Next Article in Journal
Study of Acetylcholinesterase and Butyrylcholinesterase (AChE/BuChE) Inhibition Using Molecular Modelling Methods
Previous Article in Journal
Molecular Docking for the Development of Alternative Therapies against Leishmaniasis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Transesterification of a Natural Epoxythymol Is Favored under Alkaline Conditions, Preserving the Enantiomeric Purity †

by
Jessica M. Lorenzo-García
1,
Antonio J. Oliveros-Ortiz
1,
Héctor M. Arreaga-González
2,
Carlos J. Cortés-García
1,
Rosa E. del Río
1,
Gabriela Rodríguez-García
1,* and
Mario A. Gómez-Hurtado
1,*
1
Instituto de Investigaciones Químico Biológicas, Universidad Michoacana de San Nicolás de Hidalgo, Morelia 58030, Michoacán, Mexico
2
Instituto de Agroindustrias, Universidad Tecnológica de la Mixteca, Huajuapan de Léon 69000, Oaxaca, Mexico
*
Authors to whom correspondence should be addressed.
Presented at the 27th International Electronic Conference on Synthetic Organic Chemistry (ECSOC-27), 15–30 November 2023; Available online: https://ecsoc-27.sciforum.net/.
Chem. Proc. 2023, 14(1), 30; https://doi.org/10.3390/ecsoc-27-16140
Published: 15 November 2023

Abstract

:
Transesterification is a synthetic chemistry strategy promoted in acid or alkaline conditions, yielding a structural diversity of organic compounds. Epoxythymols comprise a class of chiral natural compounds with biological relevance, and the literature describes their chiral purity loss during acid transesterification reactions. This work reports the basic transesterification of the natural derivative (8S)-10-benzoylxy-8,9-epoxy-6-hydroxythymol under alkaline conditions. Herein, the formation of (8S)-10-benzoylxy-6-isobutyryloxy-8,9-epoxythymol isobutyrate is gained, avoiding the loss of optical purity. 1H NMR-BINOL experiments revealed the enantiomeric purity of the product reaction. These results highlight that the implemented strategy promotes transesterification, preserving optical purity.

1. Introduction

Transesterification is a widely used chemical reaction to yield esterified derivatives. Several strategies can be used depending on the reaction medium and the starting material stability; however, acidic or basic conditions or catalysts with such properties are always considered. Sometimes, anhydrous conditions are used with sensitive substrates [1]. This type of reaction has been described as a result of spontaneous transformations of natural epoxythymol derivatives during or after their isolation [2,3,4]. This compound family originates from terpene biosynthesis and is found in 42 genera of the Asteraceae family, characterized by chirality due to the oxirane ring located at the C-8/C-9 position [5]. These compounds are important due to their biological potential as cytotoxic and antibacterial agents [6,7,8,9,10,11]. The literature suggested that these derivatives may be susceptible to the escalemization process [12] through acid transesterification reactions. Herein, escalemization proceeds via an intramolecular reaction mechanism involving the cleavage of the oxirane ring, thereby altering chirality [13]. Structural variation around the basic epoxythymol skeleton avoiding chiral alterations will enable the generation of novel derivatives with potential biological activity.
In the present work, the transesterification under basic conditions of (8S)-10-benzoylxy-8,9-epoxy-6-hydroxythymol isobutyrate (1) isolated from Ageratina glabrata is described, allowing the formation of (8S)-10-benzoylxy-6-isobutyryloxy-8,9-epoxythymol isobutyrate (2). After the reaction process, the optical purity of the reaction product was saved, which was validated through 1H NMR-BINOL experiments. These results demonstrate that the implemented strategy promotes transesterification while preserving the chirality of molecules containing asymmetric oxirane rings, such as those natural epoxythymol compounds (Figure 1).

2. Material and Methods

2.1. Experimental Section

All chemical reagents were acquired from Sigma-Aldrich and used as purchased. Solvents were distilled before use. Melting points were determined using a Fisher-Johns apparatus and were not corrected. Dichroic Circularly Polarized Light (CPL) and UV spectra were obtained using a Jasco CD-2095 circular dichroism detector, employing an ethanolic solution with 0.027 mmol of compound 1. The results were processed using WinDaq software and plotted in Excel. The IR spectra were acquired using a Thermo Scientific Nicolet iS10 spectrophotometer employing the ATR (Attenuated Total Reflectance) technique. The 1H and 13C NMR spectra at 400 MHz and 100 MHz, respectively, were measured in a Varian Mercury 400 spectrometer, using solutions of CDCl3 and tetramethylsilane (TMS) as the internal reference. Chemical shifts are reported in ppm, and coupling constants (J) are given in Hertz. The NMR spectra were processed using MestReNova software. The enantiomeric purity of thymol derivatives was determined by 1H NMR using 6.0 and 3.0 mg of sample (1 and 2, respectively) and (S)-BINOL as the chiral solvating reagent, in quantities of 30.0 and 15.0 mg, respectively, and dissolving in all cases with 0.7 mL of CDCl3. The enantiomeric ratio was established by analyzing the splitting of signals and their intensities. Purification of compounds was achieved by column chromatography using silica gel (230–400 mesh) as the stationary phase.

2.2. Plant Material

Specimens of Ageratina glabrata were collected during the flowering stage in February 2018, near km 4.5 of the federal road 200 from Pátzcuaro-Santa Clara del Cobre, Michoacán, Mexico, at N 19°29.516′ W 101°35.273′ and 2285 m above sea level. A voucher specimen (No. 226133) was deposited in the Herbarium of the Institute of Ecology, A. C., Regional Center of El Bajío, Pátzcuaro, Michoacán, Mexico.

2.3. Extraction and Isolation

A batch of dried leaves (1.5 kg) was macerated using hexanes (10 L) for 3 days, filtered, and concentrated under reduced pressure. This procedure was performed three times. Afterward, the same procedure was performed with dichloromethane (10 L). Macerates yielded 35 g (2.3%) of the hexanes extract and 147 g (9.8%) of the CH2Cl2 extract, respectively.
(+)-(8S)-10-benzoyloxy-6-hydroxy-8,9-epoxythymol isobutyrate (1) was isolated following the reported methodology [13]. Colorless crystals, m.p. 112–114 °C; UV (EtOH) λmax (log ε) 230 (0.75), 275 (0.38) nm; ECD (EtOH) λmaxε): 226 (−185); 1H NMR (CDCl3, 400 MHz) δ: 7.97 (2H, dd, J = 7.8, 1.5 Hz, H-3″, H-7″), 7.55 (1H, tt, J = 7.8, 1.5 Hz, H-5″), 7.41 (2H, t, J = 7.8 Hz, H-4″, H-6″), 6.93 (1H, s, H-5), 6.81 (1H, s, H-2), 4.76 (1H, d, J = 12.3 Hz, H-10a), 4.47 (1H, d, J = 12.3 Hz, H-10b), 3.10 (1H, d, J = 5.3 Hz, H-9a), 2.85 (1H, d, J = 5.3 Hz, H-9b), 2.82 (1H, hept, J = 6.8 Hz, H-2′), 2.21 (3H, s, H-7), 1.30 (3H, d, J = 6.8 Hz, H-3′), 1.29 (3H, d, J = 6.8 Hz, H-4′). 13C NMR (CDCl3, 100 MHz,) δ: 175.0 (C, C-1′), 166.0 (C, C-1″), 151.7 (C, C-6), 141.7 (C, C-3), 133.2 (CH, C-5″), 129.7 (CH, C-3″, C-7″), 129.5 (C, C-4), 128.4 (CH, C-4″, C-6″), 127.1 (C, C-2″), 125.8 (C, C-1), 124.7 (CH, C-2), 114.5 (CH, C-5), 65.5 (CH2, C-10), 57.0 (C, C-8), 51.0 (CH2, C-9), 34.1 (CH, C-2′), 19 (CH3, C-3′, C-4′), 15.7 (CH3, C-7).

2.4. Transesterification Reaction under Basic Conditions

A batch of 100 mg of 1 was dissolved in 60 mL of benzene, previously refluxed for 4 h using a Dean–Stark trap to remove moisture. Subsequently, 38 mg (0.95 mmol, 1.5 eq) of NaOH was added and stirred for 2 h under reflux. Afterward, the crude was tempered, poured over wet ice, and extracted with AcOEt (ethyl acetate) (3 × 50 mL). The organic phase was dried with Na2SO4 and evaporated under vacuum to obtain 45.6 mg of yellow oil, which was purified via column chromatography using silica gel as the stationary phase and hexanes-AcOEt (4:1) as the eluent, yielding 29.9 mg of a yellowish oily liquid.
(8S)-10-benzoyloxy-6-isobutyryloxy-8,9-epoxythymol isobutyrate (2). Yellow oil. 1H NMR (CDCl3, 400 MHz) δ: 7.97 (2H, dd, J = 7.4, 1.4 H-3″, H-7″), 7.54 (1H, tt, J = 7.4, 1.4 Hz H-5″), 7.42 (2H, t, J = 7.4 Hz, H-4″, H-6″), 7.21 (1H,s, H-5), 6.94 (1H, s, H-2), 4.75 (1H, d, J = 12.4 Hz, H-10a), 4.46 (1H, d, J = 12.4 Hz, H-10b), 3.10 (1H, d, J = 5.2 Hz, H-9a), 2.87 (1H, d, J = 5.2 Hz, H-9b), 2.82 (1H, hept, J = 7.0 Hz, H-2′), 2.82 (1H, sept, J = 7.0 Hz, H-2‴), 2.16 (1H, s, H-7), 1.34 (3H, d, J = 7.0 Hz, H-3′), 1.34 (3H, d, J = 7.0 Hz, H-3‴), 1.30 (3H, d, J = 2.8 Hz, H-4′), 1.30 (3H, d, J = 2.8 Hz, H-4‴).

2.5. Epimerization Reaction of Epoxythymol 1

A suspension of 1 g of silica gel in 60 mL of benzene was refluxed for 4 h using a Dean–Stark trap to remove moisture. A sample (100 mg) of 1 (enantiomerically pure) was added, and the mixture was refluxed for 6 h. The crude reaction solution was filtered, evaporated under vacuum, and purified via column chromatography using hexanes-AcOEt (4:1) to yield 15.3 mg (15%) of 1 as a scalemic mixture [13].

3. Results and Discussion

Compound 1 was obtained as colorless crystals (m.p. 112–114 °C), whose spectroscopic data (see Supplementary Materials) were compared with an authentic sample and those reported values [14]. In the 1H NMR spectrum, the doublet signals from CH2-10 at δ 4.76 and 4.46 are highlighted since these protons confirmed the optical purity of 1 after the 1H NMR-BINOL analysis (Figure 2a). It follows that compound 1 was subjected to transesterification using NaOH/benzene in a Dean–Stark trap. After the reaction, the crude product was purified via column chromatography, resulting in a yellowish oil. The 1H NMR spectrum (see Supplementary Materials) revealed a pattern of signals similar to the starting material, where the singlet signals of the aromatic protons H-5 and H-2 appeared at δ 7.21 and 6.94, respectively. Additionally, a new set of signals from an isobutyrate moiety at δ 2.82 (1H, hept J = 7.0 Hz) and δ 1.34 (6-H, d, J = 7.0 Hz) appeared. This result suggested incorporating an isobutyrate group at position C-6 of the starting molecule, indicating the formation of 10-benzoylxy-6-isobutyryloxy-8,9-epoxythymol (2). This product suggests the concomitant generation of 10-benzoyloxy-6-hydroxy-8,9-epoxythymol (3), whose isolation and identification are currently under experimentation.
The literature mentions the transesterification of epoxythymol derivatives, where the reaction mechanism leads to racemization or escalemization. Therefore, it was decided to evaluate the enantiomeric purity of 2 using 1H NMR-BINOL analysis, as reported in [13]. In this spectrum (Figure 2b), the enantiomeric purity of 2 was revealed when only one set of signals for CH2-10 was observed.
A deliberate escalemization of 1 was achieved to validate the above enantiopurity analysis. Thus, compound 1 was subjected to reflux with benzene and silica gel for 6 h using a Dean–Stark trap, according to the methodology reported by Arreaga-González et al. [13]. In this spectrum (Figure 2c), two sets of signals for the AB system (δ 4.57) of methylene CH2-10 were observed and assigned to each enantiomer of 1 in a 75:25 ratio (8S:8R). Based on this, it can be proposed that the transesterification reaction proceeds without affecting the optical purity of the stereogenic center C-(8S).
According to the experimental results, a mechanistic pathway is proposed to the obtention of 2. It involves a concerted intermolecular transesterification process (Scheme 1), where NaOH facilitates the activation of O-6 through deprotonation (I). The produced phenoxide ion promotes a nucleophilic attack to the carbonyl of the isobutyrate group of a neighboring molecule (II) to create the new ester bond, concomitantly generating 10-benzoyloxy-6-hydroxy-8,9-epoxythymol (3).

4. Conclusions

The transesterification with NaOH/benzene using a Dean–Stark trap allowed for the optical pure epoxythymol derivative 2 to be obtained. Presumably, transesterification takes place intermolecularly, thus yielding the expected product, 10-benzoyloxy-6-hydroxy-8,9-epoxythymol; therefore, this method is suitable for generating various optically pure epoxythymol derivatives.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ecsoc-27-16140/s1, NMR, 1H NMR-BINOL, ECD of compound 1. NMR, 1H NMR-BINOL of compound 2.

Author Contributions

Conceptualization, M.A.G.-H. and R.E.d.R.; methodology, J.M.L.-G. and H.M.A.-G.; validation, J.M.L.-G. and A.J.O.-O.; formal analysis, G.R.-G.; investigation, M.A.G.-H.; resources, M.A.G.-H. and R.E.d.R.; data curation, M.A.G.-H., R.E.d.R. and G.R.-G.; writing—original draft preparation, M.A.G.-H.; writing—review and editing, G.R.-G. and C.J.C.-G.; funding acquisition, M.A.G.-H. and R.E.d.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by CIC-UMSNH.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Acknowledgments

We thank CIC-UMSNH for the financial support. J.M.L.-G. and A.J.O.O. are grateful to CONACYT-Mexico for scholarships 828930 and 956977, respectively.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Otera, J. Transesterification. Chem. Rev. 1993, 93, 1449–1470. [Google Scholar] [CrossRef]
  2. Bohlmann, F.; Niedballa, U.; Schulz, J. Über einige Thymolderivate aus Gaillardia- und Helenium-Arten. Chem. Berichte 1969, 102, 864–871. [Google Scholar] [CrossRef]
  3. Tori, M.; Ohara, Y.; Nakashima, K.; Sono, M. Thymol Derivatives from Eupatorium fortunei. J. Nat. Prod. 2001, 64, 1048–1051. [Google Scholar] [CrossRef] [PubMed]
  4. Jiang, H.-X.; Li, Y.; Pan, J.; Gao, K. Terpenoids from Eupatorium fortunei Turcz. Helv. Chim. Acta 2006, 89, 558–566. [Google Scholar] [CrossRef]
  5. Talavera-Alemán, A.; Rodríguez-García, G.; López, Y.; García-Gutiérrez, H.A.; Torres-Valencia, J.M.; del Río, R.E.; Cerda-García-Rojas, C.M.; Joseph-Nathan, P.; Gómez-Hurtado, M.A. Systematic Evaluation of Thymol Derivatives Possessing Stereogenic or Prostereogenic Centers. Phytochem. Rev. 2016, 15, 251–277. [Google Scholar] [CrossRef]
  6. Boukhris, M.; Bouaziz, M.; Feki, I.; Jemai, H.; El Feki, A.; Sayadi, S. Hypoglycemic and Antioxidant Effects of Leaf Essential Oil of Pelargonium graveolens L’Hér. in Alloxan Induced Diabetic Rats. Lipids Health Dis. 2012, 11, 81. [Google Scholar] [CrossRef] [PubMed]
  7. Oz, M.; Lozon, Y.; Sultan, A.; Yang, K.-H.S.; Galadari, S. Effects of Monoterpenes on Ion Channels of Excitable Cells. Pharmacol. Ther. 2015, 152, 83–97. [Google Scholar] [CrossRef] [PubMed]
  8. Akerele, O. Las Plantas Medicinales: Un Tesoro Que No Debemos Desperdiciar. Foro Mund. De La Salud 1993, 14, 390–395. [Google Scholar]
  9. Pérez-Vásquez, A.; Linares, E.; Bye, R.; Cerda-García-Rojas, C.M.; Mata, R. Phytotoxic Activity and Conformational Analysis of Thymol Analogs from Hofmeisteria schaffneri. Phytochemistry 2008, 69, 1339–1347. [Google Scholar] [CrossRef]
  10. Liang, H.; Bao, F.; Dong, X.; Tan, R.; Zhang, C.; Lu, Q.; Cheng, Y. Antibacterial Thymol Derivatives Isolated from Centipeda minima. Molecules 2007, 12, 1606–1613. [Google Scholar] [CrossRef]
  11. Bustos-Brito, C.; Sánchez-Castellanos, M.; Esquivel, B.; Calderón, J.S.; Calzada, F.; Yepez-Mulia, L.; Hernández-Barragán, A.; Joseph-Nathan, P.; Cuevas, G.; Quijano, L. Structure, Absolute Configuration, and Antidiarrheal Activity of a Thymol Derivative from Ageratina cylindrica. J. Nat. Prod. 2014, 77, 358–363. [Google Scholar] [CrossRef] [PubMed]
  12. Reist, M.; Testa, B.; Carrupt, P.-A.; Jung, M.; Schurig, V. Racemization, Enantiomerization, Diastereomerization, and Epimerization: Their Meaning and Pharmacological Significance. Chirality 1995, 7, 396–400. [Google Scholar] [CrossRef]
  13. Arreaga-González, H.M.; Pardo-Novoa, J.C.; del Río, R.E.; Rodríguez-García, G.; Torres-Valencia, J.M.; Manríquez-Torres, J.J.; Cerda-García-Rojas, C.M.; Joseph-Nathan, P.; Gómez-Hurtado, M.A. Methodology for the Absolute Configuration Determination of Epoxythymols Using the Constituents of Ageratina glabrata. J. Nat. Prod. 2018, 81, 63–71. [Google Scholar] [CrossRef] [PubMed]
  14. Bustos-Brito, C.; Vázquez-Heredia, V.J.; Calzada, F.; Yépez-Mulia, L.; Calderón, J.S.; Hernández-Ortega, S.; Esquivel, B.; García-Hernández, N.; Quijano, L. Antidiarrheal Thymol Derivatives from Ageratina glabrata. Structure and Absolute Configuration of 10-Benzoyloxy-8,9-epoxy-6-hydroxythymol Isobutyrate. Molecules 2016, 21, 1132. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Formulas of epoxythymol derivatives 1 and 2.
Figure 1. Formulas of epoxythymol derivatives 1 and 2.
Chemproc 14 00030 g001
Figure 2. Optical purity analysis by 1H NMR-BINOL methodology of (a) compound 1, (b) transesterification product 2, and (c) delivered escalimizated epoxythymol 1, evidencing a 75:25 (8S:8R) ratio.
Figure 2. Optical purity analysis by 1H NMR-BINOL methodology of (a) compound 1, (b) transesterification product 2, and (c) delivered escalimizated epoxythymol 1, evidencing a 75:25 (8S:8R) ratio.
Chemproc 14 00030 g002
Scheme 1. Proposed reaction mechanism for the transesterification of 1.
Scheme 1. Proposed reaction mechanism for the transesterification of 1.
Chemproc 14 00030 sch001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lorenzo-García, J.M.; Oliveros-Ortiz, A.J.; Arreaga-González, H.M.; Cortés-García, C.J.; del Río, R.E.; Rodríguez-García, G.; Gómez-Hurtado, M.A. Transesterification of a Natural Epoxythymol Is Favored under Alkaline Conditions, Preserving the Enantiomeric Purity. Chem. Proc. 2023, 14, 30. https://doi.org/10.3390/ecsoc-27-16140

AMA Style

Lorenzo-García JM, Oliveros-Ortiz AJ, Arreaga-González HM, Cortés-García CJ, del Río RE, Rodríguez-García G, Gómez-Hurtado MA. Transesterification of a Natural Epoxythymol Is Favored under Alkaline Conditions, Preserving the Enantiomeric Purity. Chemistry Proceedings. 2023; 14(1):30. https://doi.org/10.3390/ecsoc-27-16140

Chicago/Turabian Style

Lorenzo-García, Jessica M., Antonio J. Oliveros-Ortiz, Héctor M. Arreaga-González, Carlos J. Cortés-García, Rosa E. del Río, Gabriela Rodríguez-García, and Mario A. Gómez-Hurtado. 2023. "Transesterification of a Natural Epoxythymol Is Favored under Alkaline Conditions, Preserving the Enantiomeric Purity" Chemistry Proceedings 14, no. 1: 30. https://doi.org/10.3390/ecsoc-27-16140

Article Metrics

Back to TopTop