Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-25T23:00:02.645Z Has data issue: false hasContentIssue false

Iron metabolism in trypanosomatids, and its crucial role in infection

Published online by Cambridge University Press:  15 February 2010

M. C. TAYLOR*
Affiliation:
Pathogen Molecular Biology Unit, Department of Infectious and Tropical Diseases, London School of Hygiene and Tropical Medicine, Keppel Street, LondonWC1E 7HT, UK
J. M. KELLY
Affiliation:
Pathogen Molecular Biology Unit, Department of Infectious and Tropical Diseases, London School of Hygiene and Tropical Medicine, Keppel Street, LondonWC1E 7HT, UK
*
*Corresponding author: Tel: 0044 207 927 2615. Fax: 0044 207 636 8739. E-mail: martin.c.taylor@lshtm.ac.uk

Summary

Iron is almost ubiquitous in living organisms due to the utility of its redox chemistry. It is also dangerous as it can catalyse the formation of reactive free radicals – a classical double-edged sword. In this review, we examine the uptake and usage of iron by trypanosomatids and discuss how modulation of host iron metabolism plays an important role in the protective response. Trypanosomatids require iron for crucial processes including DNA replication, antioxidant defence, mitochondrial respiration, synthesis of the modified base J and, in African trypanosomes, the alternative oxidase. The source of iron varies between species. Bloodstream-form African trypanosomes acquire iron from their host by uptake of transferrin, and Leishmania amazonensis expresses a ZIP family cation transporter in the plasma membrane. In other trypanosomatids, iron uptake has been poorly characterized. Iron-withholding responses by the host can be a major determinant of disease outcome. Their role in trypanosomatid infections is becoming apparent. For example, the cytosolic sequestration properties of NRAMP1, confer resistance against leishmaniasis. Conversely, cytoplasmic sequestration of iron may be favourable rather than detrimental to Trypanosoma cruzi. The central role of iron in both parasite metabolism and the host response is attracting interest as a possible point of therapeutic intervention.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2010

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.)

References

REFERENCES

Adak, S. and Datta, A. K. (2005). Leishmania major encodes an unusual peroxidase that is a close homologue of plant ascorbate peroxidase: a novel role of the transmembrane domain. The Biochemical Journal 390, 465474.CrossRefGoogle ScholarPubMed
Ajayi, W. U., Chaudhuri, M. and Hill, G. C. (2002). Site-directed mutagenesis reveals the essentiality of the conserved residues in the putative diiron active site of the trypanosome alternative oxidase. Journal of Biological Chemistry 277, 81878193.Google Scholar
Andersson, M. E. and Nordlund, P. (1999). A revised model of the active site of alternative oxidase. FEBS Letters 449, 1722.CrossRefGoogle ScholarPubMed
Arantes, J. M., Pedrosa, M. L., Martins, H. R., Veloso, V. M., de Lana, M., Bahia, M. T., Tafuri, W. L. and Carneiro, C. M. (2007). Trypanosoma cruzi: treatment with the iron chelator desferrioxamine reduces parasitemia and mortality in experimentally infected mice. Experimental Parasitology 117, 4350.Google Scholar
Berthold, D. A. and Stenmark, P. (2003). Membrane-bound diiron carboxylate proteins. Annu Rev Plant Biol 54, 497517.Google Scholar
Bienen, E. J., Webster, P. and Fish, W. R. (1991). Trypanosoma (Nannomonas) congolense: changes in respiratory metabolism during the life cycle. Experimental Parasitology 73, 403412.Google Scholar
Biggs, T. E., Baker, S. T., Botham, M. S., Dhital, A., Barton, C. H. and Perry, V. H. (2001). Nramp1 modulates iron homoeostasis in vivo and in vitro: evidence for a role in cellular iron release involving de-acidification of intracellular vesicles. European Journal of Immunology 31, 20602070.Google Scholar
Bisti, S., Konidou, G., Boelaert, J., Lebastard, M. and Soteriadou, K. (2006). The prevention of the growth of Leishmania major progeny in BALB/c iron-loaded mice: a process coupled to increased oxidative burst, the amplitude and duration of which depend on initial parasite developmental stage and dose. Microbes and Infection 8, 14641472.Google Scholar
Bisti, S., Konidou, G., Papageorgiou, F., Milon, G., Boelaert, J. R. and Soteriadou, K. (2000). The outcome of Leishmania major experimental infection in BALB/c mice can be modulated by exogenously delivered iron. European Journal of Immunology 30, 37323740.3.0.CO;2-D>CrossRefGoogle ScholarPubMed
Bisti, S. and Soteriadou, K. (2006). Is the reactive oxygen species-dependent-NF-kappaB activation observed in iron-loaded BALB/c mice a key process preventing growth of Leishmania major progeny and tissue-damage? Microbes and Infection 8, 14731482.CrossRefGoogle ScholarPubMed
Bitter, W., Gerrits, H., Kieft, R. and Borst, P. (1998). The role of transferrin-receptor variation in the host range of Trypanosoma brucei. Nature, London 391, 499502.Google Scholar
Blackwell, J. M., Fakiola, M., Ibrahim, M. E., Jamieson, S. E., Jeronimo, S. B., Miller, E. N., Mishra, A., Mohamed, H. S., Peacock, C. S., Raju, M., Sundar, S. and Wilson, M. E. (2009). Genetics and visceral leishmaniasis: of mice and man. Parasite Immunology 31, 254266.Google Scholar
Borges, V. M., Vannier-Santos, M. A. and de Souza, W. (1998). Subverted transferrin trafficking in Leishmania-infected macrophages. Parasitology Research 84, 811822.Google Scholar
Breidbach, T., Scory, S., Krauth-Siegel, R. L. and Steverding, D. (2002). Growth inhibition of bloodstream forms of Trypanosoma brucei by the iron chelator deferoxamine. International Journal for Parasitology 32, 473479.Google Scholar
Bucheton, B., Abel, L., Kheir, M. M., Mirgani, A., El-Safi, S. H., Chevillard, C. and Dessein, A. (2003). Genetic control of visceral leishmaniasis in a Sudanese population: candidate gene testing indicates a linkage to the NRAMP1 region. Genes Immun 4, 104109.Google Scholar
Carvalho, S., Cruz, T., Santarem, N., Castro, H., Costa, V. and Tomas, A. M. (2009). Heme as a source of iron to Leishmania infantum amastigotes. Acta Tropica 109, 131135.Google Scholar
Chang, C. S. and Chang, K. P. (1985). Heme requirement and acquisition by extracellular and intracellular stages of Leishmania mexicana amazonensis. Molecular and Biochemical Parasitology 16, 267276.Google Scholar
Chaudhuri, M., Ott, R. D. and Hill, G. C. (2006). Trypanosome alternative oxidase: from molecule to function. Trends in Parasitology 22, 484491.Google Scholar
Chaudhuri, M., Sharan, R. and Hill, G. C. (2002). Trypanosome alternative oxidase is regulated post-transcriptionally at the level of RNA stability. Journal of Eukaryotic Microbiology 49, 263269.Google Scholar
Clarkson, A. B. Jr., Bienen, E. J., Pollakis, G. and Grady, R. W. (1989). Respiration of bloodstream forms of the parasite Trypanosoma brucei brucei is dependent on a plant-like alternative oxidase. Journal of Biological Chemistry 264, 1777017776.CrossRefGoogle ScholarPubMed
Cliffe, L. J., Kieft, R., Southern, T., Birkeland, S. R., Marshall, M., Sweeney, K. and Sabatini, R. (2009). JBP1 and JBP2 are two distinct thymidine hydroxylases involved in J biosynthesis in genomic DNA of African trypanosomes. Nucleic Acids Research 37, 14521462.Google Scholar
Condo, I., Ventura, N., Malisan, F., Rufini, A., Tomassini, B. and Testi, R. (2007). In vivo maturation of human frataxin. Hum Mol Genet 16, 15341540.CrossRefGoogle ScholarPubMed
Coustou, V., Besteiro, S., Biran, M., Diolez, P., Bouchaud, V., Voisin, P., Michels, P. A., Canioni, P., Baltz, T. and Bringaud, F. (2003). ATP generation in the Trypanosoma brucei procyclic form: cytosolic substrate level is essential, but not oxidative phosphorylation. Journal of Biological Chemistry 278, 4962549635.CrossRefGoogle Scholar
d'Ieteren, G. D., Authie, E., Wissocq, N. and Murray, M. (1998). Trypanotolerance, an option for sustainable livestock production in areas at risk from trypanosomosis. Rev Sci Tech 17, 154175.Google Scholar
Das, N. K., Biswas, S., Solanki, S. and Mukhopadhyay, C. K. (2009). Leishmania donovani depletes labile iron pool to exploit iron uptake capacity of macrophage for its intracellular growth. Cell Microbiol 11, 8394.Google Scholar
Dolai, S., Yadav, R. K., Pal, S. and Adak, S. (2008). Leishmania major ascorbate peroxidase overexpression protects cells against reactive oxygen species-mediated cardiolipin oxidation. Free Radic Biol Med 45, 15201529.Google Scholar
Dormeyer, M., Reckenfelderbaumer, N., Ludemann, H. and Krauth-Siegel, R. L. (2001). Trypanothione-dependent synthesis of deoxyribonucleotides by Trypanosoma brucei ribonucleotide reductase. Journal of Biological Chemistry 276, 1060210606.Google Scholar
Dormeyer, M., Schoneck, R., Dittmar, G. A. and Krauth-Siegel, R. L. (1997). Cloning, sequencing and expression of ribonucleotide reductase R2 from Trypanosoma brucei. FEBS Letters 414, 449453.Google Scholar
Dufernez, F., Yernaux, C., Gerbod, D., Noel, C., Chauvenet, M., Wintjens, R., Edgcomb, V. P., Capron, M., Opperdoes, F. R. and Viscogliosi, E. (2006). The presence of four iron-containing superoxide dismutase isozymes in trypanosomatidae: characterization, subcellular localization, and phylogenetic origin in Trypanosoma brucei. Free Radic Biol Med 40, 210225.Google Scholar
Dupuy, J., Volbeda, A., Carpentier, P., Darnault, C., Moulis, J. M. and Fontecilla-Camps, J. C. (2006). Crystal structure of human iron regulatory protein 1 as cytosolic aconitase. Structure 14, 129139.Google Scholar
Fairlamb, A. H., Henderson, G. B., Bacchi, C. J. and Cerami, A. (1987). In vivo effects of difluoromethylornithine on trypanothione and polyamine levels in bloodstream forms of Trypanosoma brucei. Molecular and Biochemical Parasitology 24, 185191.Google Scholar
Fast, B., Kremp, K., Boshart, M. and Steverding, D. (1999). Iron-dependent regulation of transferrin receptor expression in Trypanosoma brucei. The Biochemical Journal 342, 691696.Google Scholar
Feng, C. G., Weksberg, D. C., Taylor, G. A., Sher, A. and Goodell, M. A. (2008). The p47 GTPase Lrg-47 (Irgm1) links host defense and hematopoietic stem cell proliferation. Cell Stem Cell 2, 8389.Google Scholar
Flohe, L., Hecht, H. J. and Steinert, P. (1999). Glutathione and trypanothione in parasitic hydroperoxide metabolism. Free Radic Biol Med 27, 966984.Google Scholar
Francisco, A. F., de Abreu Vieira, P. M., Arantes, J. M., Pedrosa, M. L., Martins, H. R., Silva, M., Veloso, V. M., de Lana, M., Bahia, M. T., Tafuri, W. L. and Carneiro, C. M. (2008). Trypanosoma cruzi: effect of benznidazole therapy combined with the iron chelator desferrioxamine in infected mice. Experimental Parasitology 120 314319.Google Scholar
Genest, P. A., ter Riet, B., Dumas, C., Papadopoulou, B., van Luenen, H. G. and Borst, P. (2005). Formation of linear inverted repeat amplicons following targeting of an essential gene in Leishmania. Nucleic Acids Research 33, 16991709.Google Scholar
Govoni, G., Vidal, S., Gauthier, S., Skamene, E., Malo, D. and Gros, P. (1996). The Bcg/Ity/Lsh locus: genetic transfer of resistance to infections in C57BL/6J mice transgenic for the Nramp1 Gly169 allele. Infection and Immunity 64, 29232929.Google Scholar
Gutierrez, M. G., Master, S. S., Singh, S. B., Taylor, G. A., Colombo, M. I. and Deretic, V. (2004). Autophagy is a defense mechanism inhibiting BCG and Mycobacterium tuberculosis survival in infected macrophages. Cell 119, 753766.Google Scholar
Hall, B. S., Gabernet-Castello, C., Voak, A., Goulding, D., Natesan, S. K. and Field, M. C. (2006). TbVps34, the trypanosome orthologue of Vps34, is required for Golgi complex segregation. Journal of Biological Chemistry 281, 2760027612.Google Scholar
Hall, B. S., Smith, E., Langer, W., Jacobs, L. A., Goulding, D. and Field, M. C. (2005). Developmental variation in Rab11-dependent trafficking in Trypanosoma brucei. Eukaryotic Cell 4, 971980.Google Scholar
Halliwell, B. and Gutteridge, J. M. C. (2007). Free Radicals in Biology and Medicine, 4th Edn. Oxford University Press, Oxford, UK.Google Scholar
Helfert, S., Estevez, A. M., Bakker, B., Michels, P. and Clayton, C. (2001). Roles of triosephosphate isomerase and aerobic metabolism in Trypanosoma brucei. The Biochemical Journal 357, 117125.Google Scholar
Hiner, A. N., Martinez, J. I., Arnao, M. B., Acosta, M., Turner, D. D., Lloyd Raven, E. and Rodriguez-Lopez, J. N. (2001). Detection of a tryptophan radical in the reaction of ascorbate peroxidase with hydrogen peroxide. European Journal of Biochemistry 268, 30913098.Google Scholar
Hofer, A., Schmidt, P. P., Graslund, A. and Thelander, L. (1997). Cloning and characterization of the R1 and R2 subunits of ribonucleotide reductase from Trypanosoma brucei. Proceedings of the National Academy of Sciences, USA 94, 69596964.Google Scholar
Horvath, A., Horakova, E., Dunajcikova, P., Verner, Z., Pravdova, E., Slapetova, I., Cuninkova, L. and Lukes, J. (2005). Downregulation of the nuclear-encoded subunits of the complexes III and IV disrupts their respective complexes but not complex I in procyclic Trypanosoma brucei. Molecular Microbiology 58, 116130.Google Scholar
Huynh, C., Sacks, D. L. and Andrews, N. W. (2006). A Leishmania amazonensis ZIP family iron transporter is essential for parasite replication within macrophage phagolysosomes. Journal of Experimental Medicine 203, 23632375.Google Scholar
Irigoin, F., Cibils, L., Comini, M. A., Wilkinson, S. R., Flohe, L. and Radi, R. (2008). Insights into the redox biology of Trypanosoma cruzi: Trypanothione metabolism and oxidant detoxification. Free Radic Biol Med 45, 733742.Google Scholar
Isobe, T., Holmes, E. C. and Rudenko, G. (2003). The transferrin receptor genes of Trypanosoma equiperdum are less diverse in their transferrin binding site than those of the broad-host range Trypanosoma brucei. Journal of Molecular Evolution 56, 377386.Google Scholar
Jabado, N., Cuellar-Mata, P., Grinstein, S. and Gros, P. (2003). Iron chelators modulate the fusogenic properties of Salmonella-containing phagosomes. Proceedings of the National Academy of Sciences, USA 100, 61276132.Google Scholar
Jeffries, T. R., Morgan, G. W. and Field, M. C. (2001). A developmentally regulated rab11 homologue in Trypanosoma brucei is involved in recycling processes. Journal of Cell Science 114, 26172626.Google Scholar
Kitajima, S., Kurioka, M., Yoshimoto, T., Shindo, M., Kanaori, K., Tajima, K. and Oda, K. (2008). A cysteine residue near the propionate side chain of heme is the radical site in ascorbate peroxidase. FEBS J 275, 470480.Google Scholar
Krauth-Siegel, R. L. and Ludemann, H. (1996). Reduction of dehydroascorbate by trypanothione. Mololecular and Biochemical Parasitology 80, 203208.Google Scholar
Krishnamurthy, G., Vikram, R., Singh, S. B., Patel, N., Agarwal, S., Mukhopadhyay, G., Basu, S. K. and Mukhopadhyay, A. (2005). Hemoglobin receptor in Leishmania is a hexokinase located in the flagellar pocket. Journal of Biological Chemistry 280, 58845891.Google Scholar
Lalonde, R. G. and Holbein, B. E. (1984). Role of iron in Trypanosoma cruzi infection of mice. Journal of Clinical Investigation 73, 470476.Google Scholar
Lara, F. A., Sant'anna, C., Lemos, D., Laranja, G. A., Coelho, M. G., Reis Salles, I., Michel, A., Oliveira, P. L., Cunha, E. S. N., Salmon, D. and Paes, M. C. (2007). Heme requirement and intracellular trafficking in Trypanosoma cruzi epimastigotes. Biochemical and Biophysical Research Communications 355, 1622.Google Scholar
Ligtenberg, M. J., Bitter, W., Kieft, R., Steverding, D., Janssen, H., Calafat, J. and Borst, P. (1994). Reconstitution of a surface transferrin binding complex in insect form Trypanosoma brucei. EMBO Journal 13, 25652573.Google Scholar
Lima, M. F. and Villalta, F. (1990). Trypanosoma cruzi receptors for human transferrin and their role. Molecular and Biochemical Parasitology 38, 245252.Google Scholar
Long, S., Jirku, M., Ayala, F. J. and Lukes, J. (2008 a). Mitochondrial localization of human frataxin is necessary but processing is not for rescuing frataxin deficiency in Trypanosoma brucei. Proceedings of the National Academy of Sciences, USA 105, 1346813473.Google Scholar
Long, S., Jirku, M., Mach, J., Ginger, M. L., Sutak, R., Richardson, D., Tachezy, J. and Lukes, J. (2008 b). Ancestral roles of eukaryotic frataxin: mitochondrial frataxin function and heterologous expression of hydrogenosomal Trichomonas homologues in trypanosomes. Molecular Microbiology 69, 94109.Google Scholar
MacMicking, J. D. (2005). Immune control of phagosomal bacteria by p47 GTPases. Current Opinion in Microbiology 8, 7482.Google Scholar
Maier, A. and Steverding, D. (1996). Low affinity of Trypanosoma brucei transferrin receptor to apotransferrin at pH 5 explains the fate of the ligand during endocytosis. FEBS Letters 396, 8789.Google Scholar
Marcondes, M. C., Borelli, P., Yoshida, N. and Russo, M. (2000). Acute Trypanosoma cruzi infection is associated with anemia, thrombocytopenia, leukopenia, and bone marrow hypoplasia: reversal by nifurtimox treatment. Microbes and Infection 2, 347352.Google Scholar
Mauricio, I. L., Stothard, J. R. and Miles, M. A. (2000). The strange case of Leishmania chagasi. Parasitology Today 16, 188189.Google Scholar
Mehta, A. and Shaha, C. (2006). Mechanism of metalloid-induced death in Leishmania spp.: role of iron, reactive oxygen species, Ca2+, and glutathione. Free Radic Biol Med 40, 18571868.Google Scholar
Merschjohann, K. and Steverding, D. (2006). In vitro growth inhibition of bloodstream forms of Trypanosoma brucei and Trypanosoma congolense by iron chelators. Kinetoplastid Biol Dis 5, 3.Google Scholar
Mohamed, H. S., Ibrahim, M. E., Miller, E. N., White, J. K., Cordell, H. J., Howson, J. M., Peacock, C. S., Khalil, E. A., El Hassan, A. M. and Blackwell, J. M. (2004). SLC11A1 (formerly NRAMP1) and susceptibility to visceral leishmaniasis in The Sudan. European Journal of Human Genetics 12, 6674.Google Scholar
Morgan, G. W., Allen, C. L., Jeffries, T. R., Hollinshead, M. and Field, M. C. (2001). Developmental and morphological regulation of clathrin-mediated endocytosis in Trypanosoma brucei. Journal of Cell Science 114, 26052615.Google Scholar
Mussmann, R., Engstler, M., Gerrits, H., Kieft, R., Toaldo, C. B., Onderwater, J., Koerten, H., van Luenen, H. G. and Borst, P. (2004). Factors affecting the level and localization of the transferrin receptor in Trypanosoma brucei. Journal of Biological Chemistry 279, 4069040698.Google Scholar
Mussmann, R., Janssen, H., Calafat, J., Engstler, M., Ansorge, I., Clayton, C. and Borst, P. (2003). The expression level determines the surface distribution of the transferrin receptor in Trypanosoma brucei. Molecular Microbiology 47, 2335.Google Scholar
Naessens, J. (2006). Bovine trypanotolerance: A natural ability to prevent severe anaemia and haemophagocytic syndrome? International Journal for Parasitology 36, 521528.Google Scholar
Nemeth, E., Preza, G. C., Jung, C. L., Kaplan, J., Waring, A. J. and Ganz, T. (2006). The N-terminus of hepcidin is essential for its interaction with ferroportin: structure-function study. Blood 107, 328333.Google Scholar
Nemeth, E., Tuttle, M. S., Powelson, J., Vaughn, M. B., Donovan, A., Ward, D. M., Ganz, T. and Kaplan, J. (2004). Hepcidin regulates cellular iron efflux by binding to ferroportin and inducing its internalization. Science 306, 20902093.Google Scholar
O'Brien, T. C., Mackey, Z. B., Fetter, R. D., Choe, Y., O'Donoghue, A. J., Zhou, M., Craik, C. S., Caffrey, C. R. and McKerrow, J. H. (2008). A parasite cysteine protease is key to host protein degradation and iron acquisition. Journal of Biological Chemistry 283, 2893428943.Google Scholar
Pal, A., Hall, B. S., Jeffries, T. R. and Field, M. C. (2003). Rab5 and Rab11 mediate transferrin and anti-variant surface glycoprotein antibody recycling in Trypanosoma brucei. The Biochemical Journal 374, 443451.Google Scholar
Peyssonnaux, C., Zinkernagel, A. S., Datta, V., Lauth, X., Johnson, R. S. and Nizet, V. (2006). TLR4-dependent hepcidin expression by myeloid cells in response to bacterial pathogens. Blood 107, 37273732.Google Scholar
Piacenza, L., Irigoin, F., Alvarez, M. N., Peluffo, G., Taylor, M. C., Kelly, J. M., Wilkinson, S. R. and Radi, R. (2007). Mitochondrial superoxide radicals mediate programmed cell death in Trypanosoma cruzi: cytoprotective action of mitochondrial iron superoxide dismutase overexpression. The Biochemical Journal 403, 323334.Google Scholar
Plewes, K. A., Barr, S. D. and Gedamu, L. (2003). Iron superoxide dismutases targeted to the glycosomes of Leishmania chagasi are important for survival. Infection and Immunity 71, 59105920.Google Scholar
Prathalingham, S. R., Wilkinson, S. R., Horn, D. and Kelly, J. M. (2007). Deletion of the Trypanosoma brucei superoxide dismutase gene sodb1 increases sensitivity to nifurtimox and benznidazole. Antimicrobial Agents and Chemotherapy 51, 755758.Google Scholar
Radtke, A. L. and O'Riordan, M. X. (2006). Intracellular innate resistance to bacterial pathogens. Cell Microbiol 8, 17201729.Google Scholar
Salahudeen, A. A., Thompson, J. W., Ruiz, J. C., Ma, H. W., Kinch, L. N., Li, Q., Grishin, N. V. and Bruick, R. K. (2009). An E3 ligase possessing an iron responsive hemerythrin domain is a regulator of iron homeostasis. Science 326, 722726.Google Scholar
Salmon, D., Geuskens, M., Hanocq, F., Hanocq-Quertier, J., Nolan, D., Ruben, L. and Pays, E. (1994). A novel heterodimeric transferrin receptor encoded by a pair of VSG expression site-associated genes in T. brucei. Cell 78, 7586.Google Scholar
Salmon, D., Hanocq-Quertier, J., Paturiaux-Hanocq, F., Pays, A., Tebabi, P., Nolan, D. P., Michel, A. and Pays, E. (1997). Characterization of the ligand-binding site of the transferrin receptor in Trypanosoma brucei demonstrates a structural relationship with the N-terminal domain of the variant surface glycoprotein. EMBO Journal 16, 72727278.Google Scholar
Salmon, D., Paturiaux-Hanocq, F., Poelvoorde, P., Vanhamme, L. and Pays, E. (2005). Trypanosoma brucei: growth differences in different mammalian sera are not due to the species-specificity of transferrin. Experimental Parasitology 109, 188194.Google Scholar
Santiago, H. C., Feng, C. G., Bafica, A., Roffe, E., Arantes, R. M., Cheever, A., Taylor, G., Vieira, L. Q., Aliberti, J., Gazzinelli, R. T. and Sher, A. (2005). Mice deficient in LRG-47 display enhanced susceptibility to Trypanosoma cruzi infection associated with defective hemopoiesis and intracellular control of parasite growth. Journal of Immunology 175, 81658172.Google Scholar
Schaible, U. E. and Kaufmann, S. H. (2004). Iron and microbial infection. Nat Rev Microbiol 2, 946953.Google Scholar
Schell, D., Borowy, N. K. and Overath, P. (1991 a). Transferrin is a growth factor for the bloodstream form of Trypanosoma brucei. Parasitology Research 77, 558560.Google Scholar
Schell, D., Evers, R., Preis, D., Ziegelbauer, K., Kiefer, H., Lottspeich, F., Cornelissen, A. W. and Overath, P. (1991 b). A transferrin-binding protein of Trypanosoma brucei is encoded by one of the genes in the variant surface glycoprotein gene expression site. EMBO Journal 10, 10611066.Google Scholar
Schofield, C. J. and Zhang, Z. (1999). Structural and mechanistic studies on 2-oxoglutarate-dependent oxygenases and related enzymes. Curr Opin Struct Biol 9, 722731.Google Scholar
Schwartz, K. J. and Bangs, J. D. (2007). Regulation of protein trafficking by glycosylphosphatidylinositol valence in African trypanosomes. Journal of Eukaryotic Microbiology 54, 2224.Google Scholar
Schwartz, K. J., Peck, R. F., Tazeh, N. N. and Bangs, J. D. (2005). GPI valence and the fate of secretory membrane proteins in African trypanosomes. Journal of Cell Science 118, 54995511.Google Scholar
Smid, O., Horakova, E., Vilimova, V., Hrdy, I., Cammack, R., Horvath, A., Lukes, J. and Tachezy, J. (2006). Knock-downs of iron-sulfur cluster assembly proteins IscS and IscU down-regulate the active mitochondrion of procyclic Trypanosoma brucei. Journal of Biological Chemistry 281, 2867928686.Google Scholar
Soares, M. J. and de Souza, W. (1991). Endocytosis of gold-labeled proteins and LDL by Trypanosoma cruzi. Parasitology Research 77, 461468.Google Scholar
Soares, M. J., Souto-Padron, T. and De Souza, W. (1992). Identification of a large pre-lysosomal compartment in the pathogenic protozoon Trypanosoma cruzi. Journal of Cell Science 102, 157167.Google Scholar
Srivastava, P., Sharma, G. D., Kamboj, K. K., Rastogi, A. K. and Pandey, V. C. (1997). Heme metabolism in promastigotes of Leishmania donovani. Molecular and Cellular Biochemistry 171, 6568.Google Scholar
Steverding, D. (1998). Bloodstream forms of Trypanosoma brucei require only small amounts of iron for growth. Parasitol Research 84, 5962.Google Scholar
Steverding, D. (2003). The significance of transferrin receptor variation in Trypanosoma brucei. Trends in Parasitology 19, 125127.Google Scholar
Steverding, D. (2006). On the significance of host antibody response to the Trypanosoma brucei transferrin receptor during chronic infection. Microbes and Infection 8, 27772782.Google Scholar
Steverding, D., Stierhof, Y. D., Chaudhri, M., Ligtenberg, M., Schell, D., Beck-Sickinger, A. G. and Overath, P. (1994). ESAG 6 and 7 products of Trypanosoma brucei form a transferrin binding protein complex. European Journal of Cell Biology 64, 7887.Google Scholar
Steverding, D., Stierhof, Y. D., Fuchs, H., Tauber, R. and Overath, P. (1995). Transferrin-binding protein complex is the receptor for transferrin uptake in Trypanosoma brucei. Journal of Cell Biology 131, 11731182.Google Scholar
Stijlemans, B., Vankrunkelsven, A., Brys, L., Magez, S. and De Baetselier, P. (2008). Role of iron homeostasis in trypanosomiasis-associated anemia. Immunobiology 213, 823835.Google Scholar
Taylor, M. C. and Kelly, J. M. (2006). pTcINDEX: a stable tetracycline-regulated expression vector for Trypanosoma cruzi. BMC Biotechnol 6, 32.Google Scholar
Temperton, N. J., Wilkinson, S. R., Meyer, D. J. and Kelly, J. M. (1998). Overexpression of superoxide dismutase in Trypanosoma cruzi results in increased sensitivity to the trypanocidal agents gentian violet and benznidazole. Molecular and Biochemical Parasitology 96, 167176.Google Scholar
Theodoropoulos, T. A., Silva, A. G. and Bestetti, R. B. (2009). Eosinophil blood count and anemia are associated with Trypanosoma cruzi infection reactivation in Chagas' heart transplant recipients. Int J Cardiol (in the Press).Google Scholar
Vainio, S., Genest, P. A., ter Riet, B., van Luenen, H. and Borst, P. (2009). Evidence that J-binding protein 2 is a thymidine hydroxylase catalyzing the first step in the biosynthesis of DNA base J. Molecular and Biochemical Parasitology 164, 157161.Google Scholar
van Luenen, H. G., Kieft, R., Mussmann, R., Engstler, M., ter Riet, B. and Borst, P. (2005). Trypanosomes change their transferrin receptor expression to allow effective uptake of host transferrin. Molecular Microbiology 58, 151165.Google Scholar
Vashisht, A. A., Zumbrennen, K. B., Huang, X., Powers, D. N., Durazo, A., Sun, D., Bhaskaran, N., Persson, A., Uhlen, M., Sangfelt, O., Spruck, C., Leibold, E. A. and Wohlschlegel, J. A. (2009). Control of iron homeostasis by an iron-regulated ubiquitin ligase. Science 326, 718721.CrossRefGoogle ScholarPubMed
Vidal, S., Tremblay, M. L., Govoni, G., Gauthier, S., Sebastiani, G., Malo, D., Skamene, E., Olivier, M., Jothy, S. and Gros, P. (1995). The Ity/Lsh/Bcg locus: natural resistance to infection with intracellular parasites is abrogated by disruption of the Nramp1 gene. Journal of Experimental Medicine 182, 655666.Google Scholar
Voyiatzaki, C. S. and Soteriadou, K. P. (1992). Identification and isolation of the Leishmania transferrin receptor. Journal of Biological Chemistry 267, 91129117.Google Scholar
Walden, W. E., Selezneva, A. I., Dupuy, J., Volbeda, A., Fontecilla-Camps, J. C., Theil, E. C. and Volz, K. (2006). Structure of dual function iron regulatory protein 1 complexed with ferritin IRE-RNA. Science 314, 19031908.Google Scholar
Wallander, M. L., Leibold, E. A. and Eisenstein, R. S. (2006). Molecular control of vertebrate iron homeostasis by iron regulatory proteins. Biochimica Biophysica Acta 1763, 668689.Google Scholar
Wilkinson, S. R. and Kelly, J. M. (2003). The role of glutathione peroxidases in trypanosomatids. Biological Chemistry 384, 517525.Google Scholar
Wilkinson, S. R., Obado, S. O., Mauricio, I. L. and Kelly, J. M. (2002 a). Trypanosoma cruzi expresses a plant-like ascorbate-dependent hemoperoxidase localized to the endoplasmic reticulum. Proceedings of the National Academy of Sciences, USA 99, 1345313458.Google Scholar
Wilkinson, S. R., Prathalingam, S. R., Taylor, M. C., Ahmed, A., Horn, D. and Kelly, J. M. (2006). Functional characterisation of the iron superoxide dismutase gene repertoire in Trypanosoma brucei. Free Radic Biol Med 40, 198209.Google Scholar
Wilkinson, S. R., Taylor, M. C., Touitha, S., Mauricio, I. L., Meyer, D. J. and Kelly, J. M. (2002 b). TcGPXII, a glutathione-dependent Trypanosoma cruzi peroxidase with substrate specificity restricted to fatty acid and phospholipid hydroperoxides, is localized to the endoplasmic reticulum. The Biochemical Journal 364, 787794.Google Scholar
Wilkinson, S. R., Temperton, N. J., Mondragon, A. and Kelly, J. M. (2000). Distinct mitochondrial and cytosolic enzymes mediate trypanothione-dependent peroxide metabolism in Trypanosoma cruzi. Journal of Biological Chemistry 275, 82208225.CrossRefGoogle ScholarPubMed
Wilson, M. E., Lewis, T. S., Miller, M. A., McCormick, M. L. and Britigan, B. E. (2002). Leishmania chagasi: uptake of iron bound to lactoferrin or transferrin requires an iron reductase. Experimental Parasitology 100, 196207.Google Scholar
Wilson, M. E., Vorhies, R. W., Andersen, K. A. and Britigan, B. E. (1994). Acquisition of iron from transferrin and lactoferrin by the protozoan Leishmania chagasi. Infection and Immunity 62, 32623269.Google Scholar
Witola, W. H., Sarataphan, N., Inoue, N., Ohashi, K. and Onuma, M. (2005). Genetic variability in ESAG6 genes among Trypanosoma evansi isolates and in comparison to other Trypanozoon members. Acta Tropica 93, 6373.CrossRefGoogle ScholarPubMed
Yabu, Y., Minagawa, N., Kita, K., Nagai, K., Honma, M., Sakajo, S., Koide, T., Ohta, N. and Yoshimoto, A. (1998). Oral and intraperitoneal treatment of Trypanosoma brucei brucei with a combination of ascofuranone and glycerol in mice. Parasitology International 47, 131137.Google Scholar
Yabu, Y., Suzuki, T., Nihei, C., Minagawa, N., Hosokawa, T., Nagai, K., Kita, K. and Ohta, N. (2006). Chemotherapeutic efficacy of ascofuranone in Trypanosoma vivax-infected mice without glycerol. Parasitology International 55, 3943.Google Scholar
Yabu, Y., Yoshida, A., Suzuki, T., Nihei, C., Kawai, K., Minagawa, N., Hosokawa, T., Nagai, K., Kita, K. and Ohta, N. (2003). The efficacy of ascofuranone in a consecutive treatment on Trypanosoma brucei brucei in mice. Parasitology International 52, 155164.Google Scholar
Yu, Z., Genest, P. A., ter Riet, B., Sweeney, K., DiPaolo, C., Kieft, R., Christodoulou, E., Perrakis, A., Simmons, J. M., Hausinger, R. P., van Luenen, H. G., Rigden, D. J., Sabatini, R. and Borst, P. (2007). The protein that binds to DNA base J in trypanosomatids has features of a thymidine hydroxylase. Nucleic Acids Research 35, 21072115.Google Scholar