Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Review Article
  • Published: 31 August 2023

Antimalarial drug discovery: progress and approaches

  • Jair L. Siqueira-Neto   ORCID: orcid.org/0000-0001-9574-8174 1 ,
  • Kathryn J. Wicht 2 , 3 ,
  • Kelly Chibale 2 , 3 ,
  • Jeremy N. Burrows   ORCID: orcid.org/0000-0001-8448-6068 4 ,
  • David A. Fidock   ORCID: orcid.org/0000-0001-6753-8938 5 &
  • Elizabeth A. Winzeler   ORCID: orcid.org/0000-0002-4049-2113 1  

Nature Reviews Drug Discovery volume  22 ,  pages 807–826 ( 2023 ) Cite this article

7726 Accesses

26 Citations

42 Altmetric

Metrics details

  • Drug development
  • Parasitic infection
  • Parasitology

Recent antimalarial drug discovery has been a race to produce new medicines that overcome emerging drug resistance, whilst considering safety and improving dosing convenience. Discovery efforts have yielded a variety of new molecules, many with novel modes of action, and the most advanced are in late-stage clinical development. These discoveries have led to a deeper understanding of how antimalarial drugs act, the identification of a new generation of drug targets, and multiple structure-based chemistry initiatives. The limited pool of funding means it is vital to prioritize new drug candidates. They should exhibit high potency, a low propensity for resistance, a pharmacokinetic profile that favours infrequent dosing, low cost, preclinical results that demonstrate safety and tolerability in women and infants, and preferably the ability to block Plasmodium transmission to Anopheles mosquito vectors. In this Review, we describe the approaches that have been successful, progress in preclinical and clinical development, and existing challenges. We illustrate how antimalarial drug discovery can serve as a model for drug discovery in diseases of poverty.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 12 print issues and online access

195,33 € per year

only 16,28 € per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

thesis on antimalarial drugs

Similar content being viewed by others

thesis on antimalarial drugs

The antimalarial efficacy and mechanism of resistance of the novel chemotype DDD01034957

thesis on antimalarial drugs

Anti-tuberculosis treatment strategies and drug development: challenges and priorities

thesis on antimalarial drugs

Preclinical characterization and target validation of the antimalarial pantothenamide MMV693183

Carter, R. & Mendis, K. N. Evolutionary and historical aspects of the burden of malaria. Clin. Microbiol. Rev. 15 , 564–594 (2002).

Article   PubMed   PubMed Central   Google Scholar  

Ashley, E. A., Pyae Phyo, A. & Woodrow, C. J. Malaria. Lancet 391 , 1608–1621 (2018).

Article   PubMed   Google Scholar  

Lal, A. A., Rajvanshi, H., Jayswar, H., Das, A. & Bharti, P. K. Malaria elimination: using past and present experience to make malaria-free India by 2030. J. Vector Borne Dis. 56 , 60–65 (2019).

World Health Organization. World Malaria Report https://www.who.int/teams/global-malaria-programme/reports/world-malaria-report-2022 (2022).

Chandramohan, D. et al. Seasonal malaria vaccination with or without seasonal malaria chemoprevention. N. Engl. J. Med. 385 , 1005–1017 (2021).

Article   CAS   PubMed   Google Scholar  

Sinnis, P. & Fidock, D. A. The RTS,S vaccine-a chance to regain the upper hand against malaria? Cell 185 , 750–754 (2022).

Datoo, M. S. et al. Efficacy and immunogenicity of R21/Matrix-M vaccine against clinical malaria after 2 years’ follow-up in children in Burkina Faso: a phase 1/2b randomised controlled trial. Lancet Infect. Dis. 22 , 1728–1736 (2022).

Greenwood, B. et al. Combining malaria vaccination with chemoprevention: a promising new approach to malaria control. Malar. J. 20 , 361 (2021).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Phillips, M. A. et al. Malaria. Nat. Rev. Dis. Primers 3 , 17050 (2017).

Wicht, K. J., Mok, S. & Fidock, D. A. Molecular mechanisms of drug resistance in Plasmodium falciparum malaria. Annu. Rev. Microbiol. 74 , 431–454 (2020).

Rasmussen, C., Alonso, P. & Ringwald, P. Current and emerging strategies to combat antimalarial resistance. Expert Rev. Anti Infect. Ther. 20 , 353–372 (2022).

Achan, J. et al. Quinine, an old anti-malarial drug in a modern world: role in the treatment of malaria. Malar. J. 10 , 144 (2011).

White, N. J. Qinghaosu (artemisinin): the price of success. Science 320 , 330–334 (2008).

Plowe, C. V. Malaria chemoprevention and drug resistance: a review of the literature and policy implications. Malar. J. 21 , 104 (2022).

van der Pluijm, R. W. et al. Determinants of dihydroartemisinin-piperaquine treatment failure in Plasmodium falciparum malaria in Cambodia, Thailand, and Vietnam: a prospective clinical, pharmacological, and genetic study. Lancet Infect. Dis. 19 , 952–961 (2019).

Dhorda, M., Amaratunga, C. & Dondorp, A. M. Artemisinin and multidrug-resistant Plasmodium falciparum — a threat for malaria control and elimination. Curr. Opin. Infect. Dis. 34 , 432–439 (2021).

Uwimana, A. et al. Emergence and clonal expansion of in vitro artemisinin-resistant Plasmodium falciparum kelch13 R561H mutant parasites in Rwanda. Nat. Med. 26 , 1602–1608 (2020). Reported the identification of artemisinin partial resistance emerging in African parasites.

Uwimana, A. et al. Association of Plasmodium falciparum kelch13 R561H genotypes with delayed parasite clearance in Rwanda: an open-label, single-arm, multicentre, therapeutic efficacy study. Lancet Infect. Dis. 21 , 1120–1128 (2021).

Balikagala, B. et al. Evidence of artemisinin-resistant malaria in Africa. N. Engl. J. Med. 385 , 1163–1171 (2021). Reports clinical evidence that mutant Kelch13 alleles emerging independently in Uganda were associated with prolonged parasite clearance half-lives in patients with P. falciparum infection treated with artesunate.

Straimer, J., Gandhi, P., Renner, K. C. & Schmitt, E. K. High prevalence of Plasmodium falciparum K13 mutations in Rwanda is associated with slow parasite clearance after treatment with artemether-lumefantrine. J. Infect. Dis. 225 , 1411–1414 (2021).

Article   PubMed Central   Google Scholar  

Tumwebaze, P. K. et al. Decreased susceptibility of Plasmodium falciparum to both dihydroartemisinin and lumefantrine in northern Uganda. Nat. Commun. 13 , 6353 (2022).

Ariey, F. et al. A molecular marker of artemisinin-resistant Plasmodium falciparum malaria. Nature 505 , 50–55 (2014).

Ashley, E. A. et al. Spread of artemisinin resistance in Plasmodium falciparum malaria. N. Engl. J. Med. 371 , 411–423 (2014).

Siddiqui, F. A., Liang, X. & Cui, L. Plasmodium falciparum resistance to ACTs: emergence, mechanisms, and outlook. Int. J. Parasitol. Drugs Drug Resist. 16 , 102–118 (2021).

Straimer, J. et al. Drug resistance. K13-propeller mutations confer artemisinin resistance in Plasmodium falciparum clinical isolates. Science 347 , 428–431 (2015).

Siddiqui, F. A. et al. Plasmodium falciparum falcipain-2a polymorphisms in Southeast Asia and their association with artemisinin resistance. J. Infect. Dis. 218 , 434–442 (2018).

Yang, T. et al. Decreased K13 abundance reduces hemoglobin catabolism and proteotoxic stress, underpinning artemisinin resistance. Cell Rep. 29 , 2917–2928 (2019).

Birnbaum, J. et al. A Kelch13-defined endocytosis pathway mediates artemisinin resistance in malaria parasites. Science 367 , 51–59 (2020). This study identified the role of proteins (including K13) in ART resistance and their functional association with reduced haemoglobin endocytosis, a mechanism important for ART activation.

Posner, G. H. et al. Mechanism-based design, synthesis, and in vitro antimalarial testing of new 4-methylated trioxanes structurally related to artemisinin: the importance of a carbon-centered radical for antimalarial activity. J. Med. Chem. 37 , 1256–1258 (1994).

Mok, S. et al. Population transcriptomics of human malaria parasites reveals the mechanism of artemisinin resistance. Science 347 , 431–435 (2015).

Hott, A. et al. Artemisinin-resistant Plasmodium falciparum parasites exhibit altered patterns of development in infected erythrocytes. Antimicrob. Agents Chemother. 59 , 3156–3167 (2015).

Xie, S. C., Ralph, S. A. & Tilley, L. K13, the cytostome, and artemisinin resistance. Trends Parasitol. 36 , 533–544 (2020).

Reyser, T. et al. Identification of compounds active against quiescent artemisinin-resistant Plasmodium falciparum parasites via the quiescent-stage survival assay (QSA). J. Antimicrob. Chemother. 75 , 2826–2834 (2020).

Connelly, S. V. et al. Restructured mitochondrial-nuclear interaction in Plasmodium falciparum dormancy and persister survival after artemisinin exposure. mBio 12 , e0075321 (2021).

Mok, S. et al. Artemisinin-resistant K13 mutations rewire Plasmodium falciparum ’s intra-erythrocytic metabolic program to enhance survival. Nat. Commun. 12 , 530 (2021).

Imwong, M. et al. Molecular epidemiology of resistance to antimalarial drugs in the Greater Mekong subregion: an observational study. Lancet Infect. Dis. 20 , 1470–1480 (2020).

Stokes, B. H. et al. Plasmodium falciparum K13 mutations in Africa and Asia impact artemisinin resistance and parasite fitness. eLife 10 , e66277 (2021).

Stokes, B. H., Ward, K. E. & Fidock, D. A. Evidence of artemisinin-resistant malaria in Africa. N. Engl. J. Med. 386 , 1385–1386 (2022). Laboratory P. falciparum Dd2 parasites edited with kelch13 mutations A675V or C469Y (observed in clinical isolates from Africa) alone showed only marginally reduced susceptibility to dihydroartemisinin, suggesting that high-level resistance to treatment with an artemisinin derivative must include additional mutations.

Demas, A. R. et al. Mutations in Plasmodium falciparum actin-binding protein coronin confer reduced artemisinin susceptibility. Proc. Natl Acad. Sci. USA 115 , 12799–12804 (2018).

Sharma, A. I. et al. Genetic background and PfKelch13 affect artemisinin susceptibility of PfCoronin mutants in Plasmodium falciparum . PLoS Genet. 16 , e1009266 (2020).

Henrici, R. C., van Schalkwyk, D. A. & Sutherland, C. J. Modification of pfap2mu and pfubp1 markedly reduces ring-stage susceptibility of Plasmodium falciparum to artemisinin in vitro. Antimicrob. Agents Chemother. https://doi.org/10.1128/AAC.01542-19 (2019).

Amambua-Ngwa, A. et al. Major subpopulations of Plasmodium falciparum in sub-Saharan Africa. Science 365 , 813–816 (2019). Important analysis of genetic population of P. falciparum clinical isolate samples in African continent revealing shared genomic haplotypes associated with antimalarial resistance.

Masserey, T. et al. The influence of biological, epidemiological, and treatment factors on the establishment and spread of drug-resistant Plasmodium falciparum . eLife https://doi.org/10.7554/eLife.77634 (2022).

Imwong, M. et al. Evolution of multidrug resistance in Plasmodium falciparum : a longitudinal study of genetic resistance markers in the Greater Mekong subregion. Antimicrob. Agents Chemother. 65 , e0112121 (2021).

Amato, R. et al. Genetic markers associated with dihydroartemisinin-piperaquine failure in Plasmodium falciparum malaria in Cambodia: a genotype-phenotype association study. Lancet Infect. Dis. 17 , 164–173 (2017).

Witkowski, B. et al. A surrogate marker of piperaquine-resistant Plasmodium falciparum malaria: a phenotype-genotype association study. Lancet Infect. Dis. 17 , 174–183 (2017).

Chugh, M. et al. Protein complex directs hemoglobin-to-hemozoin formation in Plasmodium falciparum . Proc. Natl Acad. Sci. USA 110 , 5392–5397 (2013).

Rathore, I. et al. Activation mechanism of plasmepsins, pepsin-like aspartic proteases from Plasmodium , follows a unique trans-activation pathway. FEBS J. 288 , 678–698 (2021).

Bopp, S. et al. Plasmepsin II-III copy number accounts for bimodal piperaquine resistance among Cambodian Plasmodium falciparum . Nat. Commun. 9 , 1769 (2018).

Dhingra, S. K. et al. A variant PfCRT isoform can contribute to Plasmodium falciparum resistance to the first-line partner drug piperaquine. mBio 8 , e00303-17 (2017).

Ross, L. S. et al. Emerging Southeast Asian PfCRT mutations confer Plasmodium falciparum resistance to the first-line antimalarial piperaquine. Nat. Commun. 9 , 3314 (2018).

Dhingra, S. K., Small-Saunders, J. L., Ménard, D. & Fidock, D. A. Plasmodium falciparum resistance to piperaquine driven by PfCRT. Lancet Infect. Dis. 19 , 1168–1169 (2019).

Hamilton, W. L. et al. Evolution and expansion of multidrug-resistant malaria in southeast Asia: a genomic epidemiology study. Lancet Infect. Dis. 19 , 943–951 (2019).

Kim, J. et al. Structure and drug resistance of the Plasmodium falciparum transporter PfCRT. Nature 576 , 315–320 (2019). The emergence of mutations in the chloroquine resistance transporter resulted in a worldwide crisis and an urgent search for new medicines. Here, the structure of PfCRT isoforms resistant to chloroquine was solved, demonstrating how these mutations confer resistance.

Agrawal, S. et al. Association of a novel mutation in the Plasmodium falciparum chloroquine resistance transporter with decreased piperaquine sensitivity. J. Infect. Dis. 216 , 468–476 (2017).

Small-Saunders, J. L. et al. Evidence for the early emergence of piperaquine-resistant Plasmodium falciparum malaria and modeling strategies to mitigate resistance. PLoS Pathog. 18 , e1010278 (2022).

van der Pluijm, R. W. et al. Triple artemisinin-based combination therapies versus artemisinin-based combination therapies for uncomplicated Plasmodium falciparum malaria: a multicentre, open-label, randomised clinical trial. Lancet 395 , 1345–1360 (2020).

van der Pluijm, R. W., Amaratunga, C., Dhorda, M. & Dondorp, A. M. Triple artemisinin-based combination therapies for malaria — a new paradigm? Trends Parasitol. 37 , 15–24 (2021).

Boni, M. F., White, N. J. & Baird, J. K. The community as the patient in malaria-endemic areas: preempting drug resistance with multiple first-line therapies. PLoS Med. 13 , e1001984 (2016).

Marwa, K. et al. Therapeutic efficacy of artemether-lumefantrine, artesunate-amodiaquine and dihydroartemisinin-piperaquine in the treatment of uncomplicated Plasmodium falciparum malaria in sub-Saharan Africa: a systematic review and meta-analysis. PLoS ONE   17 , e0264339 (2022).

Chotsiri, P. et al. Piperaquine pharmacokinetics during intermittent preventive treatment for malaria in pregnancy. Antimicrob. Agents Chemother. https://doi.org/10.1128/AAC.01150-20 (2021).

Qiu, D. et al. A G358S mutation in the Plasmodium falciparum Na + pump PfATP4 confers clinically-relevant resistance to cipargamin. Nat. Commun. 13 , 5746 (2022).

Schmitt, E. K. et al. Efficacy of cipargamin (KAE609) in a randomized, phase II dose-escalation study in adults in sub-Saharan Africa with uncomplicated Plasmodium falciparum malaria. Clin. Infect. Dis. 74 , 1831–1839 (2022).

Duffey, M. et al. Assessing risks of Plasmodium falciparum resistance to select next-generation antimalarials. Trends Parasitol. 37 , 709–721 (2021).

Tse, E. G., Korsik, M. & Todd, M. H. The past, present and future of anti-malarial medicines. Malar. J. 18 , 93 (2019).

Guantai, E. & Chibale, K. How can natural products serve as a viable source of lead compounds for the development of new/novel anti-malarials? Malar. J. 10 , S2 (2011).

Plouffe, D. et al. In silico activity profiling reveals the mechanism of action of antimalarials discovered in a high-throughput screen. Proc. Natl Acad. Sci. USA 105 , 9059–9064 (2008).

Guiguemde, W. A. et al. Chemical genetics of Plasmodium falciparum . Nature 465 , 311–315 (2010).

Miguel-Blanco, C. et al. Hundreds of dual-stage antimalarial molecules discovered by a functional gametocyte screen. Nat. Commun. 8 , 15160 (2017).

Gamo, F. J. et al. Thousands of chemical starting points for antimalarial lead identification. Nature 465 , 305–310 (2010).

Hovlid, M. L. & Winzeler, E. A. Phenotypic screens in antimalarial drug discovery. Trends Parasitol. 32 , 697–707 (2016).

Delves, M. J. et al. A high throughput screen for next-generation leads targeting malaria parasite transmission. Nat. Commun. 9 , 3805 (2018).

Antonova-Koch, Y. et al. Open-source discovery of chemical leads for next-generation chemoprotective antimalarials. Science https://doi.org/10.1126/science.aat9446 (2018). A significant example of one of the first screening campaigns against the parasite liver stage that revealed novel scaffolds and compounds with previously unidentified mechanisms of action.

Abraham, M. et al. Probing the open global health chemical diversity library for multistage-active starting points for next-generation antimalarials. ACS Infect. Dis. https://doi.org/10.1021/acsinfecdis.9b00482 (2020).

Guiguemde, W. A. et al. Global phenotypic screening for antimalarials. Chem. Biol. 19 , 116–129 (2012).

Xie, S. C. et al. Design of proteasome inhibitors with oral efficacy in vivo against Plasmodium falciparum and selectivity over the human proteasome. Proc. Natl Acad. Sci. USA https://doi.org/10.1073/pnas.2107213118 (2021).

Favuzza, P. et al. Dual plasmepsin-targeting antimalarial agents disrupt multiple stages of the malaria parasite life cycle. Cell Host Microbe 27 , 642–658 (2020).

Forte, B. et al. Prioritization of molecular targets for antimalarial drug discovery. ACS Infect. Dis. 7 , 2764–2776 (2021). Describes the most important criteria for selecting targets for antimalarial drug discovery and analyses several drug targets in the MalDA pipeline as well as strategies for identifying additional drug targets.

Chughlay, M. F. et al. Chemoprotective antimalarial activity of P218 against Plasmodium falciparum : a randomized, placebo-controlled volunteer infection study. Am. J. Trop. Med. Hyg. 104 , 1348–1358 (2021).

Baldwin, J. et al. High-throughput screening for potent and selective inhibitors of Plasmodium falciparum dihydroorotate dehydrogenase. J. Biol. Chem. 280 , 21847–21853 (2005).

Phillips, M. A. et al. A long-duration dihydroorotate dehydrogenase inhibitor (DSM265) for prevention and treatment of malaria. Sci. Transl Med. 7 , 296ra111 (2015).

Llanos-Cuentas, A. et al. Antimalarial activity of single-dose DSM265, a novel Plasmodium dihydroorotate dehydrogenase inhibitor, in patients with uncomplicated Plasmodium falciparum or Plasmodium vivax malaria infection: a proof-of-concept, open-label, phase 2a study. Lancet Infect. Dis. 18 , 874–883 (2018).

Agarwal, A. et al. Discovery of a selective, safe and novel anti-malarial compound with activity against chloroquine resistant strain of Plasmodium falciparum . Sci. Rep. 5 , 13838 (2015).

Pavadai, E. et al. Identification of new human malaria parasite Plasmodium falciparum dihydroorotate dehydrogenase inhibitors by pharmacophore and structure-based virtual screening. J. Chem. Inf. Model. 56 , 548–562 (2016).

Uddin, A. et al. Target-based virtual screening of natural compounds identifies a potent antimalarial with selective falcipain-2 inhibitory activity. Front. Pharmacol. 13 , 850176 (2022).

Dascombe, M. J. et al. Mapping antimalarial pharmacophores as a useful tool for the rapid discovery of drugs effective in vivo: design, construction, characterization, and pharmacology of metaquine. J. Med. Chem. 48 , 5423–5436 (2005).

de Sousa, A. C. C., Combrinck, J. M., Maepa, K. & Egan, T. J. Virtual screening as a tool to discover new beta-haematin inhibitors with activity against malaria parasites. Sci. Rep. 10 , 3374 (2020).

Ruggeri, C. et al. Identification and validation of a potent dual inhibitor of the P. falciparum M1 and M17 aminopeptidases using virtual screening. PLoS ONE   10 , e0138957 (2015).

Godinez, W. J. et al. Design of potent antimalarials with generative chemistry. Nat. Mach. Intell. 4 , 180–186 (2022).

Article   Google Scholar  

Yang, T. et al. MalDA, accelerating malaria drug discovery. Trends Parasitol. 37 , 493–507 (2021).

Summers, R. L. et al. Chemogenomics identifies acetyl-coenzyme A synthetase as a target for malaria treatment and prevention. Cell Chem. Biol. 29 , 191–201 (2022). Reports P. falciparum acetyl-coenzyme A synthase as an essential protein for parasite viability associated with histone acetylation and shows it can be inhibited by small molecules (MMV019721 and MMV084978).

Schalkwijk, J. et al. Antimalarial pantothenamide metabolites target acetyl-coenzyme A biosynthesis in Plasmodium falciparum .  Sci. Transl Med. 11 , eaas9917 (2019).

Paquet, T. et al. Antimalarial efficacy of MMV390048, an inhibitor of Plasmodium phosphatidylinositol 4-kinase.  Sci. Transl Med. 9 , eaad9735 (2017). Describes the development of MMV390048, the first candidate drug discovered and developed entirely in Africa, the first antimalarial kinase inhibitor tested in humans, and the first antimalarial with a phase I trial performed in Africa.

Vanaerschot, M. et al. Inhibition of resistance-refractory P. falciparum kinase PKG delivers prophylactic, blood stage, and transmission-blocking antiplasmodial activity. Cell Chem. Biol. 27 , 806–816.e8 (2020).

Dziekan, J. M. et al. Cellular thermal shift assay for the identification of drug-target interactions in the Plasmodium falciparum proteome. Nat. Protoc. 15 , 1881–1921 (2020).

Begolo, D., Erben, E. & Clayton, C. Drug target identification using a trypanosome overexpression library. Antimicrob. Agents Chemother. 58 , 6260–6264 (2014).

Melief, E. et al. Construction of an overexpression library for Mycobacterium tuberculosis . Biol. Methods Protoc. 3 , bpy009 (2018).

Allman, E. L., Painter, H. J., Samra, J., Carrasquilla, M. & Llinás, M. Metabolomic profiling of the malaria box reveals antimalarial target pathways. Antimicrob. Agents Chemother. 60 , 6635–6649 (2016).

Ganesan, S. M., Falla, A., Goldfless, S. J., Nasamu, A. S. & Niles, J. C. Synthetic RNA-protein modules integrated with native translation mechanisms to control gene expression in malaria parasites. Nat. Commun. 7 , 10727 (2016).

Hoepfner, D. et al. Selective and specific inhibition of the Plasmodium falciparum lysyl-tRNA synthetase by the fungal secondary metabolite cladosporin. Cell Host Microbe 11 , 654–663 (2012).

Baragana, B. et al. Lysyl-tRNA synthetase as a drug target in malaria and cryptosporidiosis. Proc. Natl Acad. Sci. USA 116 , 7015–7020 (2019).

Kato, N. et al. Diversity-oriented synthesis yields novel multistage antimalarial inhibitors. Nature 538 , 344–349 (2016).

Tye, M. A. et al. Elucidating the path to Plasmodium prolyl-tRNA synthetase inhibitors that overcome halofuginone-resistance. Nat. Commun. 13 , 4976 (2022). Reports the development of an aminoacyl tRNA synthetase biochemical assay based on TR-FRET and the design of P. falciparum prolyl-tRNA synthetase high-affinity inhibitors that can simultaneously engage all three substrate-binding pockets of the enzyme.

Sonoiki, E. et al. Antimalarial benzoxaboroles target Plasmodium falciparum Leucyl-tRNA synthetase. Antimicrob. Agents Chemother. 60 , 4886–4895 (2016).

Xie, S. C. et al. Reaction hijacking of tyrosine tRNA synthetase as a new whole-of-life-cycle antimalarial strategy. Science 376 , 1074–1079 (2022). Reports a novel biochemical feature of P. falciparum aminoacyl tRNA synthetases and identifies and characterizes a potent inhibitor with activity across the parasite lifecycle that acts via reaction hijacking.

Istvan, E. S. et al. Cytoplasmic isoleucyl tRNA synthetase as an attractive multistage antimalarial drug target. Sci. Transl Med. 15 , eadc9249 (2023).

Sharma, M. et al. Structural basis of malaria parasite phenylalanine tRNA-synthetase inhibition by bicyclic azetidines. Nat. Commun. 12 , 343 (2021).

Jain, V. et al. Structure of prolyl-tRNA synthetase-halofuginone complex provides basis for development of drugs against malaria and toxoplasmosis. Structure 23 , 819–829 (2015).

Ndagi, U., Kumalo, H. M. & Mhlongo, N. N. A consequence of drug targeting of aminoacyl-tRNA synthetases in Mycobacterium tuberculosis . Chem. Biol. Drug Des. 98 , 421–434 (2021).

Jain, V. et al. Targeting prolyl-tRNA synthetase to accelerate drug discovery against malaria, leishmaniasis, toxoplasmosis, cryptosporidiosis, and coccidiosis. Structure 25 , 1495–1505.e6 (2017).

Tandon, S. et al. Deciphering the interaction of benzoxaborole inhibitor AN2690 with connective polypeptide 1 (CP1) editing domain of Leishmania donovani leucyl-tRNA synthetase. J. Biosci. 45 , 63 (2020).

Mishra, S. et al. Conformational heterogeneity in apo and drug-bound structures of Toxoplasma gondii prolyl-tRNA synthetase. Acta Crystallogr. F. Struct. Biol. Commun. 75 , 714–724 (2019).

Radke, J. B. et al. Bicyclic azetidines target acute and chronic stages of Toxoplasma gondii by inhibiting parasite phenylalanyl t-RNA synthetase. Nat. Commun. 13 , 459 (2022).

Alam, M. M. et al. Validation of the protein kinase PfCLK3 as a multistage cross-species malarial drug target. Science https://doi.org/10.1126/science.aau1682 (2019).

Ducati, R. G. et al. Genetic resistance to purine nucleoside phosphorylase inhibition in Plasmodium falciparum . Proc. Natl Acad. Sci. USA 115 , 2114–2119 (2018).

Xie, S. C. et al. Target validation and identification of novel boronate inhibitors of the Plasmodium falciparum proteasome. J. Med. Chem. 61 , 10053–10066 (2018).

Stokes, B. H. et al. Covalent Plasmodium falciparum -selective proteasome inhibitors exhibit a low propensity for generating resistance in vitro and synergize with multiple antimalarial agents. PLoS Pathog. 15 , e1007722 (2019).

Zhan, W. et al. Development of a highly selective Plasmodium falciparum proteasome inhibitor with anti-malaria activity in humanized mice. Angew. Chem. Int. Ed. Engl. 60 , 9279–9283 (2021).

LaMonte, G. M. et al. Development of a potent inhibitor of the Plasmodium proteasome with reduced mammalian toxicity. J. Med. Chem. 60 , 6721–6732 (2017).

Li, H. et al. Structure- and function-based design of Plasmodium -selective proteasome inhibitors. Nature 530 , 233–236 (2016).

Yoo, E. et al. Defining the determinants of specificity of Plasmodium proteasome inhibitors. J. Am. Chem. Soc. 140 , 11424–11437 (2018).

Deni, I. et al. Mitigating the risk of antimalarial resistance via covalent dual-subunit inhibition of the Plasmodium proteasome. Cell Chem. Biol. 30 , 470–485 (2023).

Wyllie, S. et al. Preclinical candidate for the treatment of visceral leishmaniasis that acts through proteasome inhibition. Proc. Natl Acad. Sci. USA 116 , 9318–9323 (2019).

Khare, S. et al. Proteasome inhibition for treatment of leishmaniasis, Chagas disease and sleeping sickness. Nature 537 , 229–233 (2016).

Lima, M. L. et al. Identification of a proteasome-targeting arylsulfonamide with potential for the treatment of Chagas’ disease. Antimicrob. Agents Chemother. 66 , e0153521 (2022).

Waller, R. F. et al. A type II pathway for fatty acid biosynthesis presents drug targets in Plasmodium falciparum . Antimicrob. Agents Chemother. 47 , 297–301 (2003).

Yeh, E. & DeRisi, J. L. Chemical rescue of malaria parasites lacking an apicoplast defines organelle function in blood-stage Plasmodium falciparum . PLoS Biol. 9 , e1001138 (2011).

Kennedy, K., Crisafulli, E. M. & Ralph, S. A. Delayed death by plastid inhibition in apicomplexan parasites. Trends Parasitol. 35 , 747–759 (2019).

Tang, Y. et al. A mutagenesis screen for essential plastid biogenesis genes in human malaria parasites. PLoS Biol. 17 , e3000136 (2019).

Zhang, M. et al. Uncovering the essential genes of the human malaria parasite Plasmodium falciparum by saturation mutagenesis. Science https://doi.org/10.1126/science.aap7847 (2018).

Bushell, E. et al. Functional profiling of a Plasmodium genome reveals an abundance of essential genes. Cell 170 , 260–272.e8 (2017).

Stanway, R. R. et al. Genome-scale identification of essential metabolic processes for targeting the Plasmodium liver stage. Cell 179 , 1112–1128 (2019).

MMV. MMV’s Pipeline of Antimalarial Drugs https://www.mmv.org/research-development/mmvs-pipeline-antimalarial-drugs (2023).

Abd-Rahman, A. N., Zaloumis, S., McCarthy, J. S., Simpson, J. A. & Commons, R. J. Scoping review of antimalarial drug candidates in phase I and II drug development. Antimicrob. Agents Chemother. 66 , e0165921 (2022).

Wu, R. L. et al. Low-dose subcutaneous or intravenous monoclonal antibody to prevent malaria. N. Engl. J. Med. 387 , 397–407 (2022).

Hameed, P. S. et al. Triaminopyrimidine is a fast-killing and long-acting antimalarial clinical candidate. Nat. Commun. 6 , 6715 (2015).

Barber, B. E. et al. Safety, pharmacokinetics, and antimalarial activity of the novel triaminopyrimidine ZY-19489: a first-in-human, randomised, placebo-controlled, double-blind, single ascending dose study, pilot food-effect study, and volunteer infection study. Lancet Infect. Dis. 22 , 879–890 (2022).

Daubenberger, C. & Burrows, J. N. Volunteer infection studies accelerate the clinical development of novel drugs against malaria. Lancet Infect. Dis. 22 , 753–754 (2022).

Dive, D. & Biot, C. Ferroquine as an oxidative shock antimalarial. Curr. Top. Med. Chem. 14 , 1684–1692 (2014).

Baragana, B. et al. A novel multiple-stage antimalarial agent that inhibits protein synthesis. Nature 522 , 315–320 (2015).

Baragana, B. et al. Discovery of a quinoline-4-carboxamide derivative with a novel mechanism of action, multistage antimalarial activity, and potent in vivo efficacy. J. Med. Chem. 59 , 9672–9685 (2016).

Rottmann, M. et al. Preclinical antimalarial combination study of M5717, a Plasmodium falciparum elongation factor 2 inhibitor, and pyronaridine, a hemozoin formation inhibitor. Antimicrob. Agents Chemother. https://doi.org/10.1128/AAC.02181-19 (2020).

McCarthy, J. S. et al. Safety, pharmacokinetics, and antimalarial activity of the novel Plasmodium eukaryotic translation elongation factor 2 inhibitor M5717: a first-in-human, randomised, placebo-controlled, double-blind, single ascending dose study and volunteer infection study. Lancet Infect. Dis. 21 , 1713–1724 (2021). Reports promising human efficacy data with M5717 (cabamiquine) that is the basis of a phase II trial combination with pyronaridine.

Meister, S. et al. Imaging of Plasmodium liver stages to drive next-generation antimalarial drug discovery. Science 334 , 1372–1377 (2011).

Kuhen, K. L. et al. KAF156 is an antimalarial clinical candidate with potential for use in prophylaxis, treatment, and prevention of disease transmission. Antimicrob. Agents Chemother. 58 , 5060–5067 (2014).

Lehane, A. M. et al. Characterization of the ATP4 ion pump in Toxoplasma gondii . J. Biol. Chem. 294 , 5720–5734 (2019).

Kreutzfeld, O. et al. Associations between varied susceptibilities of PfATP4 inhibitors and genotypes in Ugandan Plasmodium falciparum isolates. Antimicrob. Agents Chemother. https://doi.org/10.1128/aac.00771-21 (2021).

Flannery, E. L. et al. Mutations in the P-type cation-transporter ATPase 4, PfATP4, mediate resistance to both aminopyrazole and spiroindolone antimalarials. ACS Chem. Biol. 10 , 413–420 (2015).

Lim, M. Y. et al. UDP-galactose and acetyl-CoA transporters as Plasmodium multidrug resistance genes. Nat. Microbiol. 1 , 16166 (2016).

LaMonte, G. et al. Mutations in the Plasmodium falciparum cyclic amine resistance locus (PfCARL) confer multidrug resistance. mBio https://doi.org/10.1128/mBio.00696-16 (2016).

Magistrado, P. A. et al. Plasmodium falciparum cyclic amine resistance locus (PfCARL), a resistance mechanism for two distinct compound classes. ACS Infect. Dis. 2 , 816–826 (2016).

Kublin, J. G. et al. Safety, pharmacokinetics, and causal prophylactic efficacy of KAF156 in a Plasmodium falciparum human infection study. Clin. Infect. Dis. 73 , e2407–e2414 (2021).

Shah, R. et al. Formulation development and characterization of lumefantrine nanosuspension for enhanced antimalarial activity. J. Biomater. Sci. Polym. Ed. 32 , 833–857 (2021).

NOVARTIS. Novartis and Medicines for Malaria Venture Announce Decision to move to Phase 3 study for Novel Ganaplacide/Lumefantrine-SDF Combination in Adults and Children with Malaria https://www.novartis.com/news/media-releases/novartis-and-medicines-malaria-venture-announce-decision-move-phase-3-study-novel-ganaplacidelumefantrine-sdf-combination-adults-and-children-malaria (2022).

Srivastava, I. K. & Vaidya, A. B. A mechanism for the synergistic antimalarial action of atovaquone and proguanil. Antimicrob. Agents Chemother. 43 , 1334–1339 (1999).

Murithi, J. M. et al. The antimalarial MMV688533 provides potential for single-dose cures with a high barrier to Plasmodium falciparum parasite resistance. Sci. Transl Med https://doi.org/10.1126/scitranslmed.abg6013 (2021). Reports a promising new antimalarial with a low resistance risk, fast parasite clerance and excellent pharmacological properties.

Taft, B. R. et al. Discovery and preclinical pharmacology of INE963, a potent and fast-acting blood-stage antimalarial with a high barrier to resistance and potential for single-dose cures in uncomplicated malaria. J. Med. Chem. 65 , 3798–3813 (2022). INE963 is a novel irresistible chemotype with fast parasite clearance and is completing phase I clinical study.

Duffy, P. E. & Patrick Gorres, J. Malaria vaccines since 2000: progress, priorities, products. NPJ Vaccines 5 , 48 (2020).

Pethrak, C. et al. New Insights into antimalarial chemopreventive activity of antifolates. Antimicrob. Agents Chemother. 66 , e0153821 (2022).

Katsuno, K. et al. Hit and lead criteria in drug discovery for infectious diseases of the developing world. Nat. Rev. Drug Discov. 14 , 751–758 (2015).

de Vries, L. E. et al. Preclinical characterization and target validation of the antimalarial pantothenamide MMV693183. Nat. Commun. 13 , 2158 (2022).

Pegoraro, S. et al. SC83288 is a clinical development candidate for the treatment of severe malaria. Nat. Commun. 8 , 14193 (2017).

MMV. MMV’s Pipeline of Antimalarial Drugs: GSK484 https://www.mmv.org/mmv-pipeline-antimalarial-drugs/gsk484 (2023).

MMV. MMV’s Pipeline of Antimalarial Drugs: IWY35 https://www.mmv.org/mmv-pipeline-antimalarial-drugs/iwy357 (2023).

Tisnerat, C., Dassonville-Klimpt, A., Gosselet, F. & Sonnet, P. Antimalarial drug discovery: from quinine to the most recent promising clinical drug candidates. Curr. Med. Chem. 29 , 3326–3365 (2022).

Smilkstein, M. J. et al. ELQ-331 as a prototype for extremely durable chemoprotection against malaria. Malar. J. 18 , 291 (2019).

Nilsen, A. et al. Quinolone-3-diarylethers: a new class of antimalarial drug. Sci. Transl Med. 5 , 177ra137 (2013).

Dola, V. R. et al. Synthesis and evaluation of chirally defined side chain variants of 7-chloro-4-aminoquinoline to overcome drug resistance in malaria chemotherapy. Antimicrob. Agents Chemother. 61 , e01152-16 (2017).

MMV. MMV’s Pipeline of Antimalarial Drugs: MMV609 https://www.mmv.org/mmv-pipeline-antimalarial-drugs/mmv609 (2023).

Jimenez-Diaz, M. B. et al. (+)-SJ733, a clinical candidate for malaria that acts through ATP4 to induce rapid host-mediated clearance of Plasmodium . Proc. Natl Acad. Sci. USA 111 , E5455–E5462 (2014).

Sulyok, M. et al. DSM265 for Plasmodium falciparum chemoprophylaxis: a randomised, double blinded, phase 1 trial with controlled human malaria infection. Lancet Infect. Dis. 17 , 636–644 (2017).

Demarta-Gatsi, C. et al. Malarial PI4K inhibitor induced diaphragmatic hernias in rat: potential link with mammalian kinase inhibition. Birth Defects Res. 114 , 487–498 (2022).

Moehrle, J. J. et al. First-in-man safety and pharmacokinetics of synthetic ozonide OZ439 demonstrates an improved exposure profile relative to other peroxide antimalarials. Br. J. Clin. Pharmacol. 75 , 524–537 (2013).

Adoke, Y. et al. A randomized, double-blind, phase 2b study to investigate the efficacy, safety, tolerability and pharmacokinetics of a single-dose regimen of ferroquine with artefenomel in adults and children with uncomplicated Plasmodium falciparum malaria. Malar. J. 20 , 222 (2021).

Ravikumar, A., Arzumanyan, G. A., Obadi, M. K. A., Javanpour, A. A. & Liu, C. C. Scalable, continuous evolution of genes at mutation rates above genomic error thresholds. Cell 175 , 1946–1957 (2018).

Christopoulos, K. A. et al. First demonstration project of long-acting injectable antiretroviral therapy for persons with and without detectable HIV viremia in an urban HIV clinic. Clin. Infect. Dis. https://doi.org/10.1093/cid/ciac631 (2022).

Schmidt, H. A., Rodolph, M., Schaefer, R., Baggaley, R. & Doherty, M. Long-acting injectable cabotegravir: implementation science needed to advance this additional HIV prevention choice. J. Int. AIDS Soc. 25 , e25963 (2022).

Song, Z., Iorga, B. I., Mounkoro, P., Fisher, N. & Meunier, B. The antimalarial compound ELQ-400 is an unusual inhibitor of the bc1 complex, targeting both Qo and Qi sites. FEBS Lett. 592 , 1346–1356 (2018).

Stickles, A. M. et al. Atovaquone and ELQ-300 combination therapy as a novel dual-site cytochrome bc1 inhibition strategy for malaria. Antimicrob. Agents Chemother. 60 , 4853–4859 (2016).

de Villiers, K. A. & Egan, T. J. Heme detoxification in the malaria parasite: a target for antimalarial drug development. Acc. Chem. Res. 54 , 2649–2659 (2021).

Heinberg, A. & Kirkman, L. The molecular basis of antifolate resistance in Plasmodium falciparum : looking beyond point mutations. Ann. NY Acad. Sci. 1342 , 10–18 (2015).

Chitnumsub, P. et al. The structure of Plasmodium falciparum hydroxymethyldihydropterin pyrophosphokinase-dihydropteroate synthase reveals the basis of sulfa resistance. FEBS J. 287 , 3273–3297 (2020).

Wang, X. et al. MEPicides: α,β-unsaturated fosmidomycin analogues as DXR inhibitors against malaria. J. Med. Chem. 61 , 8847–8858 (2018).

Herman, J. D. et al. The cytoplasmic prolyl-tRNA synthetase of the malaria parasite is a dual-stage target of febrifugine and its analogs. Sci. Transl Med. 7 , 288ra277 (2015).

Keller, T. L. et al. Halofuginone and other febrifugine derivatives inhibit prolyl-tRNA synthetase. Nat. Chem. Biol. 8 , 311–317 (2012).

Derbyshire, E. R., Mazitschek, R. & Clardy, J. Characterization of Plasmodium liver stage inhibition by halofuginone. ChemMedChem 7 , 844–849 (2012).

Dharia, N. V. et al. Genome scanning of Amazonian Plasmodium falciparum shows subtelomeric instability and clindamycin-resistant parasites. Genome Res. 20 , 1534–1544 (2010).

Collins, K. A. et al. DSM265 at 400 milligrams clears asexual stage parasites but not mature gametocytes from the blood of healthy subjects experimentally infected with Plasmodium falciparum . Antimicrob. Agents Chemother. https://doi.org/10.1128/AAC.01837-18 (2019).

White, J. et al. Identification and mechanistic understanding of dihydroorotate dehydrogenase point mutations in Plasmodium falciparum that confer in vitro resistance to the clinical candidate DSM265. ACS Infect. Dis. 5 , 90–101 (2019).

Belard, S. & Ramharter, M. DSM265: a novel drug for single-dose cure of Plasmodium falciparum malaria. Lancet Infect. Dis. 18 , 819–820 (2018).

Palmer, M. J. et al. Potent antimalarials with development potential identified by structure-guided computational optimization of a pyrrole-based dihydroorotate dehydrogenase inhibitor series. J. Med. Chem. 64 , 6085–6136 (2021).

Ndayisaba, G. et al. Hepatic safety and tolerability of cipargamin (KAE609), in adult patients with Plasmodium falciparum malaria: a randomized, phase II, controlled, dose-escalation trial in sub-Saharan Africa. Malar. J. 20 , 478 (2021).

McCarthy, J. S. et al. Defining the antimalarial activity of cipargamin in healthy volunteers experimentally infected with blood-stage Plasmodium falciparum . Antimicrob. Agents Chemother. https://doi.org/10.1128/AAC.01423-20 (2021).

Ashton, T. D. et al. Optimization of 2,3-dihydroquinazolinone-3-carboxamides as antimalarials targeting PfATP4. J. Med. Chem. https://doi.org/10.1021/acs.jmedchem.2c02092 (2023).

Liang, X. et al. Discovery of 6′-chloro-N-methyl-5′-(phenylsulfonamido)-[3,3′-bipyridine]−5-carboxamide (CHMFL-PI4K-127) as a novel Plasmodium falciparum PI(4)K inhibitor with potent antimalarial activity against both blood and liver stages of Plasmodium . Eur. J. Med. Chem. 188 , 112012 (2020).

Brunschwig, C. et al. UCT943, a next-generation Plasmodium falciparum PI4K inhibitor preclinical candidate for the treatment of malaria. Antimicrob. Agents Chemother. 62 , e00012-18 (2018).

Cabrera, D. G. et al. Plasmodial kinase inhibitors: license to cure? J. Med. Chem. 61 , 8061–8077 (2018).

Sinxadi, P. et al. Safety, tolerability, pharmacokinetics, and antimalarial activity of the novel Plasmodium phosphatidylinositol 4-kinase inhibitor MMV390048 in healthy volunteers. Antimicrob. Agents Chemother. https://doi.org/10.1128/aac.01896-19 (2020).

Khandelwal, A. et al. Translation of liver stage activity of M5717, a Plasmodium elongation factor 2 inhibitor: from bench to bedside. Malar. J. 21 , 151 (2022).

Ng, C. L. et al. CRISPR-Cas9-modified pfmdr1 protects Plasmodium falciparum asexual blood stages and gametocytes against a class of piperazine-containing compounds but potentiates artemisinin-based combination therapy partner drugs. Mol. Microbiol. 101 , 381–393 (2016).

de Lera Ruiz, M. et al. The invention of WM382, a highly potent PMIX/X dual inhibitor toward the treatment of malaria. ACS Med. Chem. Lett. 13 , 1745–1754 (2022).

Lowe, M. A. et al. Discovery and characterization of potent, efficacious, orally available antimalarial plasmepsin X inhibitors and preclinical safety assessment of UCB7362. J. Med. Chem. 65 , 14121–14143 (2022).

Nasamu, A. S. et al. Plasmepsins IX and X are essential and druggable mediators of malaria parasite egress and invasion. Science 358 , 518–522 (2017).

Schwertz, G. et al. Antimalarial inhibitors targeting serine hydroxymethyltransferase (SHMT) with in vivo efficacy and analysis of their binding mode based on X-ray cocrystal structures. J. Med. Chem. 60 , 4840–4860 (2017).

Zaki, M. E. A., Al-Hussain, S. A., Masand, V. H., Akasapu, S. & Lewaa, I. QSAR and pharmacophore modeling of nitrogen heterocycles as potent human N-myristoyltransferase (Hs-NMT) inhibitors. Molecules https://doi.org/10.3390/molecules26071834 (2021).

Schlott, A. C., Holder, A. A. & Tate, E. W. N-Myristoylation as a drug target in malaria: exploring the role of N-myristoyltransferase substrates in the inhibitor mode of action. ACS Infect. Dis. 4 , 449–457 (2018).

de Vries, L. E., Lunghi, M., Krishnan, A., Kooij, T. W. A. & Soldati-Favre, D. Pantothenate and CoA biosynthesis in Apicomplexa and their promise as antiparasitic drug targets. PLoS Pathog. 17 , e1010124 (2021).

Bopp, S. et al. Potent acyl-CoA synthetase 10 inhibitors kill Plasmodium falciparum by disrupting triglyceride formation. Nat. Commun. 14 , 1455 (2023).

Chhibber-Goel, J., Yogavel, M. & Sharma, A. Structural analyses of the malaria parasite aminoacyl-tRNA synthetases provide new avenues for antimalarial drug discovery. Protein Sci. 30 , 1793–1803 (2021).

Novoa, E. M. et al. Analogs of natural aminoacyl-tRNA synthetase inhibitors clear malaria in vivo. Proc. Natl Acad. Sci. USA 111 , E5508–E5517 (2014).

von Bredow, L. et al. Synthesis, antiplasmodial, and antileukemia activity of dihydroartemisinin-HDAC inhibitor hybrids as multitarget drugs. Pharmaceuticals https://doi.org/10.3390/ph15030333 (2022).

Jublot, D. et al. A histone deacetylase (HDAC) inhibitor with pleiotropic in vitro anti-toxoplasma and anti- Plasmodium activities controls acute and chronic toxoplasma infection in mice. Int. J. Mol. Sci. https://doi.org/10.3390/ijms23063254 (2022).

Wang, M. et al. Drug repurposing of quisinostat to discover novel Plasmodium falciparum HDAC1 inhibitors with enhanced triple-stage antimalarial activity and improved safety. J. Med. Chem. 65 , 4156–4181 (2022).

Zhang, H. et al. Design, synthesis, and optimization of macrocyclic peptides as species-selective antimalaria proteasome inhibitors. J. Med. Chem. 65 , 9350–9375 (2022).

Almaliti, J. et al. Development of potent and highly selective epoxyketone-based Plasmodium proteasome inhibitors. Chemistry https://doi.org/10.1002/chem.202203958 (2023).

Kabeche, S. et al. Nonbisphosphonate inhibitors of Plasmodium falciparum FPPS/GGPPS. Bioorg Med. Chem. Lett. 41 , 127978 (2021).

Gisselberg, J. E., Herrera, Z., Orchard, L. M., Llinás, M. & Yeh, E. Specific inhibition of the bifunctional farnesyl/geranylgeranyl diphosphate synthase in malaria parasites via a new small-molecule binding site. Cell Chem. Biol. 25 , 185–193.e5 (2018).

Venkatramani, A., Gravina Ricci, C., Oldfield, E. & McCammon, J. A. Remarkable similarity in Plasmodium falciparum and Plasmodium vivax geranylgeranyl diphosphate synthase dynamics and its implication for antimalarial drug design. Chem. Biol. Drug Des. 91 , 1068–1077 (2018).

Gabriel, H. B. et al. Single-target high-throughput transcription analyses reveal high levels of alternative splicing present in the FPPS/GGPPS from Plasmodium falciparum . Sci. Rep. 5 , 18429 (2015).

Jordão, F. M. et al. Cloning and characterization of bifunctional enzyme farnesyl diphosphate/geranylgeranyl diphosphate synthase from Plasmodium falciparum . Malar. J. 12 , 184 (2013).

Ricci, C. G. et al. Dynamic structure and inhibition of a malaria drug target: geranylgeranyl diphosphate synthase. Biochemistry 55 , 5180–5190 (2016).

Istvan, E. S. et al. Plasmodium Niemann-Pick type C1-related protein is a druggable target required for parasite membrane homeostasis. eLife https://doi.org/10.7554/eLife.40529 (2019).

Moustakim, M. et al. Discovery of a PCAF bromodomain chemical probe. Angew. Chem. Int. Ed. Engl. 56 , 827–831 (2017).

Swift, R. P., Rajaram, K., Liu, H. B. & Prigge, S. T. Dephospho-CoA kinase, a nuclear-encoded apicoplast protein, remains active and essential after Plasmodium falciparum apicoplast disruption. EMBO J. 40 , e107247 (2021).

Fletcher, S. et al. Biological characterization of chemically diverse compounds targeting the Plasmodium falciparum coenzyme A synthesis pathway. Parasit. Vectors 9 , 589 (2016).

Salazar, E. et al. Characterization of Plasmodium falciparum adenylyl cyclase-β and its role in erythrocytic stage parasites. PLoS ONE   7 , e39769 (2012).

Jiang, X. An overview of the Plasmodium falciparum hexose transporter and its therapeutic interventions. Proteins https://doi.org/10.1002/prot.26351 (2022).

Barbieri, D. et al. The phosphodiesterase inhibitor tadalafil promotes splenic retention of Plasmodium falciparum gametocytes in humanized mice. Front. Cell Infect. Microbiol. 12 , 883759 (2022).

Ding, S. et al. Probing the B- & C-rings of the antimalarial tetrahydro-beta-carboline MMV008138 for steric and conformational constraints. Bioorg Med. Chem. Lett. 30 , 127520 (2020).

Imlay, L. S. et al. Plasmodium IspD (2-C-methyl-D-erythritol 4-phosphate cytidyltransferase), an essential and druggable antimalarial target. ACS Infect. Dis. 1 , 157–167 (2015).

Nerlich, C., Epalle, N. H., Seick, P. & Beitz, E. Discovery and development of inhibitors of the plasmodial FNT-type lactate transporter as novel antimalarials. Pharmaceuticals https://doi.org/10.3390/ph14111191 (2021).

Walloch, P., Hansen, C., Priegann, T., Schade, D. & Beitz, E. Pentafluoro-3-hydroxy-pent-2-en-1-ones potently inhibit FNT-type lactate transporters from all five human-pathogenic Plasmodium species. ChemMedChem 16 , 1283–1289 (2021).

Peng, X. et al. Structural characterization of the Plasmodium falciparum lactate transporter PfFNT alone and in complex with antimalarial compound MMV007839 reveals its inhibition mechanism. PLoS Biol. 19 , e3001386 (2021).

Hapuarachchi, S. V. et al. The malaria parasite’s lactate transporter PfFNT is the target of antiplasmodial compounds identified in whole cell phenotypic screens. PLoS Pathog. 13 , e1006180 (2017).

Golldack, A. et al. Substrate-analogous inhibitors exert antimalarial action by targeting the Plasmodium lactate transporter PfFNT at nanomolar scale. PLoS Pathog. 13 , e1006172 (2017).

Frydrych, J. et al. Nucleotide analogues containing a pyrrolidine, piperidine or piperazine ring: Synthesis and evaluation of inhibition of plasmodial and human 6-oxopurine phosphoribosyltransferases and in vitro antimalarial activity. Eur. J. Med. Chem. 219 , 113416 (2021).

Wang, X. et al. Identification of Plasmodium falciparum mitochondrial malate: quinone oxidoreductase inhibitors from the pathogen box. Genes https://doi.org/10.3390/genes10060471 (2019).

Sonoiki, E. et al. A potent antimalarial benzoxaborole targets a Plasmodium falciparum cleavage and polyadenylation specificity factor homologue. Nat. Commun. 8 , 14574 (2017).

Yoo, E. et al. The antimalarial natural product salinipostin A identifies essential alpha/beta serine hydrolases involved in lipid metabolism in P. falciparum parasites. Cell Chem. Biol. 27 , 143–157 (2020).

Burrows, J. N. et al. New developments in anti-malarial target candidate and product profiles. Malar. J. 16 , 26 (2017).

Wouters, O. J., McKee, M. & Luyten, J. Estimated research and development investment needed to bring a new medicine to market, 2009-2018. J. Am. Med. Assoc. 323 , 844–853 (2020).

Baird, J. K. 8-Aminoquinoline therapy for latent malaria. Clin. Microbiol. Rev. https://doi.org/10.1128/CMR.00011-19 (2019).

MMV. Keeping the Promise https://www.mmv.org/sites/default/files/uploads/docs/publications/KeepingThePromise-Report_2021.pdf (2021).

de Jong, R. M. et al. Monoclonal antibodies block transmission of genetically diverse Plasmodium falciparum strains to mosquitoes. NPJ Vaccines 6 , 101 (2021).

Rathore, D., Sacci, J. B., de la Vega, P. & McCutchan, T. F. Binding and invasion of liver cells by Plasmodium falciparum sporozoites. Essential involvement of the amino terminus of circumsporozoite protein. J. Biol. Chem. 277 , 7092–7098 (2002).

Casares, S., Brumeanu, T. D. & Richie, T. L. The RTS,S malaria vaccine. Vaccine 28 , 4880–4894 (2010).

Gaudinski, M. R. et al. A monoclonal antibody for malaria prevention. N. Engl. J. Med. 385 , 803–814 (2021).

Kayentao, K. et al. Safety and efficacy of a monoclonal antibody against malaria in Mali. N. Engl. J. Med. 387 , 1833–1842 (2022).

Wang, L. T. et al. A potent anti-malarial human monoclonal antibody targets circumsporozoite protein minor repeats and neutralizes sporozoites in the liver. Immunity 53 , 733–744 (2020).

van der Boor, S. C. et al. Safety, tolerability, and Plasmodium falciparum transmission-reducing activity of monoclonal antibody TB31F: a single-centre, open-label, first-in-human, dose-escalation, phase 1 trial in healthy malaria-naive adults. Lancet Infect. Dis. 22 , 1596–1605 (2022).

Hernandez, I. et al. Pricing of monoclonal antibody therapies: higher if used for cancer? Am. J. Manag. Care 24 , 109–112 (2018).

PubMed   Google Scholar  

Mathews, E. S. & Odom John, A. R. Tackling resistance: emerging antimalarials and new parasite targets in the era of elimination. F1000Res https://doi.org/10.12688/f1000research.14874.1 (2018).

Macintyre, F. et al. Injectable anti-malarials revisited: discovery and development of new agents to protect against malaria. Malar. J. 17 , 402 (2018).

Bakshi, R. P. et al. Long-acting injectable atovaquone nanomedicines for malaria prophylaxis. Nat. Commun. 9 , 315 (2018).

Menze, B. D. et al. Marked aggravation of pyrethroid resistance in major malaria vectors in Malawi between 2014 and 2021 is partly linked with increased expression of P450 alleles. BMC Infect. Dis. 22 , 660 (2022).

Paton, D. G. et al. Exposing Anopheles mosquitoes to antimalarials blocks Plasmodium parasite transmission. Nature 567 , 239–243 (2019). Reported that adding the antimalarial atovaquone to surfaces could lead to trans-dermal drug uptake into infected mosquitoes and block P. falciparum parasite transmission.

Paton, D. G. et al. Using an antimalarial in mosquitoes overcomes Anopheles and Plasmodium resistance to malaria control strategies. PLoS Pathog. 18 , e1010609 (2022).

Adolfi, A. et al. Efficient population modification gene-drive rescue system in the malaria mosquito Anopheles stephensi. Nat. Commun. 11 , 5553 (2020).

Hammond, A. et al. Gene-drive suppression of mosquito populations in large cages as a bridge between lab and field. Nat. Commun. 12 , 4589 (2021).

Burrows, J. et al. A discovery and development roadmap for new endectocidal transmission-blocking agents in malaria. Malar. J. 17 , 462 (2018).

Miglianico, M. et al. Repurposing isoxazoline veterinary drugs for control of vector-borne human diseases. Proc. Natl Acad. Sci. USA 115 , E6920–E6926 (2018).

Khaligh, F. G. et al. Endectocides as a complementary intervention in the malaria control program: a systematic review. Syst. Rev. 10 , 30 (2021).

Llanos-Cuentas, A. et al. Tafenoquine versus primaquine to prevent relapse of Plasmodium vivax malaria. N. Engl. J. Med. 380 , 229–241 (2019).

Lacerda, M. V. G. et al. Single-dose tafenoquine to prevent relapse of Plasmodium vivax malaria. N. Engl. J. Med. 380 , 215–228 (2019).

Camarda, G. et al. Antimalarial activity of primaquine operates via a two-step biochemical relay. Nat. Commun. 10 , 3226 (2019).

Pybus, B. S. et al. The metabolism of primaquine to its active metabolite is dependent on CYP 2D6. Malar. J. 12 , 212 (2013).

Silvino, A. C. R. et al. Novel insights into Plasmodium vivax therapeutic failure: CYP2D6 activity and time of exposure to malaria modulate the risk of recurrence. Antimicrob. Agents Chemother. https://doi.org/10.1128/aac.02056-19 (2020).

Roth, A. et al. A comprehensive model for assessment of liver stage therapies targeting Plasmodium vivax and Plasmodium falciparum . Nat. Commun. 9 , 1837 (2018).

Maher, S. P. et al. Probing the distinct chemosensitivity of Plasmodium vivax liver stage parasites and demonstration of 8-aminoquinoline radical cure activity in vitro. Sci. Rep. 11 , 19905 (2021).

Download references

Acknowledgements

The University of Cape Town, Medicines for Malaria Venture (MMV09_0002, RD-17–0004 and RD-18–0001), Bill & Melinda Gates Foundation (OPP1066878 and INV-040482), South African Medical Research Council (SAMRC), Strategic Health Innovation Partnerships (SHIP) unit of the SAMRC, South African Technology Innovation Agency (TIA), Celgene, Merck KGaA (M3409), National Institutes of Health (NIH, 1R01AI152092–01 and 5R01 AI143521–04), and South African Research Chairs Initiative of the Department of Science and Innovation (DSI), administered through the South African National Research Foundation (NRF), are gratefully acknowledged for support (K.C. and K.W.). K.C. is the Neville Isdell Chair in African-centric Drug Discovery and Development and thanks Neville Isdell for generously funding the Chair. J.L.S.-N. is supported by the Bill & Melinda Gates Foundation (INV-007124) and the NIH (1 R01 AI151639 01). D.A.F. gratefully acknowledges funding from the Medicines for Malaria Venture (RD008/15), the Bill & Melinda Gates Foundation (INV-033538), the US Department of Defense (E01 W81XWH2210520) and the NIH (R01 AI109023, R37 AI050234, R01 AI124678). E.A.W. is supported by grants from the NIH (R01 AI152533) and the Bill & Melinda Gates Foundation (OPP1054480). The authors acknowledge Tim Wells for critically reviewing this manuscript.

Author information

Authors and affiliations.

University of California, San Diego, La Jolla, CA, USA

Jair L. Siqueira-Neto & Elizabeth A. Winzeler

Holistic Drug Discovery and Development (H3D) Centre, University of Cape Town, Rondebosch, South Africa

Kathryn J. Wicht & Kelly Chibale

South African Medical Research Council Drug Discovery and Development Research Unit, Department of Chemistry and Institute of Infectious Disease and Molecular Medicine, University of Cape Town, Rondebosch, South Africa

Medicines for Malaria Venture, Geneva, Switzerland

Jeremy N. Burrows

Department of Microbiology and Immunology and Center for Malaria Therapeutics and Antimicrobial Resistance, Division of Infectious Diseases, Department of Medicine, Columbia University Irving Medical Center, New York, NY, USA

David A. Fidock

You can also search for this author in PubMed   Google Scholar

Contributions

The authors contributed equally to all aspects of the article.

Corresponding author

Correspondence to Elizabeth A. Winzeler .

Ethics declarations

Competing interests.

J.N.B. is employed by the Medicines for Malaria Venture, which has a stake in developing many of the drugs cited in this Review.

Peer review

Peer review information.

Nature Reviews Drug Discovery thanks Paul Gilson, Philip Rosenthal and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Related links

Medicines for Malaria Venture (MMV): https://www.mmv.org/

The Drugs for Neglected Diseases initiative (DNDi): https://dndi.org/advocacy/transparency-rd-costs/

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Cite this article.

Siqueira-Neto, J.L., Wicht, K.J., Chibale, K. et al. Antimalarial drug discovery: progress and approaches. Nat Rev Drug Discov 22 , 807–826 (2023). https://doi.org/10.1038/s41573-023-00772-9

Download citation

Accepted : 17 July 2023

Published : 31 August 2023

Issue Date : October 2023

DOI : https://doi.org/10.1038/s41573-023-00772-9

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

thesis on antimalarial drugs

  • Open access
  • Published: 25 August 2021

Antiplasmodial, antimalarial activities and toxicity of African medicinal plants: a systematic review of literature

  • Elahe Tajbakhsh 1 ,
  • Tebit Emmanuel Kwenti 2 , 3 , 4 ,
  • Parya Kheyri 5 ,
  • Saeed Nezaratizade 5 ,
  • David S. Lindsay 6 &
  • Faham Khamesipour   ORCID: orcid.org/0000-0003-0678-2528 7 , 8  

Malaria Journal volume  20 , Article number:  349 ( 2021 ) Cite this article

12k Accesses

23 Citations

3 Altmetric

Metrics details

Malaria still constitutes a major public health menace, especially in tropical and subtropical countries. Close to half a million people mainly children in Africa, die every year from the disease. With the rising resistance to frontline drugs (artemisinin-based combinations), there is a need to accelerate the discovery and development of newer anti-malarial drugs. A systematic review was conducted to identify the African medicinal plants with significant antiplasmodial and/or anti-malarial activity, toxicity, as wells as assessing the variation in their activity between study designs (in vitro and in vivo).

Key health-related databases including Google Scholar, PubMed, PubMed Central, and Science Direct were searched for relevant literature on the antiplasmodial and anti-malarial activities of African medicinal plants.

In total, 200 research articles were identified, a majority of which were studies conducted in Nigeria. The selected research articles constituted 722 independent experiments evaluating 502 plant species. Of the 722 studies, 81.9%, 12.4%, and 5.5% were in vitro, in vivo , and combined in vitro and in vivo , respectively. The most frequently investigated plant species were Azadirachta indica, Zanthoxylum chalybeum, Picrilima nitida, and Nauclea latifolia meanwhile Fabaceae, Euphorbiaceae, Annonaceae, Rubiaceae, Rutaceae, Meliaceae, and Lamiaceae were the most frequently investigated plant families. Overall, 248 (34.3%), 241 (33.4%), and 233 (32.3%) of the studies reported very good, good, and moderate activity, respectively. Alchornea cordifolia, Flueggea virosa, Cryptolepis sanguinolenta, Zanthoxylum chalybeum, and Maytenus senegalensis gave consistently very good activity across the different studies. In all, only 31 (4.3%) of studies involved pure compounds and these had significantly (p = 0.044) higher antiplasmodial activity relative to crude extracts. Out of the 198 plant species tested for toxicity, 52 (26.3%) demonstrated some degree of toxicity, with toxicity most frequently reported with Azadirachta indica and Vernonia amygdalina . These species were equally the most frequently inactive plants reported. The leaves were the most frequently reported toxic part of plants used. Furthermore, toxicity was observed to decrease with increasing antiplasmodial activity.

Conclusions

Although there are many indigenous plants with considerable antiplasmodial and anti-malarial activity, the progress in the development of new anti-malarial drugs from African medicinal plants is still slothful, with only one clinical trial with Cochlospermum planchonii ( Bixaceae ) conducted to date. There is, therefore, the need to scale up anti-malarial drug discovery in the African region.

Malaria still constitutes a major public health menace, especially in tropical and subtropical countries. Various species of Plasmodium , transmitted through the bite of an infected female Anopheles mosquito, cause malaria, including Plasmodium falciparum, Plasmodium malariae, Plasmodium ovale, Plasmodium vivax, and Plasmodium knowlesi . Among these species, P. falciparum is the most virulent, responsible for the highest morbidity and mortality. It is also the predominant species in sub-Saharan Africa (SSA), a region with the highest number of malaria cases and deaths in the world. According to the World Health Organization (WHO), there were 228 million cases, and 405,000 malaria attributed deaths in 2018 [ 1 ]. In SSA, children and pregnant women are the most at-risk groups [ 1 , 2 , 3 ].

Malaria can be treated using chemotherapy but there is widespread resistance to many of the drugs. The first case of resistance to artemisinins was reported in Cambodia in 2006 and has then spread to most of South-East Asia [ 4 , 5 ]. The safety of chemoprophylaxis is also a major concern; for instance, primaquine, atovaquone, and doxycycline are contraindicated in pregnant women and children [ 6 ]. All these shortcomings necessitate the discovery and production of new drugs to treat malaria.

In the past 50 years, natural compounds including plant products, have played a major role in drug discovery and have provided value to the pharmaceutical industry [ 7 ]. For instance, therapeutics for various infectious diseases, cancer, and other debilitation diseases caused by metabolic disorders have all benefitted from many drug classes that were initially developed based on active compounds from plant sources [ 8 ]. Furthermore, quinine and artemisinin, and their synthetic derivatives which are the mainstay of anti-malarial chemotherapy, were also derived from plant sources. In malaria-endemic areas, especially in Africa, many people rely on herbal medicines as the first line of treatment [ 9 ]. The common reasons for their preference vary from the cost of standard drugs, availability and accessibility, perceived effectiveness, low side effect, and faith in traditional medicines [ 10 ].

Reviews of the antiplasmodial and anti-malarial activities of medicinal plants are needed to drive research into the discovery and production of new anti-malarial drugs. Only a few reviews of the antiplasmodial or anti-malarial activity of medicinal plants have been published in the scientific literature [ 11 , 12 , 13 , 14 , 15 , 16 ]. These reviews focused only on studies with high antiplasmodial or anti-malarial activity and hardly report on their toxicity. The purpose of this study was to review medicinal plants with moderate to very good antiplasmodial and anti-malarial activities, as well as assess the variation in the activities between different methods. Furthermore, the toxicity of plant species is highlighted.

The literature was reviewed in search of scientific articles reporting antiplasmodial activities (IC 50 , ED 50 , LD 50 , and parasite suppression rate) of medicinal plants used in Africa to treat malaria. The current study conforms to the Preferred Reporting Items for Systematic Reviews and Meta-analysis (PRISMA) guidelines [ 17 ].

Search strategy and selection criteria

Relevant articles were searched in health-related electronic databases including PubMed, PubMed Central, Google Scholar, and ScienceDirect using the keywords: Traditional herbs or Medicinal plants or Antiplasmodial activity or Antimalarial activity or Herbal medicine or Plasmodium .

The search was limited to studies published in English or containing at least an abstract written in English until May 2020. The titles and abstracts were subsequently examined by two reviewers, independently (parallel method) to identify articles reporting the antiplasmodial activity of medicinal plants. In the case of any discrepancy in their reports, a third reviewer was brought in to resolve the issue. Relevant papers were equally manually cross-checked to identify further references. The following data were extracted from the selected articles by the reviewers: plant species, plant family, place of collection of plant, parts of the plant used, type of study (whether in vitro , in vivo, or human), the extraction solvent used, IC 50 or ED 50 values, parasite suppression rate, isolated compounds, interaction with known malarial drugs (whether synergistic or antagonistic), and toxicity. Articles that did not report antiplasmodial or anti-malarial activity of medicinal plants as well as review articles were excluded. The entire selection process is presented in Fig.  1 .

figure 1

Flowchart of the selection process for publications included in this review

In this study, antiplasmodial activity pertains to studies performed in vitro using different strains of Plasmodium falciparum , meanwhile, anti-malarial activity is reserved for in vivo studies performed using mice and various parasite models (including Plasmodium berghei, Plasmodium yoelii, and Plasmodium chabaudi ) and reporting parasite suppression rate.

Categorization of antiplasmodial and anti-malarial activities

For in vitro studies, the antiplasmodial activity of an extract was considered very good if IC 50  < 5 µg/ml, good 5 µg/ml ≤ IC 50  < 10 µg/ml, and moderate 10 µg/ml ≤ IC 50  < 20 µg/ml [ 18 ]. For in vivo studies, the anti-malarial activity of an extract is considered very good if the suppression is ≥ 50% at 100 mg/kg body weight/day, good if the suppression is ≥ 50% at 250 mg/kg body weight/day, and moderate if the suppression is ≥ 50% at 500 mg/kg body weight/day [ 18 ]. Antiplasmodial activities of 20 µg/ml and above for in vitro studies and anti-malarial ≥ 50% at > 500 mg/kg body weight/day for in vivo studies, were considered inactive.

Risk of bias in individual studies

The level of risk of bias for the study was likely to be high mainly because of differences in the studies and the methods used to determine the antiplasmodial or anti-malarial activity. The stains of Plasmodium used to assess the antiplasmodial or anti-malarial activity of the medicinal plants equally varied between studies. Furthermore, the extraction solvent, as well as the extraction yield of the plants in the different studies, was not the same, which may have accounted for the variation in the antiplasmodial and anti-malarial activities for the same plants but in the different studies.

The PRISMA flowchart (Fig.  1 ) presents a four-phase study selection process in the present systematic review study. A total of 25,159 titles were identified in the initial search. After the title and abstract screening, 228 full-text articles were retrieved. Of these, a final 200 articles were identified for the review.

For this review, the evaluation of the individual plant species was considered as an independent study, so it is common for one article to have more than one study depending on the number of plant species evaluated. In all, there were 722 independent studies. Five hundred and ninety-on (81.9%) of the independent studies were in vitro (Table 1 ), 90 (12.4%) were in vivo (Table 2 ) and 40 (5.5%) were both in vitro and in vivo (Table 3 ). There was only one human study (clinical trial) conducted so far (Table 4 ). The selected research articles were from 31 African countries. Out of the 200 research articles reviewed, most of them were from Nigeria 58 (29.0%), Kenya 24 (12.0%), Ethiopia 13 (6.5%), Cameroon 12 (6.0%), Ivory Coast 11 (5.5%), D.R. Congo 10 (5.0%), and Burkina Faso 7 (3.5%) (Fig.  2 ). The studies cover the period from 1989 to 2020.

figure 2

Distribution of the research articles on the antiplasmodial activity of indigenous plants according to African countries

Family and species distribution of plants evaluated

From 722 studies, the most frequent plant families studied included Fabaceae 47 (6.5%), Euphorbiaceae 45 (6.2%), Annonaceae 37 (5.1%), Rubiaceae 37 (5.1%), Rutaceae 37 (5.1%), Meliaceae 30 (4.2%), and Lamiaceae 12 (1.7%). Five hundred and two (502) plant species were investigated in this study. Of them, the most investigated were: Azadirachta indica , Zanthoxylum chalybeum , Picrilima nitida , and Nauclea latifolia . The most frequent parts of the plants tested were the leaves, roots, root barkss, stems, and the whole plant. A majority of the studies used the crude extracts of the plants compared to pure compounds (95.7% vs. 4.3%). In descending order, methanol 322 (44.7%), dichloromethane 207 (28.7%), ethanol 103 (14.3%), water 85 (11.7%) and ethyl acetate 62 (8.6%) were the most frequent extraction solvent used.

In vitro and in vivo activities of the plants evaluated

Overall, 248 (34.3%) of the studies reported activity that was very good (IC 50 values < 5 µg/ml or suppression rate of ≥ 50% at 100 mg/kg body weight/day), 241 (33.4%) reported good activity and 233 (32.3%) reported moderate activity. For the in vitro studies, a majority 228 (38.6%) reported very good activity; 206 (34.9%) reported good activity and 187 (31.6%) reported moderate activity. Meanwhile for the in vivo studies, a majority 19 (21.1%) reported moderate activity, 16 (17.8%) reported very good activity and 13 (14.4%) reported good activity. For studies reporting both the in vitro and in vivo activity, a majority of 17 (42.5%) reported only moderate activity, 13 (32.5%) studies reported very good activity and 10 (25.0%) reported good activity. Among the plants with very good activity, only one species demonstrated very good activity both in vitro and in vivo (Table 3 ).

Among the studies, the most frequent plant species demonstrating very good antiplasmodial activity were: Alchornea cordifolia [3/3, 100%], Flueggea virosa [3/3, 100%], Cryptolepis sanguinolenta [¾, 75%], Zanthoxylum chalbeum [4/5, 80%] and Maytenus senegalensis [3/6, 50%]. Plant families with the most active species include Rutaceae [13/25, 52.0%], Apocynaceae [13/26, 50%], Celastraceae [7/15, 46.7%], Annonaceae [17/37, 45.9%], Euphorbiaceae [21/48, 43.8], Combretaceae [7/16, 43.8%], Fabaceae [18/47, 38.3%], Lamiaceae [8/23, 34.8%], Asteraceae [23/69, 33.3%], and Rubiaceae [8/37, 21.6%]. The fractions are derived from the count of studies reporting very good antiplasmodial activity (numerator) divided by the total number of studies that assessed the activity of that plant species (denominator).

Azadirachta indica and Vernonia amygdalina were the most frequently reported inactive species (Additional file 1 : Table S1). Furthermore, Fabaceae, Rubiaceae, Euphorbiaceae , and Asteraceae were the plant families containing the most frequently reported inactive plants. A majority of 95.7% (691/722) of the studies used the crude extract of the plants. The antiplasmodial and/or anti-malarial activity was significantly higher (p = 0.044) in studies using pure compounds compared to those using crude preparations.

Toxicity of plants evaluated for their antiplasmodial and anti-malarial activity

Out of the 198 plants evaluated in toxicity assays, 52 (26.3%) were found to demonstrate some degree of toxicity. The most frequently reported plants with toxicity were Azadirachta indica and Vernonia amygdalina. Plant families harboring the most toxic species were Lamiaceae, Anacardiaceae, Moraceae, Meliaceae, Asteraceae, and Fabaceae . Approximately 33% of the plants tested demonstrated some toxicity in vitro and 26.7% had some degree of toxicity in vivo. Among plants with very good, good, and moderate antiplasmodial activity, 17.8%, 28.3%, and 35.4% had some degree of toxicity, respectively. The leaf was the plant part with the most frequently reported toxicity. Albino mice and Vero E6 cells were the most commonly used assays for the assessment of the toxicity of the plants.

Resistance to the frontline anti-malarial drugs is increasing and is now a global concern. With this rising rate of resistance, there is a need to accelerate research into the discovery and development of new anti-malarial drugs. Unfortunately, from this study, it is evident that the progress into the discovery of a new anti-malarial drug in Africa is slothful. Despite a considerable number of plant species that have demonstrated significant antiplasmodial activity in vitro, fewer plants have been evaluated in vivo and only one clinical trial with Cochlospermum planchonii ( Bixaceae ) has been conducted so far. This reinforces the need for basic and clinical research in the region. Van Wyk [ 213 ] had also arrived at the same conclusion.

This review revealed research articles from 31 African countries. Most of the articles were from Nigeria. This is suggestive that Nigeria is leading the podium in research on anti-malarial drug discovery and development, deservedly so, because she is probably the most affected country in the world. It is noteworthy that South Africa which is generally more technologically advanced than Nigeria had very few (8) articles. The African region is the most affected in the world recording the greatest number of cases and malaria attributed deaths. However, the distribution of malaria in Africa is not even, with sub-Saharan Africa harboring disproportionately the greatest number of cases. This is suggestive that research to identify new anti-malarial drugs may be related to the burden of the disease, thus the government policy to control the disease. There is, therefore, the need for policy-driven research into new anti-malarial all across the African region. In this review, IC 50 values of < 20 µg/ml were considered as the cutoff of significant anti-malarial activity. This cutoff is considered the minimum to qualify as a first-pass “hit” in anti-malarial drugs screening [ 214 ]. Five hundred and two (502) plant species from 169 families were observed to have moderate to very good anti-malarial activity. The most investigated plant families were Euphorbiaceae, Fabaceae, Rubiaceae, and Annonaceae . However, the plant families containing the most active plants were Apocynaceae, Celestraceae, and Rutaceae . This finding suggests that more emphasis should be given to plants in these families for anti-malarial drug discovery. Besides, the most investigated plant species were Azadirachta indica, Nauclea latifolia, Picralima nitida, and Zanthoxylum chalybeum . Alchornea cordifolia, Flueggea virosa, Crytolepis sanguinolenta, and Zanthoxylum chalybeum were the only plant species with consistently very good antiplasmodial and anti-malarial activities between studies. This is very surprising that no clinical trial using any of these plants has been conducted. Further studies on these plant species should be performed.

This study revealed that overall, a majority of the plants investigated had very good antiplasmodial activity in vitro. That activity decreases as you move to in vivo in most studies, with a majority of plants demonstrating only moderate activity. For example, Gathirwa et al. [ 146 ] showed that the activity of Uvaria acuminate decreased from good activity in vitro to inactive in vivo. However, a few studies show that plant activity could also increase from in vitro to in vivo. For example, Ngbolua et al. [ 211 ] showed that the activity of Vernonia ambigua increased from in vitro to in vivo analysis. Other examples include studies by Muthaura et al. [ 20 ] using Boscia angustifolia , Kweyamba et al. [ 162 ] using Commiphora Africana, and Ajaiyeoba et al. [ 204 ] using Annona senegalensis . This suggests that plants could still have significant anti-malarial activity in vivo although they failed to in vitro. Most investigators usually progress to in vivo studies only when they observe significant antiplasmodial activity in vitro. This may explain the findings of a smaller number of in vivo studies in the current study. The investigation of the anti-malarial activities of plants should continue in vivo despite the dismal performance of the plants in vitro.

The current study revealed substantial inter-study variation in the antiplasmodial activity of several plant species. For example, considerable variation in the antiplasmodial activity was observed for Senna occidentalis, Adansonia digitata, Acanthospermum hispidum, Rotheca myricoides, Anogeissus leocarpus, Annona muricata, Ageratum conyzoides, Albizia coriaria, Ekebergia capensis, Flueggea virosa, Lippia javanica, Maytenus senegalensis, Morinda lucida, Picralima nitida, Trichilia emetica, Vernonia amydalina, and Vernonia colorata . The factors that could have accounted for these differences may include differences in the extraction solvent thus the extraction yield and extracted metabolite. With dichloromethane, mainly the apolar metabolites are extracted. In contrast, with methanol, from polar to moderate apolar metabolites are extracted.

Most (95.7%) of the studies used crude extract for their investigation and rarely the pure compounds (Additional file 1 : Table S2 presents a summary of active compounds that have been identified from some of the plants). The finding of a majority of studies in Africa using only the crude extract of plants may be attributed to the absence of the necessary infrastructure to process the plant materials to get the pure compounds. Furthermore, there may be geographical differences in the areas where the plants were collected and this may also affect the activity of the same plant species. For example, despite using the same extraction solvent, the antiplasmodial activity of Acacia nilotica was moderate in South Africa and very good in Sudan. There was also variation between the different assay types. For example, the activities of Vernonia ambigua [ 211 ] and Annona senegalensis [ 204 ] have been reported to increase from inactive in vitro to very good in vivo. However, a few plant species including Alchornea cordifolia, and Zanthoxylum chalybeum , were observed to be consistently very good between studies. These plant species should be exploited further for their antiplasmodial activity. The activities of the plants were equally observed to increase with the isolation of the active compounds thus reinforcing the need for research into identifying the active compounds of African medicinal plants. The marked difference in the antiplasmodial activity of the crude extract of Artemisia annua and the pure compounds points out the issue that even the compounds which show only low potency and may be discarded from the initial screen for further development may still have active components with therapeutic potential [ 215 ]. The strain of the Plasmodium used may also be another factor accounting for the inter-study variation observed; studies using chloroquine-sensitive strains of the parasite like P. falciparum 3D7, D6, NF54 tend to report higher antiplasmodial activity compared to studies using chloroquine-resistant strains like P. falciparum W2, Dd5, K1 or D10.

This study revealed that only a few (26.3%) of the plants demonstrated some degree of toxicity. The families hosting the most toxic plant species were Lamiaceae, Anacardiaceae, Moraceae, and Meliaceae . The most toxic plants were Azadirachta indica and Vernonia amygdalina . The former [ 168 ] is one of the few plant species that demonstrated very good antiplasmodial activity in some studies. Other plants with high toxicity but very good antiplasmodial/anti-malarial activities include Arenga engleri [ 25 ], Celtis integrifolia [ 52 ], Ficus platyhylla [ 50 ], Gutenbergia cordifolia [ 21 ], Helchrysum cymosum [ 97 ], Microglossa pyrifolia [ 92 ], Opilia celtidifolia [ 52 ], Quassia Africana [ 103 ], Rumex abyssinicus [ 92 ], Clausena anisota [ 157 ], Icacina senegalensis [ 171 ], Abutilon grandiflorum [ 200 ], and Lannea schweinfurthii [ 205 ]. The isolation of the active compounds, which has to be done, could eliminate the toxicity, if not all, to a certain degree. For example, Salvia radula crude extract (of aerial parts) has been shown to demonstrate some degree of toxicity, but betulafolientriol oxide isolated from the plant was very active with little or no toxicity against human kidney epithelial cells [ 120 ]. There was also considerable variation in the toxicity between the assay types (in vitro or in vivo). As many as 32.8% of the plants demonstrated some level of toxicity in vitro meanwhile 26.7% were toxic in vivo. Since it is customary to evaluate toxicity at the in vitro level and toxic plants are discarded before in vivo evaluation, that may explain while fewer plants were toxic in vivo. Toxicity varied within the same plant species from study to study and could be attributed to differences in the study design as well as differences in the parts of the plants used for testing. From this study, the most toxicity was observed with the leaves. Also, a relationship could be established between toxicity and antiplasmodial activity; as the activity of the plant increases, the toxicity, on the other hand, was observed to decrease. Furthermore, albino mice and Vero E6 cells were the most commonly used assays in the evaluation of toxicity. Unfortunately, the authors could nt make a meaningful relationship between the type of assay and toxicity because of the fewer studies assessing the toxicity of the medicinal plants.

This study, however, is limited in that the analyses may have been compounded by the substantial inter-study variation in the methodologies used by different independent studies for the extraction of plant material, the overall extraction yield, the diversity of extracted metabolites as well as the geographical variations in the different sites used in the plant collection. However, the study has provided important baseline data that may be exploited by researchers in the field for the discovery and development of new anti-malarial drugs.

This study has revealed the slothful progress in the discovery and development of new anti-malarial drugs from African medicinal plants. Despite the encouraging activities demonstrated by the plants in vitro, fewer plants have been evaluated in vivo and just one clinical trial has been conducted so far with Cochlospermum planchonii ( Bixaceae ). The study also revealed considerable inter-study variation in the antiplasmodial activities of the plants, however, the activity of some plants including Alchornea cordifolia, Azadirachta indica, and Zanthoxylum chalybeum was consistently very good. The study demonstrates a relationship between antiplasmodial activity and toxicity whereby the toxicity of the plants decreases as the antiplasmodial activity increases. Besides, the active compounds were identified in just a handful of the plants. Therefore, there is a need for a policy-driven approach in the discovery and development of new anti-malarial drugs to subvert the rising resistance to the frontline anti-malarial drugs in the world.

Availability of data and materials

Not applicable.

Abbreviations

Preferred Reporting Items for Systematic Reviews and Meta-analysis

Not specified

Selectivity Index

Median lethal dose

Half-maximal inhibitory concentration

50% Cytotoxic concentration

Lethal concentration

WHO. World malaria report 2019. Geneva: World Health Organization; 2019. Accessed on 28/06/2021 at https://www.who.int/publications-detail/world-malaria-report-2019 .

Kwenti ET. Malaria and HIV coinfection in sub-Saharan Africa: prevalence, impact, and treatment strategies. Res Rep Trop Med. 2018;9:123–36.

PubMed   PubMed Central   Google Scholar  

Kwenti ET, Kukwah TA, Kwenti TDB, Nyassa BR, Dilonga MH, Enow-Orock G, et al. Comparative analysis of IgG and IgG subclasses against Plasmodium falciparum MSP-1 19 in children from five contrasting bioecological zones of Cameroon. Malar J. 2019;18:16.

Article   PubMed   PubMed Central   Google Scholar  

Dondorp AM, Nosten F, Yi P, Das D, Phyo AP, Tarning J, et al. Artemisinin resistance in Plasmodium falciparum malaria. N Engl J Med. 2009;361:455–67.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Murray CJ, Rosenfeld LC, Lim SS, Andrews KG, Foreman KJ, Haring D, et al. Global malaria mortality between 1980 and 2010: a systematic analysis. Lancet. 2012;379:413–31.

Article   PubMed   Google Scholar  

Nagendrappa PB, Annamalai P, Naik M, Mahajan V, Mathur A, Susanta G, et al. A prospective comparative field study to evaluate the efficacy of a traditional plant-based malaria prophylaxis. J Intercult Ethnopharmacol. 2017;6:36–41.

Newman DJ, Cragg GM. Natural products as sources of new drugs over the 30 years from 1981 to 2010. J Nat Prod. 2012;75:311–35.

Cragg GM, Grothaus PG, Newman DJ. Impact of natural products on developing new anti-cancer agents. Chem Rev. 2009;109:3012–43.

Article   CAS   PubMed   Google Scholar  

Willcox ML. A clinical trial of “AM”, a Ugandan herbal remedy for malaria. J Public Health Med. 1999;21:318–24.

Suswardany DL, Sibbritt DW, Supardi S, Pardosi JF, Chang S, Adams J. A cross-sectional analysis of traditional medicine use for malaria alongside free antimalarial drugs treatment amongst adults in high-risk malaria endemic provinces of Indonesia. PLoS One. 2017;12:e0173522.

Article   PubMed   PubMed Central   CAS   Google Scholar  

Ibrahima HA, Imama IA, Bellob AM, Umara U, Muhammada S, Abdullahia SA. The potential of Nigerian medicinal plants as antimalarial agent: a review. Int J Sci Technol. 2012;2:600–5.

Google Scholar  

Zofou D, Kuete V, Titanji VPK. Antimalarial and other antiprotozoal products from African Medicinal plants. In: Medicinal plant research in Africa: pharmacology and chemistry. Kuete V, Ed. Chapt. 17. Amsterdam, Elsevier, 2013;661–709.

Lawal B, Shittu OK, Kabiru AY, Jigam AA, Umar MB, Berinyuy EB, et al. Potential antimalarials from African natural products: a review. J Intercult Ethnopharmacol. 2015;4:318–43.

Van Wyk BE. A review of commercially important African medicinal plants. J Ethnopharmacol. 2015;176:118–34.

Kaur R, Kaur H. Plant derived antimalarial agents. J Med Plants Studies. 2017;5:346–63.

Lemma MT, Ahmed AM, Elhady MT, Ngo HT, Vu TL, Sang TK, et al. Medicinal plants for in vitro antiplasmodial activities: a systematic review of literature. Parasitol Int. 2017;66:713–20.

Moher D, Liberati A, Tetzlaff J, Altman DG. Preferred Reporting Items for Systematic Reviews and Meta-Analyses: the PRISMA statement. PLoS Med. 2009;6:1000097.

Article   Google Scholar  

Deharo E, Bourdy G, Quenevo C, Munoz V, Ruiz G, Sauvain M. A search for natural bioactive compounds in Bolivia through multidisciplinary approach. Part V. Evaluation of the antimalarial activity of plants used by the Tacana Indians. J Ethnopharmacol. 2001;77:91–8.

Becker JVW, van der Merwe MM, van Brummelen AC, Pillay P, Crampton BG, Mmutlane EM, et al. In vitro anti-plasmodial activity of Dicoma anomala subsp gerrardii (Asteraceae): identification of its main active constituent, structure-activity relationship studies, and gene expression profiling. Malar J. 2011;10:295.

Gessler MC, Nkunya MH, Mwasumbi LB, Heinrich M, Tanner M. Screening Tanzanian medicinal plants for antimalarial activity. Acta Trop. 1994;56:65–77.

Koch A, Tamez P, Pezzuto J, Soejarto D. Evaluation of plants used for antimalarial treatment by the Maasai of Kenya. J Ethnopharmacol. 2005;101:95–9.

Clarkson C, Maharaj VJ, Crouch NR, Grace OM, Pillay P, Matsabisa MG, et al. In vitro antiplasmodial activity of medicinal plants native to or naturalised in South Africa. J Ethnopharmacol. 2004;92:177–91.

El Tahir A, Satti GM, Khalid SA. Antiplasmodial activity of selected Sudanese medicinal plants with emphasis on Maytenus senegalensis (Lam.) Exell. J Ethnopharmacol. 1999;64:227–33.

Muthaura CN, Keriko JM, Mutai C, Yenesew A, Gathirwa JW, Irungu BN, et al. Antiplasmodial potential of traditional phytotherapy of some remedies used in treatment of malaria in Meru-Tharaka Nithi County of Kenya. J Ethnopharmacol. 2015;175:315–23.

Prozesky EA, Meyer JJ, Louw AI. In vitro antiplasmodial activity and cytotoxicity of ethnobotanically selected South African plants. J Ethnopharmacol. 2001;76:239–45.

Muthaura CN, Keriko JM, Mutai C, Yenesew A, Gathirwa JW, Irungu BN, et al. Antiplasmodial potential of traditional antimalarial phytotherapy remedies used by the Kwale community of the Kenyan Coast. J Ethnopharmacol. 2015;170:148–57.

Sanon S, Ollivier E, Azas N, Mahiou V, Gasquet M, Ouattara CT, et al. Ethnobotanical survey and in vitro antiplasmodial activity of plants used in traditional medicine in Burkina Faso. J Ethnopharmacol. 2003;86:143–7.

Zirihi-Guédé N, Mambu L, Guédé-Guina F, Bodo B, Grellier P. In vitro antiplasmodial activity and cytotoxicity of 33 West African plants used for treatment of malaria. J Ethnopharmacol. 2005;98:281–5.

Koukouikila-Koussounda F, Abena AA, Nzoungani A, Mombouli JV, Ouamba JM, Kun J, et al. In vitro evaluation of antiplasmodial activity of extracts of Acanthospermum hispidum DC ( Asteraceae ) and Ficus thonningii Blume ( Moraceae ), two plants used in traditional medicine in the Republic of Congo. Afr J Tradit Complement Altern Med. 2012;10:270–6.

Owuor BO, Ochanda JO, Kokwaro JO, Cheruiyot AC, Yeda RA, Okudo CA, et al. In vitro antiplasmodial activity of selected Luo and Kuria medicinal plants. J Ethnopharmacol. 2012;144:779–81.

Malebo HM, Tanja W, Cal M, Swaleh SAM, Omolo MO, Hassanali A, et al. Antiplasmodial, anti-trypanosomal, anti-leishmanial and cytotoxicity activity of selected Tanzanian medicinal plants. Tanzan J Health Res. 2009;11:226–34.

CAS   PubMed   Google Scholar  

Annan K, Sarpong K, Asare C, Dickson R, Amponsah KI, Gyan B, et al. In vitro anti-plasmodial activity of three herbal remedies for malaria in Ghana: Adenia cissampeloides (Planch.) Harms., Termina liaivorensis A. Chev, and Elaeis guineensis Jacq. Pharmacognosy Res. 2012;4:225–9.

Lekana-Douki JB, Liabagui SLO, Bongui JB, Zatra R, Lebibi J, Toure-Ndouo FS. In vitro antiplasmodial activity of crude extracts of Tetrapleura tetraptera and Copaifera religiosa . BMC Res Notes. 2011;4:506.

Ahmed EHM, Nour BYM, Mohammed YG, Khalid HS. Antiplasmodial activity of some medicinal plants used in Sudanese fFolk-medicine. Environ Health Insights. 2010;4:1–6.

Article   PubMed Central   Google Scholar  

Kuria KA, Chepkwony H, Govaerts C, Roets E, Busson R, De Witte P, et al. The antiplasmodial activity of isolates from Ajuga remota . J Nat Prod. 2002;65:789–93.

Lasisi AA, Olayiwola MA, Balogun SA, Akinloye OA, Ojo DA. Phytochemical composition, cytotoxicity and in vitro antiplasmodial activity of fractions from Alafia barteri olive (Hook F. Icon)-Apocynaceae. J Saudi Chem Soc. 2016;20:2–6.

Article   CAS   Google Scholar  

Bapela MJ, Meyer JJ, Kaiser M. In vitro antiplasmodial screening of ethnopharmacologically selected South African plant species used for the treatment of malaria. J Ethnopharmacol. 2014;156:370–3.

Banzouzi JT, Prado R, Menan H, Valentin A, Roumestan C, Mallie M, et al. In vitro antiplasmodial activity of extracts of Alchornea cordifolia and identification of an active constituent: ellagic acid. J Ethnopharmacol. 2002;81:399–401.

Mustofa, Valentin A, Benoit-Vical F, Pélissier Y, Koné-Bamba D, Mallié M. Antiplasmodial activity of plant extracts used in West African traditional medicine. J Ethnopharmacol. 2000;73:145–51.

Musuyu Muganza D, Fruth BI, Nzunzu Lami J, Mesia GK, Kambu OK, Tona GL, et al. In vitro antiprotozoal and cytotoxic activity of 33 ethonopharmacologically selected medicinal plants from Democratic Republic of Congo. J Ethnopharmacol. 2012;141:301–8.

Abdissa D, Geleta G, Bacha K, Abdissa N. Phytochemical investigation of Aloe pulcherrima roots and evaluation for its antibacterial and antiplasmodial activities. PLoS ONE. 2017;12:e0173882.

Iyiola OA, Tijani AY, Lateef KM. Antimalarial activity of ethanolic stem barks extract of Alstonia boonei in mice. Asian J Biol Sci. 2011;4:235–43.

Zirihi-Guédé N, N’guessan K, Etien Dibié T, Grellier P. Ethnopharmacological study of plants used to treat malaria, in traditional medicine, by Bete populations of Issia (Côte d’Ivoire). J Pharm Sci Res. 2010;2:216–27.

Lumpu SL, Kikueta CM, Tshodi ME, Mbenza AP, Kambu OK, Mbamu BM, et al. Antiprotozoal screening and cytotoxicity of extracts and fractions from the leaves, stem barks and root barks of Alstonia congensis . J Ethnopharmacol. 2013;148:724–7.

Hout S, Chea A, Bun SS, Elias R, Gasquet M, Timon-David P, et al. Screening of selected indigenous plants of Cambodia for antiplasmodial activity. J Ethnopharmacol. 2006;107:12–8.

Lusakibanza M, Mesia G, Tona G, Karemere S, Lukuka A, Tits M, et al. In vitro and in vivo antimalarial and cytotoxic activity of five plants used in Congolese traditional medicine. J Ethnopharmacol. 2010;129:398–402.

Yamthe LRT, Fokou PVT, Mbouna CDJ, Keumoe R, Ndjakou BL, Djouonzo PT, et al. Extracts from Annona muricata L. and Annona reticulata L. (Annonaceae) potently and selectively inhibit Plasmodium falciparum . Medicines (Basel). 2015;2:55–66.

Ménan H, Banzouzi J-T, Hocquette A, Pélissier Y, Blache Y, Koné M, et al. Antiplasmodial activity and cytotoxicity of plants used in West African traditional medicine for the treatment of malaria. J Ethnopharmacol. 2006;105:131–6.

Boyom FF, Fokou PV, Yamthe LR, Mfopa AN, Kemgne EM, Mbacham WF, et al. Potent antiplasmodial extracts from Cameroonian Annonaceae. J Ethnopharmacol. 2011;134:717–24.

Shuaibu MN, Wuyep PA, Yanagi T, Hirayama K, Tanaka T, Kouno I. The use of microfluorometric method for activity-guided isolation of antiplasmodial compound from plant extracts. Parasitol Res. 2008;102:1119–27.

Vonthron-Sénécheau C, Weniger B, Ouattara M, Bi FT, Kamenan A, Lobstein A, et al. In vitro antiplasmodial activity and cytotoxicity of ethnobotanically selected Ivorian plants. J Ethnopharmacol. 2003;87:221–5.

Sanon S, Gansane A, Ouattara LP, Traore A, OueDRaogo IN, Tiono A, et al. In vitro antiplasmodial and cytotoxic properties of some medicinal plants from western Burkina Faso. Afr J Lab Med. 2013;2:81.

Chukwujekwu JC, van Staden J, Smith P, Meyer JJM. Antibacterial, anti-inflammatory, and antimalarial activities of some Nigerian medicinal plants. S Afr J Bot. 2005;71:316–25.

Kraft C, Jenett-Siems K, Siems K, Jakupovic J, Mavi S, Bienzle U, et al. In vitro antiplasmodial evaluation of medicinal plants from Zimbabwe. Phytother Res. 2003;17:123–8.

Boyom FF, Kemgne EM, Tepongning R, Ngouana V, Mbacham WF, Tsamo E, et al. Antiplasmodial activity of extracts from seven medicinal plants used in malaria treatment in Cameroon. J Ethnopharmacol. 2009;123:483–8.

Waako PJ, Katuura E, Smith P, Folb P. East African medicinal plants as a source of lead compounds for the development of new antimalarial drugs. Afr J Ecol. 2007;45:102–6.

Benoit F, Valentin A, Pelissier Y, Diafouka F, Marion C, Kone-Bamba D, et al. In vitro antimalarial activity of vegetal extracts used in West African traditional medicine. Am J Trop Med Hyg. 1996;54:67–71.

El-Tahir A, Satti GM, Khalid SA. Antiplasmodial activity of selected Sudanese medicinal plants with emphasis on Acacia nilotica . Phytother Res. 1999;13:474–8.

MacKinnon S1, Durst T, Arnason JT, Angerhofer C, Pezzuto J, SanchezVindas PE, et al. Antimalarial activity of tropical Meliaceae extracts and Gedunin derivatives. J Nat Prod. 1997;60:336–41.

Connelly MP, Fabiano E, Patel IH, Kinyanjui SM, Mberu EK, Watkins WM. Antimalarial activity in crude extracts of Malawian medicinal plants. Ann Trop Med Parasitol. 1996;90:597–602.

Ngwira KJ, Maharaj VJ, Mgani QA. In vitro antiplasmodial and HIV-1 neutralization activities of root and leaf extracts from Berberis holstii . J Herb Med. 2015;5:30–5.

Jansen O, Angenot L, Tits M, Nicolas JP, De Mol P, Nikiéma JB, et al. Evaluation of 13 selected medicinal plants from Burkina Faso for their antiplasmodial properties. J Ethnopharmacol. 2010;130:143–50.

Benoit-Vical F, Soh PN, Saléry M, Harguem L, Poupat C, Nongonierma R. Evaluation of Senegalese plants used in malaria treatment: focus on Chrozophora senegalensis . J Ethnopharmacol. 2008;116:43–8.

Ogunlanaa OO, Kimb H-S, Watayab Y, Olagunjuc JO, Akindahunsid AA, Tan NH. Antiplasmodial flavonoid from young twigs and leaves of Caesalpinia bonduc (Linn) Roxb. J Chem Pharm Res. 2015;7:931–7.

Weniger B, Lagnika L, Vonthron-Sénécheau C, Adjobimey T, Gbenou J, Moudachirou M, et al. Evaluation of ethnobotanically selected Benin medicinal plants for their in vitro antiplasmodial activity. J Ethnopharmacol. 2004;90:279–84.

Melariri P, Campbell W, Etusim P, Smith P. Antiplasmodial properties and bioassay-guided fractionation of ethyl acetate extracts from Carica papaya leaves. J Parasitol Res. 2011;2011:104954.

Kayembe JS, Taba KM, Ntumba K, Tshiongo MTC, Kazadi TK. In vitro antimalarial activity of 20 quinones isolated from four plants used by traditional healers in the Democratic Republic of Congo. J Med Plant Res. 2010;4:991–4.

CAS   Google Scholar  

Ramalhete C, Lopes D, Mulhovo S, Rosário VE, Ferreira MJU. Antimalarial activity of some plants traditionally used in Mozambique. Workshop Plantas Medicinais e Fitoterapêuticas nos Trópicos.IICT /CCCM, 29, 30 e 31 de Outubro de 2008.

Tona L, Cimanga RK, Mesia K, Musuamba CT, De Bruyne T, Apers S, et al. In vitro antiplasmodial activity of extracts and fractions from seven medicinal plants used in the Democratic Republic of Congo. J Ethnopharmacol. 2004;93:27–32.

Gbeassor M, Kossou Y, Amegbo K, de Souza C, Koumaglo K, Denke A. Antimalarial effects of eight African medicinal plants. J Ethnopharmacol. 1989;25:115–8.

Afoulous S, Ferhout H, Raoelison EG, Valentin A, Moukarzel B, Couderc F, et al. Chemical composition and anticancer, antiinflammatory, antioxidant and antimalarial activities of leaf essential oil of Cedrelopsis grevei . Food Chem Toxicol. 2013;56:352–62.

Irungu BN, Rukunga GM, Mungai GM, Muthaura CN. In vitro antiplasmodial and cytotoxicity activities of 14 medicinal plants from Kenya. S Afr J Bot. 2007;73:204–7.

do Céu de Madureira M, Paula Martins A, Gomes M, Paiva J, Proença da Cunha A, do Rosário V. Antimalarial activity of medicinal plants used in traditional medicine in S. Tomé and Prı́ncipe islands. J Ethnopharmacol. 2002;81:23–9.

Rukunga GM, Gathirwa JW, Omar SA, Muregi FW, Muthaura CN, Kirira PG, et al. Anti-plasmodial activity of the extracts of some Kenyan medicinal plants. J Ethnopharmacol. 2009;121:282–5.

Lacroix D, Prado S, Kamoga D, Kasenene J, Namukobe J, Krief S, et al. Antiplasmodial and cytotoxic activities of medicinal plants traditionally used in the village of Kiohima Uganda. J Ethnopharmacol. 2011;133:850–5.

Muregi FW, Chhabra SC, Njagi EN, Lang’at-Thoruwa CC, Njue WM, Orago AS, et al. Anti-plasmodial activity of some Kenyan medicinal plant extracts singly and in combination with chloroquine. Phytother Res. 2004;18:379–84.

Adia MM, Emami SN, Byamukama R, Faye I, Borg-Karlson AK. Antiplasmodial activity and phytochemical analysis of extracts from selected Ugandan medicinal plants. J Ethnopharmacol. 2016;186:14–9.

Lamien-Meda A, Kiendrebeogo M, Compaoré M, Meda RN, Bacher M, Koenig K, et al. Quality assessment and antiplasmodial activity of West African Cochlospermum species . Phytochemistry. 2015;119:51–61.

Benoit F, Valentin A, Pélissier Y, Marion C, Dakuyo Z, Mallié M, et al. Antimalarial activity in vitro of Cochlospermum tinctorium tubercle extracts. Trans R Soc Trop Med Hyg. 1995;89:217–8.

Zofou D, Kengne ABO, Tene M. Ngemenya MN, Tane P, Titanji VP. In vitro antiplasmodial activity and cytotoxicity of crude extracts and compounds from the stem barks of Kigelia africana (Lam.) Benth ( Bignoniaceae ). Parasitol Res. 2011;108:1383–90.

Paulo A, Gomes ET, Houghton PJ. New alkaloids from Cryptolepis sanguinolenta . J Nat Prod. 1995;58:1485–91.

Kirby GC, Paine A, Warhurst DC, Noamese BK, Phillipson JD. In vitro and in vivo antimalarial activity of cryptolepine, a plant-derived indoloquinoline. Phytother Res. 1995;9:359–63.

Cimanga K, De Bruyne T, Pieters L, Vlietinck AJ, Turger CA. In vitro and in vivo antiplasmodial activity of cryptolepine and related alkaloids from Cryptolepis sanguinea . J Nat Prod. 1997;60:688–91.

Grellier P, Ramiaramanana L, Millerioux V, Deharo E, Schrével J, Frappier F, et al. Antimalarial activity of cryptolepine and isocryptolepine, alkaloids isolated from Cryptolepis sanguinolenta . Phytother Res. 1996;10:317–21.

Zofou D, Tematio EL, Ntie-Kang F, Tene M, Ngemenya MN, Tane P, et al. New antimalarial hits from Dacryodes edulis (Burseraceae) - Part I: Isolation, in vitro activity, in silico “drug-likeness” and pharmacokinetic profiles. PLoS ONE. 2013;8:e79544.

Nafuka SN, Mumbengegwi DR. Phytochemical analysis and in vitro anti-plasmodial activity of selected ethnomedicinal plants used to treat malaria associated symptoms in Northern Namibia. Int Sci Technol J Namibia. 2013;2:78–93.

Jansen O, Tits M, Nicolas ALJP, De Mol P, Nikiema J-B, Frédérich M. Anti-plasmodial activity of Dicoma tomentosa ( Asteraceae ) and identification of urospermal A-15-O-acetate as the main active compound. Malar J. 2012;11:289.

Olasehinde GI, Ojurongbe O, Adeyeba AO, Fagade OE, Valecha N, Ayanda IO, et al. In vitro studies on the sensitivity pattern of Plasmodium falciparum to antimalarial drugs and local herbal extracts. Malar J. 2014;13:63.

Bickii J, Tchouya GRF, Tchouankeu JC, Tsamo E. Antimalarial activity in crude extracts of some Cameroonian medicinal plants. Afr J Tradit Complement Altern Med. 2007;4:107–11.

Liu Y, Murakami N, Ji H, Abreu P, Zhang S. Antimalarial flavonol glycosides from Euphorbiahirta. Pharm Biol. 2007;45:278–81.

Kaou AM, Mahiou-Leddet V, Hutter S, Aïnouddine S, Hassani S, Yahaya I, et al. Antimalarial activity of crude extracts from nine African medicinal plants. J Ethnopharmacol. 2008;116:74–83.

Muganga R, Angenot L, Tits M, Frédérich M. Antiplasmodial and cytotoxic activities of Rwandan medicinal plants used in the treatment of malaria. J Ethnopharmacol. 2010;128:52–7.

Traore-Keita F, Gasquet M, Di Giorgio C, Ollivier E, Delmas F, Keita A, et al. Antimalarial activity of four plants used in traditional medicine in Mali. Phytother Res. 2000;14:45–7.

Ancolio C, Azas N, Mahiou V, Ollivier E, Di Giorgio C, Keita A, et al. Antimalarial activity of extracts and alkaloids isolated from six plants used in traditional medicine in Mali and Sao Tome. Phytother Res. 2002;16:646–9.

Nyambati GK, Lagat ZO, Maranga RO, Samuel M, Ozwara H. In vitro anti-plasmodial activity of Rubia cordifolia , Harrizonia abyssinica , Leucas calostachys Olive and Sanchus schweinfurthii medicinal plants. J Appl Pharm Sci. 2013;3:57–62.

Afoulous S, Ferhout H, Raoelison EG, Valentin A, Moukarzel B, Couderc F, et al. Helichrysum gymnocephalum essential oil: chemical composition and cytotoxic, antimalarial and antioxidant activities, attribution of the activity origin by correlations. Molecules. 2011;16:8273–91.

Van Vuuren SF, Viljoen AM, Van Zyl RL, Van Heerden FR, Baser KHC. The antimicrobial, antimalarial and toxicity profiles of helihumulone, Leaf essential oil and extracts of Helichrysum cymosum (L.) D. Don subsp. cymosum . S Afr J Bot. 2006;72:287–90.

Boyom FF. Composition and anti-plasmodial activities of essential oils from some Cameroonian medicinal plants. Phytochemistry. 2003;64:1269–75.

Fotie J, Bohle DS, Leimanis ML, Georges E, Rukunga G, Nkengfack AE. Lupeol long-chain fatty acid esters with antimalarial activity from Holarrhena floribunda . J Nat Prod. 2006;69:62–7.

Sarr SO, Perrotey S, Fall I, Ennahar S, Zhao M, Diop YM, et al. Icacina senegalensis ( Icacinaceae ), traditionally used for the treatment of malaria, inhibits in vitro Plasmodium falciparum growth without host cell toxicity. Malar J. 2011;10:85.

Bickii J, Njifutie N, Foyere JA, Basco LK, Ringwald P. In vitro antimalarial activity of limonoids from Khaya grandifoliola C.D.C. (Meliaceae). J Ethnopharmacol. 2000;69:27–33.

Wube AA, Bucar F, Asres K, Gibbons S, Rattray L, Croft SL. Antimalarial compounds from Kniphofia foliosa roots. Phytother Res. 2005;19:472–6.

Mbatchi SF, Mbatchi B, Banzouzi JT, Bansimba T, Nsonde Ntandou GF, Ouamba JM, et al. In vitro antiplasmodial activity of 18 plants used in Congo Brazzaville traditional medicine. J Ethnopharmacol. 2006;104:168–74.

Oketch-Rabah HA, Dossaji SF, Mberu EK. Antimalarial activity of some Kenyan medicinal plants. Pharm Biol. 1999;37:329–34.

Ramalhete C, da Cruz FP, Mulhovo S, Sousa IJ, Fernandes MX, Prudêncio M, et al. Dual-stage triterpenoids from an African medicinal plant targeting the malaria parasite. Bioorg Med Chem. 2014;22:3887–90.

BenoitVical F, Valentin A, Cournac V, Pélissier Y, Mallié M, Bastide JM. In vitro antiplasmodial activity of stem and root extracts of Nauclea latifolia S.M. (Rubiaceae). J Ethnopharmacol. 1998;61:173–8.

Mesia K, Cimanga RK, Dhooghe L, Cos P, Apers S, Totté J, et al. Antimalarial activity and toxicity evaluation of a quantified Nauclea pobeguinii extract. J Ethnopharmacol. 2010;131:6.

Gbeassor M, Kedjagni AY, Koumaglo K, de Soma C, Agbo K, Aklikokou K, et al. In vitro antimalarial activity of six medicinal plants. Phytother Res. 1990;4:115–7.

Karim T, Béourou S, Touré AO, Ouattara K, Meité S, Ako A, et al. Antioxidant activities and estimation of the phenols and flavonoids content in the extracts of medicinal plants used to treat malaria in Ivory Coast. Int J Curr Microbiol App Sci. 2015;4:862–74.

Koudouvo K, Karou SD, Ilboudo DP, Kokou K, Essien K, Aklikokou K, et al. In vitro antiplasmodial activity of crude extracts from Togolese medicinal plants. Asian Pac J Trop Med. 2011;4:129–32.

Appiah-Opong R, Nyarko AK, Dodoo D, Gyang FN, Koram KA, Ayisi NK. Antiplasmodial activity of extracts of Tridax procumbens and Phyllanthus amarus in in vitro Plasmodium falciparum culture systems. Ghana Med J. 2011;45:143–50.

CAS   PubMed   PubMed Central   Google Scholar  

Komlaga G, Cojean S, Dickson RA, Mehdi A, Beniddir MA, SuyyaghAlbouz S, et al. Antiplasmodial activity of selected medicinal plants used to treat malaria in Ghana. Parasitol Res. 2016;115:3185–95.

Falodun A, Imieje V, Erharuyi O, Ahomafor J, Jacob MR, Khan SI, et al. Evaluation of three medicinal plant extracts against Plasmodium falciparum and selected microganisms. Afr J Tradit Complement Altern Med. 2014;11:142–6.

François G, Aké Assi L, Holenz J, Bringmann G. Constituents of Picralima nitida display pronounced inhibitory activities against asexual erythrocytic forms of Plasmodium falciparum in vitro. J Ethnopharmacol. 1996;54:113–7.

Gbedema SY, Bayor MT, Annan K, Wright CW. Clerodane diterpenes from Polyalthia longifolia (sonn.) Thw. Var. Pendula: potential antimalarial agents for drug resistant Plasmodium falciparum infection. J Ethnopharmacol. 2015;169:176–82.

Annan K, Ekuadzi E, Asare C, Sarpong K, Pistorius D, Oberer L, et al. Antiplasmodial constituents from the stem barks of Polyalthia longifolia var pendula. Phytochem Lett. 2015;11:28–31.

Orwa JA, Ngeny L, Mwikwabe NM, Ondicho J, Jondiko IJ. Antimalarial and safety evaluation of extracts from Toddalia asiatica (L) Lam. (Rutaceae). J Ethnopharmacol. 2013;145:587–90.

Karou D, Dicko MH, Sanon S, Simpore J, Traore AS. Antimalarial activity of Sida acuta Burm. F. (Malvaceae) and Pterocarpus erinaceus Poir. ( Fabaceae ). J Ethnopharmacol. 2003;89:291–4.

Muregi FW, Chhabra SC, Njagi EN, Lang’at-Thoruwa CC, Njue WM, Orago AS, et al. In vitro antiplasmodial activity of some plants used in Kisii, Kenya against malaria and their chloroquine potentiation effects. J Ethnopharmacol. 2003;84:235–9.

Kamatou GPP, Van Zyl RL, Davids H, Van Heerden FR, Lourens ACU, Viljoen AM. Antimalarial and anticancer activities of selected South African Salvia species and isolated compounds from S. radula . S Afr J Bot. 2008;74:238–43.

Bah S, Jäger AK, Adsersen A, Diallo D, Paulsen BS. Antiplasmodial and GABA(A)-benzodiazepine receptor binding activities of five plants used in traditional medicine in Mali. West Africa J Ethnopharmacol. 2007;110:451–7.

Carraz M, Jossang A, Franetich J-F, Siau A, Ciceron L, Hannoun L, et al. A plant-derived morphinan as a novel lead compound active against malaria liver stages. PLoS Med. 2006;3:e513.

Niass O, Sarr SO, Dieye B, Diop A, Diop YM. In vitro assessment of the antiplasmodial activity of three extracts used in local traditional medicine in Saloum (Senegal). Eur Sci J. 2016;12:157–65.

Kigondu EV, Rukunga GM, Keriko JM, Tonui WK, Gathirwa JW, Kirira PG, et al. Anti-parasitic activity and cytotoxicity of selected medicinal plants from Kenya. J Ethnopharmacol. 2009;123:504–9.

Gakunju DM, Mberu EK, Dossaji SF, Gray AI, Waigh RD, Waterman PG, et al. Potent antimalarial activity of the alkaloid nitidine, isolated from a Kenyan herbal remedy. Antimicrob Agents Chemother. 1995;39:2606–9.

Omoregie ES, Pal A, Sisodia B. In vitro antimalarial and cytotoxic activities of leaf extracts of Vernonia amygdalina (Del.). Niger J Basic Appl Sci. 2011;19:121–6.

Shaa KK, Oguche S, Watila IM, Ikpa TF. In vitro antimalarial activity of the extracts of Vernonia amygdalina commonly used in traditional medicine in Nigeria. Sci World J. 2011;6:5–9.

Toyang NJ, Krause MA, Fairhurst RM, Tane P, Bryant J, Verpoorte R. Antiplasmodial activity of sesquiterpene lactones and a sucrose ester from Vernonia guineensis Benth (Asteraceae). J Ethnopharmacol. 2013;147:618–21.

Goodman CD, Austarheim I, Mollard V, Mikolo B, Malterud KE, McFadden GI, et al. Natural products from Zanthoxylum heitzii with potent activity against the malaria parasite. Malar J. 2016;15:481.

Randrianarivelojosia M, Rasidimanana VT, Rabarison H, Cheplogoi PK, Ratsimbason M, Mulholland DA, et al. Plants traditionally prescribed to treat tazo (malaria) in the eastern region of Madagascar. Malar J. 2003;2:25.

Okokon JE, Augustine NB, Mohanakrishnan D. Antimalarial, antiplasmodial and analgesic activities of root extract of Alchornea laxiflora . Pharm Biol. 2017;55:1022–31.

Alli LA, Adesokan AA, Salawu OA, Akanji MA, Tijani AY. Anti-plasmodial activity of aqueous root extract of Acacia nilotica . Afr J Biochem Res. 2011;5:214–9.

Jigam AA, Akanya HO, Dauda BEN, Okogun JO. Polygalloyltannin isolated from the roots of Acacia nilotica Del. ( Leguminoseae ) is effective against Plasmodium berghei in mice. J Med Plant Res. 2010;4:1169–75.

Adeoye AO, Bewaji CO. Chemopreventive, and remediation effect of Adansonia digitata L Baobab ( Bombacaceae ) stem barks extracts in mouse model malaria. J Ethnopharmacol. 2018;210:31–8.

Musila MF, Dossaji SF, Nguta JM, Lukhoba CW, Munyao JM. In vivo antimalarial activity, toxicity and phytochemical screening of selected antimalarial plants. J Ethnopharmacol. 2013;146:557–61.

Ukwe VC, Epueke EA, Ekwunife OI, Okoye TC, Akudor GC, Ubaka CM. Antimalarial activity of aqueous extract and fractions of leaves of Ageratum conyzoides in mice infected with Plasmodium berghei . Int J Pharm Sci. 2010;2:33–8.

Rukunga GM, Muregi FW, Tolo FM, Omar SA, Mwitari P, Muthaura CN, et al. The antiplasmodial activity of spermine alkaloids isolated from Albizia gummifera . Fitoterapia. 2007;78:455–9.

Oladosu IA, Balogun SO, Ademowo GO. Phytochemical screening, antimalarial and histopathological studies of Allophylus africanus and Tragia benthamii . Chin J Nat Med. 2013;11:371–6.

Teka T, Bisrat D, Yeshak MY, Asres K. Antimalarial activity of the chemical constituents of the leaves latex of Aloe pulcherrima Gilbert and Sebsebe. Molecules. 2016;21:1415.

Article   PubMed Central   CAS   Google Scholar  

Akpan EJ, Okokon JE, Etuk IC. Antiplasmodial and antipyretic studies on root extracts of Anthocleista djalonensis against Plasmodium berghei . Asian Pac J Trop Dis. 2012;2:36–42.

Ene AC, Ameh DA, Kwanashie HO, Agomo PU, Atawodi SE. Preliminary in vivo antimalarial screening of petroleum ether, chloroform, and methanol extracts of fifteen plants grown in Nigeria. J Pharmacol Toxicol. 2008;3:254–60.

Christian AG, Mfon AG, Dick EA, David-Oku E, Akpan JL, Chukwuma EB. Antimalarial potency of the leaf extract of Aspilia africana (Pers.) C.D. Adams. Asian Pac J Trop Med. 2012;2:126–9.

Gathirwa JW, Rukunga GM, Mwitari PG, Mwikwabe NM, Kimani CW, Muthaura CN, et al. Traditional herbal antimalarial therapy in Kilifi district. Kenya J Ethnopharmacol. 2011;134:434–42.

Tepongning RN, Mbah JN, Avoulou FL, Jerme MM, Ndanga EKK, Fekam FB. Hydroethanolic extracts of Erigeron floribundus and Azadirachta indica reduced Plasmodium berghei parasitemia in Balb/c mice. Evid Based Complement Alternat Med. 2018;2018:5156710.

Akin-Osanaiye BC, Nok AJ, Ibrahim S, Inuwa HM, Onyike E, Amlabu E, et al. Antimalarial effect of neem leaves and neem stem barks extracts on Plasmodium berghei infected in the pathology and treatment of malaria. Int J Res Biochem Biophy. 2013;3:7–14.

Asrade S, Mengesha Y, Moges G, Gelayee DA. In vivo antiplasmodial activity evaluation of the leaves of Balanites rotundifolia (Van Tiegh) Blatter ( Balanitaceae ) against Plasmodium berghei . J Exp Pharmacol. 2017;9:59–66.

Otegbade OO, Ojo JA, Adefokun DI, Abiodun OO, Thomas BN, Ojurongbe O. Ethanol extract of Blighia sapida stem barks show remarkable prophylactic activity in experimental Plasmodium berghei –infected mice. Drug Target Insights. 2017;11:1177392817728725.

Christian AG, Thecla EC, Dick EA, Chile AE, Chimsorom CK, Ckukwu ND, et al. In vivo antiplasmodial activity of Bombax buonopozense root barks aqueous extract in mice infected by Plasmodium berghei . J Tradit Chin Med. 2017;37:431–5.

Muluye AB, Melese E, Adinew GM. Antimalarial activity of 80 % methanolic extract of Brassica nigra (L.) Koch. ( Brassicaceae ) seeds against Plasmodium berghei infection in mice. BMC Compl Alternative Med. 2015;15:367.

Eyasu M, Shibeshi W, Gida M. In vivo antimalarial activity of hydromethanolic leaf extract of Calpurnia aurea ( Fabaceae ) in mice infected with chloroquine-sensitive Plasmodium berghei . Int J Pharmacol. 2013;2:131–42.

Onaku LO, Attama AA, Okore VC, Tijani AY, Ngene AA, Esimone CO. Antagonistic antimalarial properties of pawpaw leaf aqueous extract in combination with artesunic acid in Plasmodium berghei -infected mice. J Vector Borne Dis. 2011;48:96–100.

Tona L, Mesia K, Ngimbi NP, Chrimwami B, Okondahoka, Cimanga K, et al. In vivo antimalarial activity of Cassia Occidentalism, Morinda morindoides and Phyllanthus niruri . Ann Trop Med Parasit. 2001;95:47–57.

Abdulrazak N, Asiya UI, Usman NS, Unata IM, Farida A. Anti-plasmodial activity of ethanolic extract of root and stem bark of Cassia sieberiana DC on mice. J Intercult Ethnopharmacol. 2015;4:96–101.

Adzu B, Abbah J, Vongtau H, Gamaniel K. Studies on the use of Cassia singueana in malaria ethnopharmacy. J Ethnopharmacol. 2003;88:261–7.

Jigam AA, Razaq UTA, Egbuta MN. In vivo antimalarial and toxicological evaluation of Chrozophora senegalensis A. Juss ( Euphorbiaceae ) extracts. J Appl Pharm Sci. 2011;1:90–4.

Ihekwereme CP, Okoye FK, Agu SC, Oli AN. Traditional consumption of the fruit pulp of Chrysophyllum albidum ( Sapotaceae ) in pregnancy may be serving as an intermittent preventive therapy against malaria infection. Anc Sci Life. 2017;36:191–5.

Okokon JE, Etebong EO, Udobang JA, Essien GE. Antiplasmodial and analgesic activities of Clausena anisate . Asian Pac J Trop Med. 2012;5:214–9.

Anato M, Ketema T. Anti-plasmodial activities of Combretum molle ( Combretaceae ) [Zwoo] seed extract in Swiss albino mice. BMC Res Notes. 2018;11:312.

Kweyamba PA, Zofou D, Efange N, Assob JN, Kitau J, Nyindo M. In vitro and in vivo studies on antimalarial activity of Commiphora africana and Dichrostachys cinerea used by the Maasai in Arusha region. Tanzania Malar J. 2019;18:119.

Elufioye TO, Agbedahunsi JM. Antimalarial activities of Tithonia diversifolia ( Asteraceae ) and Crossopteryx febrifuga ( Rubiaceae ) on mice in vivo . J Ethnopharmacol. 2004;93:167–71.

Obey JK, Ngeiywa MM, Kiprono P, Omar S, vonWright A, Kauhanen J, et al. Antimalarial activity of Croton macrostachyus stem barks extracts against Plasmodium berghei in vivo . J Pathog. 2018;2018:393854.

Tona Ngimbi NP, Tsakala M, Mesia K, Cimanga K, Apers S, De Bruyne T, et al. Antimalarial activity of 20 crude extracts from nine African medicinal plants used in Kinshasa Congo. J Ethnopharmacol. 1999;68:193–203.

Mzena T, Swai H, Chacha M. Antimalarial activity of Cucumis metuliferus and Lippia kituiensis against Plasmodium berghei infection in mice. Res Rep Trop Med. 2018;9:81–8.

Amelo W, Nagpal P, Makonnen E. Antiplasmodial activity of solvent fractions of methanolic root extract of Dodonaea angustifolia in Plasmodium berghei infected mice. BMC Complement Altern Med. 2014;14:462.

Ogbonna DN, Sokari TG, Agomuoh AA. Antimalarial activities of some selected traditional herbs from South Eastern Nigeria against Plasmodium species. Res J Parasit. 2008;3:25–31.

Gounoue Kamkumo R, Tsakem Nangap JM, Tchokouaha Yamthe LR, Ngueguim Tsofack F, Tsouh Fokou PV, Tchatat Tali MB, et al. Antimalarial activity of the aqueous extract of Euphorbia cordifolia Elliot in Plasmodium berghei -infected mice. Asian Pac J Trop Med. 2020;13:176–84.

Oluwakanyinsola AS, Adeniyi YT, Babayi H, Angela CN, Anagbogu RA, Agbakwuru VA. Antimalarial activity of ethanolic stem barks extract of Faidherbia albida (Del) a Chev (Mimosoidae) in mice. Arch Appl Sci Res. 2010;2:261–8.

Nguta JM, Mbaria JM. Brine shrimp toxicity and antimalarial activity of some plants traditionally used in treatment of malaria in Msambweni district of Kenya. J Ethnopharmacol. 2013;148:988–92.

Oluwatosin A, Tolulope A, Ayokulehin K, Okorie P, Aderemi K, Falade C, et al. Antimalarial potential of kolaviron, a biflavonoid from Garcinia kola seeds, against Plasmodium berghei infection in Swiss albino mice. Asian Pac J Trop Med. 2014;7:97–104.

Okokon JE, Ita BN, Udokpoh AE. The in vivo antimalarial activities of Uvariae chamae and Hippocratea africana . Ann Trop Med Parasitol. 2006;100:585–90.

David-Oku E, Ifeoma OO, Christian AG, Dick EA. Evaluation of the antimalarial potential of Icacina senegalensis Juss ( Icacinaceae ). Asian Pac J Trop Med. 2014;7:S469–72.

Birru EM, Geta M, Gurmu AE. Antiplasmodial activity of Indigofera spicata root extract against Plasmodium berghei infection in mice. Malar J. 2017;16:198.

Onyeto CA, Akah PA, Nworu CS, Okoye TC, Okorie NA, Mbaoji FN, et al. Anti-plasmodial and antioxidant activities of methanol extract of the fresh leaves of Lophira lanceolata ( Ochnaceae ). Afr J Biotechnol. 2014;13:1731–8.

Christian AG, Akanimo EG, Ahunna AG, Nwakaego EM, Chimsorom CK. Antimalarial potency of the methanol leaf extract of Maerua crassifolia Forssk ( Capparaceae ). Asian Pac J Trop Dis. 2014;4:35–9.

Malebo HM, Tanja W, Cal M, Swaleh SAM, Omolo MO, Hassanali A, et al. Antiplasmodial, anti-trypanosomal, anti-leishmanial and cytotoxicity activity of selected Tanzanian medicinal plants. Tanzan J Health Res. 2015;4:226–34.

Okafor AI, Nok AJ, Inuwa HM. Antiplasmodial activity of aqueous leaf extract of Mucuna Pruriens Linn in mice infected with Plasmodium berghei (NK-65 Strain). J Appl Pharm Sci. 2013;3(4 Suppl 1):S52–5.

Edagha IA, Peter AI, Aquaisua AN. Histopathological effect of Nauclea latifolia ethanolic leaf extract and artemether/lumefantrine on the hippocampus of P berghei-infected mice. Int J Brain Cognitive Sci. 2017;6:9–16.

Nworu CS, Ejikeme TI, Ezike AC, Ndu O, Akunne TC, Onyeto CA, et al. Anti-plasmodial and anti-inflammatory activities of cyclotide-rich extract and fraction of Oldenlandia affinis (R.& S.) D.C. (Rubiaceae). Afri Health Sci. 2017;17:827–43.

Ramanitrahasimbola D, Rasoanaivo P, Ratsimamanga-Urverg S, Federici E, Palazzino G, Galeffi C, et al. Biological activities of the plant-derived bisindole voacamine with reference to malaria. Phytother Res. 2001;15:30–3.

Ajala TO1, Igwilo CI, Oreagba IA, Odeku OA. The antiplasmodial effect of the extracts and formulated capsules of Phyllanthus amarus on Plasmodium yoelii infection in mice. Asian Pac J Trop Med. 2011;4:283–7.

Ifeoma O, Samuel O, Itohan AM, Adeola SO. Isolation, fractionation and evaluation of the antiplasmodial properties of Phyllanthus niruri resident in its chloroform fraction. Asian Pac J Trop Med. 2013;6:169–75.

Adinew GM. Antimalarial activity of methanolic extract of Phytolacca dodecandra leaves against Plasmodium berghei infected Swiss albino mice. Int J Pharmacol Clin Sci. 2014;3:39–45.

Okokon JE, Antia BS, Igboasoiyi AC, Essien EE, Mbagwu HO. Evaluation of antiplasmodial activity of ethanolic seed extract of Picralima nitida . J Ethnopharmacol. 2007;111:464–7.

Madara AA, Ajayi JA, Salawu OA, Tijani AY. Antimalarial activity of ethanolic leaf extract of Piliostigma thonningii Schum (Caesalpiniacea) in mice infected with Plasmodium berghei berghei . Afr J Biotechnol. 2010;9:3475–80.

Christian AG, Ahunna AG, Nwakaego EM, Chimsorom CK, Chile AE. Antimalarial potential of the ethanolic leaf extract of Pseudocedrala kotschyi . J Acute Dis. 2015;4:23–7.

Okokon JE, Ettebong E, Antia BS. In vivo antimalarial activity of ethanolic leaf extract of Stachytarpheta cayennensis . Indian J Pharmacol. 2008;40:111–3.

Adegbolagun OM, Emikpe BO, Woranola IO, Ogunremi Y. Synergistic effect of aqueous extract of Telfaria occidentalis on the biological activities of artesunate in Plasmodium berghei infected mice. Afr Health Sci. 2013;13:970–6.

Muregi FW, Ishih A, Miyase T, Suzuki T, Kino H, Amano T, et al. Antimalarial activity of methanolic extracts from plants used in Kenyan ethnomedicine and their interactions with chloroquine (CQ) against a CQ-tolerant rodent parasite, in mice. J Ethnopharmacol. 2007;111:190–5.

Olanlokun JO, Oluwole MD, Afolayan AJ. In vitro antiplasmodial activity and prophylactic potentials of extract and fractions of Trema orientalis (Linn.) stem barks. BMC Complement Altern Med. 2017;17:407.

Fadare DA, Abiodun OO, Ajaiyeoba EO. In vivo antimalarial activity of Trichilia megalantha harms extracts and fractions in animal models. Parasitol Res. 2013;112:2991–5.

François G, Steenackers T, Timperman G, Aké Assi L, Haller RD, Bär S, et al. Retarded development of exoerythrocytic stages of the rodent malaria parasite Plasmodium berghei in human hepatoma cells by extracts from dioncophyllaceae and ancistrocladaceae species. Int J Parasitol. 1997;27:29–32.

Akuodor GC, Idris-Usman M, Anyalewechi N, Odo E, Ugwu CT, Akpan JL, et al. In vivo antimalarial activity of ethanolic leaf extract of Verbena hastata against Plasmodium berghei in mice. J Herb Med Toxicol. 2010;4:17–23.

Njan AA, Adzu B, Agaba AG, Byarugaba D, Díaz-Llera S, Bangsberg DR. The analgesic and antiplasmodial activities and toxicology of Vernonia amygdalina . J Med Food. 2008;11:574–81.

Iwalokun BA. Enhanced antimalarial effects of chloroquine by aqueous Vernonia amygdalina leaf extract in mice infected with chloroquine resistant and sensitive Plasmodium berghei strains. Afr Health Sci. 2008;8:25–35.

Abosi AO, Raseroka BH. In vivo antimalarial activity of Vernonia amygdalina . Br J Biomed Sci. 2003;60:89–91.

Dame ZT, Petros B, Mekonnen Y. Evaluation of anti- Plasmodium berghei activity of crude and column fractions of extracts from Withania somnifera . Turk J Biol. 2013;37:147–50.

Adugna M, Feyera T, Taddese W, Admasu P. In vivo antimalarial activity of crude extract of aerial part of Artemisia abyssinica against Plasmodium berghei in mice. Glob J Pharmacol. 2014;8:460–8.

Deressa T, Mekonnen Y, Animut A. In vivo antimalarial activities of Clerodendrum myricoides , Dodonea angustifolia , and Aloe debrana against Plasmodium berghei . Ethiop J Health Dev. 2010;24:25–9.

Muthaura CN, Rukunga GM, Chhabra SC, Omar SA, Guantai AN, Gathirwa JW, et al. Antimalarial activity of some plants traditionally used in treatment of malaria in Kwale district of Kenya. J Ethnopharmacol. 2007;112:545–51.

Beha E, Jung A, Wiesner J, Rimpler H, Lanzer M, Heinrich M. Antimalarial activity of extracts of Abutilon grandiflorum G Don - a traditional Tanzanian medicinal plant. Phytother Res. 2004;18:236–40.

Ajaiyeoba EO, Abiodun OO, Falade MO, Ogbole NO, Ashidi JS, Happi CT, et al. In vitro cytotoxicity studies of 20 plants used in Nigerian antimalarial ethnomedicine. Phytomedicine. 2006;13:295–8.

Adebayo JO, Balogun EA, Malomo SO, Soladoye AO, Olatunji LA, Kolawole OM, et al. Antimalarial activity of Cocos nucifera Husk fibre: further studes. Evid Based Complement Altern Med. 2013;2013:742476.

Falade MO, Akinboye DO, Gbotosho GO, Ajaiyeoba EO, Happi TC, Abiodun OO, et al. In vitro and in vivo antimalarial activity of Ficus thonningii Blume ( Moraceae ) and Lophira alata Banks ( Ochnaceae ), identified from the ethnomedicine of the Nigerian middle belt. J Parasit Res. 2014;2014:972853.

Iwalewa EO, Omisore NO, Adewunmi CO, Gbolade AA, Ademowo OG, Nneji C, et al. Anti-protozoan activities of Harungana madagascariensis stem barks extract on trichomonads and malaria. J Ethnopharmacol. 2008;117:507–11.

Gathirwa JW, Rukunga GM, Njagi EN, Omar SA, Mwitari PG, Guantai AN, et al. The in vitro anti-plasmodial and in vivo antimalarial efficacy of combinations of some medicinal plants used traditionally for treatment of malaria by the Meru community in Kenya. J Ethnopharmacol. 2008;115:223–31.

Simelane MBC, Shonhai A, Shode FO, Smith P, Singh M, Opoku AR. Anti-plasmodial activity of some Zulu medicinal plants and of some triterpenes isolated from them. Molecules. 2013;18:12313.

Steele JC, Warhurst DC, Kirby GC, Simmonds MSJ. In vitro and in vivo evaluation of betulinic acid as an antimalarial. Phytother Res. 1999;13:115–9.

Builders MI, Wannang NN, Ajoku GA, Builders PF, Orisadipe A, Aguiyi JC. Evaluation of the antimalarial potential of Vernonia ambigua Kotschy and Peyr ( Asteraceae ). Int J Pharmacol. 2011;7:238–47.

Ngbolua KN, Rakotoarimanana H, Rafatro H, Ratsimamanga US, Mudogo V, Mpiana PT, et al. Comparative antimalarial and cytotoxic activities of two Vernonia species: Vernonia amygdalina from the Democratic Republic of Congo and Vernonia cinerea subsp vialis endemic to Madagascar. Int J Biol Chem Sci. 2011;5:345–53.

Benoit-Vical F, Imbert C, Bonfils JP, Sauvaire Y. Antiplasmodial and antifungal activities of iridal, a plant triterpenoid. Phytochemistry. 2003;62:747–51.

Applequist WL, Ratsimbason M, Kuhlman A, Ratonandrasana S, Rasamison V, Kingston DG. Antimalarial use of Malagasy plants is poorly correlated with performance in antimalarial bioassays. Econ Bot. 2017;71:75–82.

Thiengsusuk A, Chaijaroenkul W, NaBangchang K. Antimalarial activities of medicinal plants and herbal formulations used in Thai traditional medicine. Parasitol Res. 2013;112:1475.

Udobang JA, Nwafor PA, Okokon JE. Analgesic and antimalarial activities of crude leaf extract and fractions of Acalypha wilkensiana . J Ethnopharmacol. 2010;127:373–8.

Mohammed T, Erko B, Giday M. Evaluation of antimalarial activity of leaves of Acokanthera schimperi and Croton macrostachyus against Plasmodium berghei in Swiss albino mice. Compl Altern Med. 2014;14:314.

Kuria KA, De Coster S, Muriuki G, Masengo W, Kibwage I, Hoogmartens J, et al. Antimalarial activity of Ajuga remota Benth ( Labiatae ) and Caesalpinia Volkensii Harms ( Caesalpiniaceae ): in vitro confirmation of ethnopharmacological use. J Ethnopharmacol. 2001;74:141–8.

Mesfin A, Giday M, Animut A, Teklehaymanot T. Ethnobotanical study of antimalarial plants in Shineile District, Somali Region, Ethiopia, and in vivo evaluation of selected ones against Plasmodium berghei . J Ethnopharmacol. 2012;139:221–7.

Hilou A, Nacoulma OG, Guiguemde TR. In vivo antimalarial activities of extracts from Amaranthus spinosus L. and Boerhaavia erecta L. in mice. J Ethnopharmacol. 2006;103:236–40.

Adebajo AC, Odediran SA, Aliyu FA, Nwafor PA, Nwoko NT, Umana US. In vivo antiplasmodial potentials of the combinations of four Nigerian antimalarial plants. Molecules. 2014;19:13136–46.

Dikasso D, Mekonnen E, Debella A, Abebe D, Urga K, Menonnen W, et al. In vivo antimalarial activity of hydroalcoholic extracts from Asparagus africanus Lam in mice infected with Plasmodium berghei. Ethiop J Health Dev. 2006;20:112–8.

Yerbanga RS, Lucantoni L, Lupidi G, Dori GU, Tepongning NR, Nikiéma JB, et al. Antimalarial plant remedies from Burkina Faso: their potential for prophylactic use. J Ethnopharmacol. 2012;140:255–60.

Karou SD, Tchacondo T, Ouattara L, Anani K, Savadogo A, Agbonon A, et al. Antimicrobial, antiplasmodial, haemolytic and antioxidant activities of crude extracts from three selected Togolese medicinal plants. Asian Pac J Trop Med. 2011;4:808–13.

Bonkian LN, Yerbanga RS, Koama B, Soma A, Cisse M, Valea I, et al. In vivo antiplasmodial activity of two Sahelian plant extracts on Plasmodium berghei ANKA infected NMRI mice. Evid Based Complement Alternat Med. 2018;24:6859632.

Ajaiyeoba E, Ashidi J, Abiodun O, Okpako L, Ogbole O, Akinboye D, et al. Antimalarial ethnobotany: in vitro antiplasmodial activity of seven plants indentified in the Nigerian middle belt. Pharm Biol. 2005;42:588–91.

Kefe A, Giday M, Mamo H, Erko B. Antimalarial properties of crude extracts of seeds of Brucea antidysenterica and leaves of Ocimum lamiifolium . BMC Complement Altern Med. 2016;16:118.

Innocent E, Moshi MJ, Masimba PJ, Mbwambo ZH, Kapingu MC, Kamuhabwa A. Screening of traditionally used plants for in vivo antimalarial activity in mice. Afr J Tradit Complement Altern Med. 2009;6:163–7.

Mengiste B, Mekonnen E, Urga K. In vivo animalarial activity of Dodonaea angustifolia seed extracts against Plasmodium berghei in mice model. MEJS. 2012;4:147–63.

Biruksew A, Zeynudin A, Alemu Y, Golassa L, Yohannes M, Debella A, et al. Zingiber officinale Roscoe and Echinops kebericho Mesfin showed antiplasmodial activities against Plasmodium berghei in a dose-dependent manner in Ethiopia. Ethiop J Health Sci. 2018;28:655.

Agbaje EO, Onabanjo AO. The effects of extracts of Enantia chlorantha in malaria. Ann Trop Med Parasitol. 1991;85:585–90.

Ajayi EIO, Adelekeb MA, Adewumia TY, Adeyemia AA. Antiplasmodial activities of ethanol extracts of Euphorbia hirta whole plant and Vernonia amygdalina leaves in Plasmodium berghei -infected mice. J Taibah Univ Sci. 2017;11:831–5.

Omole AR, Malebo MH, Nondo SOR, Katani S, Mbugi H, Midiwo J, et al. In vivo anti-plasmodial activity of crude extracts of three medicinal plants used traditionally for malaria treatment in Kenya. Eur J Med Plants. 2018;24:1–7.

Nureye D, Assefa S, Nedi T, Engidawork E. In vivo antimalarial activity of the 80% methanolic root barks extract and solvent fractions of Gardenia ternifolia Schumach & Thonn ( Rubiaceae ) against Plasmodium berghei . Evid Based Complement Alternat Med. 2018;2018:9217835.

Beaufay C, Hérent MF, Quetin-Leclercq J, Bero J. In vivo antimalarial activity and toxicity studies of triterpenic esters isolated from Keetia leucantha and crude extracts. Malar J. 2017;16:406.

Bankole AE, Adekunle AA, Sowemimo AA, Umebese CE, Abiodun O, Gbotosho GO. Phytochemical screening and in vivo antimalarial activity of extracts from three medicinal plants used in malaria treatment in Nigeria. Parasitol Res. 2016;115:299–305.

Jansen O, Tchinda AT, Loua J, Esters V, Cieckiewicz E, Ledoux A. Antiplasmodial activity of Mezoneuron benthamianum leaves and identification of its active constituents. J Ethnopharmacol. 2017;203:20–6.

Udobre AS, Udobang JA, Udoh AE, Anah VU, Akpan AE, Charles GE. Effect of methanol leaf extract of Nauclea latifolia on albino mice infected with Plasmodium berghei berghei . Afr J Pharmacol Ther. 2013;2:83–7.

Mesia K, Tona L, Mampunza MM, Ntamabyaliro N, Muanda T, Muyembe T, et al. Antimalarial efficacy of a quantified extract of Nauclea pobeguinii stem barks in human adult volunteers with diagnosed uncomplicated falciparum malaria .Part 2: a clinical phase IIB trial. Planta Med. 2012;78:853–60.

Okeola VO, Adaramoye OA, Nneji CM, Falade CO, Farombi EO, Ademowo OG. Antimalarial and antioxidant activities of methanolic extract of Nigella sativa seeds (black cumin) in mice infected with Plasmodium yoelii nigeriensis . Parasitol Res. 2011;108:1507–12.

Girma S, Giday M, Erko B, Mamo H. Effect of crude leaf extract of Osyris quadripartita on Plasmodium berghei in Swiss albino mice. BMC Complement Alternative Med. 2015;15:184.

Kabiru AY, Ibikunle GF, Innalegwu DA, Bola BM, Madaki FM. In vivo antiplasmodial and analgesic effect of crude ethanol extract of Piper guineense leaf extract in Albino Mice. Scientifica (Cairo). 2016: 8687313.

Hiben MG, Sibhat GG, Fanta BS, Gebrezgi HD, Tesema SB. Evaluation of Senna singueana leaf extract as an alternative or adjuvant therapy for malaria. J Tradit Complement Med. 2015;6:112–7.

Tadesse SA, Wubneh ZB. Antimalarial activity of Syzygium guineense during early and established Plasmodium infection in rodent models. BMC Complement Alternative Med. 2017;17:21.

Adepiti AO, Iwalewa EO. Evaluation of the combination of Uvaria chamae (P Beauv) and amodiaquine in murine malaria. J Ethnopharmacol. 2016;193:30–5.

Masaba SC. The antimalarial activity of Vernonia amygdalina Del ( Compositae ). Trans R Soc Trop Med Hyg. 2000;94:694–5.

Omoregie ES, Pal A. Antiplasmodial, antioxidant and immunomodulatory activities of ethanol extract of Vernonia amygdalina del. leaves in Swiss mice. Avicenna J Phytomed. 2016;6:236–7.

Challand S, Willcox M. A Clinical trial of the traditional medicine Vernonia amygdalina in the treatment of uncomplicated Malaria. J Altern Complement Med. 2009;15:1231–7.

Ajayi BB, Ogunsola JO, Olatoye OI, Antia RE, Agbedea S. Effects of pituitary extract, ovaprim, and bitter leaves ( Vernonia amygdalina ) on the histopathology of African catfish ( Clarias gariepinus ) Aquacult. Fish. 2018;3:232–7.

Download references

Acknowledgements

I would like to express my special appreciation and thanks to Professor Dr. Wanderley de Souza for his helpful comments.

Author information

Authors and affiliations.

Department of Microbiology, Shahrekord Branch, Islamic Azad University, Shahrekord, Iran

Elahe Tajbakhsh

Department of Biomedical Science, Faculty of Health Sciences, Regional Hospital Buea, Buea, Cameroon

Tebit Emmanuel Kwenti

Department of Public Health and Hygiene, Faculty of Health Sciences, University of Buea, Yaoundé, Cameroon

Department of Medical Laboratory Sciences, Faculty of Health Sciences, University of Buea, Yaoundé, Cameroon

Young Researchers and Elite Club, Shahrekord Branch, Islamic Azad University, Shahrekord, Iran

Parya Kheyri & Saeed Nezaratizade

Department of Biomedical Sciences and Pathobiology, Center for One Health Research, Virginia Maryland College of Veterinary Medicine, Virginia Tech, 1410 Prices Fork Road, Blacksburg, VA, 24061-0342, USA

David S. Lindsay

Shahrekord Branch, Islamic Azad University, Shahrekord, Iran

Faham Khamesipour

Shahid Beheshti University of Medical Sciences, Tehran, Iran

You can also search for this author in PubMed   Google Scholar

Contributions

All authors contributed equally to the study. All authors read and approved the manuscript.

Corresponding author

Correspondence to Faham Khamesipour .

Ethics declarations

Ethics approval and consent to participate, consent for publication, competing interests.

The authors declare that they have no competing interests.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: table s1..

In vitro and in vivo studies reporting inactive antiplasmodial or antimalarial activity. Table S2. List of active compounds identified from plants.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Tajbakhsh, E., Kwenti, T.E., Kheyri, P. et al. Antiplasmodial, antimalarial activities and toxicity of African medicinal plants: a systematic review of literature. Malar J 20 , 349 (2021). https://doi.org/10.1186/s12936-021-03866-0

Download citation

Received : 16 February 2021

Accepted : 27 July 2021

Published : 25 August 2021

DOI : https://doi.org/10.1186/s12936-021-03866-0

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Medicinal plants
  • Antiplasmodial activity
  • Antimalarial activity

Malaria Journal

ISSN: 1475-2875

thesis on antimalarial drugs

  • Skip to main content
  • Accessibility information

thesis on antimalarial drugs

  • Enlighten Enlighten

Enlighten Theses

  • Latest Additions
  • Browse by Year
  • Browse by Subject
  • Browse by College/School
  • Browse by Author
  • Browse by Funder
  • Login (Library staff only)

In this section

Drug action and resistance in malaria parasites: experimental genetics models and biochemical features of fast acting novel antimalarials

Simwela, Nelson Victor (2020) Drug action and resistance in malaria parasites: experimental genetics models and biochemical features of fast acting novel antimalarials. PhD thesis, University of Glasgow.


Resistance to antimalarial drugs inevitably follows their deployment in malaria endemic parts of the world. For instance, current malaria control efforts which significantly rely on artemisinin combination therapies (ACTs) are being threatened by the emergence of resistance to artemisinins and ACTs. Understanding the role of genetic determinants of artemisinin resistance is therefore important for implementation of mitigation strategies. Moreover, elucidating the mode of action for drugs that are in advanced stages of development is specifically critical as drug resistance mechanisms can be prospectively predicted and possible means of surveillance put in place.

In the present work, CRISPR-Cas9 genome editing has been used to engineer candidate artemisinin resistance mutations (Kelch13 and UBP-1) in the rodent malaria parasite Plasmodium berghei. The role of these mutations in mediating artemisinin (and chloroquine) resistance under both in vitro and in vivo conditions has been assessed which up until now, has either remained un-validated (UBP-1) or debated (Kelch13, under in vivo conditions) in human infecting Plasmodium falciparum. The results have provided an in vivo model for understanding and validating artemisinin resistance phenotypes which just like their Plasmodium falciparum equivalents do not just mediate resistance phenotypes, but also carry accompanying fitness costs.

In addition to the above findings, biochemical and drug inhibition studies have been carried out to demonstrate that small molecule inhibitors targeting ubiquitin hydrolases (to which UBP-1 is a class member) display activity in human and rodent infecting malaria parasites in vitro and in vivo. These inhibitors also show evidence of ability to potentiate artemisinin action which can be exploited to overcome the emerging resistance as combination partner drugs. Untargeted metabolomic screens have also been used to characterize the mode of action of lead antimalarial drug candidates that are emerging from the Novartis Institute of Tropical Diseases drug discovery pipeline. A common biochemical and metabolic profile of these compounds which display a very fast parasite killing rate is presented and can hopefully be used to identify compounds that can achieve a similar feat. Moreover, these profiles have pointed to possible mode of action for novel drugs whose mechanistic mode of parasite killing is still unknown or disputed.

Item Type: Thesis (PhD)
Qualification Level: Doctoral
Keywords: Antimalarial drugs, mode of action, resistance, genetics models, biochemistry, metabolomics.
Subjects: >
Colleges/Schools: > >
Funder's Name: ,
Supervisor's Name: Waters, Professor Andy and Barrett, Professor Mike
Date of Award: 2020
Depositing User:
Unique ID: glathesis:2020-81876
Copyright: Copyright of this thesis is held by the author.
Date Deposited: 17 Dec 2020 17:02
Last Modified: 16 Sep 2022 15:52
Thesis DOI:
URI:

Actions (login required)

View Item

Downloads per month over past year

View more statistics

-

The University of Glasgow is a registered Scottish charity: Registration Number SC004401

Design, synthesis, and in silico-in vitro antimalarial evaluation of 1,2,3-triazole-linked dihydropyrimidinone quinoline hybrids

  • Open access
  • Published: 27 February 2023
  • Volume 34 , pages 2065–2082, ( 2023 )

Cite this article

You have full access to this open access article

thesis on antimalarial drugs

  • Rasheed A. Adigun 1 ,
  • Frederick P. Malan 1 ,
  • Mohammed O. Balogun 2 &
  • Natasha October 1  

2595 Accesses

8 Citations

Explore all metrics

In response to the malaria parasite’s resistance towards quinoline-based antimalarial drugs, we have employed quinoline-containing compounds in combination with dihydropyrimidinone (DHPM) analogues as resistance reversal agents (RAs) and investigated their antimalarial activities based on DHPM’s resistance reversal abilities. The present study employed click chemistry to link DHPM and quinoline compounds which offered several synthetic advantages over the previously used amide coupling for the same hybrids. Among the synthesised compounds, 4 hybrids with the 7-chloroquinoline moiety showed antimalarial activity below 1 µM while compounds with the mefloquine moiety showed lower antimalarial activity than chloroquine (CQ) and the 7-chloroquinoline hybrids. Among the tested hybrids for the IC 50 determination, four compounds displayed good antimalarial activity with increased sensitivity against the CQ-resistant K1 strain between 421 and 567 nM and showed higher activity between 138 and 245 nM against the NF54 CQ-sensitive strain, while three compounds have IC 50 values greater than 5 µM. Additionally, in silico molecular docking and molecular dynamics studies were conducted to investigate the binding affinities of all the synthesised compounds as glutathione reductase protein competitive inhibitors. Further optimisation of the compound with the highest binding affinity generated 16 compounds with higher binding affinities than the flavine adenine dinucleotide (FAD) cofactor.

Similar content being viewed by others

thesis on antimalarial drugs

Synthesis, characterization and antimalarial activity of isoquinoline derivatives

Synthesis of new chloroquine derivatives as antimalarial agents.

thesis on antimalarial drugs

Spirofused tetrahydroisoquinoline-oxindole hybrids as a novel class of fast acting antimalarial agents with multiple modes of action

Avoid common mistakes on your manuscript.

Introduction

Recent data indicates that the challenge to overcome malaria still lingers [ 1 ], regardless of all the significant efforts and advances in drug developments and medical technologies. The rise of drug or multidrug resistance in Plasmodia strains has made malaria the most important tropical parasitic disease [ 2 ]. Therefore, developing new antimalarial drugs is necessary to rectify this situation. Despite the development of many antimalarial leads, quinoline compounds remain important in malarial chemotherapy [ 3 ].

For many years, chloroquine (CQ) from the quinoline family was the first-line antimalarial drug. Despite the huge success recorded in using quinoline for malaria therapeutics, these drugs have been challenged by Plasmodium parasites due to their rapid development of resistance [ 4 , 5 ]. Drug resistance in malaria occurs when the parasite strain continues to multiply or survive even though the therapeutic has been given in the recommended tolerable dose. This situation is not malaria-specific but is also observed in cancer and other parasitic diseases like the case with antibiotics. As for the case of chloroquine (CQ) resistance, there is strong evidence through different research that resistant malaria parasites accumulate less CQ in their food vacuole than sensitive parasites [ 6 ]. This could result from an efflux pumping system, reduced CQ uptake by the parasite, or a combination of both processes [ 7 ]. This implies that CQ concentration does not reach the parasite’s lethal dose, thereby resulting in its continuous survival.

The use of resistance reversal agents (RAs) is among the strategies to reverse this drug resistance in malaria parasites. Resistance reversal refers to the use of a compound that can restore the sensitivity of resistant strains to a drug while it might have little or no therapeutic action against the parasite [ 8 ]. In the context of malaria, resistance RAs ensure antimalarial drugs accumulate within the parasite’s food vacuole. An interesting feature of this strategy is the ability to combine two biologically active compounds in a hybrid, maintaining the properties of both without the setback of differing elimination half-life (t 1/2 ) or pharmacokinetic properties associated with separate drugs that are used in combination therapy. Furthermore, a major advantage of this approach over the rest is the ability of the RA to competitively bind to the P. falciparum chloroquine resistance transporter ( Pf CRT) protein, thereby inhibiting the efflux of the CQ from the digestive food vacuole of the Plasmodium parasite. In contrast, other strategies lack this feature to bind to the mutated CQ-resistant P. falciparum [ 9 , 10 ].

Previously, it was discovered that concurrent exposure of resistant strains of P. falciparum to both CQ and verapamil reversed resistance in vitro [ 4 , 11 ]. Since verapamil is a calcium channel blocker used to treat hypertension and other cardiovascular diseases, researchers have used different calcium channel blockers and other drugs in combination with quinoline-based antimalarial drugs in attempts to reverse antimalarial drug resistance. It was observed that the RA combinations were able to increase the sensitivity of cloned CQ-resistant P. falciparum strains in vitro but did not affect the sensitivity of CQ-sensitive strains [ 6 , 11 , 12 , 13 ].

Before Peyton and co-workers coined the term “Reversed Chloroquine” (RCQ) to describe a hybrid compound containing CQ and an RA, the use of RAs in antimalarial chemotherapy involved the administration of the two agents as separate doses [ 14 ]. They discovered that covalently linking a CQ-like moiety to imipramine was effective against both sensitive and resistant strains of the Plasmodium parasite. The hybrid molecule delivered the RA in a 1:1 ratio to the antimalarial quinoline, which would lower the dose required if the two drugs, was to be given in separate forms, with resultant lower cost and toxicity [ 4 ]. This arrangement would favour the drug accumulation in the parasite’s digestive vacuole and also interfere with the CQ export by the mutated CQ-resistant P. falciparum [ 14 ].

In our previous work, we covalently linked CQ to dihydropyrimidinone (DHPM) using an amide bond linker at position 3 of the DHPM. The resulting hybrids could reverse resistance in the K1 strain of P. falciparum in vitro [ 15 ]. Since DHPMs present various sites of functionalisation that could be used for structure–activity relationship (SAR), several synthetic strategies could be employed to link quinoline-based antimalarials to various DHPM scaffolds.

The previous report of the synthesis of DHPM-based hybrids of chloroquine (CQ) moiety used amide coupling to link the CQ to DHPM, specifically at N3 of the reversal agent [ 15 ]. The synthetic route to these hybrids involved the protection of N1 nitrogen with a methyl group. Furthermore, the synthesis of the amide carbamate ester intermediate involved overnight reflux with a large excess of phenyl chloroformate and NaH, 10.0, and 11.0 molar equivalents, respectively. This synthetic procedure generated a lot of waste. However, we report on the use of click chemistry as an alternative route to synthesise DHPM-based quinoline hybrids and their antimalarial activities. This project employs a triazole covalent linker between quinoline-based compounds and different DHPMs using the C6 of the DHPM as the point of attachment. Two series of quinoline-based compounds related to chloroquine and mefloquine antimalarial drugs were chosen (Scheme 1 )

scheme 1

General structure of the target DHPM hybrids based on CQ ( III ) and mefloquine ( IV )

It also has been demonstrated that quinoline derivatives block the P. falciparum glutathione reductase ( Pf GR) enzyme [ 16 ]. The host erythrocytes undergo significant structural and physiological changes when invaded by P. falciparum to aid the parasites’ survival and growth. The oxidative stress caused by these changes affects the malaria parasites. As a result, an effective antioxidant is necessary to inhibit reactive oxygen species from harming the parasites [ 17 ]. It is thought that reduced glutathione (GSH) is a crucial antioxidant during erythrocytic infection. As an oxidoreductase, glutathione reductase in P. falciparum converts oxidised glutathione (GS-SG) to reduced GSH using hydrogen from nicotinamide adenine dinucleotide phosphate (NADPH). The transfer of hydrogen from NADPH to GS-SG is the most essential task of the FAD cofactor in Pf GR because, without it, the parasite would experience oxidative stress [ 18 , 19 ]. Thus, P. falciparum’s glutathione reductase enzyme becomes a crucial target in antimalarial treatment [ 20 ]. Therefore, in addition to the in vitro antimalarial testing, the binding affinities and the molecular stabilities of the synthesised compounds within the glutathione reductase enzyme were estimated using molecular docking and molecular dynamics in silico methods.

Results and discussion

The synthesis of the chlorinated DHPM and the subsequent azidation to form azides 9a-d was reported earlier [ 21 ]. The terminal alkynes were installed on the quinoline-based compounds using different protocols. Compound 2 was synthesised via a nucleophilic substitution reaction from 4,7-dichloroquinoline 1 according to the reported protocol (Scheme 2 ) [ 22 ]. This involved heating a mixture of 1 and propargylamine in a sealed tube under a nitrogen atmosphere.

scheme 2

Synthesis of quinolines 2 and 3 . Reagent and conditions: (i) propargylamine, 110 °C, N 2 , 18 h. (ii) propargyl alcohol, NaH, DMF, 0 °C, 50 °C

Formation of product 2 was confirmed by 1 H NMR spectroscopy with the observance of the terminal alkyne’s proton at δ H 2.66 ppm (t, 4 J HH  = 2.5 Hz). The method reported by Kaval et al. was adapted to synthesise the oxyquinoline 3 (Scheme 2 ) [ 23 ]. The formation of compound 3 was confirmed from the 1 H NMR spectrum with the terminal alkyne’s proton appearing at δ H 2.62 ppm (t, 4 J HH  = 2.4 Hz) and the Single crystal X-ray diffraction (SCXRD) technique (Fig.  1 ).

figure 1

Molecular structure of compound 3 . Thermal ellipsoids are drawn at a 70% probability level

The quinolinol 6 was synthesised via Conrad-Limpach synthesis of quinolines, which was readily converted to the alkynes 7 and 8 (Scheme 3 ).

scheme 3

Synthesis of 2,8-bis(trifluoromethyl)quinolines. Reagents and conditions: (i) polyphosphoric acid, 150 °C, 3 h; (ii) K 2 CO 3 , dry DMF, 30 °C, 30 min, propargyl bromide, 6 h; (iii) butyn-1-ol, PPh 3 , DEAD, dry THF, 0 °C, 25 °C, 12 h

Aniline 4 was quantitatively cyclised to the quinolinol 6 by heating it at 150 C with ethyl 4,4,4-trifluoroacetoacetate ( 5 ) for 3 h in the presence of polyphosphoric acid (PPA) as the catalyst [ 24 ]. The nucleophilic alkylation of 6 to 7 was performed using the Williamson ether synthesis [ 25 ]. To extend the carbon chain linker, compound 8 was synthesised from the quinolinol 6 via the Mitsunobu reaction [ 26 ] as Williamson’s nucleophilic alkylation of the quinolinol 6 with excess 4-Bromo-1-butyne failed to give the desired product. From the mechanism of Conrad-Limpach synthesis of the quinolines [ 27 , 28 ], it is possible to obtain two tautomers for the quinolinol 6 : the enol (a) and the keto (b) (Fig.  2 ). The simultaneous presence of both tautomers was confirmed from the SCXRD data (Fig.  2 ). It corroborated the earlier isolation by Sarojini et al. [ 29 ], but the subsequent deprotonations only involved the enol (a) as was observed from the SCXRD of compound 7 (Fig.  3 ). The involvement of the enol form for the subsequent deprotonations points to its preference over the keto form. The more stable enol tautomer is not unexpected for two reasons: enol stability by conjugation with a neighbouring pi system and aromaticity [ 30 ]. The two factors are present as the quinoline moiety is highly conjugated in the two rings, and they both contain aromatic systems for additional resonance stability.

figure 2

Molecular structure of compound 6 showing the two tautomers, the keto ( a ) and the enol ( b ) forms. Thermal ellipsoids are drawn at a 50% probability level

figure 3

Molecular structure of compound 7 . Thermal ellipsoids are drawn at a 50% probability level

The structure of compound 8 was confirmed from the 1 H NMR spectrum with the observation of all the diagnostic peaks. The terminal alkyne proton appeared at δ H 2.10 ppm (t, 4 J HH  = 2.7 Hz), while the two methylene protons were assigned to the signals δ H 2.90 ppm (td, 3 J HH  = 6.7 Hz, 4 J HH  = 2.7 Hz) and 4.41 ppm (t, 3 J HH  = 6.7 Hz). All the aromatic protons were also observed at the expected peak positions. The hybrid compounds were synthesised by click chemistry reactions (Schemes 4 and 5 ) as adapted from Guantai et al. [ 31 ]. The reaction was first attempted in a mixture of tert -butanol and water using copper(I) iodide (CuI) as the catalyst. This gave three spots on the TLC, and no attempt was made to separate them because of their close R f values. The catalyst was changed to freshly prepared 1-M solution of copper(II) sulphate pentahydrate (CuSO 4 .5H 2 O) and sodium ascorbate which reductively produces the required Cu(I) in situ. Initially, this approach seemed to work for the reaction where only one prominent spot was observed on the TLC. However, a difficult-to-filter precipitate also formed from the solution. Another setback was the persistent appearance of the starting materials even after long hours of stirring at 25 °C. The solvent was changed to DMF with the catalyst freshly prepared in distilled water. The reaction mixture was stirred for 12 h, and the TLC analysis showed a single spot for the product with complete consumption of the starting materials.

scheme 4

Synthesis of the triazoles 10a-h. Reagents and conditions: CuSO 4 , sodium ascorbate, DMF, H 2 O, 25 °C, 12 h

scheme 5

Synthesis of the triazoles 11a-h. Reagents and conditions: CuSO 4 , sodium ascorbate, DMF, H 2 O, 25 °C, 12 h

The 1 H NMR spectra obtained for 11a-h correlated well with the proposed structures, and all the diagnostic peaks were observed (see the supplementary information for details). Additionally, the predicted connectivity of the compounds was confirmed from the SCXRD structure elucidations of 11d and 11 g (Fig.  4 ).

figure 4

Molecular structures of compounds 11d ( a ) and 11g ( b ). Thermal ellipsoids are drawn at a 50% probability level

Antimalarial activity of DHPM-quinoline triazole hybrids

The percent growth inhibitory activities of DHPMs 9a-d and hybrids 10a-h and 11a-h were evaluated at 1-µM and 5-µM concentrations in an in vitro antiplasmodial assay (Fig.  5 , Table 1 ). An NF54 CQ-sensitive (CQS) P. falciparum asexual strain was used. The most potent compounds 10a-d were further evaluated for the IC 50 against NF54 CQS, and K1 CQ-resistant (CQR) strains, while the IC 50 of the moderately potent compounds 11c , 11e, and 11 g were determined against the K1 CQR strain only. In all the assays, CQ was used as the reference drug (Table 2 ).

figure 5

In vitro activity of DHPM-CQ triazole hybrids at 1 μM and 5 μM against asexual P. falciparum (NF54 and K1)

Among the compounds tested against the NF54 strain to determine the percentage inhibition, the reversal agents (RAs) 9a-d displayed minimal inhibition (Fig.  5 , Table 1 ). This was expected as they lack the antiplasmodial quinoline moiety, and their role in the hybrids is to inhibit the underlying process responsible for drug resistance. Within the same series, compounds 10a-d with 4-amino functionality displayed the highest inhibitions and were comparable to CQ at the two concentrations tested. Compounds 10e-f with the 4-oxyquinoline moiety exhibited very low inhibitions, even at 5 µM. The decrease in the activity of these oxygen-containing hybrids 10e-f could be attributed to the absence of the 4-amino functionality, which is present in compounds 10a-d . This agrees with previous reports that quinolines containing an amino group at the 4-position show higher antiplasmodial activities than those containing oxygen or lacking the 4-amino group [ 32 , 33 , 34 , 35 ]. This stresses the importance of the 4-amino moiety in maintaining the basic property required for the CQ and its analogues to accumulate in the food vacuole of P. falciparum [ 34 ]. Compounds 11a-h with the mefloquine moiety showed lower antiplasmodial activity than CQ and the hybrids 10a-h . Among these hybrids, compounds 11a , 11c , 11e, and 11 g , with unsubstituted acidic protons at N1 and N3, showed better activities than compounds 11b , 11d , 11f, and 11 h which instead have a methyl group at N1. The generally lower activities of hybrids 11a-h could also be ascribed to the lack of 4-amino functionality in the quinoline ring [ 34 ].

Compounds 10a-d displayed good antiplasmodial activities against the CQR K1 strain. The IC 50 s for this strain were between 421 and 567 nM, but much greater potencies of 138–245 nM were observed against the NF54 CQS strain. Compounds 11c , 11e , and 11 g were at least three orders of magnitude weaker. Hybrids 10b and 10d , containing chloroquine moiety with a methyl group at N1 of the DHPM, showed better antiplasmodial activities than hybrids 10a and 10c with unsubstituted acidic protons at N1 and N3 in both NF54 and K1 strains. This higher antiplasmodial activity could result from the decreased acidity associated with the removal of the acidic proton, which increases the basic property of the quinoline hybrid needed for its accumulation in the acidic food vacuole of the P. falciparum [ 36 ]. In addition, replacing the acidic proton with a methyl group increases the lipophilicity and in turn increases the antimalarial activities [ 37 ]. However, other substitutions, particularly the para -methoxy and the para -fluoro groups on the DHPM aromatic, do not seem to influence the activities of the hybrid compounds. Hybrids 10a-d also showed reduced Resistance indices (RI) factor ranging from 2.3 to 3.6 as compared to CQ with an RI factor of 8.5. The reduced RI factor suggests that the addition of the DHPM to the quinoline scaffold was able to change the sensitivity of the K1 strain observed.

In silico study

Using the Maestro tool from the Schrödinger software, we performed the in silico study utilising the crystal structure of glutathione reductase (PDB ID: 1ONF). The X-ray crystallographic structure of P. falciparum glutathione reductase was retrieved from the Protein Data Bank (PDB, https://www.rcsb.org/ ). The raw protein structure was prepared using the protein preparation wizard of the Schrodinger suite. This was used to assign bond orders, remove water molecules beyond 3 Å, and add H atoms. In addition, both missing side chains and missing loops were added using the Prime. While refining, water orientations were sampled using the PROPKA at a pH of 7.0 for the energy optimisation process and the restrained minimisation was done to confine the root-mean-square deviation (RMSD) of heavy atoms to 0.30 Å using an OPLS4 force field [ 38 ]. The active site of the protein residues surrounding the FAD cofactor (the bound ligand) was taken as the binding site, and these residues were used for the preparation of the docking grid using the receptor grid generation panel of the Schrodinger suite. No volume was excluded, and likewise, no constraint was defined. The input ligand structures of the synthesised compounds were built using the LigPrep module of the Schrödinger suite. For each 2D structure, the LigPrep process creates 3D structures with minimised energy, proper bond lengths, and angles. At a physiological pH of 7.0 ± 0.2, the possible ionisation states for each ligand structure were generated using the EPIK tool. Lastly, the Glide ligand docking module and the related default force field OPLS4 were used to execute the molecular docking in an extra precision (XP) mode.

The docking simulations of the synthesised compounds were run in the binding site of P. falciparum glutathione reductase crystal structure (PDB ID: 1ONF) co-crystallised with FAD cofactor, and the outcomes are shown in Table 3 . The binding energy of the ligand–protein complex corresponds to the intensity and affinity of the ligand–protein interaction. The docked complexes were rated according to their lowest energy value (kcal/mol). The negative value of the binding energy increases with the increasing strength of the interaction and vice versa. As expected, the FAD cofactor showed the highest binding affinity at −9.644 kcal/mol, followed by compound 10a at −8.1 kcal/mol. Accidentally, no correlation was observed between the docking scores and the antiplasmodial IC 50 values of the synthesised compounds. As it is well known with competitive inhibitors, it is always a challenge to design compounds that could competitively bind at the active site with higher binding affinity than the cofactor [ 39 ]; however, rational optimisations could be a saving grace. Though all the synthesised compounds showed insignificantly lower binding affinity than the FAD cofactor, all of them displayed considerable binding energy, and five of the synthesised compounds showed higher binding affinity than the chloroquine standard.

figure 6

Binding pose and electronic interactions of compound 10a within the active site of glutathione reductase. 3D and 2D structures are shown in a and b , respectively

Analysis of the binding pose of compound 10a showed different electronic interactions as shown in Fig.  6 . The compound was able to make different interactions within the active site of glutathione reductase. The + NH 3 of the Lys151 side chain was able to simultaneously make two π-cation interactions with the phenyl group on the DHPM ring (at 4.69 Å) and the DHPM ring itself (at 4.94 Å), while the benzene ring of the quinoline moiety was able to form another π-cation interaction with + NH 3 side chain of Lys32 (at 3.67 Å). In addition, several H-bond interactions were noted. These include H-bond interactions between N3H of the DHPM and the carboxylic OH of the Asp167 side chain at 1.90 Å, C = O of DHPM, and one NH of the guanidino group of Arg272 side chain at 1.89 A. Most importantly, the nitrogen of the pyridine ring of the quinoline moiety was able to form hydrophobically packed H-bond interaction with NH of Ala110 backbone at 1.92 Å (Fig.  7 ). This type of H-bond is very crucial in maintaining the ligand stability within a protein because of the difficulty involved in breaking the H-bond formed in a hydrophobic space [ 38 ]. Another notable feature is the halogen interaction between the chlorine atom on the quinoline and the C = O backbone of Asn278 at 3.36 Å.

figure 7

Hydrophobically packed H-bond (labelled yellow dotted lines) between N of the pyridine ring of compound 10a and Ala110 within the hydrophobic region of glutathione reductase (interacting residues in CPK representation)

Optimisation of compound 10a

The synthesised compounds must exhibit stronger interactions than the FAD cofactor in order to function as competitive inhibitors at the active site of glutathione reductase. To achieve this, the protein–ligand complex of compound 10a with the highest binding affinity was loaded onto the workspace and rationally optimised using the Ligand Designer tool in the Maestro platform. Analysis of the workspace showed the 4-methoxy substituent on the phenyl ring of DHPM in a solvent-exposed region. The 4-methoxy group was therefore truncated, leading to an increase in the binding affinity from −8.1 to −8.2 kcal/mol. The growth space of compound 10a , the ligand–protein interaction, and the truncated 10a are shown in Fig.  8 a, b, and c, respectively. The light blue regions show the cavity space within the binding pocket, while the deep blue regions show the solvent-exposed areas.

figure 8

The growth space of compound 10a within the glutathione reductase active site ( a ), the ligand-protein interaction ( b ), and the truncated 10a ( c )

The truncated 10a was then loaded into the workspace with the protein crystal structure. The pathfinder bonds were activated to identify possible points of R group attachments. With this, the acetate group at position 5 of the DHPM was selected and enumerated to be replaced with the default R groups in Maestro Ligand Designer. The methyl group of the acetate moiety was subsequently replaced and this generated 916 optimised derivatives of compound 10a . The top 20 ligands are shown in Table 4 .

From the generated 916 new ligands, the highest scoring ligand was LD_350 with a docking score of −11.0 kcal/mol, while ligand LD_753 had the lowest docking score of −0.3 kcal/mol. The mean was found to be −8.2 kcal/mol, while the median was −8.4 kcal/mol. This indicates that more than 50% of the optimised ligands had higher binding affinity than the original compound 10a at −8.1 kcal/mol. Interestingly, the optimisation process generated 16 ligands with higher binding affinities than the FAD cofactor, as seen in Table 4 . In comparison, 664 ligands had higher binding affinities than the original compound 10a .

figure 9

The interaction diagrams of the top two ligands LD 350 ( a ) and LD 762 ( b )

The interaction diagrams of the top two ligands (LD 350 and LD 762) are shown in Fig.  9 , showing the changes, especially in the electronic interactions from the added R groups. In LD 350, for example, all the original electronic interactions in compound 10a were preserved, while the added R group could generate additional H-bond interactions. These are the H-bond between 2-OH of the phenyl ring and C = O backbone of Asp167 at 2.05 Å, the H-bond between 4-OH of the phenyl ring and C = O side chain of Glu168 at 1.55 Å and H-bond between the C = O at the benzyl position and NH side chain of Asn171 at 2.01 Å. These interactions alone with the truncated 4-methoxy group on the DHPM contributed −2.9 kcal/mol of binding energy, which is a 35% increase in binding affinity compared to the original compound 10a .

Molecular dynamics (MD) simulation

Because molecular docking lacks precise information on the explicit biological water system and the inherent flexibility of the receptor, studies of molecular dynamics that imitate similar biological settings can help to understand the stability of the ligand–protein complex. Based on the docking findings, we ran molecular dynamics simulations for compounds 10a , LD_350, and LD_762 in this work. The MD simulation was carried out using the Schrödinger Desmond MD simulation software (version 2021–1). To achieve the biological solvation system, the ligand–protein complex was placed in a TIP3P solvation model and then subsequently neutralised by adding 6Cl − counter ions. The energy minimisation step was also completed to confirm that the system has no steric conflict. The simulation was run for 50 ns at 300 K and 1 bar under an “isothermal-isobaric ensemble” (NPT) condition. The “Nose–Hoover chain thermostat” and “Martyna-Tobias-Klein barostat” approaches were ensembled, respectively, for isothermal-isobaric conditions [ 40 , 41 ]. At 50-ps intervals, simulation trajectories were obtained, and the resulting trajectories were evaluated.

The MD simulation utilises protein equilibration, flexibility, and the average distance between backbone atoms to calculate the RMSD to estimate the fluctuation of the total protein–ligand complex. A protein–ligand complex with a lower RMSD value indicates a more stable interaction. In the case of compound 10a , both the protein and the ligand RMSD values lie at approximately 3.0 Å, though the RMSD of the protein averagely lies above 3.0. In contrast, the ligand averagely lies below 3.0 Å except at approximately 35 and 43 ns, where it went slightly above 3.0 Å (See SI Fig. S1 a). In addition, the protein and the ligand attained equilibration at less than 5 ns, and no apparent separation between them could be seen throughout the 50-ns MD simulation. Indicating the stability of the original docked structures, the approximately constant RMSD of Cα atoms and minor fluctuations of the ligand atoms during the MD simulation show that the conformation of the ligand did not change or varied very little. Comparing this with the RMSD of the optimised ligands, the LD_350 complex (See SI Fig. S1 b), for example, showed lower RMSD values at the beginning of the simulation for the ligand but was only able to achieve equilibration at around 37 ns. Though the RMSD of the ligand increased from ~ 1.8 Å to ~ 4.2 Å during the equilibration, no separation was observed after the equilibration till the end of the simulation period. The RMSD diagram of LD_762 revealed that the protein–ligand complex achieved equilibration at ~ 18 ns, and no separation was observed till the end of the simulation period (See SI Fig. S1 c). In this case, the protein showed the slightest deviation and had a maximum RMSD value of 2.4 Å till the end of the simulation, while the RMSD value of the ligand was higher at ~ 5.6 Å at the end of the simulation.

The molecular dynamics simulation identified the major binding interactions between the simulated ligands and glutathione reductase protein. One of the most significant advantages of MD simulation is its ability to precisely identify stable binding interactions from those revealed by molecular docking studies. In most cases, the flexibility of the protein and the ligand in MD, as obtained in a physiological condition, gives rise to the observed differences. In addition, the contact time of each interaction of the ligand with different protein residues as a function of the total simulation period is shown in percentage (See SI Fig. S2 ). In the case of compound 10a , for example, the two π-cation interactions between + NH 3 of Lys151 and the DHPM rings, as shown in molecular docking, were not preserved. The π-cation interaction of the amino side of Lys32 was preserved at 45% of the simulation time but with the triazole ring rather than the benzene ring of the quinoline moiety. Likewise, the halogen bond interaction between Cl of the quinoline moiety and Asn278 was not preserved.

Interestingly, most of the H-bond interactions involving compound 10a were preserved, and MD was able to identify new H-bond interactions. H-bond interactions involving N3H of DHPM and Asp167 and that of the N of the pyridine ring in quinoline with Ala110 were preserved at 96 and 92%, respectively. In comparison, that of the C = O of the DHPM with Arg272 was lost, but the same C = O was able to form two new H-bond interactions with Thr38 and Asn150 simultaneously at 72 and 94%, respectively. As for the newly identified H-bond interactions, these involved N1H of the DHPM with Gly37 and N3 of the triazole ring with Gly149 at 62 and 32%, respectively. The same trend was observed with the optimised ligands. Analysing LD_350, for instance, the MD did not preserve any of the identified H-bond interactions from the added R-group except for the newly formed intramolecular H-bond interaction between the 2-OH on the benzene ring and its C = O (See SI Fig. S2 b). The π-cation interactions between the quinoline rings and Lys32, as well as the DHPM ring and Lys151, are also noteworthy. These were preserved from the molecular docking interactions except for the newly identified π-cation interaction between the pyridine ring of the quinoline moiety and Lys32. Interestingly, the changes in either atomic distances or bond angles beyond the allowed limits resulted in the observed lost interactions in all the cases.

Using the root mean square fluctuation (RMSF) approach, we identified the protein regions exhibiting residue fluctuations across the simulation time. This clarifies how the flexibility of the protein is impacted by ligand binding. As expected, the graphs show the high flexibility of the protein’s N and C termini [ 39 ]. Although there were non-terminal residues with greater RSMF values, they were either extremely close to the terminal or very far from the protein’s binding pocket, as demonstrated by Gly67 in the 10a -protein complex, which had an RMSF value of 3.29 Å. All the protein residues interacting with the ligands generally displayed variations at a mean value of 1.0 Å (See SI Fig. S3 a, b, and c).

Click chemistry has been successfully used to link DHPM to quinoline-based compounds. The use of click chemistry offers several advantages over the previous amide coupling method. These include a better functional group tolerance, ease of synthesis, lesser waste generation, and better atom economy. From the antimalarial results obtained, we have been able to show that the hybrids containing DHPM and 4-aminoquinoline moieties with triazole linker at position 6 of the DHPM showed good antimalarial activity with IC 50 values below 1 µM and decreased resistance indices than CQ. In general, hybrids with 4-amino functionality displayed better antimalarial activities than those containing oxygen at position 4 of the quinoline ring. The hybrids with mefloquine moiety could be modified in the future to include a 4-amino group on the quinoline ring for comparative antimalarial activities. Apart from position 6 of the DHPM, other positions on the DHPM ring could also be explored to link the quinolines using triazole linkers. Using molecular docking and dynamics approaches, we further explored the synthesised compounds as glutathione reductase inhibitors. Five of the synthesised compounds showed higher binding affinity than the CQ standard, though none showed a higher binding affinity than the FAD cofactor. Rational optimisation of the compound with the highest binding affinity generated 916 ligands, with 664 having higher binding affinities than the original compound and 16 ligands higher than the FAD cofactor. The rational optimisation approach was necessary to produce ligands that could serve as competitive inhibitors of glutathione reductase.

Experimental

A Gallenkamp melting-point apparatus was used for the melting point determination in open capillary tubes and is uncorrected. 1 H and 13 C NMR spectra were recorded on either a Bruker Avance 400 (at 400.21 MHz for 1 H and 100.64 MHz for 13 C) or 300 (at 300.13 MHz for 1 H and 75.48 MHz for 13 C) spectrometers using CDCl 3 or DMSO-d 6 as solvents at room temperature. Chemical shifts were recorded as part per million (ppm) using tetramethylsilane as an internal standard. 2D NMR experiments were recorded on Bruker Avance 400. Mass analysis was performed on Waters® Synapt G2 High Definition Mass Spectrometry (HDMS) system with flow injection analysis (FIA) using electrospray ionisation (ESI) probe. The MS data were acquired and processed on MassLynx™ software (version 4.1). FT-IR measurements were made on a Bruker Alpha Platinum-ATR spectrometer as neat. All reagents and solvents were purchased from Sigma-Aldrich and used without further purification.

X-ray crystallography

Single-crystal diffraction experiments of compounds 3 and 6 were performed using Quazar multi-layer optics monochromated Cu Kα radiation ( k  = 1.54178 Å) on a Bruker D8 Venture kappa geometry diffractometer with duo Iμs sources, a Photon 100 CMOS detector and APEX III control software [ 42 ]. Data reduction was performed using the SAINT + [ 42 ], and the intensities were corrected for absorption using the SADABS [ 42 ]. Single crystals of 7 , 11d , and 11 g were analysed on a Rigaku XtaLAB Synergy R diffractometer with a rotating-anode X-ray source and a HyPix CCD detector. Data reduction and absorption were carried out using the CrysAlis Pro (version 1.171.40.23a) software package [ 43 ]. All X-ray diffraction measurements were performed at 150(1) K, using an Oxford Cryogenics Cryostat. All structures were solved by direct methods with SHELXT-2016 [ 44 ] using the SHELXL-2016 algorithm [ 45 ]. All H atoms were placed in geometrically idealised positions and constrained to ride on their parent atoms. For data collection and refinement parameters, see the SI (Tables S1 , S2 ). The X-ray crystallographic coordinates for all structures have been deposited at the Cambridge Crystallographic Data Centre (CCDC), with deposition numbers CCDC 2076919–2076923. The data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif .

Synthesis of 7-chloro-N-(prop-2-yn-1-yl)quinolin-4-amine (2)

thesis on antimalarial drugs

Synthesis of 7-chloro-4-(prop-2-yn-1-yloxy)quinoline (3)

thesis on antimalarial drugs

Synthesis of 2,8-bis(trifluoromethyl)quinolin-4-ol (6)

thesis on antimalarial drugs

Colourless crystals: yield 26.64 g, 79%; 1 H NMR (400 MHz, CDCl 3 ): δ H 8.62 (d, 3 J HH  = 8.3 Hz, 1H, Ar H ), 8.10 (d, 3 J HH  = 7.4 Hz, 1H, Ar H ), 7.61 (t, 3 J HH  = 7.8 Hz, 1H, Ar H ), 7.45 (s, 1H, py H ), and 6.95 (bs, 1H, OH).

Synthesis of 4-(prop-2-yn-1-yloxy)-2,8-bis(trifluoromethyl)quinoline (7)

thesis on antimalarial drugs

Synthesis of 4-(but-3-yn-1-yloxy)-2,8-bis(trifluoromethyl)quinoline (8)

thesis on antimalarial drugs

General procedure for the coupling of azido DHPMs to acetylenic quinolines

The general procedure for the click chemistry used for coupling azido DHPHs to acetylenic quinolines is described below using 10 h as an example.

9d (0.19 g, 0.60 mmol) and 3 (0.13 g, 0.60 mmol) were dissolved in 3.0 mL DMF. To this mixture was added freshly prepared 1-M aqueous solution of Na ascorbate (240.0 µL, 0.24 mmol) and CuSO 4 (120.0 µL. 0.12 mmol) in sequential. The reaction mixture was stirred at ambient temperature overnight. The mixture was poured into distilled water (25.0 mL) and extracted with EtOAc (3 × 10.0 mL). The organic layers were combined and washed with brine (10.0 mL), distilled water (5 × 10.0 mL), dried with MgSO 4 , and filtered. The filtrate was concentrated in vacuo to give the crude product, which was purified on column chromatography with SiO 2 using 6% MeOH in DCM as the eluent. See the supplementary information for the full characterisations of the final compounds.

Availability of data and materials

Some of the datasets generated during and/or analysed during the current study are in the manuscripts while the rest are available from the corresponding authors on reasonable request.

WHO, World Malaria Report (2019)   https://www.who.int/malaria ,   https://www.who.int/publications-detail/world-malaria-report-2019

Menezes CMS, Ferreira EI (2005) Modulating agents in resistant malaria. Drug Des Rev Online 2:409–418. https://doi.org/10.2174/1567269054546433

Article   CAS   Google Scholar  

Hu Y-QQ, Gao C, Zhang S, Xu L, Xu Z, Feng LS, Wu X, Zhao F (2017) Quinoline hybrids and their antiplasmodial and antimalarial activities. Eur J Med Chem 139:22–47. https://doi.org/10.1016/j.ejmech.2017.07.061

Article   CAS   PubMed   Google Scholar  

Van DA, Schalkwyk JC, Walden PJ (2001) Smith, Reversal of chloroquine resistance in Plasmodium falciparum using combinations of chemosensitizer. Antimicrob Agents Chemother 45:3171–3174. https://doi.org/10.1128/AAC.45.11.3171

Article   Google Scholar  

Wright DJ, O’Reilly M, Tisi D (2018) Engineering and purification of a thermostable, high-yield, variant of PfCRT, the Plasmodium falciparum chloroquine resistance transporter. Protein Expr Purif 141:7–18. https://doi.org/10.1016/j.pep.2017.08.005

Martin SK, Oduola AM, Milhous WK (1987) Reversal of chloroquine resistance in Plasmodium falciparum by verapamil. Science 235:899–901. https://doi.org/10.1128/AAC.45.11.3171

Alibert-Franco S, Pradines B, Mahamoud A, Davin-Regli A, Pages J-M (2008) Efflux mechanism, an attractive target to combat multidrug resistant Plasmodium falciparum and Pseudomonas aeruginosa. Curr Med Chem 16:301–317. https://doi.org/10.2174/092986709787002619

Egan TJ, Kaschula CH (2007) Strategies to reverse drug resistance in malaria. Curr Opin Infect Dis 20:598–604. https://doi.org/10.1097/QCO.0b013e3282f1673a

Warhurst DC (2003) Polymorphism in the Plasmodium falciparum chloroquine-resistance transporter protein links verapamil enhancement of chloroquine sensitivity with the clinical efficacy of amodiaquine. Malar J 2:1–12. https://doi.org/10.1186/1475-2875-2-1

Henry M, Alibert S, Orlandi-Pradines E, Bogreau H, Fusai T, Rogier C, Barbe J, Pradines B (2006) Chloroquine resistance reversal agents as promising antimalarial drugs. Curr Drug Targets 7:935–948. https://doi.org/10.2174/138945006778019372

Krogstad DJ, Gluzman IY, Kyle DE, Oduola AM, Martin SK, Milhous WK, Schlesinger PH (1987) Efflux of chloroquine from Plasmodium falciparum: mechanism of chloroquine resistance. Science 238:1283–1285. https://doi.org/10.1126/science.3317830

Millet J, Torrentino-Madamet M, Alibert S, Rogier C, Santelli-Rouvier C, Mosnier J, Baret E, Barbe J, Parzy D, Pradines B (2004) Dihydroethanoanthracene derivatives as in vitro malarial chloroquine resistance reversal agents. Antimicrob Agents Chemother 48:2753–2756. https://doi.org/10.1128/aac.48.7.2753-2756.2004

Article   CAS   PubMed   PubMed Central   Google Scholar  

Kyle DE, Oduola AMJJ, Martin SK, Milhous WK (1990) Plasmodium falciparum: modulation by calcium antagonists of resistance to chloroquine, desethylchloroquine, quinine, and quinidine in vitro. Trans R Soc Trop Med Hyg 84:474–478. https://doi.org/10.1016/0035-9203(90)90004-X

Burgess SJ, Selzer A, Kelly JX, Smilkstein MJ, Riscoe MK, Peyton DH (2006) A chloroquine-like molecule designed to reverse resistance in Plasmodium falciparum. J Med Chem 49:5623–5625. https://doi.org/10.1021/jm060399n

October N, Watermeyer ND, Yardley V, Egan TJ, Ncokazi K, Chibale K (2008) Reversed chloroquines based on the 3,4-dihydropyrimidin-2(1H)-one scaffold: synthesis and evaluation for antimalarial, β-haematin inhibition, and cytotoxic activity. ChemMedChem 3:1649–1653. https://doi.org/10.1002/cmdc.200800172

Kaur K, Jain M, Reddy RP, Jain R (2010) Quinolines and structurally related heterocycles as antimalarials. Eur J Med Chem 45:3245–3264. https://doi.org/10.1016/j.ejmech.2010.04.011

Becker K, Rahlfs S, Nickel C, Schirmer RH (2003) Glutathione–functions and metabolism in the malarial parasite Plasmodium falciparum. Biol Chem 384:551–566

Kamaria P, Kawathekar N (2014) Ligand-based 3D-QSAR analysis and virtual screening in exploration of new scaffolds as Plasmodium falciparum glutathione reductase inhibitors. Med Chem Res 23:25–33. https://doi.org/10.1007/s00044-013-0603-7

Fagan RL, Palfey BA (2010) 7.03 - Flavin-dependent enzymes, In: Ben HW, Liu LBT, Mander CNPII (ed) Comprehensive Natural Products II Chemical Biology, Elsevier, Oxford. p 37–113. https://doi.org/10.1016/B978-008045382-8.00135-0

Pastrana-Mena R, Dinglasan RR, Franke-Fayard B, Vega-Rodríguez J, Fuentes-Caraballo M, Baerga-Ortiz A, Coppens I, Jacobs-Lorena M, Janse CJ, Serrano AE (2010) Glutathione reductase-null malaria parasites have normal blood stage growth but arrest during development in the mosquito. J Biol Chem 285:27045–27056

Adigun RA, Malan FP, Balogun MO, October N (2021) Substitutional effects on the reactivity and thermal stability of dihydropyrimidinones. J Mol Struct 1223:129193. https://doi.org/10.1016/j.molstruc.2020.129193

Fisher GM, Tanpure RP, Douchez A, Andrews KT, Poulsen S-A (2014) Synthesis and evaluation of antimalarial properties of novel 4-aminoquinoline hybrid compounds. Chem Biol Drug Des 84:462–472. https://doi.org/10.1111/cbdd.12335

Kaval N, Ermolat'ev D, Appukkuttan P, Dehaen W, Kappe CO, Van der Eycken E (2005) The application of “click chemistry” for the decoration of 2(1h)-pyrazinone scaffold: generation of templates. J Comb Chem 7:490–502. https://doi.org/10.1021/cc0498377

Eswaran S, Vasudeva A, Chowdhury IH, Pal NK, Thomas KD (2010) New quinoline derivatives: synthesis and investigation of antibacterial and antituberculosis properties. Eur J Med Chem 45:3374–3383. https://doi.org/10.1016/j.ejmech.2010.04.022

Mao J, Yuan H, Wang ÂY, Wan B, Pieroni M, Huang Q, Van Breemen RB, Kozikowski PA, Scott GF (2009) From serendipity to rational antituberculosis drug discovery of mefloquine-isoxazole carboxylic acid esters. J Med Chem 52:6966–6978. https://doi.org/10.1021/jm900340a

Lilienkampf A, Jialin M, Baojie W, Yuehong W, Franzblau SG, Kozikowski AP (2009) Structure-activity relationships for a series of quinoline-based compounds active against replicating and nonreplicating Mycobacterium tuberculosis. J Med Chem 52:2109–2118. https://doi.org/10.1021/jm900003c

Brouet JC, Gu S, Peet NP, Williams JD (2009) Survey of solvents for the Conrad-Limpach synthesis of 4-hydroxyquinolones. Synth Commun 39:1563–1569. https://doi.org/10.1080/00397910802542044

Li JJ (2014) Name reactions: a collection of detailed mechanisms and synthetic applications, 5th ed. Springer, Science and Business Media

Sarojini BK, Narayana B, Mayekar AN, Yathirajan HS, Bolte M (2007) A 1:1 cocrystal of 2,8-bis-(trifluoro-methyl)quinolin-4-ol and 2,8-bis-(tri-fluoromethyl)quinolin-4(1H)-one, Acta Crystallogr. Sect E Struct Reports Online 63:7–13. https://doi.org/10.1107/S1600536807053482

Keto-enol tautomerism : key points - Master Organic Chemistry (nd)  https://www.masterorganicchemistry.com/2022/06/21/keto-enol-tautomerism-key-points/ (accessed 25 January 2023)

Guantai EM, Ncokazi K, Egan TJ, Gut J, Rosenthal PJ, Smith PJ, Chibale K (2010) Design, synthesis and in vitro antimalarial evaluation of triazole-linked chalcone and dienone hybrid compounds, Bioorganic. Med Chem 18:8243–8256. https://doi.org/10.1016/j.bmc.2010.10.009

Chu XM, Wang C, Wang WL, Liang LL, Liu W, Gong KK, Sun KL (2019) Triazole derivatives and their antiplasmodial and antimalarial activities. Eur J Med Chem 166:206–223. https://doi.org/10.1016/j.ejmech.2019.01.047

Chopra R, Chibale K, Singh K (2018) Pyrimidine-chloroquinoline hybrids: synthesis and antiplasmodial activity. Eur J Med Chem 148:39–53. https://doi.org/10.1016/j.ejmech.2018.02.021

Natarajan JK, Alumasa JN, Yearick K, Ekoue-Kovi KA, Casabianca LB, De Dios AC, Wolf C, Roepe PD (2008) 4-N-, 4-S-, and 4-O-chloroquine analogues: influence of side chain length and quinolyl nitrogen pKa on activity vs chloroquine resistant malaria. J Med Chem 51:3466–3479. https://doi.org/10.1021/jm701478a

Taleli L, De Kock C, Smith PJ, Pelly SC, Blackie MAL, Van Otterlo WAL (2015) In vitro antiplasmodial activity of triazole-linked chloroquinoline derivatives synthesised from 7-chloro-N-(prop-2-yn-1-yl)quinolin-4-amine. Bioorg Med Chem 23:4163–4171. https://doi.org/10.1016/j.bmc.2015.06.044

Slater AFG (1993) Chloroquine: mechanism of drug action and resistance in plasmodium falciparum. Pharmacol Ther 57:203–235. https://doi.org/10.1016/0163-7258(93)90056-J

Joshi MC, Wicht KJ, Taylor D, Hunter R, Smith PJ, Egan TJ (2013) In vitro antimalarial activity, β-haematin inhibition and structure-activity relationships in a series of quinoline triazoles. Eur J Med Chem 69:338–347. https://doi.org/10.1016/j.ejmech.2013.08.046

Adigun RA, Malan FP, Balogun MO, October N (2022) Rational optimization of dihydropyrimidinone-quinoline hybrids as Plasmodium falciparum glutathione reductase inhibitors. ChemMedChem 17:1–15. https://doi.org/10.1002/cmdc.202200034

Mensah JO, Ampomah GB, Gasu EN, Adomako AK, Menkah ES, Borquaye LS (2022) Allosteric modulation of the main protease (MPro) of SARS-CoV-2 by casticin—insights from molecular dynamics simulations. Chem Africa 5:1305–1320. https://doi.org/10.1007/s42250-022-00411-7

Kalibaeva G, Ferrario M, Ciccotti G (2003) Constant pressure-constant temperature molecular dynamics: a correct constrained NPT ensemble using the molecular virial. Mol Phys 101:765–778

Martyna GJ (1994) Remarks on ‘“Constant-temperature molecular dynamics with momentum conservation”,.’ Phys Rev E 50:3234

APEX3 (including SAINT and SADABS) (2016) Bruker AXS Inc. Madison, WI

Rigaku Oxford Diffraction (2018) CrysAlisPro software system

Sheldrick GM (2015) SHELXT - integrated space-group and crystal-structure determination. Acta Crystallogr Sect A Found Crystallogr 71:3–8. https://doi.org/10.1107/S2053273314026370

Sheldrick GM (2015) Crystal structure refinement with SHELXL, Acta Crystallogr. Sect C Struct Chem C 71:3–8

Google Scholar  

Download references

Acknowledgements

The authors would also like to thank the Centre for High Performance Computing (CHPC, Cape Town, South Africa) for access to CHPC Lengau Cluster and Schrödinger molecular docking software.

Open access funding provided by University of Pretoria. This work was supported by the National Research Foundation (South Africa) (Grant number: 105152) and the University of Pretoria, South Africa (doctoral financial support to Rasheed A. Adigun).

Author information

Authors and affiliations.

Department of Chemistry, University of Pretoria, Lynnwood Road, 0002, Hatfield, Pretoria, South Africa

Rasheed A. Adigun, Frederick P. Malan & Natasha October

Biopolymer Modification & Therapeutics Lab, Chemicals Cluster, Council for Scientific and Industrial Research, Meiring Naude Road, 0001, Brummeria, Pretoria, South Africa

Mohammed O. Balogun

You can also search for this author in PubMed   Google Scholar

Contributions

All authors contributed to the study. Conception and design, material preparation, data collection, and analysis were performed by Rasheed A. Adigun, Frederick P. Malan, Mohammed O. Balogun, and Natasha October. The first draught of the manuscript was written by Rasheed A. Adigun, and all authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Rasheed A. Adigun or Natasha October .

Ethics declarations

Ethics approval.

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file1 (PDF 1.98 MB)

Rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Adigun, R.A., Malan, F.P., Balogun, M.O. et al. Design, synthesis, and in silico-in vitro antimalarial evaluation of 1,2,3-triazole-linked dihydropyrimidinone quinoline hybrids. Struct Chem 34 , 2065–2082 (2023). https://doi.org/10.1007/s11224-023-02142-y

Download citation

Received : 26 December 2022

Accepted : 07 February 2023

Published : 27 February 2023

Issue Date : December 2023

DOI : https://doi.org/10.1007/s11224-023-02142-y

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Antimalarial
  • Click chemistry
  • Dihydropyrimidinone
  • Resistance reversal

Advertisement

  • Find a journal
  • Publish with us
  • Track your research

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Int J Health Sci (Qassim)
  • v.15(6); Nov-Dec 2021

Pharmacokinetic and molecular docking studies to design antimalarial compounds targeting Actin I

Vandana guleria.

1 Department of Biotechnology, Faculty of Applied Sciences and Biotechnology, Shoolini University of Biotechnology and Management Sciences, Solan, Himachal Pradesh, India

2 Department of Biotechnology, Vignan’s Foundation for Science, Technology and Research, Vadlamudi, Guntur, Andhra Pradesh, India

Bhanu Sharma

Shweta chauhan, varun jaiswal.

3 Department of Engineering, School of Electrical and Computer Science Engineering, Shoolini University of Biotechnology and Management Sciences, Solan, Himachal Pradesh, India

4 Department of Food and Nutrition, College of BioNano Technology, Gachon University, Seongnam-si 13120, Gyeonggi-do, Korea

Associated Data

The data used in this study are available and will be provided by the corresponding author on a reasonable request.

Malaria is an ancient disease that still causes more than 200 million of cases 7 with high mortality globally. Identification of new drug targets and development of novel antimalarial drugs with unique mode of action encounter the drug resistance and reduce the mortality by Plasmodium parasites. Actin protein is one of the key proteins in Plasmodium falciparum playing multifarious important roles including transport, cell motility, cell division, and shape determination. This study investigated Actin I as a drug target, in silico screening of diverse molecules through molecular docking was considered. Further, pharmacokinetic parameters of the selected molecules from the docking and interaction studies were planned to propose the lead molecules.b

Molecules were selected according to score and protein ligand interaction and selected molecules were subjected for pharmacokinetic studies to investigate important drug parameters.

The docked molecules were ranked according to the binding score and good interaction pattern was observed with Actin I within top 20 scoring molecules. The selected molecules also had optimum pharmacokinetic parameters.

Conclusion:

The current study provides a set of hit molecules which can be further explored through in vitro and in vivo experiments for the development of potential drugs against malaria, there by encountering drug resistance and establishing Actin I as an important drug target.

Introduction

Malaria is a life-threatening disease caused by parasites belonging to five species of genus Plasmodium falciparum which are transmitted through the bite of female Anopheles mosquito. In 2017 as per the WHO report, 219 million cases of malaria were reported in 87 countries and the estimated number of deaths was found to be 435,000. Despite such drastic figures there is still no widely used efficacious vaccine available for malaria parasites.[ 1 ] Although, antimalarial drugs are available from long back and the popular antimalarial medications used against malarial parasites includes Chloroquine, Chloroguanide (Proguanil), Sulfadoxine/pyrimethamine, Quinine, Mefloquine, Halofantrine, Artemisinin (ART), and Atovaquone. In 21 st century, the drug resistance has emerged to almost all classes of antimalarial drugs and therefore a sudden leap has been witnessed in malaria-related mortality, especially in Africa.[ 1 ]

Both drug resistance and unavailability of highly effective vaccine makes malaria a highly challenging disease and a great public health burden. The identification of new/novel drug targets as well as development of novel antimalarial drugs with unique mode of action can be the possible answer to drug resistance in malaria and subsequently it can control the mortality due to malaria.[ 2 , 3 ] Being highly studied diseases several potential drug targets have been suggested in research which can be further developed as a drug target in experiments. Actin I is also such kind of drug target in malaria. Recently, our group has discovered that Actin I protein is expressed in almost all stages of P. falciparum which are present in human which justifies its importance in plasmodium lifecycle and its drug target potential.[ 4 ] Although in different literature actin is proposed as the drug target in malaria, it is not yet studied for structure based drug design and subsequently not developed as an important drug target.[ 4 , 5 ] The main reasons for the selection of the Actin I in the current study are the Actin I protein is novel drug target which is important in different stages of P. falciparum and can be a probable answer to drug resistance.[ 4 ] Although the Actin I protein is conserved in both, there is difference between human and P. falciparum actin protein. In human, three isoforms of actin have been identified, alpha, beta, and gamma whereas in P. falciparum there exists only two isoforms, that is, Actin I and Actin II.[ 6 , 7 ] This protein is an essential part of the parasite motor machinery and responsible for gliding motility of the parasite.[ 8 , 9 ] In nature, there exists six different mammalian actin isoforms which differ from each other by a maximum of 6% of the sequence, and are practically identical across the species. Commonly most actins have the capacity to form long filaments but in P. falciparum the presence of regular actin filaments is uncertain. In P. falciparum Actin I is abundant and expressed throughout the lifecycle of the parasite. In vitro, apicomplexan actins form short, ~100-nm long filaments, which suffer from rapid tread milling.[ 10 , 11 ] Recently, specific antibodies revealed filament-like structures in motile forms of P. falciparum .[ 12 ] The P. falciparum Actin I fibers are refined from merozoites and are thereby distinctive for their helical symmetry from canonical actin structures.[ 13 , 14 ]

The most prevalent forms of actin protein are ATP-G-actin and ADP-F-actin. The structural determination of G-actin reveals that it has two lobes separated by a cleft, which is an ATPase fold. The functional form of G-actin exhibits only when the cleft comprises of an ATP or ADP. The F-actin polymer is a filamentous protein which is also referred as microfilament with levorotatory helix structure. The cell creeping process helps actin cytoskeleton to encounter dynamic assembly and disassembly, which further controls central attachment of assembly/disassembly bulge advancement, and contractile fiber association. Moreover cell relocation and adhesion is inhibited by the disturbance of the actin cytoskeleton.[ 15 - 17 ] The assembly and disassembly of local Actin I filament regulates the dynamic formation of lamellipodia. The lamellipodia is slim, sheet-like film distensions of the motile cells. During movement, cells grow the film forward to explore their environment. The branching and elongation are the two examples of actin fiber assembly, which advance the improvement of the actin “work” in the cell distension.[ 16 , 18 ] The nucleation promoting factors such as neuronal Wiskott - Aldrich syndrome Protein (N-WASP) and WASP-family verprolin-homologous protein (WAVE) prohibit the movement of Arp2/3 complex, which are in turn modulated by upstream regulators.[ 16 ] The Actin I fibers are organized into two kinds of arrays: Bundles and web like systems. Similarly, the Actin I fibers cross-connecting the proteins help to stabilize and keep up these structures and are divided into two classes: Packaging proteins and web-forming proteins. Actin I plays an important role in the intracellular blood stages of P. falciparum , such as endocytosis and vesicle trafficking,[ 18 , 19 ] ring stage morphology,[ 20 ] and spatial positioning of genes in the nucleus.[ 21 ] As per the recent studies, the actin elements have appeared to assume significant roles in the inheritance of intracellular organelles, for example, the apicoplast, mitochondria and secretory vesicles during intracellular parasite replication.[ 21 - 24 ] The Actin I association has also been well studied in zoite and asexual blood stages. In this intracellular stage, the parasite experiences a wonderful morphological change during which the parasite embraces a crescent or falci from shape in anticipation for transmission by means of the Anopheles mosquito.[ 25 , 26 ] In Actin I structure the C terminus acquires the shape of a-helix and it interacts with the bottom of the sub domain 1. Actin I protein consist of cleft named hydrophobic cleft between the sub-domains I and III. This cleft is major regulatory site and regarded as hotspot for the protein binding.[ 7 ] The binding pocket of Actin I has two parameters, the “mouth” and “phosphate clamp.” The “mouth” is considered as the distance between the Ca atoms of Gln59and Glu207. The distance between the Caatoms of Gly15 and Asp157 makes the “phosphate clamp.” As compared to the canonical actins, in Actin I there is a smaller cleft where large hydrophobic residues have adopted different conformations together withTrp357. Actin I shows significant structural deviations in specific regions which are involved in binding of proteins.[ 7 , 27 ] This may provide opportunities for structure-based drug design targeted at the P. falciparum act in-regulatory boundaries. The importance of Actin I protein is established in most of the stages of parasite. The crucial function of Actin I in asexual, blood, liver, mosquito and sexual stages is reported in the literature which also correlates with our earlier finding of its expression in nearly all stages of P. falciparum .[ 4 ] The active site residues of protein consist of amino acids, that is, Asn 17(A), Gly 16(A), Lys 19 (A), Gly 303(A), Glu 215(A), Lys 214(A), Val 160 (A), and Gly 159(A).

In current study, Actin I protein was taken for molecular docking studies because it is a new potential drug target and relatively not explored as drug target in drug discovery in spite of its conservation in different species of P. falciparum . In P. falciparum intracellular functions of Actin I have been recommended, that is, endocytosis,[ 20 ] secretion, and antigenic variation.[ 18 , 28 , 29 ] Molecular docking is an important tool for drug discovery and can be used to model the interaction between small drug like molecule and a protein at atomic level, which permit us to describe the enactment of small molecules in the binding sites of target proteins as well as to irradiate the biochemical processes. Docking is very useful and one of the most commonly used methods in structure-based drug design, due to its capability to predict the binding-conformation of small molecule ligands to the appropriate target binding site.[ 4 , 30 - 32 ] In the current study, the library of small molecules is screened against Actin I protein. The top scoring molecules from docking study were subjected for protein-ligand interaction studies and prediction of pharmacokinetics properties. Favorable pharmacokinetics property is also essential parameter for the ligand to be considered in drug discovery process. Therefore, the best molecules according to these three studies can be strongly suggested for further wet lab experiments of drug design.

Materials and Methods

Retrieval of protein and ligand libraries.

The crystal structure of Actin I of P. falciparum (PDB ID: 4CBU) was downloaded from protein databank in pdb format [ https://www.rcsb.org/structure/4CBU ]. For virtual screening organic molecule library from NPACT database consisting of total 1574 molecules ( https://zinc.docking.org/browse/catalogs/natural-products ) was downloaded from zinc database in mol2 format.

Target protein preparation

The crystal structure of P. falciparum Actin I (PDB ID: 4CBU) is of 1.3 A ° resolution with bound ligand ATP and have two chains; chain A and chain G. It was observed that the binding pocket is located in chain-A and chain-G was away from binding pocket. Hence, chain-G was removed through UCSF Chimera before the protein preparation and in further docking studies only chain-A was considered. The bound ligand molecule was also removed before docking. The protein was prepared for docking studies by assigning hydrogen, polaraties, calculating Gasteiger charges to protein structures, and converting protein structures from the pdb format to pdbqt format using Auto-Dock tool1.5.4.

Preparation of ligand libraries for virtual screening

The natural occurring plant-based compound activity target database library was downloaded from ZINC database ( https://zinc.docking.org/ )in MOL2 format and was further used for performing docking studies. The library was in a single molecular file and all the ligand molecules were extracted from it with the help of in-house developed PERL scripts. Subsequently all the ligand molecules were converted into PDB format from MOL2 format by using open babel with the help of in-house developed PERL scripts. Antimalarial drugs including Chloroquine, Chloroguanide, Quinine, Mefloquine, and Halofantrine were also used as positive control to target ACTIN I and structure of these molecules were taken from PubChem ( https://pubchem.ncbi.nlm.nih.gov/ ).

Molecular docking

All the ligands within the organic molecular library were used to perform molecular docking studies[ 33 , 34 ] against Actin I to find the potential hit molecules for further drug discovery experiments. In the current research for performing the docking studies the version 4.2 of AutoDock was used. AutoDock uses a Lamarckian Genetic Algorithm (LGA) and is based on a semi empirical free energy force field.[ 35 ] The docking grid was set manually through visualization of the protein and grid was defined to cover entire binding site of Actin I.

The size of grid was defined as X=54, Y=54 and Z=54 and the coordinates used for docking of ligands library with 4CBU were X=9.175, Y=20.366, and Z=18.583. Also for molecular docking for every molecule ten runs of LGA were performed. The main interacting residues in the binding site were Asn17, Lys19, Tyr338, Lys337, Lys214, Gln215, and other nearby residues. Finally, molecular docking was carried out on target proteins with all ligands from the library using self-developed PERL script, which was used for screening of these large number of ligand molecules one by one [ Figure 1 ]. The top scoring compounds were selected on the basis of binding energy of ligands with the receptor. The lower is the binding energy of ligand with the receptor the more strongly it will bound to the target receptor.

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g001.jpg

The methodology followed for docking and screening

Interaction study and visualization of docked complex

The UCSF Chimera1.8.1 was used to visualize and analyze docked complexes of ligand and protein.[ 36 ] The interaction such as hydrogen bonds and hydrophobic interaction was also analyzed using Ligplot+.[ 37 , 38 ] The generated plots have shown the hydrophobic interaction patterns and hydrogen bonding between the main chain/side-chain atoms of the protein with the ligand.

ADME-toxicity prediction

The pharmacokinetic profile prediction of potential compounds was done using pkCSM ( http://biosig.unimelb.edu.au/pkcsm/prediction ) with incorporated prediction of toxicity, metabolism, distribution, absorption, and excretion.[ 39 ] Some other properties such as molecular descriptors and drug likeliness properties were also analyzed with the help of pkCSM. The drug likeliness properties were predicted according to the Lipinski Rules of five.[ 40 ]

In the present study, considering the importance of Actin I protein as a drug target in P. falciparum molecular docking of actin I of P. falciparum was carried out. The organic molecular library was docked into the active site of Actin I to find the potential ligand molecules which can be used in further drug design experiments and also establish the Actin I as an important drug target in malaria. Pharmacokinetics properties of top ranked molecules were also calculated computationally so that molecules with better pharmacokinetics properties can be further shortlisted for wet lab experiments.

Docking analysis

Molecular docking of all the ligands within the library was found to be successful on the basis that all the ligands were docked in the active site of the receptor. The 20 top scoring molecules as per estimated free energy of binding (EFEB) were selected for further investigations. The protein ligand complexes of top scoring molecules were analyzed for interactions, the orientation of the docked compound, and interacting active site residues were visualized. The selected 20 molecules had EFEB in the range of −13.20–−10.51 [ Table 1 ]. The molecule, that is, ZINC30724344 shows least EFEB of −13.20 [ Table 1 ]. The four other top scoring molecules ZINC13515285, ZINC03943903, ZINC95098956, and ZINC03984030 showed the EFEB −12.61, −12.57, −12.37, and −12.23, respectively [ Table 1 ] [Supplementary Figures ​ Figures1 1 - ​ -3]. 3 ]. Despite these molecules the other molecules also had EFEB better than-10 and also show good interaction with the target molecule [ Table 1 ] [ Supplementary Table 2 ] [ Supplementary Table 3 ].

Binding affinity of selected potential molecules

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g002.jpg

The top scoring molecule, that is, ZINC30724344 strongly interacts with the Glu215, Lys19 and Asn17 residues of the binding cavity [ Figure 2 ] [ Table 1 ]. The second-best scoring molecule as per the EFEB, that is, ZINC13515285 forms hydrogen bonds with the Lys337, Asn17, Lys19, Lys214, Tyr307, Thr304, and Gly303 [ Figure 3 ] [ Table 1 ].

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g003.jpg

(a and b) Interactions of ZINC30724344 with Actin I

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g004.jpg

(a and b) Interaction of ZINC13515285 with Actin I

The molecule ZINC03943903 strongly interacts with the Lys19, Ca1370, Lys337, Gly303, Gly16, Asn17, Asp158, Thr304, and Lys214. All other selected molecules also show good patterns of hydrogen bond and hydrophobic interactions [ Table 1 ].

Among the five antimalarial drugs which were used for control docking Mefloquine exhibit the best estimated free energy of −8.5 kcal/mol with Actin I and interact strongly with Thr278, Ala273 [ Supplementary Table 1 ]. The other drugs which were used as positive controls, that is, Chloroquine, Chloroguanide, Quinine, and Halofantrine show the EFEB −8.2, −7.1, −7.6, and −8.0, respectively [ Supplementary Table 1 ].

Molecular parameters

The drug molecules which are having molecular weight less than 500Da can be easily transported, diffused and absorbed as compared to bigger molecules.[ 41 ] The physicochemical properties were calculated and the molecular weight of 10 hit molecules out of 20 was less than 500 Da. Most of the molecules [ Table 1 ] showed log P < 5 [ Table 1 ]. The 13 out of 20 selected hit molecules have less than 5 hydrogen bond donors and also have less than 10 hydrogen bond acceptors [ Table 2 ].

Physicochemical properties of potential compounds

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g005.jpg

ADMET properties

The selected 20 top scoring molecules were subjected to ADMET properties prediction. In ADMET, the absorption properties such as intestinal solubility (% absorbed), water solubility (log mol/L), and skin permeability (logKp) values were calculated. The other properties such as Ames test, BBB penetrability, and Caco2 (cell line of human colorectal adenocarcinoma cells) permeability was included under ADMET. For human intestinal absorption, the compounds which have value less than 30% are poorly absorbed [ 41 , 42 ] Out of the 20 selected hit molecules only three molecules depicted the intestinal absorption value less than 30% and other molecules displayed good absorption potential in human intestine [ Table 3] . The selected top 20 scoring molecules also showed good skin permeability values as compared to standard value (−2.5 logKp) [ 41 ] and only few can penetrate to Caco2 as compared to (>0.90 as per pkCSM) standard value [Tables ​ [Tables3 3 and ​ and4 4 ].

Absorption profile of potential compounds

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g006.jpg

Distribution/excretion profile of compounds

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g007.jpg

The human microsomal P450s aromatase catalyze the metabolism of a wide variety of compounds including drugs and xenobiotic. In ADMET under metabolism profile, the top scoring molecules were tested whether they were acting as an inhibitor or substrate for different isoforms of cytochrome P450 as it is important for drug metabolism. Many drugs can be inactivated by cytochrome P450 whereas some can be activated also. The inhibitors of cytochrome P450 can affect metabolism of drugs. Hence, from the predicted values, it was observed that the ten molecules out of 20 can acts as a substrate for one of the isoforms of P450, that is, CYP3A4 and none of the molecule acts as a substrate for other isoforms, that is, CYP2D6. The predictive values of the selected hit compounds suggest that only few compounds were acting as inhibitors for CYP3A4, CYP2C19, CYP2C9 and CYP3A4 isoforms of the cytochrome [ Table 5 ].

Metabolism profile of potential compounds

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g008.jpg

The toxicity and mutagenicity of the selected molecules were depicted with the help of Ames test data, only single molecule, that is, ZINC95099527 displayed the positive results and remaining 19 molecules were note found to be toxic according to Ames test data [ Table 6 ]. Furthermore, for a given molecule maximum recommended tolerated dose (MRTD) ≤0.477 was considered as low and MRTD high if greater than 0.477 as per pkCSM, although most of selected molecules show low MRTD [ Table 6 ].

Toxicity profile of the potential compounds

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g009.jpg

Malaria is still a major global health problem, although antimalarial drugs are available drug resistance has emerged thereby increasing morbidity and mortality rate. The research for the development of efficient licensed malaria vaccine is still not over despite many rigorous efforts have been made in the last decade. It was previously suggested in the literature that the proteins such as actin, tubulin, and histone are involved in the structural assembly of the pathogen.[ 43 , 44 ] Hence, the potential new drug target Actin I is one of the most important structural proteins which is found in P. falciparum and plays important role in its motility apparatus, sexual, asexual and blood stages of the organism.[ 39 ] It is expressed in all human stages and little is known about its potential as a drug target. In the current repertoire, we have considered library of 1574 natural organic molecules of diverse nature which were docked with Actin I protein of P. falciparum employ in one of the most important docking software.[ 6 ] AutoDock 4.2. The selected molecules were further subjected to their pharmacokinetic studies to make them potentially promising drug candidates.[ 45 ] These approaches were widely utilized to design drugs against a number of pathologies such as cardiovascular diseases and cancer.[ 46 - 48 ] The current study is the first report to explore the potential of receptor protein Actin I as an efficient drug target in structure-based drug design.

The prospect of virtual screening techniques was employed to screen the 20 top scoring molecules on the basis of their binding affinity. Although EFEB of ligand protein interaction is relative score and direct correlation of this score is not studied in different docking studies, EFEB of selected ligands in current study was better than the ligands found to be active in recent in silico and in vitro studies conducted by our group in falcipains 2 and 3 of P. falciparum .[ 49 ] In addition, when the docked complexes of these selected hit molecules were subjected to post docking analysis, they displayed good interaction patterns (hydrogen as well as hydrophobic interactions). The compound with zinc ID ZINC30724344 was the top scoring molecule and also depicted good interactions indicating a potent promising candidate. Similar trend was also discovered for other remaining 19 molecules which were also prioritized as hit molecules displaying good binding affinities and interaction patterns [ Table 1 ]. Furthermore, the 20 top scoring molecules interact strongly with the target protein forming number of hydrogen bonds [ Table 1 ] [Figures ​ [Figures2 2 and ​ and3]. 3 ]. In most of these top scoring molecules Lys337, Asn17, and Lys19 were found to be most common interacting residues forming hydrogen bonding. The top scoring molecules exhibit better binding affinity in terms of EFEB as compared to the antimalarial drugs which were used as positive control [ Supplementary Table 1 ].

Further, these top scoring molecules were subjected to ADMET prediction and physicochemical analysis to ensure drug likeness of hit molecules as part of Lipinski’s rule of five. The ADMET properties are important criteria because significant number of molecules failed in the clinical trial due to poor ADMET properties.[ 50 ] The top scoring molecule from docking studies ZINC30724344 exhibited molecular weight of 592.696 Da, logP of5.01246, HBA of 6 and HBD value as 2. Furthermore, this hit molecule was non-toxic as per Ames test data and displayed good intestinal absorption (84.303%) and skin permeability (logKp-2.735) values. The molecular weight and logP were observed to be slightly higher as per the standard rules, but natural origin and all other parameters justified that the above potential hit molecule can be developed against malaria which can serve as a roadmap for perusing drug discovery process. The remaining molecules having zinc ID ZINC38145808, ZINC95099570, ZINC14883289, and ZINC02138728also showed high binding affinities and these molecules also justified the different essential parameters to be considered as hit molecules for the drug discovery [ Table 1 ] [Figures ​ [Figures2 2 and ​ and3]. 3 ]. Although current results are very promising, computational studies require further wet lab experimental verification including target validation. Moreover, current findings warrants that these selected hit molecules can be subjected to further wet lab experimentations for the efficacy of hits compound in vitro and in vivo studies against plasmodium and subsequently modification accordingly so that these molecules can be used as potential hit molecules for anti-malarial drug discovery.

First time virtual screening methodology was implemented successfully for Actin I of P. falciparum using structure-based drug design. Library of diverse molecules was screened to find potential hit molecules on the basis of EFEB. Protein ligand interaction and ADMET properties of top selected scoring molecules were found to be favorable. Among the selected top scoring molecules, the five molecules, that is, ZINC30724344, ZINC13515285, ZINC03943903, ZINC95098956, and ZINC03984030 looks more prominent as a potential candidate for further drug development. The current findings provide a suitable starting point for further in vitro and in vivo analyses to exploit Actin I as an optimal drug target which can encounter drug resistance.

Authors’ Declaration Statements

Ethics approval and consent to participate.

Not applicable (as no human subject was used in the study)

Consent for publication

Availability of data and material, competing interests.

None declared.

Funding Statement

This research received no specific grant from any funding agency

Authors’ contributions

The research “Conceptualization, V.J. and S.C.; Methodology, V.G.; Software, B.S. and V.G.; Validation, T.P., and B.S.; Formal Analysis, B.S. and V.G.; Data Curation, T.P. and B.S.; Writing – Original Draft Preparation, S.C. and V.G.; Writing – Review & Editing, V.J., T.P., B.S. and S.C.; Visualization, V.G. and B.S.; Supervision, V.J. and T.P.; Project Administration, V.J.;

Acknowledgement

Authors would like to acknowledge Shoolini University, Solan Himachal Pradesh, India.

ORCID link of the submitting author: https://orcid.org/0000-0001-8068-0672

Supplementary Data

Supplementary table 1.

Docking of Actin I with different Antimalarial drugs:

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g010.jpg

Supplementary Table 2

Top 20 molecules ki (Values)

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g011.jpg

Supplementary Table 3

Top 20 molecules structures

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g012.jpg

Supplementary Figure 1

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g013.jpg

Interactions of ZINC03984030 with Actin I

Supplementary Figure 2

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g014.jpg

Interactions of ZINC95098956 with Actin I

Supplementary Figure 3

An external file that holds a picture, illustration, etc.
Object name is IJHS-15-4-g015.jpg

Interactions of ZINC03943903 with Actin I

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Pharmacokinetic interactions of antimalarial agents

Affiliation.

  • 1 Division of Infectious Diseases, Tropical Medicine and AIDS, Academic Medical Center, Amsterdam, The Netherlands.
  • PMID: 11432537
  • DOI: 10.2165/00003088-200140050-00003

Combination of antimalarial agents has been introduced as a response to widespread drug resistance. The higher number of mutations required to express complete resistance against combinations may retard the further development of resistance. Combination of drugs, especially with the artemisinin drugs, may also offer complete and rapid eradication of the parasite load in symptomatic patients and thus reduce the chance of survival of resistant strains. The advantages of combination therapy should be balanced against the increased chance of drug interactions. During the last decade, much of the pharmacokinetics and metabolic pathways of antimalarial drugs have been elucidated, including the role of the cytochrome P450 (CYP) enzyme complex. Change in protein binding is not a significant cause of interactions between antimalarial agents. CYP3A4 and CYP2C19 are frequently involved in the metabolism of antimalarial agents. Quinidine is a potent inhibitor of CYP2D6, but it appears that this enzyme does not mediate the metabolism of any other antimalarial agent. The new combinations proguanil-atovaquone and chlorproguanil-dapsone do not show significant interactions. CYP2B6 and CYP3A4 are involved in the metabolism of artemisinin and derivatives, but further studies may reveal involvement of more enzymes. Artemisinin may induce CYP2C19. Several artemisinin drugs suffer from auto-induction of the first-pass effect, resulting in a decline of bioavailability after repeated doses. The mechanism of this effect is not yet clear, but induction by other agents cannot be excluded. The combination of artemisinin drugs with mefloquine and the fixed combination artemether-lumefantrine have been studied widely, and no significant drug interactions have been found. The artemisinin drugs will be used at an increasing rate, particularly in combination with other agents. Although clinical studies have so far not shown any significant interactions, drug interactions should be given appropriate attention when other combinations are used.

PubMed Disclaimer

Similar articles

  • Artemisinin derivatives and malaria: useful, in combination with other antimalarials. [No authors listed] [No authors listed] Prescrire Int. 2008 Aug;17(96):162-8. Prescrire Int. 2008. PMID: 19492494
  • Clinical pharmacology of artemisinin-based combination therapies. German PI, Aweeka FT. German PI, et al. Clin Pharmacokinet. 2008;47(2):91-102. doi: 10.2165/00003088-200847020-00002. Clin Pharmacokinet. 2008. PMID: 18193915 Review.
  • Investigation of pre-clinical pharmacokinetic parameters of atovaquone nanosuspension prepared using a pH-based precipitation method and its pharmacodynamic properties in a novel artemisinin combination. Kathpalia H, Juvekar S, Mohanraj K, Apsingekar M, Shidhaye S. Kathpalia H, et al. J Glob Antimicrob Resist. 2020 Sep;22:248-256. doi: 10.1016/j.jgar.2020.02.018. Epub 2020 Feb 29. J Glob Antimicrob Resist. 2020. PMID: 32119990
  • Treatment of falciparum malaria in the age of drug resistance. Shanks GD. Shanks GD. J Postgrad Med. 2006 Oct-Dec;52(4):277-80. J Postgrad Med. 2006. PMID: 17102546 Review.
  • Ketoconazole, a cytochrome P(450) inhibitor can potentiate the antimalarial action of α/β arteether against MDR Plasmodium yoelii nigeriensis. Tripathi R, Rizvi A, Pandey SK, Dwivedi H, Saxena JK. Tripathi R, et al. Acta Trop. 2013 May;126(2):150-5. doi: 10.1016/j.actatropica.2013.01.012. Epub 2013 Feb 4. Acta Trop. 2013. PMID: 23391499
  • Non-compartmental and population pharmacokinetic analysis of dapsone in healthy NIGERIANS: A pilot study. Kotila OA, Ajayi DT, Masimirembwa C, Thelingwani R, Odetunde A, Falusi AG, Babalola CP. Kotila OA, et al. Br J Clin Pharmacol. 2023 Nov;89(11):3454-3459. doi: 10.1111/bcp.15862. Epub 2023 Aug 15. Br J Clin Pharmacol. 2023. PMID: 37489004
  • Human Cytochrome P450 1, 2, 3 Families as Pharmacogenes with Emphases on Their Antimalarial and Antituberculosis Drugs and Prevalent African Alleles. Chamboko CR, Veldman W, Tata RB, Schoeberl B, Tastan Bishop Ö. Chamboko CR, et al. Int J Mol Sci. 2023 Feb 8;24(4):3383. doi: 10.3390/ijms24043383. Int J Mol Sci. 2023. PMID: 36834793 Free PMC article. Review.
  • Modulation of Cytochrome P450, P-glycoprotein and Pregnane X Receptor by Selected Antimalarial Herbs-Implication for Herb-Drug Interaction. Fasinu PS, Manda VK, Dale OR, Egiebor NO, Walker LA, Khan SI. Fasinu PS, et al. Molecules. 2017 Nov 23;22(12):2049. doi: 10.3390/molecules22122049. Molecules. 2017. PMID: 29168799 Free PMC article.
  • Proguanil and cycloguanil are organic cation transporter and multidrug and toxin extrusion substrates. van der Velden M, Bilos A, van den Heuvel JJMW, Rijpma SR, Hurkmans EGE, Sauerwein RW, Russel FGM, Koenderink JB. van der Velden M, et al. Malar J. 2017 Oct 23;16(1):422. doi: 10.1186/s12936-017-2062-y. Malar J. 2017. PMID: 29061131 Free PMC article.
  • Risks of Hemolysis in Glucose-6-Phosphate Dehydrogenase Deficient Infants Exposed to Chlorproguanil-Dapsone, Mefloquine and Sulfadoxine-Pyrimethamine as Part of Intermittent Presumptive Treatment of Malaria in Infants. Poirot E, Vittinghoff E, Ishengoma D, Alifrangis M, Carneiro I, Hashim R, Baraka V, Mosha J, Gesase S, Chandramohan D, Gosling R. Poirot E, et al. PLoS One. 2015 Nov 23;10(11):e0142414. doi: 10.1371/journal.pone.0142414. eCollection 2015. PLoS One. 2015. PMID: 26599634 Free PMC article. Clinical Trial.
  • Br J Clin Pharmacol. 1998 Dec;46(6):553-61 - PubMed
  • J Pharm Pharmacol. 1995 Nov;47(11):957-63 - PubMed
  • Lancet. 1996 Jun 1;347(9014):1511-4 - PubMed
  • Antimicrob Agents Chemother. 1999 Jun;43(6):1334-9 - PubMed
  • Drugs. 1996 Dec;52(6):818-36 - PubMed

Publication types

  • Search in MeSH

Related information

  • Cited in Books
  • PubChem Compound
  • PubChem Substance

LinkOut - more resources

Full text sources.

  • Ovid Technologies, Inc.
  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

Advances in Anti-Malarial Drug Discovery

  • Download PDF Copy

Benedette Cuffari, M.Sc.

Malaria, which is caused by several protozoan Plasmodium species that transmit the disease through mosquitoes, continues to affect individuals throughout the world, particularly in sub-Saharan Africa, where 95% and 96% of cases and deaths occur, respectively 1 .

Common symptoms associated with malaria can include fever, headache, malaise, weakness, gastrointestinal distress, dizziness, confusion, disorientation, and, in severe cases, coma.

Image Credit: Jarun Ontakrai/Shutterstock.com

Introduction

Despite the World Health Organization's (WHO) Global Technical Strategy (GTS) goal to eradicate malaria by 2030, case numbers have increased over the past several years. In 2021 alone, over 247 million malaria cases were reported, 619,000 of which resulted in death, a significant rise of 33 million cases and 181,000 deaths compared to those reported in 2015.

Several antimalarial drugs are currently available to limit the detrimental effects of this disease and reduce associated mortality. However, many of these drugs are associated with limitations that prevent their widespread availability. Resistance to malaria medications also necessitates continuous monitoring and the identification of new candidate drugs with superior efficacy profiles.

Current malaria therapies

Suspected malaria cases must be quickly identified and treated to reduce the risk of mortality, as this disease can evolve rapidly into lethal symptoms. Some of the most widely used treatments for uncomplicated Plasmodium falciparum malaria include artemisinin (ART)-based combination treatments (ACTs), which include aremether-lumefantrine (AL) and artesunate-amodiaquine, which comprise 75% and 24% of the African market share, respectively 1 .

Other less frequently utilized treatment options include combinations of dihydroartemisinin-piperaquine (DHA-PPQ), atovaquone-proguanil (Malarone), and quinine with doxycycline or clindamycin. The choice of treatment will depend on the patient’s age, severity of symptoms, and immune status.

Despite the availability of these treatments, there remains an urgent need to treat asymptomatic patients, as these individuals are still capable of spreading disease through dormant Plasmodium vivax parasites.

Furthermore, patients with severe malaria often cannot be treated with oral medications and, as a result, require injectable treatments like artesunate or quinine.

Novel small molecules

To date, many different drug candidates are being investigated in preclinical exploratory phases, as well as human volunteer and malaria patient exploratory phases. For example, several novel small molecules have been evaluated in patients with malaria, including ZY-19489 and ferroquine. Recent studies on the combination treatment of ZY-19489 and ferroquine indicate that single-dose and long-term efficacy may be possible in children 1 .

Other notable small molecules that have been identified and studied in humans include cabamiquine, cipargamin, and ganaplacide. The latter is considered the most advanced antimalarial, as it is highly potent and has a favorable pharmacokinetic and safety profile.

The combination of ganaplacide with lumefantrine, a partner drug traditionally used in ACTs, has been shown to reduce the risk of resistance while also maintaining efficacy in large patient cohorts 1 .

Monoclonal antibodies

The United States National Institutes of Health (NIH) is currently investigating the prophylactic efficacy of two monoclonal antibodies, CIS43LS and L9LS. CIS43LS targets the circumsporozoite protein (CSP), which is crucial for parasites' ability to invade and infect hepatocytes.

Both phase I and phase II clinical trials have demonstrated that CIS43LS is associated with over 88% efficacy for at least six months 1 . As compared to CIS43, the parent antibody of CIS43LS, L9LS, has been shown to be three times more potent.

MMA01 and TB-21F are other monoclonal antibodies being investigated for malaria prophylaxis. Despite their potential, monoclonal antibodies are expensive to produce and often require specific storage conditions; therefore, their practical use in low—and medium-income countries will be a challenge.

Targeting mosquitos

For several decades, insecticidal nets have been successful in mitigating malaria transmission; however, this approach has been shown to increase the risk of mosquito resistance.

Related Stories

  • Surface Plasmon Resonance Sensors: Transforming Drug Development and Clinical Diagnostics
  • Clinical Trials Supported by Genetics Show Higher Success Rates
  • Stem cell-based kidney chip revolutionizes disease modeling and research

As an alternative vector control method, researchers have investigated the potential of treating mosquitos with antimalarials to support malaria control and elimination.

Some of the different methods that have been investigated include genetically modifying mosquitos so that they are resistant to parasitic infection and systemic treatment of mosquitos with insecticides like ivermectin, lota liner, or isooxazoline 1 .

What Role do Natural Products Play in Modern Drug Discovery?

Challenges in Anti-Malarial Drug Discovery

Non-compliance with the three-day regimen for ART, as well as its short half-life, has contributed to resistance to this drug and, as a result, its limited efficacy in treating malaria. Both total and partial resistance to other malaria drugs have also been reported throughout the world.

‘’The development of plasmodium parasites resistance to many drugs shows that more innovative research is still required, particularly on discovery of small molecules with novel mechanisms of action and targeted delivery approach of drugs to specific organs/tissues in order to increase efficacy and reduce adverse drug interactions. 3 ”

Malaria primarily affects sub-Saharan African populations, particularly those residing in the poorest areas of this region. Agencies like the Global Fund have historically provided these communities with free anti-malaria medications; however, limited funding has been available to support the production and subsequent distribution of these medications on a larger scale.

Furthermore, many pharmaceutical companies have excluded malaria and other tropical diseases from their portfolio due to the high cost of drug development and limited commercial return. 

Strategies to Overcome Challenges

Despite the lack of innovative drug discovery by pharmaceutical companies, several programs have been developed to support the discovery and delivery of new antimalarials to high-risk populations.

The Medicines for Malaria Venture (MMV), for example, is funded by governments and philanthropic organizations worldwide to ensure the dissemination of current anti-malarial medications while also supporting research endeavors for next-generation treatments 1 .

The Malaria Drug Accelerator (MaIDA), formed in 2012, primarily focuses on discovering novel malaria targets and advancing phenotypic drug screening to ultimately identify highly efficacious drug candidates 2 . To date, MaIDA has facilitated the discovery and prioritization of several malaria targets and tool compound characteristics capable of their inhibition.

The Biggest Challenges in Drug Discovery

Future Directions in Anti-Malarial Drug Discovery

Recent advancements in artificial intelligence (AI) indicate that this technology is capable of leveraging existing data to identify hit compounds more efficiently and rapidly than traditional screening platforms.

DeepMalaria, a deep-learning program, has recently been trained on 13,446 antiplasmodial hit compounds from the GlaxoSmithKline (FSK) database that are currently being investigated for their anti-malaria potential.

In a 2023 study, DeepMalaria successfully identified 72.32% of active molecules with increasing accuracy in its ability to distinguish more potent compounds 4 . Although these results are promising, additional research is needed to determine the potential role of AI-based approaches in malaria drug discovery projects.

Multiomics technologies have been widely used over the past several years to understand better the biological mechanisms involved in Plasmodium falciparum infection. For example, genomics studies have provided important insights into the various stages of liver-stage infection with this parasite, which has been crucial for the identification of potential target genes involved in the development of severe malaria 5 .

Thus, future studies will continue to utilize omics-based methodologies to identify novel malaria targets and those involved in other protozoan diseases.

  • Siqueira-Neto, J. L., Wicht, K. J., Chibale, K., et al. (2023). Antimalarial drug discovery: progress and approaches. Nature Reviews Drug Discovery 22 ; 807-826. doi:10.1038/s41573-023-00772-9.
  • Yang, T., Ottilie, S., Istvan, E. S., et al. (2021). MaIDA, Accelerating Malaria Drug Discovery. Trends in Parasitology37 (6). doi:10.1016/j.pt.2021.01.009.
  • Umumararungu, T., Nkuranga, J. B., Habarurema, G., et al. (2023). Recent developments in antimalarial drug discovery. Bioorganic & Medicinal Chemistry 88-89. doi:10.1016/j.bmc.2023.117339.
  • Arshadi, A. K., Salem, M., Collins, J., et al. (2019). DeepMalaria: Artificial Intelligence Driven Discovery of Potent Antiplasmodials. Frontiers in Pharmacology 10 ; 1536. doi:10.3389/fphar.2019.01526.
  • Pandey, S. K., Anand, U., Siddiqui, W. A., & Tripathi, R. (2023). Drug Development Strategies for Malaria: With the Hope for New Antimalarial Drug Discovery – An Update. Advances in Medicine. doi:10.1155/2023/5060665.

Further Reading

  • All Drug Discovery Content
  • Exploring the Structure-Activity Relationship (SAR) of Drugs
  • Hit to Lead (H2L) Process in Drug Discovery
  • Hot Melt Extrusion in the Pharmaceutical and Food Industries
  • How is Additive Manufacturing Used in the Pharmaceutical Industry?

Last Updated: Sep 3, 2024

Benedette Cuffari

Benedette Cuffari

After completing her Bachelor of Science in Toxicology with two minors in Spanish and Chemistry in 2016, Benedette continued her studies to complete her Master of Science in Toxicology in May of 2018. During graduate school, Benedette investigated the dermatotoxicity of mechlorethamine and bendamustine; two nitrogen mustard alkylating agents that are used in anticancer therapy.

Please use one of the following formats to cite this article in your essay, paper or report:

Cuffari, Benedette. (2024, September 03). Advances in Anti-Malarial Drug Discovery. AZoLifeSciences. Retrieved on September 06, 2024 from https://www.azolifesciences.com/article/Advances-in-Anti-Malarial-Drug-Discovery.aspx.

Cuffari, Benedette. "Advances in Anti-Malarial Drug Discovery". AZoLifeSciences . 06 September 2024. <https://www.azolifesciences.com/article/Advances-in-Anti-Malarial-Drug-Discovery.aspx>.

Cuffari, Benedette. "Advances in Anti-Malarial Drug Discovery". AZoLifeSciences. https://www.azolifesciences.com/article/Advances-in-Anti-Malarial-Drug-Discovery.aspx. (accessed September 06, 2024).

Cuffari, Benedette. 2024. Advances in Anti-Malarial Drug Discovery . AZoLifeSciences, viewed 06 September 2024, https://www.azolifesciences.com/article/Advances-in-Anti-Malarial-Drug-Discovery.aspx.

Suggested Reading

CARA Benchmark Set to Revolutionize Data-Driven Drug Discovery Methods

Cancel reply to comment

  • Trending Articles
  • Latest Interviews
  • Editorial Highlight

Emerging Technologies in Forensic Science: From CRISPR to AI

Driving Support for Pharmaceutical Testing Through Continuous Innovation

In this interview, Kyle James from ERWEKA highlights the company's commitment to supporting pharmaceutical sciences through advanced equipment and continuous innovation.

Driving Support for Pharmaceutical Testing Through Continuous Innovation

From Memory to Metaplasticity: Unraveling the Brain's Secrets

Explore the groundbreaking work of Mark Bear, a leading figure in neuroscience, as he shares insights on synaptic plasticity, learning, and the future of neurological research.

From Memory to Metaplasticity: Unraveling the Brain's Secrets

Revolutionizing Mass Spectrometry: A Deep Dive into Plasmion’s SICRIT® Technology

Advancements in Direct MS with Plasmion’s SICRIT®.

Revolutionizing Mass Spectrometry: A Deep Dive into Plasmion’s SICRIT® Technology

Latest in Genomics

ERS Genomics and Medicines Discovery Catapult Forge CRISPR Licensing Partnership

Newsletters you may be interested in

Drug Discovery

Your AI Powered Scientific Assistant

Hi, I'm Azthena, you can trust me to find commercial scientific answers from AZoNetwork.com.

A few things you need to know before we start. Please read and accept to continue.

  • Use of “Azthena” is subject to the terms and conditions of use as set out by OpenAI .
  • Content provided on any AZoNetwork sites are subject to the site Terms & Conditions and Privacy Policy .
  • Large Language Models can make mistakes. Consider checking important information.

Great. Ask your question.

Azthena may occasionally provide inaccurate responses. Read the full terms .

While we only use edited and approved content for Azthena answers, it may on occasions provide incorrect responses. Please confirm any data provided with the related suppliers or authors. We do not provide medical advice, if you search for medical information you must always consult a medical professional before acting on any information provided.

Your questions, but not your email details will be shared with OpenAI and retained for 30 days in accordance with their privacy principles.

Please do not ask questions that use sensitive or confidential information.

Read the full Terms & Conditions .

Provide Feedback

thesis on antimalarial drugs

IMAGES

  1. (PDF) Antimalarial drugs: Mode of action and status of resistance

    thesis on antimalarial drugs

  2. Thesis

    thesis on antimalarial drugs

  3. (PDF) In Search of a Potent Antimalarial Agent: Antiplasmodial

    thesis on antimalarial drugs

  4. (PDF) Antimalarial herbal drugs: a review of their interactions with

    thesis on antimalarial drugs

  5. (PDF) Antimalarial drug review

    thesis on antimalarial drugs

  6. Antimalarial Drugs I: Biological Background, Experimental Methods, and

    thesis on antimalarial drugs

VIDEO

  1. Malaria || Pathogenesis of Malaria || Pharmacotheraphetics

  2. Antimalarial Drugs

  3. Antimalarial drugs

  4. Medicinal Chemistry III

  5. antimalarial topic

  6. Antimalarial Drugs

COMMENTS

  1. PDF Optimization of Antimalarial Drug Regimens in A Cytocdial Murine

    improve the efficacy of antimalarial drug regimens. Developing a novel murine model of artemisinin-tolerant Plasmodium parasites will further accelerate the development of next-generation antimalarials and serve as a platform on which currently approved antimalarials can be evaluated and improved. _____ Thesis Advisory Committee _____

  2. Recent Progress in the Development of New Antimalarial Drugs with Novel

    So, protease is a promising target for antimalarial drugs development. 46. In P. falciparum food vacuole several aspartic proteases (plasmepsin I, II, III, IV) and cysteine proteases (falcipain-1, falcipain-2/, falcipain-3) have been isolated which used for degradation of hemoglobin as displayed in Figure 2. Figure 2.

  3. The past, present and future of anti-malarial medicines

    Great progress has been made in recent years to reduce the high level of suffering caused by malaria worldwide. Notably, the use of insecticide-treated mosquito nets for malaria prevention and the use of artemisinin-based combination therapy (ACT) for malaria treatment have made a significant impact. Nevertheless, the development of resistance to the past and present anti-malarial drugs ...

  4. Antimalarial drug discovery: progress and approaches

    Abstract. Recent antimalarial drug discovery has been a race to produce new medicines that overcome emerging drug resistance, whilst considering safety and improving dosing convenience. Discovery ...

  5. PDF Report on antimalarial drug efficacy, resistance and response

    Monitoring antimalarial drug efficacy and resistance is important for the early detection of resistance which in turn enables timely action to prevent its spread and limit the impact on global health. Measurement of drug efficacy and resistance in malaria is complex. This report provides an overview of the tools currently used to

  6. Antimalarial Drug Discovery: Approaches and Progress towards New

    For antimalarial drugs, it might include pharmacokinetic and pharmocodynamics parameters, oral bioavailability, cost, potency, and activity against different lifecycle stages ... University in 1997 and joined the California Institute of Technology in October 1998 to conduct his doctoral thesis research in the laboratory of Professor Robert H ...

  7. Antiplasmodial, antimalarial activities and toxicity of African

    Background Malaria still constitutes a major public health menace, especially in tropical and subtropical countries. Close to half a million people mainly children in Africa, die every year from the disease. With the rising resistance to frontline drugs (artemisinin-based combinations), there is a need to accelerate the discovery and development of newer anti-malarial drugs. A systematic ...

  8. Drug action and resistance in malaria parasites: experimental genetics

    Resistance to antimalarial drugs inevitably follows their deployment in malaria endemic parts of the world. For instance, current malaria control efforts which significantly rely on artemisinin combination therapies (ACTs) are being threatened by the emergence of resistance to artemisinins and ACTs. Understanding the role of genetic determinants of artemisinin resistance is therefore important ...

  9. Recent developments in the synthesis of hybrid antimalarial drug

    In this 21 st century, Malaria remains a global burden and causes massive economic trouble to disease-endemic nations. The control and eradication of malaria is a major challenge that requires an urgent need to develop novel antimalarial drugs. To overcome the aforementioned situation, several researchers have given significant effort to develop hybrid antimalarial agents in the search for new ...

  10. Antimalarial Drug Resistance: Trends, Mechanisms, and Strategies to

    Antimalarial drug resistance ascends due to genetic mutations that occur by a change in single nucleotide polymorphisms or gene amplifications . Table 2 summarizes the mechanism of action and associated genetic markers of antimalarial drug resistance, while Fig. 2 depicts the mechanism of resistance. Certain other factors of concern are overall ...

  11. Design, synthesis, and in silico-in vitro antimalarial evaluation of 1

    Recent data indicates that the challenge to overcome malaria still lingers [], regardless of all the significant efforts and advances in drug developments and medical technologies.The rise of drug or multidrug resistance in Plasmodia strains has made malaria the most important tropical parasitic disease [].Therefore, developing new antimalarial drugs is necessary to rectify this situation.

  12. Antimalarial Drug Resistance: Literature Review and Activities and

    The latter would seem intuitive from initial low prevalence of innate resistance observed in the field for some antimalarials such as mefloquine and atovaquone, presumably owing to the mutant parasites being outcompeted by the wild type before start of drug use. 117, 147, 148 More compelling evidence for a fitness cost of resistance has been ...

  13. Antimalarial Drug Resistance: Literature Review and Activities and

    Antimalarial drugs are key tools for the control and elimination of malaria. Recent decreases in the global malaria burden are likely due, in part, to the deployment of artemisinin-based ...

  14. An Efficient and Sustainable Synthesis of the Antimalarial Drug

    An 11-step, 8-pot synthesis of the antimalarial drug tafenoquine succinate was achieved in 42% overall yield using commercially available starting materials. Compared to the previous manufacturing processes that utilize environmentally egregious organic solvents and toxic reagents, the current route features a far greener (as measured by Sheldon's E Factors) and likely more economically ...

  15. PDF Potent in Vitro Antimalarial Activity of Artemisinin Dimers

    A thesis submitted to Johns Hopkins University in conformity with the requirements for the degree of Master of Science Baltimore, Maryland ... sodium artesunate are the fastest acting antimalarial drugs available. However, they possess a short plasma half-life of less than 1 hour. Also, there is recrudescence of parasitemia when used as ...

  16. PDF Safety, efficacy and pharmacokinetics profile of antimalarial drugs in

    antimalarial cure rate. Another factor that affects antimalarial effectiveness in pregnancy is parasite resistance against sulfadoxine-pyrimethamine (SP), a drug that is used for intermittent preventive treatment of malaria in pregnancy (IPTp). The objectives of the thesis were to assess the magnitude of drugs exposure during pregnancy in

  17. In silico study, synthesis, and evaluation of the antimalarial activity

    The widespread emergence of resistance to first-line drugs has further bolstered an urgent need for a new and cost-effective antimalarial(s). Thus, the present study enumerates the synthesis of novel hybrid dimethoxy pyrazole 1,3,5-triazine derivatives 7(a-j) and their in silico results short-listed three compounds with good binding energies ...

  18. (PDF) Design of Antimalarial Agents: A review

    Quinine, an old anti-malarial drug in a modern . world. Role in the treatment of malaria. Malaria J 2011;10:1-12, 10. Giao PT, de Vries PJ. Pharmacokinetic interactions of antimalarial .

  19. Pharmacokinetic and molecular docking studies to design antimalarial

    The identification of new/novel drug targets as well as development of novel antimalarial drugs with unique mode of action can be the possible answer to drug resistance in malaria and subsequently it can control the mortality due to malaria.[2,3] Being highly studied diseases several potential drug targets have been suggested in research which ...

  20. PDF Antimalarial drug concentrations in the blood: a monitoring tool for

    Antimalarial drug concentrations in the blood: a monitoring tool for evaluating population drug usage in relation to diagnosis and treatment environment INAUGURALDISSERTATION ... These three years of PhD thesis have been an amazing journey from both a professional and personal side. I have enjoyed discovering various aspects of the amazing and ...

  21. Antimalarial drugs: Mode of action and status of resistance

    African region. An estimated 3.3 billion people are at risk. of malaria. An estimated 655, 000 pers ons died of. malaria in 2010. 86% of the victims were children under 5. years of age, and 91% of ...

  22. Antimalarial drug sulfadoxine induces gametocytogenesis in ...

    Background: The spread of antimalarial drug resistance parasites is a major obstacle in eliminating malaria in endemic areas. This increases the urgency for developing novel antimalarial drugs with improved profiles to eliminate both sensitive and resistant parasites in populations. The invention of the drug candidates needs a model for ...

  23. PDF Synthesis and Evaluation of Antimalarial Agents As Inhibitors of

    The drugs in use act on different stages of the malaria life cycle. However, most antimalarial drugs target the asexual stage of malaria, but not the gametocytes. Hence, new strategies are ... The second part of this thesis involved the synthesis of pyrrolo[2,3-d]pyrimidine derivatives that are potential inhibitors of PfCDPK4 and PfCDPK1. These ...

  24. Pharmacokinetic interactions of antimalarial agents

    Combination of antimalarial agents has been introduced as a response to widespread drug resistance. The higher number of mutations required to express complete resistance against combinations may retard the further development of resistance. Combination of drugs, especially with the artemisinin drug …

  25. Advances in Anti-Malarial Drug Discovery

    ''The development of plasmodium parasites resistance to many drugs shows that more innovative research is still required, particularly on discovery of small molecules with novel mechanisms of action and targeted delivery approach of drugs to specific organs/tissues in order to increase efficacy and reduce adverse drug interactions. 3 "