Hostname: page-component-848d4c4894-jbqgn Total loading time: 0 Render date: 2024-06-29T15:20:20.490Z Has data issue: false hasContentIssue false

Messenger RNAs with large numbers of upstream open reading frames are translated via leaky scanning and reinitiation in the asexual stages of Plasmodium falciparum

Published online by Cambridge University Press:  26 May 2020

Chhaminder Kaur
Affiliation:
Department of Biosciences & Bioengineering, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India
Mayank Kumar
Affiliation:
Department of Biosciences & Bioengineering, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India
Swati Patankar*
Affiliation:
Department of Biosciences & Bioengineering, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India
*
Author for correspondence: Swati Patankar, E-mail: patankar@iitb.ac.in

Abstract

The genome of Plasmodium falciparum has one of the most skewed base-pair compositions of any eukaryote, with an AT content of 80–90%. As start and stop codons are AT-rich, the probability of finding upstream open reading frames (uORFs) in messenger RNAs (mRNAs) is high and parasite mRNAs have an average of 11 uORFs in their leader sequences. Similar to other eukaryotes, uORFs repress the translation of the downstream open reading frame (dORF) in P. falciparum, yet the parasite translation machinery is able to bypass these uORFs and reach the dORF to initiate translation. This can happen by leaky scanning and/or reinitiation.

In this report, we assessed leaky scanning and reinitiation by studying the effect of uORFs on the translation of a dORF, in this case, the luciferase reporter gene, and showed that both mechanisms are employed in the asexual blood stages of P. falciparum. Furthermore, in addition to the codon usage of the uORF, translation of the dORF is governed by the Kozak sequence and length of the uORF, and inter-cistronic distance between the uORF and dORF. Based on these features whole-genome data was analysed to uncover classes of genes that might be regulated by uORFs. This study indicates that leaky scanning and reinitiation appear to be widespread in asexual stages of P. falciparum, which may require modifications of existing factors that are involved in translation initiation in addition to novel, parasite-specific proteins.

Type
Research Article
Copyright
Copyright © The Author(s), 2020. Published by Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

*

Current address: Amity Institute of Biotechnology, Amity University, Panvel, Maharashtra 410206, India.

References

Aitken, CE and Lorsch, JR (2012) A mechanistic overview of translation initiation in eukaryotes. Nature Structural & Molecular Biology 19, 568576.CrossRefGoogle ScholarPubMed
Amulic, B, Salanti, A, Lavstsen, T, Nielsen, MA and Deitsch, KW (2009) An upstream open reading frame controls translation of var2csa, a gene implicated in placental malaria. PLoS Pathogens 5, e1000256.CrossRefGoogle ScholarPubMed
Andreev, DE, O'Connor, PBF, Zhdanov, AV, Dmitriev, RI, Shatsky, IN, Papkovsky, DB and Baranov, PV (2015) Oxygen and glucose deprivation induces widespread alterations in mRNA translation within 20 minutes. Genome Biology 16, 90.CrossRefGoogle ScholarPubMed
Aurrecoechea, C, Brestelli, J, Brunk, BP, Dommer, J, Fischer, S, Gajria, B, Gao, X, Gingle, A, Grant, G, Harb, OS, Heiges, M, Innamorato, F, Iodice, J, Kissinger, JC, Kraemer, E, Li, W, Miller, JA, Nayak, V, Pennington, C, Pinney, DF, Roos, DS, Ross, C, Stoeckert, CJ, Treatman, C and Wang, H (2009) PlasmoDB: a functional genomic database for malaria parasites. Nucleic Acids Research 37, D539D543.CrossRefGoogle ScholarPubMed
Balaji, S, Babu, MM, Iyer, LM and Aravind, L (2005) Discovery of the principal specific transcription factors of Apicomplexa and their implication for the evolution of the AP2-integrase DNA binding domains. Nucleic Acids Research 33, 39944006.CrossRefGoogle ScholarPubMed
Bancells, C and Deitsch, KW (2013) A molecular switch in the efficiency of translation reinitiation controls expression of var2csa, a gene implicated in pregnancy-associated malaria: var2csa is expressed by translation reinitiation. Molecular Microbiology 90, 472488.CrossRefGoogle ScholarPubMed
Bunnik, EM, Chung, D-WD, Hamilton, M, Ponts, N, Saraf, A, Prudhomme, J, Florens, L and Le Roch, KG (2013) Polysome profiling reveals translational control of gene expression in the human malaria parasite Plasmodium falciparum. Genome Biology 14, R128.CrossRefGoogle ScholarPubMed
Calvo, SE, Pagliarini, DJ and Mootha, VK (2009) Upstream open reading frames cause widespread reduction of protein expression and are polymorphic among humans. Proceedings of the National Academy of Sciences 106, 75077512.CrossRefGoogle ScholarPubMed
Caro, F, Ahyong, V, Betegon, M and DeRisi, JL (2014) Genome-wide regulatory dynamics of translation in the Plasmodium Falciparum asexual blood stages. eLife 3, e04106. doi: 10.7554/eLife.04106CrossRefGoogle ScholarPubMed
Chan, S, Frasch, A, Mandava, CS, Ch'ng, J-H, Quintana, MDP, Vesterlund, M, Ghorbal, M, Joannin, N, Franzén, O, Lopez-Rubio, J-J, Barbieri, S, Lanzavecchia, A, Sanyal, S and Wahlgren, M (2017) Regulation of PfEMP1–VAR2CSA translation by a Plasmodium translation-enhancing factor. Nature Microbiology 2, 17068.CrossRefGoogle ScholarPubMed
Chaubey, S, Kumar, A, Singh, D and Habib, S (2005) The apicoplast of Plasmodium Falciparum is translationally active. Molecular Microbiology 56, 8189.CrossRefGoogle ScholarPubMed
Cheung, Y-N, Maag, D, Mitchell, SF, Fekete, CA, Algire, MA, Takacs, JE, Shirokikh, N, Pestova, T, Lorsch, JR and Hinnebusch, AG (2007) Dissociation of eIF1 from the 40S ribosomal subunit is a key step in start codon selection in vivo. Genes & Development 21, 12171230.CrossRefGoogle ScholarPubMed
Child, SJ, Miller, MK and Geballe, AP (1999) Translational control by an upstream open reading frame in the HER-2/ neu Transcript. Journal of Biological Chemistry 274, 2433524341.CrossRefGoogle ScholarPubMed
Coulson, RMR, Hall, N and Ouzounis, CA (2004) Comparative genomics of transcriptional control in the human malaria parasite Plasmodium falciparum. Genome Research 14, 15481554.CrossRefGoogle ScholarPubMed
Cuchalová, L, Kouba, T, Herrmannová, A, Dányi, I, Chiu, W-L and Valásek, L (2010) The RNA recognition motif of eukaryotic translation initiation factor 3 g (eIF3 g) is required for resumption of scanning of posttermination ribosomes for reinitiation on GCN4 and together with eIF3i stimulates linear scanning. Molecular and Cellular Biology 30, 46714686.CrossRefGoogle Scholar
de Breyne, S and Ohlmann, T (2018) Focus on translation initiation of the HIV-1 mRNAs. International Journal of Molecular Sciences 20(1), 101. doi: 10.3390/ijms20010101CrossRefGoogle Scholar
Deitsch, K, Driskill, C and Wellems, T (2001) Transformation of malaria parasites by the spontaneous uptake and expression of DNA from human erythrocytes. Nucleic Acids Research 29, 850853.CrossRefGoogle ScholarPubMed
Dzikowski, R, Li, F, Amulic, B, Eisberg, A, Frank, M, Patel, S, Wellems, TE and Deitsch, KW (2007) Mechanisms underlying mutually exclusive expression of virulence genes by malaria parasites. EMBO reports 8, 959965.CrossRefGoogle ScholarPubMed
Ferreira, JP, Overton, KW and Wang, CL (2013) Tuning gene expression with synthetic upstream open reading frames. Proceedings of the National Academy of Sciences 110, 1128411289.CrossRefGoogle ScholarPubMed
Ferreira, JP, Noderer, WL, Diaz de Arce, AJ and Wang, CL (2014) Engineering ribosomal leaky scanning and upstream open reading frames for precise control of protein translation. Bioengineered 5, 186192.CrossRefGoogle ScholarPubMed
Fichera, ME and Roos, DS (1997) A plastid organelle as a drug target in apicomplexan parasites. Nature 390, 407409.CrossRefGoogle ScholarPubMed
Fijalkowska, D, Verbruggen, S, Ndah, E, Jonckheere, V, Menschaert, G and Van Damme, P (2017) eIF1 modulates the recognition of suboptimal translation initiation sites and steers gene expression Via uORFs. Nucleic Acids Research 45, 79978013.CrossRefGoogle ScholarPubMed
Foth, BJ, Zhang, N, Chaal, BK, Sze, SK, Preiser, PR and Bozdech, Z (2011) Quantitative time-course profiling of parasite and host cell proteins in the human malaria parasite Plasmodium falciparum. Molecular & Cellular Proteomics: MCP 10(8), M110.006411. doi: 10.1074/mcp.M110.006411.CrossRefGoogle Scholar
Gardner, MJ, Hall, N, Fung, E, White, O, Berriman, M, Hyman, RW, Carlton, JM, Pain, A, Nelson, KE, Bowman, S, Paulsen, IT, James, K, Eisen, JA, Rutherford, K, Salzberg, SL, Craig, A, Kyes, S, Chan, M-S, Nene, V, Shallom, SJ, Suh, B, Peterson, J, Angiuoli, S, Pertea, M, Allen, J, Selengut, J, Haft, D, Mather, MW, Vaidya, AB, Martin, DMA, Fairlamb, AH, Fraunholz, MJ, Roos, DS, Ralph, SA, McFadden, GI, Cummings, LM, Subramanian, GM, Mungall, C, Venter, JC, Carucci, DJ, Hoffman, SL, Newbold, C, Davis, RW, Fraser, CM and Barrell, B (2002) Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419, 498511.CrossRefGoogle ScholarPubMed
Goodman, CD, Pasaje, CFA, Kennedy, K, McFadden, GI and Ralph, SA (2016) Targeting protein translation in Organelles of the Apicomplexa. Trends in Parasitology 32, 953965.CrossRefGoogle ScholarPubMed
Haimov, O, Sehrawat, U, Tamarkin-Ben Harush, A, Bahat, A, Uzonyi, A, Will, A, Hiraishi, H, Asano, K and Dikstein, R (2018) Dynamic interaction of eukaryotic initiation factor 4G1 (eIF4G1) with eIF4E and eIF1 underlies scanning-dependent and -independent translation. Molecular and Cellular Biology 38, e00139–18.CrossRefGoogle ScholarPubMed
Hasenkamp, S, Russell, KT and Horrocks, P (2012) Comparison of the absolute and relative efficiencies of electroporation-based transfection protocols for Plasmodium falciparum. Malaria Journal 11, 210.CrossRefGoogle ScholarPubMed
Hinnebusch, AG (1993) Gene-specific translational control of the yeast GCN4 gene by phosphorylation of eukaryotic initiation factor 2. Molecular Microbiology 10, 215223.CrossRefGoogle ScholarPubMed
Hinnebusch, AG (2005) Translational regulation of GCN4 and the general amino acid control of yeast. Annual Review of Microbiology 59, 407450.CrossRefGoogle ScholarPubMed
Hinnebusch, AG, Ivanov, IP and Sonenberg, N (2016) Translational control by 5′-untranslated regions of eukaryotic mRNAs. Science (New York, N.Y.) 352, 14131416.CrossRefGoogle ScholarPubMed
Hood, HM, Neafsey, DE, Galagan, J and Sachs, MS (2009) Evolutionary roles of upstream open reading frames in mediating gene regulation in fungi. Annual Review of Microbiology 63, 385409.CrossRefGoogle ScholarPubMed
Hronová, V, Mohammad, MP, Wagner, S, Pánek, J, Gunišová, S, Zeman, J, Poncová, K and Valášek, LS (2017) Does eIF3 promote reinitiation after translation of short upstream ORFs also in mammalian cells? RNA biology 14, 16601667.CrossRefGoogle ScholarPubMed
Hughes, S, Mellstrom, K, Kosik, E, Tamanoi, F and Brugge, J (1984) Mutation of a termination codon affects src initiation. Molecular and Cellular Biology 4, 17381746.CrossRefGoogle ScholarPubMed
Jackson, KE, Habib, S, Frugier, M, Hoen, R, Khan, S, Pham, JS, Ribas de Pouplana, L, Royo, M, Santos, MAS, Sharma, A and Ralph, SA (2011) Protein translation in Plasmodium Parasites. Trends in Parasitology 27, 467476.CrossRefGoogle ScholarPubMed
Johnstone, TG, Bazzini, AA and Giraldez, AJ (2016) Upstream ORFs are prevalent translational repressors in vertebrates. The EMBO Journal 35, 706723.CrossRefGoogle ScholarPubMed
Kozak, M (1978) How do eucaryotic ribosomes select initiation regions in messenger RNA? Cell 15, 11091123.CrossRefGoogle ScholarPubMed
Kozak, M (1984) Selection of initiation sites by eucaryotic ribosomes: effect of inserting AUG triplets upstream from the coding sequence for preproinsulin. Nucleic Acids Research 12, 38733893.CrossRefGoogle ScholarPubMed
Kozak, M (1987) Effects of intercistronic length on the efficiency of reinitiation by eucaryotic ribosomes. Molecular and Cellular Biology 7, 34383445.CrossRefGoogle ScholarPubMed
Kozak, M (1999) Initiation of translation in prokaryotes and eukaryotes. Gene 234, 187208.CrossRefGoogle ScholarPubMed
Kozak, M (2001) Constraints on reinitiation of translation in mammals. Nucleic Acids Research 29, 52265232.CrossRefGoogle ScholarPubMed
Kumar, M, Srinivas, V and Patankar, S (2015) Upstream AUGs and upstream ORFs can regulate the downstream ORF in Plasmodium falciparum. Malaria Journal 14, 512.CrossRefGoogle ScholarPubMed
Kyes, SA, Christodoulou, Z, Raza, A, Horrocks, P, Pinches, R, Rowe, JA and Newbold, CI (2003) A well-conserved Plasmodium Falciparum var gene shows an unusual stage-specific transcript pattern. Molecular Microbiology 48, 13391348.CrossRefGoogle ScholarPubMed
Lasonder, E, Ishihama, Y, Andersen, JS, Vermunt, AMW, Pain, A, Sauerwein, RW, Eling, WMC, Hall, N, Waters, AP, Stunnenberg, HG and Mann, M (2002) Analysis of the Plasmodium Falciparum proteome by high-accuracy mass spectrometry. Nature 419, 537542.CrossRefGoogle ScholarPubMed
Lavstsen, T, Salanti, A, Jensen, AT, Arnot, DE and Theander, TG (2003) Sub-grouping of Plasmodium Falciparum 3D7 var genes based on sequence analysis of coding and non-coding regions. Malaria Journal 2, 27.CrossRefGoogle ScholarPubMed
Le Roch, KG, Johnson, JR, Florens, L, Zhou, Y, Santrosyan, A, Grainger, M, Yan, SF, Williamson, KC, Holder, AA, Carucci, DJ, Yates, JR and Winzeler, EA (2004) Global analysis of transcript and protein levels across the Plasmodium Falciparum life cycle. Genome Research 14, 23082318.CrossRefGoogle ScholarPubMed
Liu, CC, Simonsen, CC and Levinson, AD (1984) Initiation of translation at internal AUG codons in mammalian cells. Nature 309, 8285.CrossRefGoogle ScholarPubMed
Loughran, G, Sachs, MS, Atkins, JF and Ivanov, IP (2012) Stringency of start codon selection modulates autoregulation of translation initiation factor eIF5. Nucleic Acids Research 40, 28982906.CrossRefGoogle ScholarPubMed
Luukkonen, BG, Tan, W and Schwartz, S (1995) Efficiency of reinitiation of translation on human immunodeficiency virus type 1 mRNAs is determined by the length of the upstream open reading frame and by intercistronic distance. Journal of Virology 69, 40864094.CrossRefGoogle ScholarPubMed
McGeachy, AM and Ingolia, NT (2016) Starting too soon: upstream reading frames repress downstream translation. The EMBO journal 35, 699700.CrossRefGoogle ScholarPubMed
Militello, KT, Dodge, M, Bethke, L and Wirth, DF (2004) Identification of regulatory elements in the Plasmodium Falciparum genome. Molecular and Biochemical Parasitology 134, 7588.CrossRefGoogle ScholarPubMed
Mohammad, MP, Munzarová Pondelícková, V, Zeman, J, Gunišová, S and Valášek, LS (2017) In vivo evidence that eIF3 stays bound to ribosomes elongating and terminating on short upstream ORFs to promote reinitiation. Nucleic Acids Research 45, 26582674.Google ScholarPubMed
Morris, DR and Geballe, AP (2000) Upstream open reading frames as regulators of mRNA translation. Molecular and Cellular Biology 20, 86358642.CrossRefGoogle ScholarPubMed
Nakamura, Y, Gojobori, T and Ikemura, T (2000) Codon usage tabulated from international DNA sequence databases: status for the year 2000. Nucleic Acids Research 28, 292.CrossRefGoogle ScholarPubMed
Nanda, JS, Cheung, Y-N, Takacs, JE, Martin-Marcos, P, Saini, AK, Hinnebusch, AG and Lorsch, JR (2009) eIF1 controls multiple steps in start codon recognition during eukaryotic translation initiation. Journal of Molecular Biology 394, 268285.CrossRefGoogle ScholarPubMed
Painter, HJ, Campbell, TL and Llinás, M (2011) The Apicomplexan AP2 family: integral factors regulating Plasmodium development. Molecular and Biochemical Parasitology 176, 17.CrossRefGoogle ScholarPubMed
Palam, LR, Baird, TD and Wek, RC (2011) Phosphorylation of eIF2 facilitates ribosomal bypass of an inhibitory upstream ORF to enhance CHOP translation. The Journal of Biological Chemistry 286, 1093910949.CrossRefGoogle ScholarPubMed
Pasaje, CFA, Cheung, V, Kennedy, K, Lim, EE, Baell, JB, Griffin, MDW and Ralph, SA (2016) Selective inhibition of apicoplast tryptophanyl-tRNA synthetase causes delayed death in Plasmodium falciparum. Scientific Reports 6, 27531.CrossRefGoogle ScholarPubMed
Passmore, LA, Schmeing, TM, Maag, D, Applefield, DJ, Acker, MG, Algire, MA, Lorsch, JR and Ramakrishnan, V (2007) The eukaryotic translation initiation factors eIF1 and eIF1A induce an open conformation of the 40S ribosome. Molecular Cell 26, 4150.CrossRefGoogle ScholarPubMed
Pease, BN, Huttlin, EL, Jedrychowski, MP, Talevich, E, Harmon, J, Dillman, T, Kannan, N, Doerig, C, Chakrabarti, R, Gygi, SP and Chakrabarti, D (2013) Global analysis of protein expression and phosphorylation of three stages of Plasmodium Falciparum intraerythrocytic development. Journal of Proteome Research 12, 40284045.CrossRefGoogle ScholarPubMed
Pestova, TV, Borukhov, SI and Hellen, CU (1998) Eukaryotic ribosomes require initiation factors 1 and 1A to locate initiation codons. Nature 394, 854859.CrossRefGoogle ScholarPubMed
Pisarev, AV, Kolupaeva, VG, Pisareva, VP, Merrick, WC, Hellen, CUT and Pestova, TV (2006) Specific functional interactions of nucleotides at key −3 and +4 positions flanking the initiation codon with components of the mammalian 48S translation initiation complex. Genes & Development 20, 624636.CrossRefGoogle ScholarPubMed
Roy, A, Cox, RA, Williamson, DH and Wilson, RJ (1999) Protein synthesis in the plastid of Plasmodium falciparum. Protist 150, 183188.CrossRefGoogle ScholarPubMed
Roy, B, Vaughn, JN, Kim, B-H, Zhou, F, Gilchrist, MA and Von Arnim, AG (2010) The h subunit of eIF3 promotes reinitiation competence during translation of mRNAs harboring upstream open reading frames. RNA (New York, N.Y.) 16, 748761.CrossRefGoogle Scholar
Rug, M and Maier, AG (2013) Transfection of Plasmodium falciparum. Methods in Molecular Biology (Clifton, N.J.) 923, 7598.CrossRefGoogle ScholarPubMed
Ryabova, LA, Pooggin, MM and Hohn, T (2006) Translation reinitiation and leaky scanning in plant viruses. Virus Research 119, 5262.CrossRefGoogle ScholarPubMed
Saraf, A, Cervantes, S, Bunnik, EM, Ponts, N, Sardiu, ME, Chung, D-WD, Prudhomme, J, Varberg, JM, Wen, Z, Washburn, MP, Florens, L and Le Roch, KG (2016) Dynamic and combinatorial landscape of histone modifications during the intraerythrocytic developmental cycle of the malaria parasite. Journal of Proteome Research 15, 27872801.CrossRefGoogle ScholarPubMed
Saul, A and Battistutta, D (1988) Codon usage in Plasmodium falciparum. Molecular and Biochemical Parasitology 27, 3542.CrossRefGoogle ScholarPubMed
Scherf, A, Hernandez-Rivas, R, Buffet, P, Bottius, E, Benatar, C, Pouvelle, B, Gysin, J and Lanzer, M (1998) Antigenic variation in malaria: in situ switching, relaxed and mutually exclusive transcription of var genes during intra-erythrocytic development in Plasmodium falciparum. The EMBO Journal 17, 54185426.CrossRefGoogle ScholarPubMed
Sharp, PM and Li, WH (1987) The codon Adaptation Index--a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Research 15, 12811295.CrossRefGoogle ScholarPubMed
Sheridan, CM, Garcia, VE, Ahyong, V and DeRisi, JL (2018) The Plasmodium Falciparum cytoplasmic translation apparatus: a promising therapeutic target not yet exploited by clinically approved anti-malarials. Malaria Journal 17, 465.CrossRefGoogle Scholar
Skinner-Adams, TS, Lawrie, PM, Hawthorne, PL, Gardiner, DL and Trenholme, KR (2003) Comparison of Plasmodium Falciparum transfection methods. Malaria Journal 2, 19.CrossRefGoogle ScholarPubMed
Sorber, K, Dimon, MT and DeRisi, JL (2011) RNA-Seq analysis of splicing in Plasmodium Falciparum uncovers new splice junctions, alternative splicing and splicing of antisense transcripts. Nucleic Acids Research 39, 38203835.CrossRefGoogle ScholarPubMed
Srinivas, V, Kumar, M, Noronha, S and Patankar, S (2016) ORFpred: a machine learning program to identify translatable small open reading frames in intergenic regions of the Plasmodium Falciparum genome. Current Bioinformatics 11, 000000.CrossRefGoogle Scholar
Vembar, SS, Macpherson, CR, Sismeiro, O, Coppée, J-Y and Scherf, A (2015) The PfAlba1 RNA-binding protein is an important regulator of translational timing in Plasmodium Falciparum blood stages. Genome Biology 16, 212.CrossRefGoogle ScholarPubMed
Vembar, SS, Droll, D and Scherf, A (2016) Translational regulation in blood stages of the malaria parasite Plasmodium Spp.: systems-wide studies pave the way. Wiley Interdisciplinary Reviews. RNA 7, 772792.CrossRefGoogle ScholarPubMed
von Arnim, AG, Jia, Q and Vaughn, JN (2014) Regulation of plant translation by upstream open reading frames. Plant Science: An International Journal of Experimental Plant Biology 214, 112.CrossRefGoogle ScholarPubMed
Wang, XC, Yang, J, Huang, W, He, L, Yu, JT, Lin, QS, Li, W and Zhou, HM (2002) Effects of removal of the N-terminal amino acid residues on the activity and conformation of firefly luciferase. The International Journal of Biochemistry & Cell Biology 34, 983991.CrossRefGoogle ScholarPubMed
Watanabe, J, Sasaki, M, Suzuki, Y and Sugano, S (2002) Analysis of transcriptomes of human malaria parasite Plasmodium Falciparum using full-length enriched library: identification of novel genes and diverse transcription start sites of messenger RNAs. Gene 291, 105113.CrossRefGoogle ScholarPubMed
Wong, W, Bai, X, Brown, A, Fernandez, IS, Hanssen, E, Condron, M, Tan, YH, Baum, J and Scheres, SHW (2014) Cryo-EM structure of the Plasmodium Falciparum 80S ribosome bound to the anti-protozoan drug emetine. eLife 3, e03080. doi: 10.7554/eLife.03080CrossRefGoogle ScholarPubMed
Wong, W, Bai, X-C, Sleebs, BE, Triglia, T, Brown, A, Thompson, JK, Jackson, KE, Hanssen, E, Marapana, DS, Fernandez, IS, Ralph, SA, Cowman, AF, Scheres, SHW and Baum, J (2017) Mefloquine targets the Plasmodium Falciparum 80S ribosome to inhibit protein synthesis. Nature Microbiology 2, 17031.CrossRefGoogle ScholarPubMed
World Malaria Report 2018 (2018) World Health Organisation.Google Scholar
Ye, Y, Liang, Y, Yu, Q, Hu, L, Li, H, Zhang, Z and Xu, X (2015) Analysis of human upstream open reading frames and impact on gene expression. Human Genetics 134, 605612.CrossRefGoogle ScholarPubMed
Young, SK, Willy, JA, Wu, C, Sachs, MS and Wek, RC (2015) Ribosome reinitiation directs gene-specific translation and regulates the integrated stress response. The Journal of Biological Chemistry 290, 2825728271.CrossRefGoogle ScholarPubMed
Zach, L, Braunstein, I and Stanhill, A (2014) Stress-induced start codon fidelity regulates arsenite-inducible regulatory particle-associated protein (AIRAP) translation. The Journal of Biological Chemistry 289, 2070620716.CrossRefGoogle ScholarPubMed
Zhang, M, Gallego-Delgado, J, Fernandez-Arias, C, Waters, NC, Rodriguez, A, Tsuji, M, Wek, RC, Nussenzweig, V and Sullivan, WJ (2017) Inhibiting the Plasmodium eIF2α Kinase PK4 Prevents Artemisinin-Induced Latency. Cell Host & Microbe 22, 766776.e4.CrossRefGoogle ScholarPubMed
Supplementary material: File

Kaur et al. supplementary material

Kaur et al. supplementary material 1

Download Kaur et al. supplementary material(File)
File 94.8 KB
Supplementary material: File

Kaur et al. supplementary material

Kaur et al. supplementary material 2

Download Kaur et al. supplementary material(File)
File 10.6 KB
Supplementary material: File

Kaur et al. supplementary material

Kaur et al. supplementary material 3

Download Kaur et al. supplementary material(File)
File 2.8 MB