You are on page 1of 6

Research Article

Cite This: ACS Sustainable Chem. Eng. 2019, 7, 9422−9427 pubs.acs.org/journal/ascecg

Proline-Mediated Knoevenagel−Doebner Condensation in Ethanol:


A Sustainable Access to p‑Hydroxycinnamic Acids
Ced́ ric Peyrot, Aureĺ ien A. M. Peru, Louis M. M. Mouterde, and Florent Allais*
URD Agro-Biotechnologies Industrielles (ABI), CEBB, AgroParisTech, 51110 Pomacle, France
*
S Supporting Information

ABSTRACT: Naturally occurring p-hydroxycinnamic acids


were obtained in good yields (50−85%) using a pyridine/
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

piperidine-free Knoevenagel−Doebner condensation of the


corresponding p-hydroxybenzaldehydes with malonic acid in
ethanol. This method uses fully renewable and cheap nontoxic
Downloaded via UNIV DE ALCALA on January 31, 2023 at 14:28:57 (UTC).

reagents and solvents. By combination of a design of


experiments (DoE) and a one variable at a time optimization
(OVAT), the different reaction parameters were optimized in
order to favor p-hydroxycinnamic acids production over that
of the two other reaction byproducts, the corresponding
diacid and vinylphenol.

KEYWORDS: Knoevenagel−Doebner, L-Proline, p-Hydroxycinnamic acids, Sinapic acid, Ferulic acid, Caffeic acid, Coumaric acid

■ INTRODUCTION
Knoevenagel−Doebner reaction is one of the most studied
such as the use of amino acids as catalysts in the presence of
ionic liquid or PEG.28−30 More recently, de Winter et al.
described the synthesis of α,β-unsaturated ketones in excellent
reactions in the recent years.1 It allows access to, among
yields through an aldol condensation in the presence of L-
others, p-hydroxycinnamic acids that are key intermediates for
proline and magnesium oxide, two nontoxic and naturally
the synthesis of a large range of (bio)active molecules: non-
occurring chemicals.31 In the present work, we aimed to
endocrine disruptive bisphenol A substitutes,2,3 antioxi- determine whether this sustainable approach could be applied
dants,4−6 anti-UV ingredients,7−9 or renewable polymers and to Knoevenagel−Doebner condensation for the synthesis of
resins.10−14 However, its use is limited by the reaction naturally occurring p-hydroxycinnamic acids from the corre-
conditions that involve large amounts of pyridine as solvent sponding p-hydroxybenzaldehydes.
and aniline or piperidine as catalyst.15 Pyridine is toxic and
may induce serious health damages. Although aniline and
piperidine are less hazardous, reducing their utilization is
desirable.
■ EXPERIMENTAL SECTION
General. Knoevenagel−Doebner condensations were performed
Few alternatives have been described in the literature. For in round-bottom flask topped by air cooling column. Evaporations
instance, ionic liquids offer high solubility and recyclability;16 were conducted under reduced pressure. 1H NMR spectra of
compounds in the indicated solvent were recorded at 300 MHz at
however their use in industrial applications is compromised 20 °C [1H NMR: (CD3)2CO residual signal at δ = 2.05 ppm; DMSO
due to their high prices.17−20 Microwave-assisted Knoevena- residual signal at δ = 2.50 ppm]. 13C NMR spectra in the indicated
gel−Doebner was also investigated, and Mouterde et al. solvent were recorded at 75 MHz at 20 °C [13C NMR: (CD3)2CO
demonstrated that such activation reduces both solvent volume residual signal at δ = 206.26 and 29.84 ppm; DMSO residual signal at
and reaction time while providing the targets in excellent yields δ = 39.52 ppm]. The entire assigned NMR spectra are presented in
(85−92%) and limiting the production of the corresponding the Supporting Information. All reported yields are uncorrected and
vinylphenols through decarboxylation.21 Even if this method- refer to purified products. All reagents were purchased from Sigma-
Aldrich or TCI and used without any purification.
ology seems to be efficient in terms of yield, its implementation Optimized Procedure for the Synthesis of p-Hydroxycin-
at the industrial scale remains, unfortunately, limited by the namic Acids. The corresponding p-hydroxybenzaldehydes (1 g, 1.0
microwave reactor size and the use of piperidine. Another equiv), malonic acid (2.0 equiv), and L-proline (0.5 equiv) were
alternative consists of performing a solvent-free Knoevenagel− mixed in ethanol (0.5 M). Stirring and thermal heating (90 °C) were
Doebner reaction that permits decrease of the amount of applied during 16 h. The reaction mixture was then concentrated
catalyst needed and reduction of the reaction time.22 Other
catalysts such as Lewis acids (NbCl5, ZnCl2), heterogeneous Received: January 30, 2019
solid bases, or amino acids have also been investigated.22−27 Revised: March 19, 2019
Combinations of these approaches have been also reported Published: April 18, 2019

© 2019 American Chemical Society 9422 DOI: 10.1021/acssuschemeng.9b00624


ACS Sustainable Chem. Eng. 2019, 7, 9422−9427
ACS Sustainable Chemistry & Engineering Research Article

Scheme 1. Knoevenagel−Doebner Condensation of p-Hydroxycinnamic Aldehydes Using Malonic Acid and Proline as Catalyst

under vacuum. Conversion yields were determined by 1H NMR degradation thus leads to the use of several equivalents of
spectroscopy, and crude products were extracted with an AcOEt/HCl malonic acid to ensure p-hydroxybenzaldehydes consumption.
mixture (1 M). Organic layers were washed with brine, dried over Furthermore, high temperatures can lead to a second
anhydrous MgSO4, and concentrated under vacuum. Crude mixture
decarboxylation that generates vinylphenols.34,35 On the basis
was then purified by column chromatography on silica gel using
appropriate eluent system (cyclohexane/AcOEt 50−50 to 100% of the thermodynamic study performed by Bermudez et al.
AcOEt). using piperidine as activating group, the proposed mechanism
p-Coumaric Acid. 1H NMR (300 MHz, (CD3)2CO): δ = 6.34 (d, shown in Scheme 1 illustrates perfectly the positive effect of
J = 15.93 Hz, 1H, H-2), 6.88 (s, 1H, H-6 or H-8), 6.91 (s, 1H, H-6 or temperature on the first decarboxylation step but also its
H-8), 7.54 (s, 1H, H-5 or H-9), 7.57 (s, 1H, H-5 or H-9), 7.61 (d, J = negative impact leading to vinylphenols formation and malonic
15.96 Hz, 1H, H-3). 13C NMR (75 MHz, (CD3)2CO): δ = 115.7 (d, acid degradation.32 Finding the right temperature that will
C-6 and C-8), 116.7 (d, C-2), 127.0 (s, C-4), 130.9 (d, C-5 and C-9), favor one product or the other is therefore quite challenging.
145.6 (d, C-3), 160.5 (s, C-7), 168.3 (s, C-1). In their work dedicated to the preparation of α,β-
Caffeic Acid. 1H NMR (300 MHz, (CD3)2CO): δ = 6.27 (d, J =
15.9 Hz, 1H, H-2), 6.86 (d, J = 8.16 Hz, 1H, H-5), 7.04 (dd, J = 2.01
unsaturated ketones through aldol condensation,31 de Winter
and 8.22 Hz, 1H, H-6), 7.16 (d, J = 2.04 Hz, 1H, H-9), 7.55 (d, J = et al. determined the following optimized conditions: 1 equiv
15.9 Hz, 1H, H-3). 13C NMR (75 MHz, (CD3)2CO): δ = 115.1 (d, of ketone, 1.2 equiv of aldehyde, 10 wt % MgO, 1 equiv of L-
C-9), 115.6 (d, C-2), 116.2 (d, C-6), 122.4 (d, C-5), 127.5 (s, C-4), proline in methanol (0.45 M) at 50 °C. These conditions were
145.9 (d, C-3), 146.2 (s, C-8), 148.6 (s, C-7), 168.2 (s, C-1). thus directly applied on malonic acid and syringaldehyde and,
Ferulic Acid. 1H NMR (300 MHz, (CD3)2CO): δ = 3.92 (s, 3H, after 4 h, resulted in a mixture of sinapic acid (10%), diacid
H-11), 6.39 (d, J = 15.93 Hz, 1H, H-2), 6.87 (d, J = 8.16 Hz, 1H, H- (27%), and unreacted syringaldehyde (63%) as measured by
5), 7.15 (dd, J = 1.86 and 8.22 Hz, 1H, H-6), 7.34 (d, J = 1.89 Hz, 1
H NMR. Consequently, it was concluded that under such
1H, H-9), 7.61 (d, J = 15.9 Hz, 1H, H-3). 13C NMR (75 MHz,
(CD3)2CO): δ = 56.2 (q, C-11), 111.2 (d, C-9), 115.8 (d, C-6), 116.0 conditions the reaction was possible but an optimization was
(d, C-2), 123.8 (d, C-5), 127.4 (s, C-4), 145.9 (d, C-3), 148.6 (s, C- required to achieve higher yields and selectivities.
7), 149.9 (s, C-8), 168.3 (s, C-1). In order to have a better understanding of the reaction, the
Sinapic Acid. 1H NMR (300 MHz, (CD3)2CO): δ = 1H NMR influence of various reaction parameters/variables (i.e.,
(300 MHz, (CD3)2CO): δ = 3.89 (s, 6H, H-11 and H-12), 6.41 (d, J temperature, concentration, equivalent of malonic acid,
= 15.87 Hz, 1H, H-2), 7.02 (s, 2H, H-5 and H-9), 7.59 (d, J = 15.87 equivalent of MgO, and equivalent of proline) was investigated
Hz, 1H, H-3). 13C NMR (75 MHz, (CD3)2CO): δ = 56.6 (q, C-11 by performing a design of experiments (DoE) (Table 1).
and C-12), 106.7 (d, C-5 and C-9), 116.1 (d, C-2), 126.1 (s, C-4), It is worth mentioning that methanol was substituted by
139.3 (s, C-7), 146.3 (d, C-3), 148.9 (s, C-8 and C-6), 168.3 (s, C-1).

■ RESULTS AND DISCUSSION


Knoevenagel−Doebner reactions are known to be really
Table 1. Variables and Their Variation Corresponding to
Defined Levels of the DoE
sensitive to reaction conditions, and it has been shown that level
each reaction parameter impacts significantly p-hydroxycin-
variable −1 0 1
namic acid production as well as that of the side products,
namely, the diacid and vinylphenol (Scheme 1). For instance, temperature (°C) 40 60 80
high temperatures are an essential factor that promotes concentration (mol L−1) 0.1 0.3 0.5
decarboxylation of p-hydroxycinnamic diacids into p-hydrox- malonic acid equivalents 1 2 3
ycinnamic acids,32 but they also induce malonic acid MgO equivalents 0.1 0.6 1.1
degradation and acetic acid generation.33 This undesired proline equivalents 0.1 0.6 1.1

9423 DOI: 10.1021/acssuschemeng.9b00624


ACS Sustainable Chem. Eng. 2019, 7, 9422−9427
ACS Sustainable Chemistry & Engineering Research Article

Scheme 2. Regression Coefficients of Quadratic Model

Scheme 3. Contour Plot of Syringaldehyde Conversion into Sinapic Acid

9424 DOI: 10.1021/acssuschemeng.9b00624


ACS Sustainable Chem. Eng. 2019, 7, 9422−9427
ACS Sustainable Chemistry & Engineering Research Article

Table 2. Optimization of the Knoevenagel-Doebner Condensation of Syringaldehyde and Malonic Acid in Ethanol
malonic acid proline MgO temp concn conversion to diacida conversion to sinapic conversion to vinylphenola
(equiv) (equiv) (equiv) (°C) (M) (%) acida(%) (%)
1 2.8 0.8 0.25 80 0.1 38 60 0
2 2.8 0.8 0 80 0.5 18 82 0
3 2 0.5 0 80 0.5 14 80 0
4 2 0.5 0 90 0.5 0 90 10
5 2 0.8 0 90 0.5 0 76 24
6 2 0.3 0 90 0.5 11 78 11
a
Conversion was determined by 1H NMR of the crude reaction mixture.

ethanol to provide a more sustainable process. All sets of not significant on the conversion rate. To summarize, the
reaction conditions tested for the DoE and the corresponding equation of the model (eq 2):
results can be found in the Supporting Information.
The following second-order polynomial equation was used Conversion into sinapic acid (%)
to determine a relationship between the variables and response
= 52.66 + 6.06· equiv of malonic acid + 11.01
surface:
·equiv of proline + 15.59 ·temp − 11.81 ·temp2
Y = α0 + ∑ αixi + ∑ αjxi 2 + ∑ αijxixj
i i ij (1) + 5.06· (equiv of malonic acid ·equiv of proline)

where Y is the conversion into sinapic acid response, xi are the + 8.16· (equiv of malonic acid ·equiv of MgO)
variables, α0 is a constant, and αi, αj, and αij are the linear, − 5.66· (equiv of proline ·concn) (2)
quadratic, and interaction coefficients. Multiple linear
regressions (MLRs) allowed the determination of the The visualization of the influence of each parameter can be
regression coefficients. The significant parameters in the readily assessed through the analysis of the 4D response
model are those whom p-value is <0.05. contour (Scheme 3). In the interval considered, the optimal
The model validation was based on the variance (ANOVA) temperature was evaluated at 80 °C and the optimal number of
for each response by analyzing R2, Q2, and lack of fit (LOF) equivalents of malonic acid, proline, and MgO at 2.8, 0.8, and
test. R2 indicates whether the regression model fits the 0.25, respectively. Applying these conditions allowed a total
experimental data (the closer to 1 the better the fit), Q2 consumption of syringaldehyde with a conversion into sinapic
estimates the future prediction precision, and LOF assesses acid and sinapic diacid of 60% and 38%, respectively (Table 2,
whether the models error is comparable to the replicate error. entry 1).
To determine the optimum set of experimental parameters to The maximum conversion into sinapic acid being relatively
reach the highest conversions, a D-optimal design was used low despite the DoE, one or more reaction parameters
and consisted of 31 experiments. Among these experiments, remained to be adjusted using the classical OVAT approach.
three replicates at the central point were performed to evaluate Magnesium oxide having an insignificant influence on the
their reproducibility. reaction, the reaction was first performed without MgO, and at
The experimental data of the D-optimal design was fitted to relatively high concentration (0.5 M) as the latter parameter
the second-order polynomial equation (eq 1). Variance has a significant impact as shown by the DoE. Doing so, the
analysis shows a good correlation of the second-order conversion into sinapic acid increased up to 82%, whereas that
polynomial model between the response (conversion) and of sinapic diacid decreased to 18% (Table 2, entry 2). This
the significant variables (p-value of <0.05). The lack of fit (p- result undoubtedly confirming that magnesium oxide has no
value of >0.05) shows the low replicate errors of the model. effect in this reaction, its use was ruled out in the next
Finally, the design gives very good coefficient of determination experiments.
and acceptable coefficient of cross-validation, R2 = 0.824 From a sustainable point of view, the use of 2.8 equiv of
(>0.5) and Q2 = 0.503 (>0.5), respectively, demonstrating a malonic acid seemed to be excessive for a reaction that could
good fit and prediction of the chosen model. All coefficients be theoretically performed using equimolar quantities. The
(α0 as constant, αi, αj, and αij of the eq 1) of the model for the reaction was thus performed using only 2 equiv of malonic acid
response are shown in Scheme 2. It is noteworthy to mention (Table 2, entry 3). Moreover, the parameter of equivalent of
that a positive coefficient corresponds to a variable with a proline/concentration having a negative effect on the
positive influence on the conversion and vice versa. formation of sinapic acid, the quantity of proline was reduced
DoE results indicates that the temperature, concentration, to 0.5 equiv. In such conditions, similar conversion into sinapic
equivalent of malonic acid, and equivalent of proline have a acid was obtained (80% vs 82%) (Table 2, entry 3).
positive impact on the conversion of syringaldehyde into Nevertheless, at this stage of the optimization, a relatively
sinapic acid whereas equivalent of MgO has a slightly negative significant quantity of sinapic diacid was still observed (14%).
one, with optimum values on the intervals of study according The DoE confirmed the strong influence of the temperature
to their respective quadratic terms (Scheme 2). Concentration on the decarboxylation of the sinapic diacid into the sinapic
interacts negatively with equivalents of proline, which suggests acid. The reaction was thus carried out at higher temperature
that the catalytic cycle is influenced by the concentration. With (90 °C) to promote the decarboxylation and drive the
p-value of >0.1, the number of equivalent of MgO, the equilibrium of the reaction toward sinapic acid. Doing so,
concentration, the concentration quadratic terms, and that of the sinapic diacid was completely consumed but the formation
the equivalent of malonic acid toward the temperature proved of the corresponding vinylphenol was also observed (10%)
9425 DOI: 10.1021/acssuschemeng.9b00624
ACS Sustainable Chem. Eng. 2019, 7, 9422−9427
ACS Sustainable Chemistry & Engineering Research Article

(Table 2, entry 4). Indeed, increasing the temperature conversion, 85% isolated yield) over that of the two other side
promotes the second decarboxylation (Scheme 1); it is products (diacid and vinylphenol). This set of optimized
therefore crucial to finely control the reaction temperature to conditions was applied to other p-hydroxybenzaldehydes
favor sinapic acid production while limiting that of the diacid providing the corresponding p-hydroxycinnamic acids in
and the vinylphenol. good yields (55−72% 1H NMR conversions, 50−69% isolated
In a previous paper, we have illustrated the impact of the yields). This sustainable route offers an interesting approach to
equivalent of catalyst (i.e., piperidine) on the vinylphenol access quickly and efficiently naturally occurring p-hydroxycin-
formation.21 Indeed, it was shown that a large amount of namic acids while limiting the toxicity risks associated with
catalyst promoted its formation. On the basis of this their production.
observation, the equivalent of proline was modulated from
0.8 to 0.3 (Table 2, entries 4−6). As previously observed with
piperidine, increasing the equivalent of catalyst (0.8) induced

*
ASSOCIATED CONTENT
S Supporting Information
vinylphenol formation (24%) (Table 2, entry 5). A decrease to The Supporting Information is available free of charge on the
0.3 equiv promoted the formation of only 11% of the ACS Publications website at DOI: 10.1021/acssusche-
corresponding vinylphenol but also limited the first decarbox- meng.9b00624.
ylation step maintaining 11% of sinapic diacid (Table 2, entry Experimental details and NMR spectra (PDF)


6). From these three sets of conditions, 0.5 equiv of proline
appeared as the best compromise for the formation of sinapic
acid while limiting that of sinapic diacid and vinylphenol.
AUTHOR INFORMATION
In summary, through the combination of the DoE and Corresponding Author
OVAT, optimal conditions were found as 1 equiv of *E-mail: florent.allais@agroparistech.fr.
syringaldehyde, 2 equiv of malonic acid, and 0.5 equiv of ORCID
proline at 90 °C during 16 h in ethanol (0.5 M) (Table 2, Louis M. M. Mouterde: 0000-0002-4096-1629
entry 4). As it is impossible to obtain exclusively sinapic acid, a Florent Allais: 0000-0003-4132-6210
purification step is inevitable. Vinylphenol and sinapic diacid Notes
having different polarities, their separation is readily achieved The authors declare no competing financial interest.


through chromatography on silica gel. Moreover, vinylphenol
being an oily compound, it can be also eliminated by ACKNOWLEDGMENTS
precipitating sinapic acid.
The authors are grateful to Agence Nationale de la Recherche
Applying these optimal conditions to the other p-
(ANR, Grant ANR-17-CE07-0046), Région Grand Est,
hydroxybenzaldehydes provided the corresponding p-hydrox-
Conseil Départemental de la Marne, and Grand Reims for
ycinnamic acids in good yields while limiting the presence of
financial support.


diacid and vinylphenol (Table 3). It is noteworthy to mention
ABBREVIATIONS
Table 3. Conversion Rate and Yields of Naturally Occurring
DoE, design of experiments; OVAT, one variable at a time


p-Hydroxycinnamic Acids
conversiona yieldb REFERENCES
substrate phenolic acid (%) (%)
(1) List, B. Emil Knoevenagel and the Roots of Aminocatalysis.
4-hydroxybenzaldehyde p-coumaric acid 55 50 Angew. Chem., Int. Ed. 2010, 49 (10), 1730−1734.
3,4-dihydroxybenzaldehyde caffeic acid 71 68 (2) Jaufurally, A. S.; Teixeira, A. R. S.; Hollande, L.; Allais, F.;
vanillin ferulic acid 72 69 Ducrot, P.-H. Optimization of the laccase-catalyzed synthesis of
syringaldehyde sinapic acid 87 85 (±)-syringaresinol and study of its thermal and antiradical activities.
a ChemistrySelect 2016, 1 (16), 5165−5171.
Conversion was determined by 1H NMR of the crude reaction
(3) Janvier, M.; Hollande, L.; Jaufurally, A. S.; Pernes, M.; Menard,
mixture. bYields were calculated from isolated product after
R.; Grimaldi, M.; Beaugrand, J.; Balaguer, P.; Ducrot, P.-H.; Allais, F.
purification.
Syringaresinol: A Renewable and Safer Alternative to Bisphenol A for
Epoxy-Amine Resins. ChemSusChem 2017, 10 (4), 738−746.
that these syntheses have been carried out at the multigram (4) Reano, A. F.; Pion, F.; Domenek, S.; Ducrot, P.-H.; Allais, F.
Chemo-enzymatic preparation and characterization of renewable
scale. Nevertheless, the previous optimization having been
oligomers with bisguaiacol moieties: promising sustainable anti-
performed on the more activated p-hydroxybenzaldehyde (i.e., radical/antioxidant additives. Green Chem. 2016, 18 (11), 3334−
syringaldehyde), optimization of the reaction conditions 3345.
should be carried out for every single p-hydroxybenzaldehydes (5) Reano, A.; Domenek, S.; Pernes, M.; Beaugrand, J.; Allais, F.
to access higher yields and selectivities.


Ferulic Acid-based Bis/trisphenols as Renewable Antioxidants for
polypropylene and poly(butylen succinate). ACS Sustainable Chem.
CONCLUSION Eng. 2016, 4 (12), 6562−6571.
An efficient and sustainable proline-mediated Knoevenagel− (6) Reano, A.; Cherubin, J.; Peru, A. M.; Wang, Q.; Clément, T.;
Doebner condensation of malonic acid and syringaldehyde in Domenek, S.; Allais, F. Structure-activity Remationships and
Structural design Optimization of a Series of p-hydroxycinnamic
ethanol has been developed and optimized to access naturally Acids-based bis- and trisphenols as novel sustainable Antiradical/
occurring sinapic acid. We successfully substituted the classical antioxidant additives. ACS Sustainable Chem. Eng. 2015, 3, 3486−
petro-sourced and toxic catalysts, such as pyridine/piperidine, 3496.
by proline, and toxic solvent by ethanol. Through a (7) Allais, F.; Martinet, S.; Ducrot, P.-H. Straightforward total
combination of DoE and OVAT, each parameter has been synthesis of 2-O-feruloyl-L-malate, 2-O-sinapoyl-L-malate and 2-O-5-
optimized to favor sinapic acid production (87% 1H NMR hydroxyferuloyl-L-malate. Synthesis 2009, No. 21, 3571−3578.

9426 DOI: 10.1021/acssuschemeng.9b00624


ACS Sustainable Chem. Eng. 2019, 7, 9422−9427
ACS Sustainable Chemistry & Engineering Research Article

(8) Dean, J. C.; Kusaka, R.; Walsh, P. J.; Allais, F.; Zwier, T. S. J. (28) Wang, Y.; Shang, Z.-C.; Wu, T.-X.; Fan, J.-C.; Chen, X.
Plant sunscreens in the UV-B: ultraviolet spectroscopy of jet-cooled Synthetic and theoretical study on proline-catalyzed Knoevenagel
sinapoyl malate, sinapic acid, and sinapate ester derivatives. J. Am. condensation in ionic liquid. J. Mol. Catal. A: Chem. 2006, 253 (1−2),
Chem. Soc. 2014, 136 (42), 14780−14795. 212−221.
(9) Baker, L. A.; Staniforth, M.; Flourat, A. L.; Allais, F.; Stavros, V. (29) Liu, X.-H.; Fan, J.-C.; Liu, Y.; Shang, Z.-C. L-Proline as an
G. Gas-Solution Phase Transient Absorption Study of the Plant efficient and reusable promoter for the synthesis of coumarins in ionic
Sunscreen Derivative Methyl Sinapate. ChemPhotoChem 2018, 2 (8), liquid. J. Zhejiang Univ., Sci., B 2008, 9 (12), 990−995.
743−748. (30) Venkatanarayana, M.; Dubey, P. K. L-Proline-catalyzed
(10) Hollande, L.; Jaufurally, A. S.; Ducrot, P.-H.; Allais, F. ADMET Knoevenagel condensation. A tandem synthesis of 3-acetylcoumar-
Polymerization of Biobased Monomers Deriving from Syringaresinol. inoindoles and their N-alkyl derivatives by using PEG-600 as the
RSC Adv. 2016, 6, 44297. reaction medium. J. Heterocycl. Chem. 2014, 51 (4), 877−882.
(11) Barbara, I.; Flourat, A. L.; Allais, F. Renewable Polymers (31) de Winter, T. M.; Balland, Y.; Neski, A. E.; Petitjean, L.;
Derived from Ferulic Acid and Biobased Diols via ADMET. Eur. Erythropel, H. C.; Moreau, M.; Hitce, J.; Coish, P.; Zimmerman, J. B.;
Polym. J. 2015, 62, 236. Anastas, P. T. Exploration of a Novel, Enamine-Solid-Base Catalyzed
(12) Oulame, M. Z.; Pion, F.; Allauddin, S.; Raju, K. V. S. N.; Aldol Condensation with C-Glycosidic Pyranoses and Furanoses. ACS
Ducrot, P.-H.; Allais, F. Renewable alternating aliphatic-aromatic Sustainable Chem. Eng. 2018, 6 (9), 11196−11199.
poly(ester-urethane)s prepared from ferulic acid and bio-based diols. (32) Bermudez, E.; Ventura, O. N.; Saenz Méndez, P. Mechanism of
Eur. Polym. J. 2015, 63, 186. the Organocatalyzed Decarboxylative Knoevenagel−Doebner Reac-
(13) Ménard, R.; Caillol, S.; Allais, F. Chemo-enzymatic Synthesis tion. A Theoretical Study. J. Phys. Chem. A 2010, 114, 13086−13092.
(33) Caires, F. J.; Lima, L. S.; Carvalho, C. T.; Giagio, R. J.;
and Characterization of Renewable Thermoplastic and Thermoset
Ionashiro, M. Thermal behaviour of malonic acid, sodium malonate
Isocyanate-free poly(hydroxy)urethanes from Ferulic Acid Deriva-
and its compounds with some bivalent transition metal ions.
tives. ACS Sustainable Chem. Eng. 2017, 5, 1446−1456.
Thermochim. Acta 2010, 497 (1−2), 35−40.
(14) Kasmi, S.; Gallos, A.; Beaugrand, J.; Paes, G.; Allais, F. Ferulic
(34) Aldabalde, V.; Risso, M.; Derrudi, M. L.; Geymonat, F.; Seoane,
acid derivatives used as biobased powders for a convenient G.; Gamenara, D.; Saenz-Mendez, P. Organocatalyzed decarbox-
plasticization of polylactic acid in continuous hot-melt process. Eur. ylation of naturally occurring cinnamic acids: potential role in
Polym. J. 2019, 110, 293−300. flavoring chemicals production. Open J. Phys. Chem. 2011, 1 (3), 85−
(15) Knoevenagel, E. Ueber eine Darstellungsweise der Glutarsaure. 93.
Ber. Dtsch. Chem. Ges. 1894, 27, 2345−2346. (35) Sinha, A. K.; Sharma, A.; Joshi, B. P. One-Pot Two-Step
(16) Forbes, D. C.; Law, A. M.; Morrison, D. W. The Knoevenagel Synthesis of 4-Vinylphenols from 4-Hydroxy Substituted Benzalde-
reaction: analysis and recycling of the ionic liquid medium. hydes under Microwave Irradiation: a New Perspective on the
Tetrahedron Lett. 2006, 47 (11), 1699−1703. Classical Knoevenagel-Doebner Reaction. Tetrahedron 2007, 63,
(17) Hangarge, R. V.; Jarikote, D. V.; Shingare, M. S. Knoevenagel 960−965.
condensation reactions in an ionic liquid. Green Chem. 2002, 4 (3),
266−268.
(18) Ouyang, F.; Zhou, Y.; Li, Z.-M.; Hu, N.; Tao, D.-J.
Tetrabutylphosphonium amino acid ionic liquids as efficient catalysts
for solvent-free Knoevenagel condensation reactions. Korean J. Chem.
Eng. 2014, 31 (8), 1377−1383.
(19) Hu, X.; Ngwa, C.; Zheng, Q. A Simple and Efficient Procedure
for Knoevenagel Reaction Promoted by Imidazolium-Based Ionic
Liquids. Curr. Org. Synth. 2015, 13 (1), 101−110.
(20) Zhao, S.; Wang, X.; Zhang, L. Rapid and efficient Knoevenagel
condensation catalyzed by a novel protic ionic liquid under ultrasonic
irradiation. RSC Adv. 2013, 3 (29), 11691−11696.
(21) Mouterde, L. M. M.; Allais, F. Microwave-Assisted
Knoevenagel-Doebner Reaction: An Efficient Method for Naturally
Occurring Phenolic Acids Synthesis. Front. Chem. 2018, 6, 426.
(22) Pasha, M. A.; Manjula, K. Lithium hydroxide: a simple and an
efficient catalyst for Knoevenagel condensation under solvent-free
Grindstone method. J. Saudi Chem. Soc. 2011, 15 (3), 283−286.
(23) Leelavathi, P.; Kumar, S. R. Niobium(V) chloride-catalyzed
Knoevenagel condensation. An efficient protocol for the preparation
of electrophilic alkenes. J. Mol. Catal. A: Chem. 2005, 240 (1−2), 99−
102.
(24) Ogiwara, Y.; Takahashi, K.; Kitazawa, T.; Sakai, N. Indium(III)-
Catalyzed Knoevenagel Condensation of Aldehydes and Activated
Methylenes Using Acetic Anhydride as a Promoter. J. Org. Chem.
2015, 80 (6), 3101−3110.
(25) Zhang, H.; Han, M.; Chen, T.; Xu, L.; Yu, L. Poly(N-
isopropylacrylamide-co-L-proline)-catalyzed Claisen-Schmidt and
Knoevenagel condensations: unexpected enhanced catalytic activity
of the polymer catalyst. RSC Adv. 2017, 7 (76), 48214−48221.
(26) He, Y.-H.; Hu, Y.; Guan, Z. Natural α-amino acid L-lysine-
catalyzed Knoevenagel condensations of α,β-unsaturated aldehydes
and 1,3-dicarbonyl compounds. Synth. Commun. 2011, 41 (11),
1617−1628.
(27) Rahmati, A.; Vakili, K. L-Histidine and L-arginine promote
Knoevenagel reaction in water. Amino Acids 2010, 39 (3), 911−916.

9427 DOI: 10.1021/acssuschemeng.9b00624


ACS Sustainable Chem. Eng. 2019, 7, 9422−9427

You might also like