Next Article in Journal
Designing a Phosphino-Thiosemicarbazone Ligand Capable of Stabilizing Coinage Metal Ions
Previous Article in Journal
Synthesis and Evaluation of Biological Activities of Schiff Base Derivatives of 4-Aminoantipyrine and Cinnamaldehydes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Synthesis, Insecticidal Activity and Computational Studies of Eugenol-Based Insecticides †

by
José Ricardo A. Coelho
1,
Tatiana F. Vieira
2,3,
Renato B. Pereira
4,
David M. Pereira
4,
Elisabete M. S. Castanheira
5,6,
António Gil Fortes
1,
Sérgio F. Sousa
2,3,
Maria José G. Fernandes
1 and
Maria Sameiro T. Gonçalves
1,*
1
Centre of Chemistry (CQ/UM), University of Minho, Campus of Gualtar, 4710-057 Braga, Portugal
2
Associate Laboratory i4HB—Institute for Health and Bioeconomy, Faculty of Medicine, University of Porto, 4200-319 Porto, Portugal
3
UCIBIO—Applied Molecular Biosciences Unit, BioSIM—Department of Biomedicine, Faculty of Medicine, University of Porto, Alameda Prof. Hernâni Monteiro, 4200-319 Porto, Portugal
4
REQUIMTE/LAQV, Laboratory of Pharmacognosy, Department of Chemistry, Faculty of Pharmacy, University of Porto, R. Jorge Viterbo Ferreira 228, 4050-313 Porto, Portugal
5
Physics Centre of Minho and Porto Universities (CF-UM-UP), University of Minho, Campus of Gualtar, 4710-057 Braga, Portugal
6
Associate Laboratory LaPMET—Laboratory of Physics for Materials and Emergent Technologies, University of Minho, Campus of Gualtar, 4710-057 Braga, Portugal
*
Author to whom correspondence should be addressed.
Presented at the 26th International Electronic Conference on Synthetic Organic Chemistry, 15–30 November 2022; Available online: https://sciforum.net/event/ecsoc-26.
Chem. Proc. 2022, 12(1), 46; https://doi.org/10.3390/ecsoc-26-13649
Published: 16 November 2022

Abstract

:
Eugenol, a natural phenolic allyl benzene that has been used as an active lead compound showing significant biological activities, including insecticidal effects on a wide variety of domestic arthropod pests, was used as the main reagent in the present work. Ester eugenol derivatives were synthesized and evaluated for their insecticidal activities against the Spodoptera frugiperda cell line. Studies of structure-based inverted virtual screening were carried out in order to identify the potential targets associated with the obtained insecticidal activity. The results indicate that the insecticide activity observed is most likely a result of the interaction of these molecules with the odorant-binding proteins and/or with acetylcholinesterase.

1. Introduction

The global population is increasing at an exponential rate, so it is necessary to ensure agricultural production that meets the actual food requirements. In crop protection, the reduction in damage caused by pathogens and pests in agricultural fields is mainly achieved through the extensive use of synthetic pesticides. To mitigate the environmental problems caused by the intensive use of conventional synthetic pesticides, biopesticides and semi-synthetic pesticides based on natural plant products are alternatives as pest-management agents [1,2].
Natural products are good substrates, due to their structural diversity and associated biological activity. Among several groups of natural insecticides (pyrethroids, neonicotinoids, avermectins, etc.), essential oils and their derivatives have also shown relevant potential as insecticides [3,4,5]. One possibility is eugenol, 4-allyl-2-methoxyphenol, a volatile phenolic bioactive compound that has been identified in several aromatic plants, among them Syzygium aromaticum, and has shown a wide range of biological activities including anti-inflammatory, antioxidant, analgesic, anticancer, antifungal, antimicrobial, antiparasitic and insecticidal activities [6,7,8,9,10,11].
Considering these facts, in the present work, eugenol derivatives were synthesized through an esterification reaction and evaluated for their effect on the viability of Sf9 cells. A structure-based inverted virtual screening protocol was employed to identify the potential proteins associated with the observed insecticidal activity.

2. Results and Discussion

2.1. Synthesis of Compounds 3ac

The reaction of eugenol, the trivial name for 4-allyl-2-methoxyphenol 1, with 4-chlorobenzoic acid 2a, 2-chlorobenzoic acid 2b and 3-amino-2-methylbenzoic acid 2c in dichloromethane, at room temperature and in the presence of 4-dimethylaminopyridine (DMAP) and N,N′-dicyclohexylcarbodiimide (DCC) gave the corresponding ester derivatives, namely 4-allyl-2-methoxyphenyl 4-chlorobenzoate 3a, 4-allyl-2-methoxyphenyl 2-chlorobenzoate 3b and 4-allyl-2-methoxyphenyl 3-amino-2-methylbenzoate 3c, respectively (Scheme 1). All compounds were obtained in 46% to 56% yields, and were characterized with the usual analytical techniques.
The 1H NMR of compounds 3ac show the signals of aromatic protons derived from the 4-allyl-2-methoxyphenyl group (δ 6.80–7.10 ppm) in addition to the protons of 4-chloro-, 2-chloro- or 3-amino-2-methylbenzoate, highlighting H-5 and H-6 displayed as doublets, double doublets or double triplets (δ 7.28–7.49 ppm, H-5; δ 7.86–8.16 ppm, H-6). The alkene protons are shown as multiplets (δ 5.10–5.17 and 5.95–6.05 ppm). The 13C NMR show the signals related to ester bonds (δ 164.08–165.68 ppm) in all compounds. The IR spectra also confirm the presence of the ester bonds through the stretching vibration bands of the carbonyl groups (υ 1718 to 1748 cm−1).

2.2. Biological Activity of Compounds 3ac

The impact of ester eugenol derivatives 3ac on the viability of Sf9 cells was evaluated at 100 μg/mL, following 24 h of exposure. As shown in Figure 1, compounds 3a and 3b, resulting from the eugenol esterification with chlorobenzoic acids, displayed similar toxicity to the starting material eugenol 1. On the other hand, the eugenol esterification with 3-amino-2-methylbenzoic acid led to increased activity (compound 3c), causing nearly 50% of viability loss.

2.3. Inverted Virtual Screening Results

In Table 1, the average score for all the eugenol derivatives is presented for each potential target using each scoring function. In each set of targets, the structure with the highest score was selected and ranked from best to worst, based on the docking programs/scoring functions’ predictions. For this study, the four scoring functions (SFs) of GOLD were used. The score of each of the GOLD SFs is dimensionless, with a higher value indicating a better binding affinity.
There is a high degree of consistency across all SFs, with odorant-binding proteins (OBPs) and acetylcholinesterases (AChEs) suggesting more likely binding. The octopamine receptor, voltage-gated sodium channel, polyphenol oxidate and N-acetylglucosamine-1-phosphate uridyltransferase (GlmU) consistently exhibit lower scores.

2.4. Molecular Dynamics Simulations and Free Energy Calculation Results

Simulations were carried out for both groups of targets predicted at the inverted VS stage odorant binding proteins and acetylcholinesterases, in complex with the three most potent eugenol derivatives. The structures chosen were the ones that presented the best score from each group (3K1E for OBP and 1QON for acetylcholinesterases—AChE). Molecular dynamics simulations were used to confirm and validate the inverted screening predictions, allowing for a more detailed analysis. Additionally, the interactions formed between the protein and ligand were further analyzed, and the most determinant residues were identified. The results are presented in Table 2.
The protein RMSD value of OBP compared to the docking pose had an average value of about 2.3 Å. Interestingly, the AChE complexes presented a higher RMSD value. However, in these cases, the standard deviation was low. It is possible that the complexes of AChE and the eugenol derivatives were optimized at the beginning of the simulation to obtain a more stable conformation. Throughout the simulation, all molecules remained bound to their targets and an induced-fit adjustment was observed. A study was also conducted to determine how buried the eugenol derivatives were in the binding pockets, measuring the solvent accessible surface area (SASA) and the percentage of potential SASA. An increase in the percentage of ligand SASA with a lower SASA indicates that the molecule is buried in the target pocket, making it less exposed to solvent. Compound 3b is the most buried in the OBP pocket (with a percentage of ligand SASA of 97% and an average SASA of 18.5 Å2). Regarding AChE, compound 3a is the most buried in the active site (with a percentage of ligand SASA of 94% and a SASA of 31.7 Å2).
The three eugenol derivatives exhibit slightly better binding affinities toward OBP, with compound 3a showing a ΔGbind of −35.6 kcal/mol when in complex with OBP and −33.5 kcal/mol when bound to AChE. When compared to the other compounds, compound 3c is the weakest binder for both OBP and AChE (with ΔGbind of −30.1 kcal/mol when bound to OBP and −29.1 kcal/mol when bound to AChE).
When bound to OBP, the compounds are stabilized primarily by electrostatic interactions with Trp114, Leu76, Gly92, Phe15, Leu80 and Tyr122. From all the compounds studied, the results seem to suggest that compounds 3a and 3b may be good candidates to be used as repellents, having OBP as their main target. Regarding AChE, the main interacting residues are Tyr71, Trp83 and Tyr374.

3. Material and Methods

3.1. Typical Procedure for the Preparation of Compounds 3ac (Illustrated for 3a)

A mixture of 4-allyl-2-methoxyphenol 1 (0.500 g, 3.05 mmol), DMAP (0.075 g, 0.61 mmol) and DCC (0.944 g, 4.56 mmol) was added to a solution of 4-chlorobenzoic acid 2a (0.716 g, 4.58 mmol) in dichloromethane (5 mL). The reaction mixture was stirred for 24 h at room temperature and was monitored by TLC (silica: dichloromethane). The white suspension obtained was filtered and the liquid phase was washed successively with 5% (w/v) HCl (2 × 5 mL), 5% NaHCO3 (w/v; 3 × 5 mL) and H2O (3 × 5 mL). After drying with anhydrous MgSO4, the organic phase was evaporated under reduced pressure. Compound 3a was obtained as a yellow solid (0.523 g; 56%). Rf = 0.61 (silica: dichloromethane), m.p. = 76–78 °C. IR (νmax): 3306, 2934, 1737, 1646, 1591, 1506, 1487, 1464, 1421, 1401, 1284, 1262, 1198, 1187, 1172, 1149, 1068, 922, 847, 751, 732 cm−1. 1H NMR (CDCl3, 400 MHz): δH 3.42 (2H, d, J = 6.8 Hz, CH2Ph), 3.81 (3H, s, OCH3), 5.10-5.17 (2H, m, CH=CH2), 5.95–6.05 (1H, m, CH=CH2), 6.82 (1H, d, J = 1.6 Hz, H-2), 6.84 (1H, dd, J = 6.0 and 1.6 Hz, H-5), 7.07 (1H, d, J = 8.0 Hz, H-6), 7.49 (2H, d, J = 8.8 Hz, H-3 and H-5 Ph-Cl), 8.16 (2H, d, J = 8.8 Hz, H-2 and H-6 Ph-Cl) ppm. 13C NMR (CDCl3, 100 MHz): δC 40.09 (CH2Ph), 55.84 (OCH3), 112.82 (C-3), 116.17 (CH=CH2), 120.72 (C-6), 122.52 (C-5), 127.95 (C-1 Ph-Cl), 128.82 (C-3 and C-5 Ph-Cl), 131.65 (C-2 and C-6 Ph-Cl), 137.01 (CH=CH2), 137.99 (C-4), 139.20 (C-4 Ph-Cl) 139.88 (C-1), 150.97 (C-2), 164.08 (C=O) ppm.

3.2. Biological Assays

The potential of compounds 3ac as biopesticides was evaluated in assays using the Sf9 (Spodoptera frugiperda) insect cell line. Cells were maintained at 28 °C and cultivated in Grace’s medium with 10% FBS. For the evaluation of viability, cells were plated at 3.0 × 104 cells/well and exposed to the molecules, after which resazurin was added, and the fluorescence was read at 560/590 nm after 60 min of incubation.

3.3. Docking and Inverted Virtual Screening Studies

The Scopus database was searched for papers reporting virtual screening studies involving targets and molecules with insecticidal activity. The publication year and relevance of the target were considered in the selection process. A total of thirteen targets were identified in the eighteen studies found, as shown in Table 3.
After preparing each PDB structure (removing water and other molecules from the crystallization process), the crystallographic ligands were extracted and saved in separate files, to provide a reference site for docking coordinates and to posteriorly perform re-docking. In the absence of crystallographic ligands, active site residues were selected based on their importance for activity. In order to evaluate the quality of the docking protocol, as well as the capability of the docking software to reproduce geometry and orientation of the crystallographic pose, re-docking was used. The docking program/scoring functions used was GOLD [29] (PLP, ASP, ChemScore and GoldScore). By comparing the predicted docking pose with the crystallographic one through RMSD calculation, the ability of the SF to correctly dock the ligand is evaluated and that is a valuable step in the validation process. A lower RMSD means a better docking prediction.
The three eugenol derivatives were prepared for the study using DataWarrior [30] and OpenBabel [31], and subsequently docked into each PDB structure, using all the four SFs as soon as the protocol was optimized. The parameters that were optimized for each program/scoring function were: the docking coordinates, the docking box dimension or radius, the exhaustiveness, the search efficiency and the number of runs. Lastly, a ranked list was prepared based on the average scores of each target.

3.4. Molecular Dynamics Simulations and Free Energy Calculations

The three eugenol derivatives (compounds 3a, 3b and 3c) in complex with the two most promising targets identified from the inverted virtual screening study (odorant-binding protein 1—3KIE and acetylcholinesterase—1QON) were further studied through 100 ns molecular dynamics simulations with the Amber18 software [32].
The pose predicted at the IVS stage (with the PLP scoring function) was selected as the starting point of each MD simulation. ANTECHAMBER with RESP HF/6-31G(d) charges (calculated with Gaussian16 [33] and the general Amber force field (GAFF) [34]) were used to assign parameters. To describe the protein targets, the ff14SB force field [35] was applied. Posteriorly, the protein–ligand complexes were placed in TIP3P water boxes with a minimum distance of 12 Å between the protein surface and the side of the box, and periodic boundary conditions were used. By adding counter-ions (Na+), the overall charge on the system was neutralized. Ewald’s particle-mesh summation method was used to calculate long-range electrostatic interactions, while for short-range electrostatic and Lennard-Jones interactions, a cut-off value of 10.0 Å was used. A time step of 2 fs was used and all bonds involving hydrogen atoms were constrained using the SHAKE algorithm.
Then, the systems were submitted to four consecutive minimization stages to remove clashes, followed by an equilibration and production run, with each minimization having a maximum of 2500 cycles. In the equilibration run, the procedure was divided into two phases; NVT ensemble, where the systems were gradually heated to 298 K using a Langevin thermostat at constant volume (50 ps), and equilibration of the system density at 298 K (subsequent 50 ps). Lastly, the production run was performed during 100 ns with an NPT ensemble at constant temperature (298 K, Langevin thermostat) and pressure (1 bar, Berendsen barostat). The last 70 ns of the simulation were considered for SASA and hydrogen bonding analysis.
In this study, the molecular mechanics/generalized Born surface area method [36] was used in conjunction with the MM/PBSA.py script [37] from Amber. Each simulation’s last 70 ns were analyzed with an interval of 100 ps and salt concentration of 0.100 mol dm−3. Additionally, the contribution of the amino acid residues was evaluated using the energy decomposition method. The MM-GBSA calculations considered 1400 conformations from each MD trajectory taken from the last 70 ns of simulation.

4. Conclusions

The eugenol derivatives obtained were fully characterized and their biological evaluation as insecticides using the Sf9 (Spodoptera frugiperda) insect cell line has shown that it was possible to obtain a molecule (compound 3c) that was more toxic by tuning its structure.
The results strongly suggest that the insecticide activity observed arises from their interaction with the odorant-binding proteins and/or with acetylcholinesterase.

Author Contributions

Conceptualization, M.J.G.F., D.M.P., M.S.T.G. and S.F.S.; methodology, J.R.A.C., M.S.T.G., S.F.S., E.M.S.C. and D.M.P.; formal analysis, J.R.A.C., M.S.T.G., M.J.G.F., R.B.P., D.M.P. and S.F.S.; investigation, J.R.A.C., T.F.V. and R.B.P.; validation, M.J.G.F., D.M.P., M.S.T.G., E.M.S.C., S.F.S. and M.S.T.G.; writing—original draft preparation, J.R.A.C., M.S.T.G., M.J.G.F., R.B.P. and S.F.S.; writing—review and editing, M.S.T.G., M.J.G.F., R.B.P., D.M.P., S.F.S. and E.M.S.C.; supervision, M.S.T.G., M.J.G.F. and A.G.F.; project administration, M.S.T.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by FCT under project PTDC/ASP-AGR/30154/2017 (PO-CI-01-0145-FEDER-030154) of COMPETE 2020, co-financed by FEDER and the EU. FCT-Portugal and FEDER-COMPETE/QREN-EU also gave financial support to the research centers CQ/UM (UIDB/00686/2021), CF-UM-UP (UIDB/04650/2021) and REQUIMTE (UIDB/50006/2021). The NMR spectrometer Bruker Avance III 400 (part of the National NMR Network) was financed by FCT and FEDER.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, X.; Cao, A.; Yan, D.; Ouyang, C.; Wang, Q.; Li, Y. Overview of mechanisms and uses of biopesticides. Int. J. Pest Manag. 2021, 67, 65–72. [Google Scholar] [CrossRef]
  2. Lengai, G.M.W.; Muthomi, J.W.; Mbega, E.R. Phytochemical activity and role of botanical pesticides in pest management for sustainable agricultural crop production. Sci. Afr. 2020, 7, e00239. [Google Scholar] [CrossRef]
  3. Salman, M.; Abbas, R.Z.; Israr, M.; Abbas, A.; Mehmood, K.; Khan, M.K.; Sindhu, Z.D.; Hussaind, R.; Saleemie, M.K.; Shaha, S. Repellent and acaricidal activity of essential oils and their components against Rhipicephalus ticks in cattle. Vet. Parasitol. 2020, 283, 109178. [Google Scholar] [CrossRef] [PubMed]
  4. Chaieb, K.; Hajlaoui, H.; Zmantar, T.; Ben Kahla-Nakbi, A.; Rouabhia, M.; Mahdouani, K.; Bakhrouf, A. The chemical composition and biological activity of clove essential oil, Eugenia Caryophyllata (Syzigium Aromaticum L. Myrtaceae): A short review. Phytother. Res. 2007, 21, 501–506. [Google Scholar] [CrossRef] [PubMed]
  5. Natal, C.M.; Fernandes, M.J.G.; Pinto, N.F.S.; Pereira, R.B.; Vieira, T.F.; Rodrigues, A.R.O.; Pereira, D.M.; Sousa, S.F.; Fortes, A.G.; Castanheira, E.M.S.; et al. New carvacrol and thymol derivatives as potential insecticides: Synthesis, biological activity, computational studies and nanoencapsulation. RSC Adv. 2021, 11, 34024–34035. [Google Scholar] [CrossRef]
  6. Ju, J.; Xie, Y.; Yu, H.; Guo, Y.; Cheng, Y.; Qian, H.; Yao, W. Analysis of the synergistic antifungal mechanism of eugenol and citral. LWT Food Sci. Technol. 2020, 123, 109128. [Google Scholar] [CrossRef]
  7. Lee, M.Y. Essential oils as repellents against arthropods. Biomed. Res. Int. 2018, 2018, 6860271. [Google Scholar] [CrossRef]
  8. Fernandes, M.J.G.; Pereira, R.B.; Pereira, D.M.; Fortes, A.G.; Castanheira, E.M.S.; Gonçalves, M.S.T. New eugenol derivatives with enhanced insecticidal activity. Int. J. Mol. Sci. 2020, 21, 9257. [Google Scholar] [CrossRef]
  9. Pereira, R.B.; Pinto, N.F.S.; Fernandes, M.J.G.; Vieira, T.F.; Rodrigues, A.R.O.; Pereira, D.M.; Sousa, S.F.; Castanheira, E.M.S.; Fortes, A.G.; Gonçalves, M.S.T. Amino alcohols from eugenol as potential semisynthetic insecticides: Chemical, biological and computational insights. Molecules 2021, 26, 6616. [Google Scholar] [CrossRef]
  10. Coelho, C.M.M.; Pereira, R.B.; Vieira, T.F.; Teixeira, C.M.; Fernandes, M.J.G.; Rodrigues, A.R.O.; Pereira, D.M.; Sousa, S.F.; Fortes, A.G.; Castanheira, E.M.S.; et al. Synthesis, computational and nanoencapsulation studies on eugenol-derived insecticides. New J. Chem. 2022, 46, 14375–14387. [Google Scholar] [CrossRef]
  11. Fernandes, M.J.G.; Pereira, R.B.; Rodrigues, A.R.O.; Vieira, T.F.; Fortes, A.G.; Pereira, D.M.; Sousa, S.F.; Gonçalves, M.S.T.; Castanheira, E.M.S. Liposomal formulations loaded with an eugenol derivative for application as insecticides: Encapsulation studies and in silico identification of protein targets. Nanomaterials 2022, 12, 3583. [Google Scholar] [CrossRef] [PubMed]
  12. Ramos, R.S.; Costa, J.S.; Silva, R.C.; Costa, G.V.; Rodrigues, A.B.L.; Rabelo, E.M.; Souto, R.N.P.; Taft, C.A.; Silva, C.H.T.P.; Rosa, J.M.C.; et al. Identification of potential inhibitors from Pyriproxyfen with insecticidal activity by virtual screening. Pharmaceuticals 2019, 12, 20. [Google Scholar] [CrossRef] [PubMed]
  13. Riva, C.; Suzanne, P.; Charpentier, G.; Dulin, F.; Halm-Lemeille, M.P.; Sopkova-de Oliveira Santos, J. In silico chemical library screening and experimental validation of novel compounds with potential varroacide activities. Pestic. Biochem. Physiol. 2019, 160, 11–19. [Google Scholar] [CrossRef] [PubMed]
  14. Correy, G.J.; Zaidman, D.; Harmelin, A.; Carvalho, S.; Mabbitt, P.D.; Viviane, C.; James, P.J.; Kotze, A.C.; Jackson, C.J.; London, N. Overcoming insecticide resistance through computational inhibitor design. Proc. Natl. Acad. Sci. USA 2019, 116, 21012–21021. [Google Scholar] [CrossRef]
  15. Liu, J.; Liu, M.; Yao, Y.; Wang, J.; Li, Y.; Li, G.; Wang, Y. Identification of novel potential β-N-acetyl-D-hexosaminidase inhibitors by virtual screening, molecular dynamics simulation and MM-PBSA calculations. Int. J. Mol. Sci. 2012, 13, 4545–4563. [Google Scholar] [CrossRef]
  16. Dong, L.; Shen, S.; Xu, Y.; Wang, L.; Yang, Q.; Zhang, J.; Lu, H. Identification of novel insect β-N-acetylhexosaminidase OfHex1 inhibitors based on virtual screening, biological evaluation, and molecular dynamics simulation. J. Biomol. Struct. Dyn. 2021, 39, 1735–1743. [Google Scholar] [CrossRef]
  17. Dong, Y.; Jiang, X.; Liu, T.; Ling, Y.; Yang, Q.; Zhang, L.; He, X. Structure-based virtual screening, compound synthesis, and bioassay for the design of chitinase inhibitors. J. Agric. Food Chem. 2018, 66, 3351–3357. [Google Scholar] [CrossRef]
  18. Hu, X.; Yin, B.; Cappelle, K.; Swevers, L.; Smagghe, G.; Yang, X.; Zhang, L. Identification of novel agonists and antagonists of the ecdysone receptor by virtual screening. J. Mol. Graph. Model. 2018, 81, 77–85. [Google Scholar] [CrossRef]
  19. Harada, T.; Nakagawa, Y.; Ogura, T.; Yamada, Y.; Ohe, T.; Miyagawa, H. Virtual screening for ligands of the insect molting hormone receptor. J. Chem. Inf. Model. 2011, 51, 296–305. [Google Scholar] [CrossRef]
  20. Min, J.; Lin, D.; Zhang, Q.; Zhang, J.; Yu, Z. Structure-based virtual screening of novel inhibitors of the uridyltransferase activity of Xanthomonas oryzae pv. oryzae GlmU. Eur. J. Med. Chem. 2012, 53, 150–158. [Google Scholar] [CrossRef]
  21. Offermann, L.R.; Chan, S.L.; Osinski, T.; Tan, Y.W.; Chew, F.T.; Sivaraman, J.; Mok, Y.K.; Minor, W.; Chruszcz, M. The major cockroach allergen Bla g 4 binds tyramine and octopamine. Mol. Immunol. 2014, 60, 86–94. [Google Scholar] [CrossRef] [PubMed]
  22. Laughlin, J.D.; Ha, T.S.; Jones, D.N.M.; Smith, D.P. Activation of pheromone-sensitive neurons is mediated by conformational activation of pheromone-binding protein. Cell 2008, 133, 1255–1265. [Google Scholar] [CrossRef] [PubMed]
  23. Oliferenko, P.V.; Oliferenko, A.A.; Poda, G.I.; Osolodkin, D.I.; Pillai, G.G.; Bernier, U.R.; Tsikolia, M.; Agramonte, N.M.; Clark, G.G.; Linthicum, K.J.; et al. Promising Aedes aegypti repellent chemotypes identified through integrated QSAR, virtual screening, synthesis, and bioassay. PLoS ONE 2013, 8, e64547. [Google Scholar] [CrossRef] [PubMed]
  24. Joshi, T.; Joshi, T.; Sharma, P.; Chandra, S.; Pande, V. Molecular docking and molecular dynamics simulation approach to screen natural compounds for inhibition of Xanthomonas oryzae pv. Oryzae by targeting peptide deformylase. J. Biomol. Struct. Dyn. 2021, 39, 823–840. [Google Scholar] [CrossRef]
  25. Fu, Y.; Liu, Y.X.; Kang, T.; Sun, Y.N.; Li, J.Z.; Ye, F. Identification of novel inhibitors of p-hydroxyphenylpyruvate dioxygenase using receptor-based virtual screening. J. Taiwan Inst. Chem. Eng. 2019, 103, 33–43. [Google Scholar] [CrossRef]
  26. Fattouch, S.; Raboudi-Fattouch, F.; Ponce, J.V.; Forment, J.V.; Lukovic, D.; Marzouki, N.; Vidal, D.R. Concentration dependent effects of commonly used pesticides on activation versus inhibition of the quince (Cydonia oblonga) polyphenol oxidase. Food Chem. Toxicol. 2010, 48, 957–963. [Google Scholar] [CrossRef]
  27. Cai, J.; Du, X.; Wang, C.; Lin, J.; Du, X. Identification of potential Helicoverpa armigera (Lepidoptera: Noctuidae) sterol carrier protein-2 inhibitors through high-throughput virtual screening. J. Econ. Entomol. 2017, 110, 1779–1784. [Google Scholar] [CrossRef]
  28. Shen, H.; Li, Z.; Jiang, Y.; Pan, X.; Wu, J.; Cristofori-Armstrong, B.; Smith, J.J.; Chin, Y.K.Y.; Lei, J.; Zhou, Q.; et al. Structural basis for the modulation of voltage-gated sodium channels by animal toxins. Science 2018, 362, eaau2596. [Google Scholar] [CrossRef]
  29. Jones, G.; Willett, P.; Glen, R.C.; Leach, A.R.; Taylor, R. Development and validation of a genetic algorithm for flexible docking 1, F.E. Cohen (Ed). J. Mol. Biol. 1997, 267, 727–748. [Google Scholar] [CrossRef]
  30. Sander, T.; Freyss, J.; von Korff, M.; Rufener, C. Data warrior: An open-source program for chemistry aware data visualization and analysis. J. Chem. Inf. Model. 2015, 55, 460–473. [Google Scholar] [CrossRef]
  31. O’Boyle, N.M.; Banck, M.; James, C.A.; Morley, C.; Vandermeersch, T.; Hutchison, G.R. Open Babel: An open chemical toolbox. J. Cheminform. 2011, 3, 33. [Google Scholar] [CrossRef]
  32. Case, D.A.; Cheatham, T.E., 3rd; Darden, T.; Gohlke, H.; Luo, R.; Merz, K.M., Jr.; Onufrie, A.; Simmerling, C.; Wang, B.; Woods, R.J. The Amber biomolecular simulation programs. J. Comput. Chem. 2005, 26, 1668–1688. [Google Scholar] [CrossRef] [PubMed]
  33. Frisch, M.J.; Trucks, G.; Schlegel, H.B.; Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; et al. Gaussian 09, Revision A.02; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  34. Wang, J.; Wolf, R.M.; Caldwell, J.W.; Kollman, P.A.; Case, D.A.; Wang, J. Development and testing of a general amber force field. J. Comp. Chem. 2004, 25, 1157–1174. [Google Scholar] [CrossRef] [PubMed]
  35. Maier, J.A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K.E.; Simmerling, C. ff14SB: Improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theory Comput. 2015, 11, 3696–3713. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, E.; Sun, H.; Wang, J.; Wang, Z.; Liu, H.; Zhang, J.Z.H.; Hou, T. End-point binding free energy calculation with MM/PBSA and MM/GBSA: Strategies and applications in drug design. Chem. Rev. 2019, 119, 9478–9508. [Google Scholar] [CrossRef] [PubMed]
  37. Miller, B.R.; McGee, T.D.; Swails, J.M.; Homeyer, N.; Gohlke, H.; Roitberg, A.E. MMPBSA.py: An efficient program for end-state free energy calculations. J. Chem. Theory Comput. 2012, 8, 3314–3321. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of esters derived from eugenol 3ac.
Scheme 1. Synthesis of esters derived from eugenol 3ac.
Chemproc 12 00046 sch001
Figure 1. Viability of Sf9 cells after incubation with the indicated molecules (100 µg/mL). *** p < 0.001.
Figure 1. Viability of Sf9 cells after incubation with the indicated molecules (100 µg/mL). *** p < 0.001.
Chemproc 12 00046 g001
Table 1. Average scores obtained with the five different scoring functions used and overall ranking.
Table 1. Average scores obtained with the five different scoring functions used and overall ranking.
TargetPDBPLPASPChemScoreGoldScoreOverall Ranking
Acetylcholinesterase1QON79.2052.5738.4563.132
4EY676.4846.5237.8244.88
1DX474.0046.6339.1660.42
Alpha-esterase-7 (αE7)5TYJ66.8038.3434.3057.036
5TYP64.8039.4034.3655.56
Beta-N-acetyl-D-hexosaminidase of Hex13NSN73.9249.5032.1265.903
3OZP68.1845.2431.8263.98
Chitinase3WL174.2744.4733.9759.114
3WQV74.0045.9234.0363.07
Ecdysone receptor1R2071.2733.4134.2457.755
1R1K67.5733.2136.1259.70
N-Acetylglucosamine-1-phosphate uridyltransferase (GlmU)2V0K55.3424.3822.7754.4713
2VD447.3326.0224.3647.33
Octopamine receptor4N7C60.6232.6933.2055.2910
Odorant-binding protein5V1382.1448.9539.8866.731
2GTE77.4246.5042.1968.77
3N7H76.7539.8131.7670.47
3K1E67.3839.8837.2661.61
Peptide deformylase5CY867.1130.0025.5864.727
p-hydroxyphenylpyruvate dioxygenase6ISD63.1437.7428.0955.428
Polyphenol oxidase1BUG52.8031.5022.1558.4212
Sterol carrier protein-2 (HaSCP-2)4UEI62.7732.8033.3952.069
Voltage-gated sodium channel6A9558.2623.5823.6059.5211
Table 2. Average RMSD values (Å), ligand RMSD (Å), average SASA (Å2), percentage of potential ligand SASA buried and average number of hydrogen bonds for the ligands for the last 70 ns of the simulation of the OBP and AChE–ligand complexes.
Table 2. Average RMSD values (Å), ligand RMSD (Å), average SASA (Å2), percentage of potential ligand SASA buried and average number of hydrogen bonds for the ligands for the last 70 ns of the simulation of the OBP and AChE–ligand complexes.
Average RMSD of the Complex (Å)Average RMSD of the Ligand (Å)Average SASA (Å2)Percentage of Potential Ligand SASA Buried (%)Average Number of H BondsΔGbind (kcal/mol)Main Contributors
OBP3a2.5 ± 0.20.6 ± 0.365.1 ± 13.7870.01 ± 0.07−35.6 ± 0.2Trp114 (−2.7 ± 0.5)
Leu76 (−1.9 ± 0.5)
Gly92 (−1.5 ± 0.5)
3b2.1 ± 0.31.3 ± 0.418.5 ± 10.6970.01 ± 0.1−32.1 ± 0.2Trp114 (−1.3 ± 0.4)
Phe15 (−1.2 ± 0.3)
Leu80 (−1.2 ± 0.3)
3c2.4 ± 0.31.5 ± 0.430.1 ± 17.0940.1 ± 0.2−30.1 ± 0.2Met19 (−1.5 ± 0.5)
Phe59 (−1.3 ± 0.5)
Tyr122 (−1.3 ± 0.4)
AChE3a3.2 ± 0.40.5 ± 0.231.7 ± 12.2940.4 ± 0.5−33.5 ± 0.1Trp83 (−2.4 ± 0.5)
Tyr71 (−1.4 ± 0.4)
Tyr374 (−1.2 ± 0.4)
3b4.0 ± 0.70.9 ± 0.274.2 ± 37.4850.1 ± 0.2−25.5 ± 0.2Tyr71 (−1.5 ± 0.7)
Tyr374 (−1.0 ± 0.4)
3c3.1 ± 0.30.9 ± 0.247.6 ± 22.1910.4 ± 0.6−29.1 ± 0.2 Trp83 (−1.9 ± 0.8)
Tyr71 (−1.5 ± 0.5)
Tyr374 (−1.4 ± 0.9)
Table 3. List of targets selected for the inverted virtual screening study.
Table 3. List of targets selected for the inverted virtual screening study.
TargetOrganismPDB TargetResolution (Å)Ref.
AcetylcholinesteraseAedes aegypti1QON2.72[12]
4EY62.40
Drosophila melanogaster1DX42.70[13]
Alpha-esterase-7 (αE7)Lucilia cuprina5TYJ1.75[14]
5TYP1.88
beta-N-acetyl-D-hexosaminidase OfHex1Ostrinia furnacalis3NSN2.10[15]
3OZP2.00[16]
ChitinaseOstrinia furnacalis3WL11.77[17]
3WQV2.04
Ecdysone receptorHeliothis virescens1R203[18]
1R1K2.9[19]
N-Acetylglucosamine-1-phosphate uridyltransferase (GlmU)Xanthomonas oryzae2V0K2.3[20]
2VD41.9
Octopamine receptorBlattella germanica4N7C1.75[21]
Odorant-binding proteinAedes aegypti5V131.84[12]
Drosophila melanogaster2GTE1.4[22]
Anopheles gambiae3N7H1.6[23]
Aedes aegypti3K1E1.85
Peptide deformylaseXanthomonas oryzae5CY82.38[24]
p-hydroxyphenylpyruvate dioxygenaseArabidopsis thaliana6ISD2.4[25]
Polyphenol oxidaseManduca sexta3HSS2.7[26]
Sterol carrier protein-2 (HaSCP-2)Helicoverpa armigera4UEISolution NMR[27]
Voltage-gated sodium channelPeriplaneta americana6A952.6[28]
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

Coelho, J.R.A.; Vieira, T.F.; Pereira, R.B.; Pereira, D.M.; Castanheira, E.M.S.; Fortes, A.G.; Sousa, S.F.; Fernandes, M.J.G.; Gonçalves, M.S.T. Synthesis, Insecticidal Activity and Computational Studies of Eugenol-Based Insecticides. Chem. Proc. 2022, 12, 46. https://doi.org/10.3390/ecsoc-26-13649

AMA Style

Coelho JRA, Vieira TF, Pereira RB, Pereira DM, Castanheira EMS, Fortes AG, Sousa SF, Fernandes MJG, Gonçalves MST. Synthesis, Insecticidal Activity and Computational Studies of Eugenol-Based Insecticides. Chemistry Proceedings. 2022; 12(1):46. https://doi.org/10.3390/ecsoc-26-13649

Chicago/Turabian Style

Coelho, José Ricardo A., Tatiana F. Vieira, Renato B. Pereira, David M. Pereira, Elisabete M. S. Castanheira, António Gil Fortes, Sérgio F. Sousa, Maria José G. Fernandes, and Maria Sameiro T. Gonçalves. 2022. "Synthesis, Insecticidal Activity and Computational Studies of Eugenol-Based Insecticides" Chemistry Proceedings 12, no. 1: 46. https://doi.org/10.3390/ecsoc-26-13649

Article Metrics

Back to TopTop