Catal. B: Enzym. 2014, 107, 1−7.
(55) Droesbeke, M. A.; Du Prez, F. E. Sustainable Synthesis of
Renewable Terpenoid-Based (Meth)Acrylates Using the CHEM21
Green Metrics Toolkit. ACS Sustainable Chem. Eng. 2019, 7, 11633−
11639.
(56) Perreux, L.; Loupy, A.; Petit, A. Nonthermal Effects of
Microwaves in Organic Synthesis. In Microwaves in Organic Synthesis;
3rd ed.; Wiley-VCH, 2013; Vol. 1, pp 127−207.
(57) Major, B.; Kelemen-Horváth, I.; Csanádi, Z.; Bélafi-Bakó, K.;
Gubicza, L. Microwave Assisted Enzymatic Esterification of Lactic
Acid and Ethanol in Phosphonium Type Ionic Liquids as CoSolvents. Green Chem. 2009, 11, 614−616.
(58) Mulalee, S.; Srisuwan, P.; Phisalaphong, M. Influences of
Operating Conditions on Biocatalytic Activity and Reusability of
Novozym 435 for Esterification of Free Fatty Acids with Short-Chain
Alcohols: A Case Study of Palm Fatty Acid Distillate. Chin. J. Chem.
Eng. 2015, 23, 1851−1856.
(59) Rajan, A.; Prasad, V. S.; Emilia Abraham, T. Enzymatic
Esterification of Starch Using Recovered Coconut Oil. Int. J. Biol.
Macromol. 2006, 39, 265−272.
(60) Roy, I.; Gupta, M. N. Non-Thermal Effects of Microwaves on
Protease-Catalyzed Esterification and Transesterification. Tetrahedron
2003, 59, 5431−5436.
(61) Huang, W.; Xia, Y.-M.; Gao, H.; Fang, Y.-J.; Wang, Y.; Fang, Y.
Enzymatic Esterification between N-Alcohol Homologs and nCaprylic Acid in Non-Aqueous Medium under Microwave Irradiation.
J. Mol. Catal. B: Enzym. 2005, 35, 113−116.
(62) Rejasse, B.; Lamare, S.; Legoy, M.-D.; Besson, T. Influence of
Microwave Irradiation on Ezymatic Properties: Applications in
Enzyme Chemistry. J. Enzyme Inhib. Med. Chem. 2007, 22, 519−527.
(63) Lukasiewicz, M.; Kowalski, S. Low Power Microwave-Assisted
Enzymatic Esterification of Starch. Starch/Staerke 2012, 64, 188−197.
(64) Dange, P. N.; Rathod, V. K. Equilibrium and Thermodynamic
Parameters for Heterogeneous Esterification of Butyric Acid with
Methanol under Microwave Irradiation. Resour. Technol. 2017, 3, 64−
70.
(65) Noppalit, S.; Simula, A.; Ballard, N.; Callies, X.; Asua, J. M.;
Billon, L. Renewable Terpene Derivative as a Biosourced Elastomeric
Building Block in the Design of Functional Acrylic Copolymers.
Biomacromolecules 2019, 20, 2241.
(66) Warren, J. C.; Cheatum, S. G. Effect of Neutral Salts on
Enzyme Activity and Structure. Biochemistry 1966, 5, 1702−1707.
(67) Khan, N. R.; Rathod, V. K. Enzyme Catalyzed Synthesis of
Cosmetic Esters and Its Intensification: A Review. Process Biochem.
2015, 50, 1793−1806.
(68) Xu, J. H.; Kawamoto, T.; Tanaka, A. Efficient Kinetic
Resolution of Dl-Menthol by Lipase-Catalyzed Enantioselective
Esterification with Acid Anhydride in Fed-Batch Reactor. Appl.
Microbiol. Biotechnol. 1995, 43, 402−407.
(69) Sarathy, S. M. Fuel Class Higher Alcohols. In Biofuels from
Lignocellulosic Biomass; John Wiley & Sons, Ltd, 2016; pp 29−57.
Research Article
■
NOTE ADDED AFTER ASAP PUBLICATION
This paper was published ASAP on May 5, 2020. Scheme 1
and Table 1 graphics were updated. The revised paper was
reposted on May 7, 2020.
7512
https://dx.doi.org/10.1021/acssuschemeng.0c02307
ACS Sustainable Chem. Eng. 2020, 8, 7503−7512
Download