Docking of New Designed Compounds Derived from 1,6-Dihydro-1,3,5-triazine-2,4-diamine Toward Quadruple Mutant Plasmodium Dihydrofolate Reductase

https://doi.org/10.22146/ijc.39943

Imam Siswanto(1*), Harno Dwi Pranowo(2), Mudasir Mudasir(3)

(1) Department of Chemistry, Faculty of Science and Technology, Universitas Airlangga, Kampus C Mulyorejo, Surabaya 60115, Indonesia
(2) Department of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Gadjah Mada, Sekip Utara, Yogyakarta 55281, Indonesia
(3) Department of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Gadjah Mada, Sekip Utara, Yogyakarta 55281, Indonesia
(*) Corresponding Author

Abstract


Resistance to the traditional antifolates is now widespread in Plasmodium falciparum and Plasmodium vivax. To find the interaction model of some compounds derived from 1,6-dihydro-1,3,5-triazine-2,4-diamine, molecular docking technique was carried out using these compounds ligand and pDHFR as the receptor. Complex ligand and the receptor from Protein Data Bank (PDB ID 1J3K) were chosen as an interaction model between a ligand and its receptor. Each compound was optimized using ab initio methods with 6-311G basis set. Partial charges of ligand were added using AM1-BCC methods. Each ligand was docked to the receptor utilizing Dock6 software. Theoretical prediction based on the binding energy between these compounds as the ligand with pDHFR as receptor resulted in 1 compound, namely 6,6-dimethyl-1-[3-(2-chloro-4,5-dibromophenoxy)propoxy]-1,6-dihydro-1,3,5-triazine-2,4-diamine possessing binding affinity better than that of WR99210 which was known to have good antimalarial activity.


Keywords


Amber score; pDHFR; molecular docking; triazine

Full Text:

Full Text PDF


References

[1] World Health Organization, 2017, World Malaria Report 2017, https://apps.who.int/iris/bitstream/10665/259492/1/9789241565523-eng.pdf.

[2] Fidock, D.A., Rosenthal, P.J., Croft, S.L., Brun, R., and Nwaka, S., 2004, Antimalarial drug discovery: Efficacy models for compound screening, Nat. Rev. Drug Discovery, 3 (6), 509–520.

[3] O’Neill, P.M., Mukhtar, A., Stocks, P.A., Randle, L.E., Hindley, S., Ward, S.A., Storr, R.C., Bickley, J.F., O’Neil, I.A., Maggs, J.L., Hughes, R.H., Winstanley, P.A., Bray, P.G., and Park, B.K., 2003, Isoquine and related amodiaquine analogues: A new generation of improved 4–aminoquinoline antimalarials, J. Med. Chem., 46 (23), 4933–4945.

[4] Valverde, E.A., Romero, A.H., Acosta, M.E., Gamboa, N., Henriques, G., Rodrigues, J.R., Ciangherotti, C., and López, S.E., 2018, Synthesis, β–hematin inhibition studies and antimalarial evaluation of new dehydroxy isoquine derivatives against Plasmodium berghei: A promising antimalarial agent, Eur. J. Med. Chem., 148, 498–506.

[5] Gatton, M.L., Martin, L.B., and Cheng, Q., 2004, Evolution of resistance to sulfadoxine–pyrimethamine in Plasmodium falciparum, Antimicrob. Agents Chemother., 48 (6), 2116–2123.

[6] Mishra, M., Mishra, V.K., Kashaw, V., Iyer, A.K., and Kashaw, S.K., 2017, Comprehensive review on various strategies for antimalarial drug discovery, Eur. J. Med. Chem., 125, 1300–1320.

[7] Sainy, J., and Sharma, R., 2017, Synthesis, antimalarial evaluation and molecular docking studies of some thiolactone derivatives, J. Mol. Struct., 1134, 350–359.

[8] Maurya, S.S., Khan, S.I., Bahuguna, A., Kumar, D., and Rawat, D.S., 2017, Antimalarial activity, heme binding and docking studies of N –substituted 4–aminoquinoline–pyrimidine molecular hybrids, Eur. J. Med. Chem., 129, 175–185.

[9] Bhat, H.R., Singh, U.P., Thakur, A., Ghosh, S.K., Gogoi, K., Prakash, A., and Singh, R.K., 2015, Synthesis, antimalarial activity and molecular docking of hybrid 4-aminoquinoline-1,3,5-triazine derivatives, Exp. Parasitol., 157, 59–67.

[10] Brandão, G.C., Missias, F.C.R., Arantes, L.M., Soares, L.F., Roy, K.K., Doerksen, R.J., de Oliveira, A.B., and Pereira, G.R., 2018, Antimalarial naphthoquinones. Synthesis via click chemistry, in vitro activity, docking to PfDHODH and SAR of lapachol–based compounds, Eur. J. Med. Chem., 145, 191–205.

[11] Debbabi, K.F., Bashandy, M.S., Al-Harbi, S.A., Aljuhani, E.H., and Al-Saidi, H.M., 2017, Synthesis and molecular docking against dihydrofolate reductase of novel pyridin–N–ethyl–N–methylbenzenesulfonamides as efficient anticancer and antimicrobial agents, J. Mol. Struct., 1131, 124–135.

[12] Ugwu, D.I., Okoro, U.C., Ukoha, P.O., Okafor, S., Ibezim, A., and Kumar, N.M., 2017, Synthesis, characterization, molecular docking and in vitro antimalarial properties of new carboxamides bearing sulphonamide, Eur. J. Med. Chem., 135, 349–369.

[13] Gálvez, J., Polo, S., Insuasty, B., Gutiérrez, M., Cáceres, D., Alzate-Morales, J.H., De–la–Torre, P., and Quiroga, J., 2018, Design, facile synthesis, and evaluation of novel spiro- and pyrazolo[1,5-c]quinazolines as cholinesterase inhibitors: Molecular docking and MM/GBSA studies, Comput. Biol. Chem., 74, 218–229.

[14] Karaman, B., and Sippl, W., 2015, Docking and binding free energy calculations of sirtuin inhibitors, Eur. J. Med. Chem., 93, 584–598.

[15] Fogel, G.B., Cheung, M., Pittman, E., and Hecht, D., 2008, In silico screening against wild–type and mutant Plasmodium falciparum dihydrofolate reductase, J. Mol. Graphics Modell., 26 (7), 1145–1152.

[16] Al–Omary, F.A.M., Abou-zeid, L.A., Nagi, M.N., Habib, E.S.E., Abdel-Aziz, A.A.M., El-Azab, A.S., Abdel-Hamide, S.G., Al-Omar, M.A., Al-Obaid, A.M., and El-Subbagh, H.I., 2010, Non–classical antifolates. Part 2: Synthesis, biological evaluation, and molecular modeling study of some new 2,6-substituted-quinazolin-4-ones, Bioorg. Med. Chem., 18 (8), 2849–2863.

[17] Algul, O., Paulsen, J.L., and Anderson, A.C., 2011, 2,4-Diamino-5-(2′-arylpropargyl)pyrimidine derivatives as new nonclassical antifolates for human dihydrofolate reductase inhibition, J. Mol. Graphics Modell., 29 (5), 608–613.

[18] Chopra, R., Chibale, K., and Singh, K., 2018, Pyrimidine-chloroquinoline hybrids: Synthesis and antiplasmodial activity, Eur. J. Med. Chem., 148, 39–53.

[19] Maurya, S.S., Bahuguna, A., Khan, S.I., Kumar, D., Kholiya, R., and Rawat, D.S., 2019, N-Substituted aminoquinoline-pyrimidine hybrids: Synthesis, in vitro antimalarial activity evaluation and docking studies, Eur. J. Med. Chem., 162, 277–289.

[20] Ma, X., and Chui, W.K., 2010, Antifolate and antiproliferative activity of 6,8,10-triazaspiro[4.5]deca-6,8–dienes and 1,3,5-triazaspiro[5.5]undeca-1,3-dienes, Bioorg. Med. Chem., 18 (2), 737–743.

[21] Adane, L., Bhagat, S., Arfeen, M., Bhatia, S., Sirawaraporn, R., Sirawaraporn, W., Chakraborti, A.K., and Bharatam, P.V, 2014, Design and synthesis of guanylthiourea derivatives as potential inhibitors of Plasmodium falciparum dihydrofolate reductase enzyme, Bioorg. Med. Chem. Lett., 24 (2), 613–617.

[22] Ma, X., Xiang, G., Yap, C.W., and Chui, W.K., 2012, 3D–QSAR Study on dihydro-1,3,5–triazines and their spiro derivatives as DHFR inhibitors by comparative molecular field analysis (CoMFA), Bioorg. Med. Chem. Lett., 22 (9), 3194–3197.

[23] Yuvaniyama, J., Chitnumsub, P., Kamchonwongpaisan, S., Vanichtanankul, J., Sirawaraporn, W., Taylor, P., Walkinshaw, M.D., and Yuthavong, Y., 2003, Insights into antifolate resistance from malarial DHFR–TS structures, Nat. Struct. Biol., 10 (5), 357–365.

[24] Wang, J., Wang, W., Kollman, P.A., and Case, D.A., 2006, Automatic atom type and bond type perception in molecular mechanical calculations, J. Mol. Graphics Modell., 25 (2), 247–260.

[25] Siswanto, I., Armunanto, R., Pranowo, H.D., and Mudasir, M., 2019, Contribution of the atomic charge of the triazine ring members to the DHFR inhibitory activity using 2D-QSAR, BMC Proc., in press.

[26] Fogel, G.B., Cheung, M., Pittman, E., and Hecht, D., 2008, Modeling the inhibition of quadruple mutant Plasmodium falciparum dihydrofolate reductase by pyrimethamine derivatives, J. Comput.-Aided Mol. Des., 22 (1), 29–38.

[27] Pettersen, E.F., Goddard, T.D., Huang, C.C., Couch, G.S., Greenblatt, D.M., Meng, E.C., and Ferrin, T.E., 2004, UCSF Chimera–A visualization system for exploratory research and analysis, J. Comput. Chem., 25 (13), 1605–1612.

[28] Allen, W.J., Balius, T.E., Mukherjee, S., Brozell, S.R., Moustakas, D.T., Lang, P.T., Case, D.A., Kuntz, I.D., and Rizzo, R.C., 2015, DOCK 6: Impact of new features and current docking performance, J. Comput. Chem., 36 (15), 1132–1156.

[29] Allen, W.J., and Rizzo, R.C., 2014, Implementation of the Hungarian algorithm to account for ligand symmetry and similarity in structure-based design, J. Chem. Inf. Model., 54 (2), 518–529.

[30] Kuntz, I.D., Blaney, J.M., Oatley, S.J., Langridge, R., and Ferrin, T.E., 1982, A geometric approach to macromolecule-ligand interactions, J. Mol. Biol., 161 (2), 269–288.

[31] Wang, R., Lu, Y., and Wang, S., 2003, Comparative evaluation of 11 scoring functions for molecular docking, J. Med. Chem., 46 (12), 2287–2303.

[32] Chen, Y.C., 2015, Beware of docking!, Trends Pharmacol. Sci., 36 (2), 78–95.

[33] Genheden, S., and Ryde, U., 2015, The MM/PBSA and MM/GBSA methods to estimate ligand-binding affinities, Expert Opin. Drug Discovery, 10 (5), 449–461.

[34] Sun, H., Li, Y., Shen, M., Tian, S., Xu, L., Pan, P., Guan, Y., and Hou, T., 2014, Assessing the performance of MM/PBSA and MM/GBSA methods. 5. Improved docking performance using high solute dielectric constant MM/GBSA and MM/PBSA rescoring, Phys. Chem. Chem. Phys., 16 (40), 22035–22045.

[35] Rastelli, G., Del Rio, A., Degliesposti, G., and Sqobba, M., 2010, Fast and accurate predictions of binding free energies using MM–PBSA and MM–GBSA, J. Comput. Chem., 31 (4), 797–810.



DOI: https://doi.org/10.22146/ijc.39943

Article Metrics

Abstract views : 3634 | views : 3803


Copyright (c) 2019 Indonesian Journal of Chemistry

Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.

 


Indonesian Journal of Chemistry (ISSN 1411-9420 /e-ISSN 2460-1578) - Chemistry Department, Universitas Gadjah Mada, Indonesia.

Web
Analytics View The Statistics of Indones. J. Chem.