Acessibilidade / Reportar erro

Chemical Composition and Larvicidal Activity of Essential Oil Extracted from Anemia tomentosa (SAV.) Sw: Study Conducted in vitro and in silico

Abstract

The essential oil extracted from Anemia tomentosa (EOAT) has shown larvicidal potential against Aedes aegypti, based on methods in vitro and in silico. Chromatographic and spectrometric techniques (gas chromatography-mass spectrometry (CG-MS), gas chromatography-flame ionization detector (GC-FID)), as well as mono and bidimensional nuclear magnetic resonance spectroscopy (NMR) were used to assess 10 essential oil components. Chemical composition of essential oil comprised 87.51% sesquiterpenes, with emphasis on presilphiperfolan-1-ol (42.13%) and silphiperfol-5-ene (19.47%). Larvicidal assay conducted in vitro with EOAT has evidenced potential cytotoxic activity up to 48 h exposure to it; mortality rate observed for A. aegypti larvae exposed to essential oil reached 100%. Study conducted in silico with chemical compounds deriving from the herein investigated plant species has evidenced its potential to inhibit acetylcholinesterase in A. aegypti. Activity of triquinane sesquiterpenes ranging from -6.8 to -8.2 kcal mol-1 stood out in comparison to that of temephos (-7.5 kcal mol-1). Chemical compounds identified in the investigated essential oil presented low human and environmental toxicity, as observed in absorption, distribution, metabolism and excretion, and toxicity (ADMETox) predictions.

Keywords:
Anemiaceae; Aedes aegypti; molecular docking; acetylcholinesterase inhibitors; sesquiterpens


Introduction

Anemia tomentosa (SAV.) Sw. (Anemiaceae) is a recurrent plant species that has highly aromatic leaves and grows in rocky regions. It is used in folk medicine to improve individuals’ digestion, as expectorant and anti-flu drug, as well as to treat bronchitis. It naturally grows in South America, mainly in mountainous regions of Northeastern and Central Brazil, as well as in Bolivia, Paraguay and Argentina.11 Alvarez, M. A.; Pharmacological Properties of Native Plants from Argentina, 1st ed.; Springer: Cham, 2019. [Crossref]
Crossref...
The current study used A. tomentosa leaves collected in the Caatinga biome, in an area located on a rocky outcrop in Jequié County, Bahia State, Brazil.

Essential oils (EO) are secondary metabolites formed by volatile and semi-volatile compounds. They are produced by more than 17,500 plant species and are stored in different plant organs, such as leaves, fruits, flowers, roots and seeds. EOs are chemically featured by hydrocarbons and terpenes found in them.22 Kammoun, A. K.; Altyar, A. E.; Gad, H. A.; J. Pharm. Biomed. Anal. 2021, 198, 113991. [Crossref]
Crossref...
,33 Lenardão, E. J.; Savegnago, L.; Jacob, R. G.; Victoria, F. N.; Martinez, D. M.; J. Braz. Chem. Soc. 2016, 27, 435. [Crossref]
Crossref...
Sesquiterpene compounds of the EOs have fifteen carbon atoms. In addition, they can have double bonds and (or) oxo/oxy functions with broad action spectrum. EO extracted from A. tomentosa presented activity against tuberculosis bacillus, among other bacteria, as well as repellent activity against mosquitoes.44 Pinto, S. C.; Leitão, G. G.; de Oliveira, D. R.; Bizzo, H. R.; Ramos, D. F.; Coelho, T. S.; Silva, P. E. A.; Lourenço, M. C. S.; Leitão, S. G.; Nat. Prod. Commun. 2009, 4, 1675. [Crossref]
Crossref...

Aedes aegypti mosquito is one of the main vectors of pathogens capable of affecting humans. Dengue, Zika and Chikungunya viruses are examples of pathogens transmitted by female individuals belonging to this species.55 Muñoz-Benavent, M.; Pérez-Cobas, A. E.; García-Ferris, C.; Moya, A.; Latorre, A.; J. Pharm. Biomed. Anal. 2021, 194, 113787. [Crossref]
Crossref...
Several studies66 Magalhães, L. A. M.; Lima, M. P.; Marques, M. O. M.; Facanali, R.; Pinto, A. C. S.; Tadei, W. P.; Molecules 2010, 15, 5734. [Crossref]
Crossref...
conducted with essential oils have evidenced their larvicidal activity, which was associated with their composition mostly comprising sesquiterpenes.

Several molecular targets have been explored in studies focused on finding inhibitors to be used to develop new sustainable larvicides associated with natural active compounds.77 Jankowska, M.; Rogalska, J.; Wyszkowska, J.; Stankiewicz, M.; Molecules 2017, 23, 34. [Crossref]
Crossref...
Acetylcholinesterase (AChE) is the main molecular target with larvicidal action, because of interference generated by larvicides that block larval movements through mechanotropic receptors that prevent larvae from feeding. This feature emphasizes its cytotoxic potential and chemical control ability.88 Balachandran, C.; Anbalagan, S.; Kandeepan, C.; Arun Nagendran, N.; Jayakumar, M.; Fathi Abd_Allah, E.; Alqarawi, A. A.; Hashem, A.; Baskar, K.; J. Asia-Pac. Entomol. 2021, 24, 645. [Crossref]
Crossref...

There is a limited number of studies in the literature focused on investigating A. tomentosa composition and the activity of its essential oil. Studies99 Pinto, S. C.; Leitão, G. G.; Castellar, A.; Delia, D. S.; Lage, C. L. S.; Henriques, A. B.; Fernandes, J.; Motta, G. S.; Bizzo, H. R.; Leitão, S. G.; J. Essent. Oil Res. 2013, 25, 198. [Crossref]
Crossref...

10 Castilho, C. V. V.; Ferra Neto, J. F.; Leitão, S. G.; Barreto, C. S.; Pinto, S. C.; da Silva, N. C. B.; Plant Cell, Tissue Organ Cult. 2018, 133, 311. [Crossref]
Crossref...
-1111 Juliani, H. R.; Zygadlo, J. A.; Scrivanti, R.; de la Sota, E.; Simon, J. E.; Flavour Fragrance J. 2004, 19, 541. [Crossref]
Crossref...
have described compositions rich in monoand sesquiterpenes, as well as highlighted major compounds, such as pinocarvone, α-bisabolol and presilphiperfolan-8-ol. Presilphiperfolan-8-ol was found in A. tomentosa essential oil analyzed in previous studies.99 Pinto, S. C.; Leitão, G. G.; Castellar, A.; Delia, D. S.; Lage, C. L. S.; Henriques, A. B.; Fernandes, J.; Motta, G. S.; Bizzo, H. R.; Leitão, S. G.; J. Essent. Oil Res. 2013, 25, 198. [Crossref]
Crossref...

10 Castilho, C. V. V.; Ferra Neto, J. F.; Leitão, S. G.; Barreto, C. S.; Pinto, S. C.; da Silva, N. C. B.; Plant Cell, Tissue Organ Cult. 2018, 133, 311. [Crossref]
Crossref...
-1111 Juliani, H. R.; Zygadlo, J. A.; Scrivanti, R.; de la Sota, E.; Simon, J. E.; Flavour Fragrance J. 2004, 19, 541. [Crossref]
Crossref...
With respect to its biological activity, the investigated essential oil has shown antimycobacterial potential to be used against Mycobacterium tuberculosis and M. smegmatis strains, as well as its repellent action against A. aegypti.44 Pinto, S. C.; Leitão, G. G.; de Oliveira, D. R.; Bizzo, H. R.; Ramos, D. F.; Coelho, T. S.; Silva, P. E. A.; Lourenço, M. C. S.; Leitão, S. G.; Nat. Prod. Commun. 2009, 4, 1675. [Crossref]
Crossref...
,1212 Gillij, Y. G.; Gleiser, R. M.; Zygadlo, J. A.; Bioresour. Technol. 2008, 99, 2507. [Crossref] [PubMed]
Crossref...

Studies carried out with essential oils have evidenced the ability of volatile compounds to kill A. aegypti larvae, which was constantly attributed to high sesquiterpene activity in the composition of EOs.1313 Dias, C. N.; Moraes, D. F. C.; Parasitol. Res. 2014, 113, 565. [Crossref]
Crossref...
,1414 de Morais, S. M.; Facundo, V. A.; Bertini, L. M.; Cavalcanti, E. S. B.; dos Anjos Jr., J. F.; Ferreira, S. A.; de Brito, E. S.; de Souza Neto, M. A.; Biochem. Syst. Ecol. 2007, 35, 670. [Crossred]
Crossred...
Caryophyllene, α-muurolene, as well as trichinan sesquiterpenoids, were identified in these oils’ composition. They play a key role in the larvicidal activity of these volatile mixes since they inhibit the activity of the AChE enzyme in these insects’ larvae.1515 Huy Hung, N.; Ngoc Dai, D.; Satyal, P.; Thi Huong, L.; Thi Chinh, B.; Quang Hung, D.; Anh Tai, T.; Setzer, W. N.; Chem. Biodiversity 2021, 18, e2100145 [Crossref] [PubMed]
Crossref...
,1616 Botelho, A. S.; Ferreira, O. O.; de Oliveira, M. S.; Cruz, J. N.; Chaves, S. H. R.; do Prado, A. F.; do Nascimento, L. D.; da Silva, G. A.; do Amarante, C. B.; Andrade, E. H. A.; Int. J. Mol. Sci. 2022, 23, 11172. [Crossref]
Crossref...

Thus, given the larvicidal potential of essential oils, mainly due to the action of sesquiterpenes in inhibiting AchE enzyme of the insects, in addition to the lack of studies focused on investigating essential oil extracted from A. tomentosa (EOAT), it is necessary assessing a new aspect of the action of EOATs, based on the following steps: essential oil identification, quantification and featuring as a potential larvicidal agent.

The aim of the current study was to perform the chemical and pharmacological assessment of essential oil extracted from species A. tomentosa grown in Jequié County, Bahia State, Brazil. Gas chromatography techniques (gas chromatography-mass spectrometry (GC MS) and gas chromatography-flame ionization detector (GC-FID)), biological assays in vitro and studies in silico were used to assess both the chemical composition and larvicidal potential of the investigated essential oil against Aedes aegypti. The 1H and 13C nuclear magnetic resonance (NMR) data have confirmed the identity of the major compounds found in EOAT.

Experimental

Plant material collection

A. tomentosa shoots were collected in Poço Dantas region (latitude (W): -13.8575, longitude (S): -40.0836), rural area of Jequié County, Bahia State, Brazil, in February 2021. Two vouchers were deposited in the Herbarium of Southwestern Bahia State University and received registration and botanical identification code of HUESB-8934.

Essential oil extraction and analyses

Fresh A. tomentosa leaves (1,300 g) were subjected to hydrodistillation for EOAT (5.1 g) extraction at crude yield close to 0.4%. The extracted essential oil presented yellow color and typical citrus-woody odor. The obtained crude oil was passed through a column filled with anhydrous sodium sulfate to rule out moisture. EOAT (density measured at 0.951 g mL-1) was stored in freezer at approximately -18 °C.

Chromatographic analyses aimed at identifying the main EOAT constituents were carried out in Shimadzu gas chromatograph, model QP2010-SE (Shimadzu company, Kyoto, Japan), coupled to mass spectrometry detector (GC-MS) equipped with a simple quadrupole ion filter detector, which operates by electron impact ionization (EI, 70 eV.). The equipment was set at scan speed of 1,000 amu s-1 and m/z fragment detection ranging from 45 to 800 Da. Chromatographic separation applied to EOAT components used an Agilent SLB-5ms poly-(5% diphenyl/95% dimethyl siloxane) phase capillary column (length: 30 m × internal diameter: 0.32 mm × film thickness: 0.32 μm). Helium provided by White Martins (Salvador, Brazil) was used as carrier gas at linear velocity of 48.4 cm s-1 and at pressure of 22.2 kPa. Samples were dissolved (1%) in dichloromethane and injected (1 µL) in triplicate, in split mode (1:40). Both the injector and the detector were adjusted to operate at 220 and 240 °C, respectively. The oven was set at heating rate of 3 °C min-1, temperature increased from 40 to 280 °C, and this temperature was kept for additional 4 min. The ion source and the interface were kept at 240 and 280 °C, respectively. Essential oil components were identified based on comparison to their retention indices (RI), which were calculated in a series of n-alkanes (C8-C36) provided by Sigma-Aldrich (São Paulo, Brazil), as well as by comparing their mass spectra to the NIST-14 library (match > 90%).

EOAT constituents were quantified in Shimadzu gas chromatograph, model GC-2010 Plus (Shimadzu, Kyoto, Japan), equipped with flame ionization detector (GC-FID). Column type, oven and inlet temperature adjustment parameters were the same as the ones adopted in GC-MS. The flow rates of gases used as carriers (column, split vent, purge and H2), FID flame (H2) and make-up (N2) by White Martins (Salvador, Brazil) were 14.0, 40 and 30 mL min 11 Alvarez, M. A.; Pharmacological Properties of Native Plants from Argentina, 1st ed.; Springer: Cham, 2019. [Crossref]
Crossref...
, respectively. Synthetic air flow of 400 mL min-1 was used to feed the FID flame. Sample split ratio was 1:10. Operating parameters were established after checking the best resolution. Injections were performed in triplicate, at volume of 1.0 μL.

Phytochemical study applied to methanolic extract

In total, 1 kg of A. tomentosa leaves was subjected to EO extraction through maceration with methanol (Synth, Salvador, Brazil) in three repetitions (48 h period, each). The aliquot of 39.8 g of Anemia tomentosa methanolic extract (ATME) was obtained after solvent evaporation.

In total, 30.0 g of ATME were subjected to preliminary fractionation in filtering column (silica gel 60 70 230 Mesh) (Synth, Salvador, Brazil), from which 4 different fractions (F1 to F4) were obtained based on polarity: hexane (Synth, Salvador, Brazil) (F1, 1.7 g), chloroform (Synth, Salvador, Brazil) (F2, 0.3 g), ethyl acetate (Synth, Salvador, Brazil) (F3, 2.1 mg) and methanol (F4, 20.7 mg), consecutively. The chromatographic profiles fractions were obtained through thin-layer chromatography (TLC) (Synth, Salvador, Brazil).

Given its best chromatographic profile and highest yield, F3 was subjected to column fractionation by liquid chromatography in a column (35 mm diameter) packed with silica gel 60 (70-230 Mesh). It was done based on using the polarity gradient with hexane:ethyl acetate solvents, at the 10:0, 9.5:0.5, 9:1, 8:2 and 7:3 ratios. In total, 95 fractions (10 mL, each) were collected. Fractions 37, 38 and 39 presented a yellow substance with oily appearance (mass = 4.6 mg), which was soluble in chloroform and dichloromethane, and presented retention factor (Rf) close to 0.5 in the TLC analysis conducted with eluent mix of hexane:ethyl acetate solvents (9:1). Later on, the aforementioned substance was identified as presilphiperfolan-1-ol (8).

Spectrometric and optical analyses

Infrared and 1H and 13C magnetic resonance spectrometric analyses were used to improve and resolve ambiguities in the process to identify chemical constituents of EOs through GC-MS. Fourier transform infrared (FTIR) spectra (4,000 to 450 cm-1) were obtained through PerkinElmer attenuated total reflectance (ATR)-FTIR spectrometer, model Spectrum Two (PerkinElmer, Brazil). One and two-dimensional 1H and 13C NMR spectra were recorded in Varian Inova 500 NMR spectrometer (Varian, USA), (500 MHz); tetramethylsilane (TMS) was used as internal standard, whereas chloroform-d was applied as solvent. Optical rotations were measured at 25 °C in Anton Paar MCP 300 polarimeter (Anton Paar, USA)

Larvicidal assay in vitro

A. aegypti 3rd and 4th instar larvae were obtained from Rockefeller-strain eggs, which were kindly provided by the Toxicology Research Laboratory of the Antibiotics Department of Federal University of Pernambuco (UFPE). They were placed in glass container filled with 30 mL of 0.1% EOAT and solubilized in deionized water and Tween 80 (0.5%) to carry out the tests. Each test used 30 larvae and comprised five repetitions (150 larvae, in total). Control solutions comprised (i) 0.5% Tween 80, and (ii) distilled and deionized H2O. Larval death (nine dead larvae, in total) was assessed 1-12 h (2 h intervals), as well as 24 and 48 h after the beginning of the experiment. Experiments were carried out under laboratory conditions and followed a completely randomized design (CRD).

Homology model prediction

AChE1 enzyme deriving from A. aegypti was subjected to homology modeling based on using Swiss-Modeller PDB and AlphaFold. Protein sequence was retrieved from the A. aegypti GenBank (ID: ABN09910.1). The best scoring model was assessed in Swiss-Modeller based on the Global QMEAN closest to 1.0 and on identity close to 100%.1717 Bienert, S.; Waterhouse, A.; de Beer, T. A. P.; Tauriello, G.; Studer, G.; Bordoli, L.; Schwede, T.; Nucleic. Acids Res. 2017, 45, D313. [Crossref]
Crossref...
PDB 6ARY was the model identified and used for homology modeling purposes. The modeled structure was subjected to energy minimization process based on using Gromacs 2018 package1818 Gromacs, version 2018; Royal The Gromacs Development Team, Sweden, 2018. [Link] accessed in July 2024
Link...
with GROMOS96 force field.1919 GROMOS, GROMOS96; Gromos Development Group, Switzerland, 1996. [Link] accessed in July 2024
Link...
,2020 Benkert, P.; Biasini, M.; Schwede, T.; Bioinformatics 2011, 27, 343. [Crossref]
Crossref...
Hydrogen atoms were added to finalized target-protein model and used for molecular docking analysis with MGL Tools.2121 Huey, R.; Morris, G. M.; Olson, A. J.; Goodsell, D. S.; J. Comput. Chem. 2007, 28, 1145. [Crossref]
Crossref...
All images and the electrostatic potential of the enzyme surface were visualized in PyMOL2222 PyMOL, version 2.1; Schrödinger, LLC, 2018. and Discovery Studio 2021 software.1717 Bienert, S.; Waterhouse, A.; de Beer, T. A. P.; Tauriello, G.; Studer, G.; Bordoli, L.; Schwede, T.; Nucleic. Acids Res. 2017, 45, D313. [Crossref]
Crossref...
,2323 BIOVIA Discovery Studio, v. 4.5; Dassault Systèmes, San Diego, USA, 2021.

Dataset, virtual screening, molecular docking and ADMETox tools

All molecules were checked and designed in Marvin Sketch software.2424 Marvin Sketch, version 23.4; ChemAxon, Hungary, 2023. [Link] accessed in July 2024
Link...
Molecular structures were downloaded in SMILES format and converted into 3D sdf format in Open babel software for docking calculation purposes.2525 Open Babel, version 3.1.1; The Open Babel Project, United States, 2023. [Link] accessed in July 2024
Link...
The acethycolinesterase (AChE) crystallographic structure (6ARY) of Anopheles gambiae was obtained from the Protein Data Bank (PDB). All docking simulations were performed in AutoDock Vina software.2626 AutoDock Vina, version 1.1.2; The Scripps Research Institute, United States, 2010. [Link] accessed in July 2024
Link...
Docking was performed between the proposed ligands (10 compounds deriving from EOAT, acetylcholine (PubChem CID 187), temephos (PubChem CID 5392) and the receptor), which were prepared and converted into pdbqt format in Autodock tools software. Docking results and the assessment of each receptor-ligand complex, such as affinity energy (kcal mol 11 Alvarez, M. A.; Pharmacological Properties of Native Plants from Argentina, 1st ed.; Springer: Cham, 2019. [Crossref]
Crossref...
) and ligand positioning inside the acethycolinesterase active site, were analyzed in PyMOL 2.1 software. Nine ligand poses were generated for each complex and returned their respective affinity energies. Subsequently, all best selected ligand-protein poses were graphically plotted in PyMOL 2.1 software. Their respective 2D interaction maps were plotted in Discovery Studio 4.5 software. ADMETox properties were assessed in Data Warrior,TM 2727 Pires, D. E. V.; Blundell, T. L.; Ascher, D. B.; J. Med. Chem. 2015, 58, 4066. [Crossref]
Crossref...
PkCSMTM 2828 Sander, T.; Freyss, J.; von Korff, M.; Rufener, C.; J. Chem. Inf. Model. 2015, 55, 460. [Crossref]
Crossref...
and ToxCastTM 2929 U. S. Environmental Protection Agency (USEPA); ToxCastTM Predicting Hazard, Characterizing Toxicity Pathways, and Prioritizing the Toxicity Testing of Environmental Chemicals, https://www.epa.gov/comptox-tools/toxicity-forecasting-toxcast, accessed in June 2024.
https://www.epa.gov/comptox-tools/toxici...
software.

Results and Discussion

Chemical composition analysis of EOAT

The herein extracted EOAT (0.4%) presented yellowish color, citrus-woody odor and density (0.951 g mL-1) typical of essential oil rich in terpenoids. In addition, its physicochemical data and organoleptic features were similar to those observed in previous studies.44 Pinto, S. C.; Leitão, G. G.; de Oliveira, D. R.; Bizzo, H. R.; Ramos, D. F.; Coelho, T. S.; Silva, P. E. A.; Lourenço, M. C. S.; Leitão, S. G.; Nat. Prod. Commun. 2009, 4, 1675. [Crossref]
Crossref...
,1111 Juliani, H. R.; Zygadlo, J. A.; Scrivanti, R.; de la Sota, E.; Simon, J. E.; Flavour Fragrance J. 2004, 19, 541. [Crossref]
Crossref...
GC-MS and GC-FID analyses (Table 1) enabled identifying and quantifying the ten main chemical constituents of the essential oil extracted from fresh A. tomentosa leaves, which recorded prevalence of 87.5% sesquiterpenes.

Table 1
Chemical constituents of EOAT identified through GC-MS, 1H and 13C NMR and FTIR

The main EOAT constituent is a triquinane sesquiterpene called presilphiperfolan-1-ol (8, tR at 33.671 min, 42.13%). The 13C NMR spectra analysis has evidenced the most intense 13C (C-OH) signal at 84.90 ppm, it referred to compound 8.44 Pinto, S. C.; Leitão, G. G.; de Oliveira, D. R.; Bizzo, H. R.; Ramos, D. F.; Coelho, T. S.; Silva, P. E. A.; Lourenço, M. C. S.; Leitão, S. G.; Nat. Prod. Commun. 2009, 4, 1675. [Crossref]
Crossref...
,3030 Coates, R. M.; Ho, Z.; Klobus, M.; Wilson, S. R.; J. Am. Chem. Soc. 1996, 118, 9249. [Crossref]
Crossref...
Compound 8 was isolated from A. tomentosa leaves (ethyl acetate extract) through liquid chromatography on silica gel column eluted with hexane:ethyl acetate (9:1). The analysis applied to spectrometric data (optical activity [α]D2525 Open Babel, version 3.1.1; The Open Babel Project, United States, 2023. [Link] accessed in July 2024
Link...
, FTIR, 1H and 13C NMR) on compound 8 has confirmed it as presilphiperfolan-1-ol31 (see Table 2).

Table 2
Structures of EOAT-constituent sesquiterpenes elucidated by GC-MS, 1H and 13C NMR (500 MHz, chloroform-d), FTIR, and comparison to spectral data available in the literature

Complementary analysis applied to the 1H and 13C NMR spectra of EOAT identified signals typical of chemical sesquiterpenes’ shifts with alcohol hydroxyls and double bond (C=C),3232 Silverstein, R. M.; Webster, F. X.; Kiemle, D. J.; Bryce, D. L.; Spectrometric Identification of Organic Compounds, 7th ed.; John Wiley & Sons: New York, US, 2014. which enabled featuring the structure of its nine sesquiterpene constituents. Shift values observed at dH 1.64 (d, J 8.5, H-8) and dc 84.90 (C-1) were assigned to presilphiperfolan-1-ol (8), which was the most abundant EOAT constituent.3131 Hong, A. Y.; Stoltz, B. M.; Angew. Chem., Int. Ed. 2014, 53, 5248. [Crossref]
Crossref...
Proton-shift values observed at dH 4.94 (H-5, s-br), 4.90 (H-5, s-br) and 5.43 (H-5 and H-9, m) were assigned to four olefinic protons in sesquiterpenes (2), (4) and (5), respectively.3333 Bohlmann, F.; Jakupovic, J.; Phytochemistry 1980, 19, 259. [Crossref]
Crossref...
,3434 Kashman, Y.; Rudi, A.; Gutman-Naveh, N.; Tetrahedron 1978, 34, 1227. [Crossref]
Crossref...
With respect to the 13C NMR spectrum, four other signals of oxygenated sesquiterpenes were also observed at dC 78.48 (C-6), 89.69 (C-7), 82.11 (C-8), and 96.93 (C-8). They were assigned to oxygenated sesquiterpenes, such as silphiperfolan-6-α-ol (6), cameroonan-7α-ol (7), prenopsan-8-ol (9) and presilphiperfolan-8α-ol (10), respectively.3535 Weyerstahl, P.; Marschall, H.; Seelmann, I.; Jakupovic, J.; Eur. J. Org. Chem. 1998, 1998, 1205. [Crossref]
Crossref...
Five pairs of 13C olefinic carbon NMR signals observed at dc 133.11 and 142.40 (C5=C6), dc 141.55 and 161.07 (C8=C7), dc 133.46 and 140.80 (C5=C6), dc 120.63 and 134.18 (C9=C10), dc 124.38 and 135.93 (C5=C4), were assigned to EOAT compounds 2, 3, 4 and 5, respectively. Compound 5 presented two double bonds3030 Coates, R. M.; Ho, Z.; Klobus, M.; Wilson, S. R.; J. Am. Chem. Soc. 1996, 118, 9249. [Crossref]
Crossref...
,3333 Bohlmann, F.; Jakupovic, J.; Phytochemistry 1980, 19, 259. [Crossref]
Crossref...
,3434 Kashman, Y.; Rudi, A.; Gutman-Naveh, N.; Tetrahedron 1978, 34, 1227. [Crossref]
Crossref...
(Table 2).

Assessing the larvicidal activity of EOAT in vitro

Results in the current study have evidenced broad larvicidal activity of EOAT (Figure 1). In total, 71.31% A. aegypti larvae were dead after 24 h experiment, whereas all larvae (100%) were dead after 48 h.

Figure 1
Cumulative mortality rate (%) observed for Aedes aegypti larvae, based on time exposed to EOAT treatment at 0.1%.

Chemical diversity of terpenes leads to variation in their polarity degree. Thus, they can increase the absorption of both lipophilic and hydrophilic substances by the membranes, and it contributes to the synergistic action resulting from the combination among oil components. Thus, lipophilicity gradient variation resulting from the mix of neutral and functionalized sesquiterpenes can explain the larvicidal activity observed for this class of compounds.1313 Dias, C. N.; Moraes, D. F. C.; Parasitol. Res. 2014, 113, 565. [Crossref]
Crossref...

Sesquiterpenes found in essential oils can show activity in larvae, at different target sites. Oviposition inhibition can be achieved through interaction with odorant-binding protein 1 (OBP1), which is one of the most important players in the olfactory system.3636 Albuquerque, B. N. L.; da Silva, M. F. R.; da Silva, P. C. B.; de Lira Pimentel, C. S.; Lino da Rocha, S. K.; Farias de Aguiar, J. C. R. O.; Neto, A. C. A.; Paiva, P. M. G.; Gomes, M. G. M.; da Silva-Júnior, E. F.; Navarro, D. M. A. F.; Ind. Crops Prod. 2022, 182, 114830. [Crossref]
Crossref...
Activities observed for these EO constituents were reported in other receptors and proteins, such as γ-aminobutyric acid (GABA) receptor. Among them, one finds death by nervous system overactivation; blockade of the protein accounting for cholesterol transport in insects (AeSCP2), which is found at high concentrations in larvae; and acetylcholinesterase (AChE) inhibition, which leads to a typical neurotoxic effect produced by organophosphorus and carbamate insecticides.3737 Andrade-Ochoa, S.; Correa-Basurto, J.; Rodríguez-Valdez, L. M.; Sánchez-Torres, L. E.; Nogueda-Torres, B.; Nevárez-Moorillón, G. V.; Chem. Cent. J. 2018, 12, 53. [Crossref]
Crossref...

Triquinane sesquiterpenes, such as (3), (7) and (10), were described in the literature3838 Fernandez, C. M. M.; Barba, E. L.; Fernandez, A. C. M.; Cardoso, B. K.; Borges, I. B.; Takemura, O. S.; Martins, L. A.; Cortez, L. E. R.; Cortez, D. A. G.; Gazim, Z. C.; J. Essent. Oil Bear. Plants 2014, 17, 813. [Crossref]
Crossref...
as having larvicidal activity. Pavela et al.3939 Pavela, R.; Maggi, F.; Mbuntcha, H.; Woguem, V.; Fogang, H. P. D.; Womeni, H. M.; Tapondjou, L. A.; Barboni, L.; Nicoletti, M.; Canale, A.; Benelli, G.; Parasitol. Res. 2016, 115, 4617. [Crossref]
Crossref...
reported these larvicidal action of the compounds against species Culex quinquefasciatus Say, which is the filariasis transmission agent. Silphiperfolanes, presilphiperfolanes and cameroonane are sesquiterpene groups deriving from caryophyllene.4040 Fraga, B. M.; Nat. Prod. Rep. 2011, 28, 1580. [Crossref]
Crossref...
The investigated EOAT was rich in compounds presenting caryophyllene skeleton. This finding corroborates the high larvicidal activity presented by this natural product.66 Magalhães, L. A. M.; Lima, M. P.; Marques, M. O. M.; Facanali, R.; Pinto, A. C. S.; Tadei, W. P.; Molecules 2010, 15, 5734. [Crossref]
Crossref...
,3838 Fernandez, C. M. M.; Barba, E. L.; Fernandez, A. C. M.; Cardoso, B. K.; Borges, I. B.; Takemura, O. S.; Martins, L. A.; Cortez, L. E. R.; Cortez, D. A. G.; Gazim, Z. C.; J. Essent. Oil Bear. Plants 2014, 17, 813. [Crossref]
Crossref...

Molecular docking of major A. tomentosa (SAV.) Sw essential oil compounds against acetylcholine-sterase (AChE) from Aedes aegypti

The best scoring model used in the current study was assessed in Swiss-Modeller, based on Global QMEAN of 0.92 (the closer to this value, the higher the accuracy of the model) and identity of 92.58%. Results have evidenced that the modeled structure maintained high homology degree to AChE deriving from A. aegypti. Molecular docking results have indicated that all sesquiterpenes described for the investigated EOAT interacted with amino acids in the active site of AChE deriving from A. aegypti. In addition, they showed affinity energy ranging from -4.0 to -8.2 kcal mol-1, and it featured their potential to inhibit the enzyme at energies higher than that of larvicide temephos (-7.5 kcal mol-1), except for (1) (-4.3 kcal mol-1), which showed similar energy to that of endogenous ligand acetylcholine (-4.0 kcal mol-1). The active site of this enzyme has ion channel shape. Moreover, it comprises the 10 most representative amino acid residues; among them, sesquiterpene compounds bind to the active site of the AChE enzyme, which presents a quite characteristic behavior at the entrance of the chanel (Figure 2).

Figure 2
In gray: the active “gorge”-shaped site of the AChE deriving from Aedes aegypti stands out; red: acetylcholine; green: temephos; blue: hexanal; yellow: silphiperfol-5-ene; magenta: presilphiperfol-7-ene, cyan: α-muurolene, orange: silphiperfolan-6α-ol; and beige: cameroonan-7α-ol.

AChE found in arthropods has important physiological functions, such as maintaining neuronal transmission and locomotion mechanisms accounting for favoring environmental exploration and feeding processes. In addition, this active site of the enzyme is evolutionarily conserved in mosquito species belonging to genera Aedes and Anopheles.4141 Engdahl, C.; Knutsson, S.; Fredriksson, S.-Å.; Linusson, A.; Bucht, G.; Ekström, F.; PLoS One 2015, 10, e0138598. [Crossref]
Crossref...
AChE is a major target for insecticides. Thus, this disruption of the receptor leads to ACh accumulation, which, in its turn, triggers neurotransmission hyperstimulation and eventual blockage in insects (Figure S1, Supplementary Information (SI) section).77 Jankowska, M.; Rogalska, J.; Wyszkowska, J.; Stankiewicz, M.; Molecules 2017, 23, 34. [Crossref]
Crossref...

Overall, EOs are almost fully composed of sesquiterpenes, which have similar chemical structures that, in their turn, present small variations. This feature justifies their close affinity energies. Most sesquiterpenes, (4) (-8.2 kcal mol-1) and (8) (-7.6 kcal mol-1), have shown better affinity than the temephos larvicide (-7.2 kcal mol-1). Results of assay conducted in silico corresponded to data deriving from assay conducted in vivo, according to which, larvicidal activity recorded 100% larval death within 48 h. Computational and biological data integration enabled inferring that biological activity may be associated with the individual potential of isolated compounds (most of them), as well as with the synergistic action observed among them. Terpenoid compounds found in EOs have great neurotoxic potential. They can cause paralysis followed by death in insects, and this factor features them as bioinsecticides. Table 3 shows the affinity energies of all ligands.

Table 3
Intermolecular interactions of sesquiterpenes found in Anemia tomentosa (SAV.) essential oil in the Aedes aegypti acetylcholinesterase site

Family Anemiaceae has been investigated as acknowledged source of terpenoid compounds, mainly of triquinane sesquiterpenes, besides being considered a standard taxonomic chemomarker for genus Anemia. These compounds present the larvicidal and anticholinesterase potential described in other species that are rich in terpenoid compounds, such as the ones belonging to genera Lantana, Pomacea, Gyraulus, Aframomum, Dichrostachys and Echinops.3939 Pavela, R.; Maggi, F.; Mbuntcha, H.; Woguem, V.; Fogang, H. P. D.; Womeni, H. M.; Tapondjou, L. A.; Barboni, L.; Nicoletti, M.; Canale, A.; Benelli, G.; Parasitol. Res. 2016, 115, 4617. [Crossref]
Crossref...
Understanding the best molecular mechanism of interaction with these residues enables building a library of analogs, as drug design strategy applied to compounds with bioinsecticidal activity against A. aegypti.

Notably, studies focused on investigating dengue vector insecticide-resistance mechanisms are of paramount importance, since this disease is a public health issue in several countries.4343 Vargas, L. D. L.; Ferreira, S. M. B.; Souza, M. D.; da Silva, C. A. L.; Shimoya-Bittencourt, W.; Rev. Cien. Med. Biol. 2022, 21, 98. [Crossref]
Crossref...
The large-scale production and the frequent use of insecticides have caused their accumulation in ecosystems, and it led to environmental contamination, as well as to toxicity in several species, including humans. Many vector control strategies are used, in an integrated manner, to fight A. aegypti. Among them, one finds the use of biopesticides, such as microorganisms, viruses, biological toxins and natural products. Combining different approaches can help maximizing vector control effectiveness and avoiding the outspread of mosquito-borne diseases.4444 Zara, A. L. S. A.; Santos, S. M.; Fernandes-Oliveira, E. S.; Carvalho, R. G.; Coelho, G. E.; Epidemiol. Serv. Saúde 2016, 25, 1. [Crossref]
Crossref...

Anticholinesterase insecticides currently used for biological vector control purposes are inhibitors capable of forming a covalent bond with a conserved serine in the active site of AChE. This active site involves a serine residue: Ser203, in humans and Ser397, in A. aegypti. It is located at the basis of a deep and narrow gorge, where it forms part of the catalytic triad along with His447 and Glu334 residues, among other aromatic residues that form the basis of the structural gorge in this active site of the enzyme (Figure 2).4141 Engdahl, C.; Knutsson, S.; Fredriksson, S.-Å.; Linusson, A.; Bucht, G.; Ekström, F.; PLoS One 2015, 10, e0138598. [Crossref]
Crossref...

All compounds found in the investigated EOAT interacted at the entrance of the AChE “gorge” and showed potential to develop new natural inhibitors as alternative to increasing resistance of the insect to organophosphate larvicides. The following factors may be associated with the high potential larvicidal activity observed for EOAT: the relative concentration of each constituent in the mix found in EO,4545 Jukic, M.; Politeo, O.; Maksimovic, M.; Milos, M.; Milos, M.; Phytother. Res. 2007, 21, 259. [Crossref]
Crossref...
spatial size of the molecules4646 Reegan, A. D.; Stalin, A.; Paulraj, M. G.; Balakrishna, K.; Ignacimuthu, S.; Al-Dhabi, N. A.; Med. Chem. Res. 2016, 25, 1411. [Crossref]
Crossref...
and the presence of fewer rings in its structure associated with higher larvicidal activity.4747 López, M. D.; Campoy, F. J.; Pascual-Villalobos, M. J.; Muñoz Delgado, E.; Vidal, C. J.; Chem. Biol. Interact. 2015, 229, 36. [Crossref]
Crossref...
Therefore, the inhibitory activity observed in AChE is associated with the synergistic potential observed among its compounds and with its selective and irreversible ability to bind to the active site of the target enzyme.

Structural differences in amino acid position may be strategies leading to selectivity of the compounds to the active site of A. aegypti AChE over the same enzyme in humans. Divergences in the “gorge” entry between mosquitoes (free cysteine) and humans (Phe295) may be a potential target for selective covalent inhibitors.4848 Pezzementi, L.; Rowland, M.; Wolfe, M.; Tsigelny, I.; Invertebr. Neurosci. 2006, 6, 47. [Crossref]
Crossref...
Catalytic serine residue has evidenced the likelyhood of finding high selective inhibition rates for mosquito enzymes over vertebrate enzymes.4949 Alout, H.; Labbé, P.; Berthomieu, A.; Djogbénou, L.; Leonetti, J.-P.; Fort, P.; Weill, M.; PLoS One 2012, 7, e47125. [Crossref]
Crossref...
Although the potential of covalent inhibitors has been investigated in several studies, inhibitors with different action modes, such as non-covalent inhibitors, are poorly explored in the literature.4848 Pezzementi, L.; Rowland, M.; Wolfe, M.; Tsigelny, I.; Invertebr. Neurosci. 2006, 6, 47. [Crossref]
Crossref...
,4949 Alout, H.; Labbé, P.; Berthomieu, A.; Djogbénou, L.; Leonetti, J.-P.; Fort, P.; Weill, M.; PLoS One 2012, 7, e47125. [Crossref]
Crossref...
Furthermore, one of the barriers hindering the development of new insecticides lies on overcoming increasing resistance of the insects to AChE inhibitors due to mutations in the active site. Despite these mutations, it is possible targeting resistant mutant enzymes with non-covalent inhibitors, such as those obtained from EO.3737 Andrade-Ochoa, S.; Correa-Basurto, J.; Rodríguez-Valdez, L. M.; Sánchez-Torres, L. E.; Nogueda-Torres, B.; Nevárez-Moorillón, G. V.; Chem. Cent. J. 2018, 12, 53. [Crossref]
Crossref...

Triquinane sesquiterpenes presented similar chemical structure since they were obtained through the same chemosynthetic route. They showed similar affinity energy features and interaction with the enzyme site.5050 Muangphrom, P.; Misaki, M.; Suzuki, M.; Shimomura, M.; Suzuki, H.; Seki, H.; Muranaka, T.; Phytochemistry 2019, 164, 144. [Crossref]
Crossref...
Thus, all molecules have formed a complex with the active site of AChE, as shown in Figure 3. The 3 ligands showing the best energies: (2) (-8.2 kcal mol-1), (4) (-8.2 kcal mol-1) and (6) (-7.9 kcal mol-1), recorded the best intermolecular interactions, with emphasis on van der Waals, π-alkyl and polar (π-alkyl) interactions with amino acid residues found in the catalytic site of the enzyme, such as Cys414, Glu415, Phe416 and Phe457 (Figure 3). Temephos interacted with 19 aa residues, acetylcholine interacted with 13 aa residues, whereas natural compounds recorded from 9 to 11 interactions. They were the ones showing trichinane structures. All compounds, except for (1), interacted with at least 8 amino acids similar to temephos, mainly with those involved in the catalytic site of the enzyme (Tyr460, Tyr456, Tyr249, Tyr461, Ile413, Gly412, Glu415, Phe416, Phe457 and Trp408). Figure S1 shows all ligand-receptor complexes (SI section).

Figure 3
2D map depicting Aedes aegypti AChE channel binding pocket-amino acid interactions. (a) AChE complex with the main ligands; (b) complex receptor-7-epi-silphiperfol-5-ene AChE enzyme; (c) complex receptoracetylcholine AChE enzyme; (d) complex receptor-temephos AChE enzyme.

Predicting ADMETox properties in silico

Predictions in silico carried out for physicochemical, pharmacokinetic and toxicological parameters of sesquiterpenes identified in EOAT have shown that these compounds presented good oral absorption and bioavailability, according to Lipinski’s rule of 5 (Ro5) (Table 4). Sesquiterpenes presented small polar surface area (159.6-181.9), high intestinal absorption rate (94.1-98.6%) and the best liposolubility profile (cLogP), in comparison to temephos (88.8%). Because they have similar chemical structures, they presented distribution volumes (VDs = 0.78) and high binding to plasma proteins as similar features. These molecules presented high distribution-to-tissues ability, as shown by the low fraction of unbound drug in blood plasma (Fu = 0.18) in comparison to temephos (Fu = 0.0), which, in its turn, can induce higher toxicity at higher concentrations in tissues as opposed to blood. Thus, triquinane sesquiterpenes presented the best distribution profile in human tissues. Sesquiterpenes did not show hepatotoxicity. They presented good renal excretion profile in comparison to Temephos. This feature may help their metabolization and excretion in case they come in contact with the human body during their use as larvicide (Table 5).5151 Zárybnický, T.; Boušová, I.; Ambrož, M.; Skálová, L.; Arch. Toxicol. 2018, 92, 1. [Crossref]
Crossref...

Table 4
Physical-chemical parameters of sesquiterpenes (A. tomentosa) and temephos
Table 5
Pharmacokinetic properties of sesquiterpenes (A. tomentosa) and temephos in silico

The herein analyzed triquinane sesquiterpenes did not interfere with the activity of enzyme CYP3A4 (it is the main enzyme in the hepatic microsomal complex accounting for metabolizing most orally administered chemicals) and it indicated easy biotransformation, as well as favored detoxification and CYP3A4 excretion from the body.4444 Zara, A. L. S. A.; Santos, S. M.; Fernandes-Oliveira, E. S.; Carvalho, R. G.; Coelho, G. E.; Epidemiol. Serv. Saúde 2016, 25, 1. [Crossref]
Crossref...
In addition to low risk of liver damage, these compounds presented higher renal clearance rate (0.43 0.99 log mg kg-1 day-1) than that observed for temephos (0.09 log mg kg-1 day-1); moreover, they may be more easily excreted. These compounds’ lipophilic features (Table 5) enable them to easily attach to, and penetrate, the larval structure. Consequently, it increases their inhibition power and has less impact on humans in comparison to temephos, which, in its turn, can be more easily absorbed through inhalation, through the skin or through ingestion.5252 Huang, H.-T.; Lin, C.-C.; Kuo, T.-C.; Chen, S.-J.; Huang, R.-N.; Planta 2019, 250, 59. [Crossref]
Crossref...
,5353 Luz, T. R. S. A.; de Mesquita, L. S. S.; do Amaral, F. M. M.; Coutinho, D. F.; Acta Trop. 2020, 212, 105705. [Crossref]
Crossref...
These phamacokinetic properties are shown in Table 5.

According to the toxicity profile, oral rat acute toxicity (LD50) and maximum tolerated toxicity values in humans (MRTD) observed for sesquiterpenes were within the safe range for human intake. Chronic oral toxicity values observed in rats (LOAEL) were higher than those observed for temephos larvicide. This finding has evidenced that natural compounds have better safety range for chronic dose use (Table 6).

Table 6
Drug likeness, drug score and toxicity values recorded for sesquiterpenes (A. tomentosa) and temephos

Because they are authentic chemical structures, drug likeness (DL) and drug score (DS) values recorded for sesquiterpenes differed from those observed for temephos. This factor has evidenced their potential to be used to develop new natural and semisynthetic compounds deriving from their structures.5353 Luz, T. R. S. A.; de Mesquita, L. S. S.; do Amaral, F. M. M.; Coutinho, D. F.; Acta Trop. 2020, 212, 105705. [Crossref]
Crossref...
None of the natural compounds presented toxicological risks, except for (1), which showed potential mutagenic, irritative and aggressive activity in the reproductive system, although its aggressive potential can be assessed in cell cultures (in vitro) and in models in vivo.5252 Huang, H.-T.; Lin, C.-C.; Kuo, T.-C.; Chen, S.-J.; Huang, R.-N.; Planta 2019, 250, 59. [Crossref]
Crossref...

Reducing the number of mosquito vectors is a consolidated strategy used to control disease transmission. The four main insecticide classes mostly used for mosquito vector control purposes comprise chlorinated hydrocarbons, organophosphates, carbamates and pyrethroids.4141 Engdahl, C.; Knutsson, S.; Fredriksson, S.-Å.; Linusson, A.; Bucht, G.; Ekström, F.; PLoS One 2015, 10, e0138598. [Crossref]
Crossref...
Environmental toxicity parameters’ prediction carried out in the US Environmental Protection Agency (EPA) software2929 U. S. Environmental Protection Agency (USEPA); ToxCastTM Predicting Hazard, Characterizing Toxicity Pathways, and Prioritizing the Toxicity Testing of Environmental Chemicals, https://www.epa.gov/comptox-tools/toxicity-forecasting-toxcast, accessed in June 2024.
https://www.epa.gov/comptox-tools/toxici...
has shown that sesquiterpene triquinane compounds may have lower environmental impact than that observed for temephos (organophosphate), as shown in Table 7.

Table 7
Environmental toxicity parameters observed for sesquiterpenes obtained from A. tomentosa essential oil, as well as for temephos

Finding selectivity for these compounds in mosquitoes is a challenge in the process to develop new bioinsecticides, since AChE of vertebrates has similarities to that of mosquito vectors (Yellow Fever and Aedes). However, their structural differences, mainly the ones associated with amino acid changes at the entry site, such as free cysteine residues at the entry site of the mosquito active receptor channel, which corresponds to Phe295 in humans, may be a potential target for selective covalent inhibitors, as well as target for the development of new biopesticides that are more selective and less aggressive to humans.4747 López, M. D.; Campoy, F. J.; Pascual-Villalobos, M. J.; Muñoz Delgado, E.; Vidal, C. J.; Chem. Biol. Interact. 2015, 229, 36. [Crossref]
Crossref...
Although the potential of covalent inhibitors has been investigated in several studies, inhibitors showing different action modes, such as non-covalent inhibitors, remain poorly explored.4949 Alout, H.; Labbé, P.; Berthomieu, A.; Djogbénou, L.; Leonetti, J.-P.; Fort, P.; Weill, M.; PLoS One 2012, 7, e47125. [Crossref]
Crossref...

Conclusions

The current study assessed the chemical composition of EOAT, as well as its potential to act as larvicide in tests conducted both in vitro and in silico. Based on GC-MS, GC-FID, and mono and bidimensional NMR analyses, sesquiterpene triquienanes, such as presylphiperfolan-1-ol, silphiperfol-5-ene and presylphiperfol-7-ene, were the major components of the investigated EO. Analyses performed with 3rd and 4th instar larvae of A. aegypti in vitro have evidenced 100% larval mortality 48 h after the beginning of the test. Assessments conducted in silico have shown high binding energies between essential oil components and the active site of the AchE enzyme. They were even higher than that observed for organophosphate larvicide temephos. Results in the current study have indicated that the investigated EOAT and its isolated constituents can be explored as potential environmentally and human safe larvicides. Further studies should be conducted with the investigated EOAT, and with its isolated constituents, to assess their effects on non-target organisms, as well as to define their effective concentrations for larvicidal activity. These results are important to help produce new safe, sustainable and biodegradable larvicides.

Supplementary Information

Supplementary information about ligand-receptor complexes of triquinane sesquiterpenes investigated in the current research is available free of charge at http://jbcs.sbq.org.br as PDF file.https://minio.scielo.br/documentstore/1678-4790/MZKrmmzLyYNWvBFsG6Lcq6Q/63d76a76778fdf92969c04e49f475717b25e5b3d.pdf

Acknowledgments

The authors are grateful for the financial support provided by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, 311419/2018-6), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brazil (CAPES, Finance Code 001) and Fundação de Amparo à Pesquisa do Estado da Bahia (FAPESB).

References

  • 1
    Alvarez, M. A.; Pharmacological Properties of Native Plants from Argentina, 1st ed.; Springer: Cham, 2019. [Crossref]
    » Crossref
  • 2
    Kammoun, A. K.; Altyar, A. E.; Gad, H. A.; J. Pharm. Biomed. Anal. 2021, 198, 113991. [Crossref]
    » Crossref
  • 3
    Lenardão, E. J.; Savegnago, L.; Jacob, R. G.; Victoria, F. N.; Martinez, D. M.; J. Braz. Chem. Soc. 2016, 27, 435. [Crossref]
    » Crossref
  • 4
    Pinto, S. C.; Leitão, G. G.; de Oliveira, D. R.; Bizzo, H. R.; Ramos, D. F.; Coelho, T. S.; Silva, P. E. A.; Lourenço, M. C. S.; Leitão, S. G.; Nat. Prod. Commun. 2009, 4, 1675. [Crossref]
    » Crossref
  • 5
    Muñoz-Benavent, M.; Pérez-Cobas, A. E.; García-Ferris, C.; Moya, A.; Latorre, A.; J. Pharm. Biomed. Anal. 2021, 194, 113787. [Crossref]
    » Crossref
  • 6
    Magalhães, L. A. M.; Lima, M. P.; Marques, M. O. M.; Facanali, R.; Pinto, A. C. S.; Tadei, W. P.; Molecules 2010, 15, 5734. [Crossref]
    » Crossref
  • 7
    Jankowska, M.; Rogalska, J.; Wyszkowska, J.; Stankiewicz, M.; Molecules 2017, 23, 34. [Crossref]
    » Crossref
  • 8
    Balachandran, C.; Anbalagan, S.; Kandeepan, C.; Arun Nagendran, N.; Jayakumar, M.; Fathi Abd_Allah, E.; Alqarawi, A. A.; Hashem, A.; Baskar, K.; J. Asia-Pac. Entomol. 2021, 24, 645. [Crossref]
    » Crossref
  • 9
    Pinto, S. C.; Leitão, G. G.; Castellar, A.; Delia, D. S.; Lage, C. L. S.; Henriques, A. B.; Fernandes, J.; Motta, G. S.; Bizzo, H. R.; Leitão, S. G.; J. Essent. Oil Res. 2013, 25, 198. [Crossref]
    » Crossref
  • 10
    Castilho, C. V. V.; Ferra Neto, J. F.; Leitão, S. G.; Barreto, C. S.; Pinto, S. C.; da Silva, N. C. B.; Plant Cell, Tissue Organ Cult. 2018, 133, 311. [Crossref]
    » Crossref
  • 11
    Juliani, H. R.; Zygadlo, J. A.; Scrivanti, R.; de la Sota, E.; Simon, J. E.; Flavour Fragrance J. 2004, 19, 541. [Crossref]
    » Crossref
  • 12
    Gillij, Y. G.; Gleiser, R. M.; Zygadlo, J. A.; Bioresour. Technol. 2008, 99, 2507. [Crossref] [PubMed]
    » Crossref» PubMed
  • 13
    Dias, C. N.; Moraes, D. F. C.; Parasitol. Res. 2014, 113, 565. [Crossref]
    » Crossref
  • 14
    de Morais, S. M.; Facundo, V. A.; Bertini, L. M.; Cavalcanti, E. S. B.; dos Anjos Jr., J. F.; Ferreira, S. A.; de Brito, E. S.; de Souza Neto, M. A.; Biochem. Syst. Ecol. 2007, 35, 670. [Crossred]
    » Crossred
  • 15
    Huy Hung, N.; Ngoc Dai, D.; Satyal, P.; Thi Huong, L.; Thi Chinh, B.; Quang Hung, D.; Anh Tai, T.; Setzer, W. N.; Chem. Biodiversity 2021, 18, e2100145 [Crossref] [PubMed]
    » Crossref» PubMed
  • 16
    Botelho, A. S.; Ferreira, O. O.; de Oliveira, M. S.; Cruz, J. N.; Chaves, S. H. R.; do Prado, A. F.; do Nascimento, L. D.; da Silva, G. A.; do Amarante, C. B.; Andrade, E. H. A.; Int. J. Mol. Sci. 2022, 23, 11172. [Crossref]
    » Crossref
  • 17
    Bienert, S.; Waterhouse, A.; de Beer, T. A. P.; Tauriello, G.; Studer, G.; Bordoli, L.; Schwede, T.; Nucleic. Acids Res. 2017, 45, D313. [Crossref]
    » Crossref
  • 18
    Gromacs, version 2018; Royal The Gromacs Development Team, Sweden, 2018. [Link] accessed in July 2024
    » Link
  • 19
    GROMOS, GROMOS96; Gromos Development Group, Switzerland, 1996. [Link] accessed in July 2024
    » Link
  • 20
    Benkert, P.; Biasini, M.; Schwede, T.; Bioinformatics 2011, 27, 343. [Crossref]
    » Crossref
  • 21
    Huey, R.; Morris, G. M.; Olson, A. J.; Goodsell, D. S.; J. Comput. Chem. 2007, 28, 1145. [Crossref]
    » Crossref
  • 22
    PyMOL, version 2.1; Schrödinger, LLC, 2018.
  • 23
    BIOVIA Discovery Studio, v. 4.5; Dassault Systèmes, San Diego, USA, 2021.
  • 24
    Marvin Sketch, version 23.4; ChemAxon, Hungary, 2023. [Link] accessed in July 2024
    » Link
  • 25
    Open Babel, version 3.1.1; The Open Babel Project, United States, 2023. [Link] accessed in July 2024
    » Link
  • 26
    AutoDock Vina, version 1.1.2; The Scripps Research Institute, United States, 2010. [Link] accessed in July 2024
    » Link
  • 27
    Pires, D. E. V.; Blundell, T. L.; Ascher, D. B.; J. Med. Chem. 2015, 58, 4066. [Crossref]
    » Crossref
  • 28
    Sander, T.; Freyss, J.; von Korff, M.; Rufener, C.; J. Chem. Inf. Model. 2015, 55, 460. [Crossref]
    » Crossref
  • 29
    U. S. Environmental Protection Agency (USEPA); ToxCastTM Predicting Hazard, Characterizing Toxicity Pathways, and Prioritizing the Toxicity Testing of Environmental Chemicals, https://www.epa.gov/comptox-tools/toxicity-forecasting-toxcast, accessed in June 2024.
    » https://www.epa.gov/comptox-tools/toxicity-forecasting-toxcast
  • 30
    Coates, R. M.; Ho, Z.; Klobus, M.; Wilson, S. R.; J. Am. Chem. Soc. 1996, 118, 9249. [Crossref]
    » Crossref
  • 31
    Hong, A. Y.; Stoltz, B. M.; Angew. Chem., Int. Ed. 2014, 53, 5248. [Crossref]
    » Crossref
  • 32
    Silverstein, R. M.; Webster, F. X.; Kiemle, D. J.; Bryce, D. L.; Spectrometric Identification of Organic Compounds, 7th ed.; John Wiley & Sons: New York, US, 2014.
  • 33
    Bohlmann, F.; Jakupovic, J.; Phytochemistry 1980, 19, 259. [Crossref]
    » Crossref
  • 34
    Kashman, Y.; Rudi, A.; Gutman-Naveh, N.; Tetrahedron 1978, 34, 1227. [Crossref]
    » Crossref
  • 35
    Weyerstahl, P.; Marschall, H.; Seelmann, I.; Jakupovic, J.; Eur. J. Org. Chem. 1998, 1998, 1205. [Crossref]
    » Crossref
  • 36
    Albuquerque, B. N. L.; da Silva, M. F. R.; da Silva, P. C. B.; de Lira Pimentel, C. S.; Lino da Rocha, S. K.; Farias de Aguiar, J. C. R. O.; Neto, A. C. A.; Paiva, P. M. G.; Gomes, M. G. M.; da Silva-Júnior, E. F.; Navarro, D. M. A. F.; Ind. Crops Prod 2022, 182, 114830. [Crossref]
    » Crossref
  • 37
    Andrade-Ochoa, S.; Correa-Basurto, J.; Rodríguez-Valdez, L. M.; Sánchez-Torres, L. E.; Nogueda-Torres, B.; Nevárez-Moorillón, G. V.; Chem. Cent. J. 2018, 12, 53. [Crossref]
    » Crossref
  • 38
    Fernandez, C. M. M.; Barba, E. L.; Fernandez, A. C. M.; Cardoso, B. K.; Borges, I. B.; Takemura, O. S.; Martins, L. A.; Cortez, L. E. R.; Cortez, D. A. G.; Gazim, Z. C.; J. Essent. Oil Bear. Plants 2014, 17, 813. [Crossref]
    » Crossref
  • 39
    Pavela, R.; Maggi, F.; Mbuntcha, H.; Woguem, V.; Fogang, H. P. D.; Womeni, H. M.; Tapondjou, L. A.; Barboni, L.; Nicoletti, M.; Canale, A.; Benelli, G.; Parasitol. Res. 2016, 115, 4617. [Crossref]
    » Crossref
  • 40
    Fraga, B. M.; Nat. Prod. Rep. 2011, 28, 1580. [Crossref]
    » Crossref
  • 41
    Engdahl, C.; Knutsson, S.; Fredriksson, S.-Å.; Linusson, A.; Bucht, G.; Ekström, F.; PLoS One 2015, 10, e0138598. [Crossref]
    » Crossref
  • 42
    Keniya, M. V.; Sabherwal, M.; Wilson, R. K.; Woods, M. A.; Sagatova, A. A.; Tyndall, J. D. A.; Monk, B. C.; Antimicrob. Agents Chemother. 2018, 62, e01134-18. [Crossref]
    » Crossref
  • 43
    Vargas, L. D. L.; Ferreira, S. M. B.; Souza, M. D.; da Silva, C. A. L.; Shimoya-Bittencourt, W.; Rev. Cien. Med. Biol. 2022, 21, 98. [Crossref]
    » Crossref
  • 44
    Zara, A. L. S. A.; Santos, S. M.; Fernandes-Oliveira, E. S.; Carvalho, R. G.; Coelho, G. E.; Epidemiol. Serv. Saúde 2016, 25, 1. [Crossref]
    » Crossref
  • 45
    Jukic, M.; Politeo, O.; Maksimovic, M.; Milos, M.; Milos, M.; Phytother. Res. 2007, 21, 259. [Crossref]
    » Crossref
  • 46
    Reegan, A. D.; Stalin, A.; Paulraj, M. G.; Balakrishna, K.; Ignacimuthu, S.; Al-Dhabi, N. A.; Med. Chem. Res. 2016, 25, 1411. [Crossref]
    » Crossref
  • 47
    López, M. D.; Campoy, F. J.; Pascual-Villalobos, M. J.; Muñoz Delgado, E.; Vidal, C. J.; Chem. Biol. Interact. 2015, 229, 36. [Crossref]
    » Crossref
  • 48
    Pezzementi, L.; Rowland, M.; Wolfe, M.; Tsigelny, I.; Invertebr. Neurosci. 2006, 6, 47. [Crossref]
    » Crossref
  • 49
    Alout, H.; Labbé, P.; Berthomieu, A.; Djogbénou, L.; Leonetti, J.-P.; Fort, P.; Weill, M.; PLoS One 2012, 7, e47125. [Crossref]
    » Crossref
  • 50
    Muangphrom, P.; Misaki, M.; Suzuki, M.; Shimomura, M.; Suzuki, H.; Seki, H.; Muranaka, T.; Phytochemistry 2019, 164, 144. [Crossref]
    » Crossref
  • 51
    Zárybnický, T.; Boušová, I.; Ambrož, M.; Skálová, L.; Arch. Toxicol. 2018, 92, 1. [Crossref]
    » Crossref
  • 52
    Huang, H.-T.; Lin, C.-C.; Kuo, T.-C.; Chen, S.-J.; Huang, R.-N.; Planta 2019, 250, 59. [Crossref]
    » Crossref
  • 53
    Luz, T. R. S. A.; de Mesquita, L. S. S.; do Amaral, F. M. M.; Coutinho, D. F.; Acta Trop. 2020, 212, 105705. [Crossref]
    » Crossref

Edited by

Editor handled this article: Paulo Cezar Vieira

Publication Dates

  • Publication in this collection
    16 Aug 2024
  • Date of issue
    2025

History

  • Received
    28 Feb 2024
  • Accepted
    17 July 2024
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br