Iron deprivation in cancer--potential therapeutic implications

Jessica L Heath, Joshua M Weiss, Catherine P Lavau, Daniel S Wechsler, Jessica L Heath, Joshua M Weiss, Catherine P Lavau, Daniel S Wechsler

Abstract

Iron is essential for normal cellular function. It participates in a wide variety of cellular processes, including cellular respiration, DNA synthesis, and macromolecule biosynthesis. Iron is required for cell growth and proliferation, and changes in intracellular iron availability can have significant effects on cell cycle regulation, cellular metabolism, and cell division. Perhaps not surprisingly then, neoplastic cells have been found to have higher iron requirements than normal, non-malignant cells. Iron depletion through chelation has been explored as a possible therapeutic intervention in a variety of cancers. Here, we will review iron homeostasis in non-malignant and malignant cells, the widespread effects of iron depletion on the cell, the various iron chelators that have been explored in the treatment of cancer, and the tumor types that have been most commonly studied in the context of iron chelation.

Figures

Figure 1
Figure 1
Intestinal absorption of iron. Ferric (Fe3+) iron present in the intestinal lumen is reduced to ferrous (Fe2+) iron, which can be transported by DMT1 across the apical surface of the enterocyte. Intracellular iron may be stored in ferritin, or transported out of the cell via ferroportin. After its release from the cell, it is oxidized by hephaestin and bound to transferrin for transport through the body.
Figure 2
Figure 2
Cellular iron metabolism. Iron laden transferrin is taken up by the cell via clathrin-mediated endocytosis of transferrin receptor-1 (TfR1). Acidification of the resultant endosome releases iron from TfR. Iron is transported out of the endosome via DMT1, and becomes part of the labile iron pool. Iron may then be used in a variety of cellular processes, stored as ferritin, or exported out of the cell via ferroportin.
Figure 3
Figure 3
Regulation of translation by intracellular iron through IRP binding to mRNA IREs. (A) mRNA is stabilized by the binding of the IRP to the 3′ IRE in the absence of iron. When iron is abundant, iron-bound IRP is displaced, allowing mRNA degradation. (B) Translation is repressed by the binding of an IRP to a 5′ IRE. Iron abundance removes the IRP and allows translation to proceed.
Figure 4
Figure 4
The chemical structure of several key iron chelators. Abbreviations: PIH––pyridoxal isonicotinoyl hydrazone; Dp44mT––di-2-pyridylketone-4,4,-dimethyl-3-thiosemicarbazone.

References

    1. Fuqua B.K., Vulpe C.D., Anderson G.J. Intestinal iron absorption. J. Trace Elem. Med. Biol. 2012;26:115–119. doi: 10.1016/j.jtemb.2012.03.015.
    1. De Domenico I., Ward D.M., Langelier C., Vaughn M.B., Nemeth E., Sundquist W.I., Ganz T., Musci G., Kaplan J. The molecular mechanism of hepcidin-mediated ferroportin down-regulation. Mol. Biol. Cell. 2007;18:2569–2578. doi: 10.1091/mbc.E07-01-0060.
    1. Aisen P. Transferrin receptor 1. Int. J. Biochem. Cell Biol. 2004;36:2137–2143. doi: 10.1016/j.biocel.2004.02.007.
    1. Dautry-Varsat A., Ciechanover A., Lodish H.F. pH and the recycling of transferrin during receptor-mediated endocytosis. Proc. Natl. Acad. Sci. USA. 1983;80:2258–2262. doi: 10.1073/pnas.80.8.2258.
    1. Graf E., Mahoney J.R., Bryant R.G., Eaton J.W. Iron-catalyzed hydroxyl radical formation. Stringent requirement for free iron coordination site. J. Biol. Chem. 1984;259:3620–3624.
    1. Cairo G., Recalcati S. Iron-regulatory proteins: Molecular biology and pathophysiological implications. Expert Rev. Mol. Med. 2007;9:1–13.
    1. Cairo G., Tacchini L., Pogliaghi G., Anzon E., Tomasi A., Bernelli-Zazzera A. Induction of ferritin synthesis by oxidative stress. Transcriptional and post-transcriptional regulation by expansion of the “free” iron pool. J. Biol. Chem. 1995;270:700–703.
    1. Meyron-Holtz E.G., Ghosh M.C., Rouault T.A. Mammalian tissue oxygen levels modulate iron-regulatory protein activities in vivo. Science. 2004;306:2087–2090. doi: 10.1126/science.1103786.
    1. Recalcati S., Alberghini A., Campanella A., Gianelli U., de Camilli E., Conte D., Cairo G. Iron regulatory proteins 1 and 2 in human monocytes, macrophages and duodenum: Expression and regulation in hereditary hemochromatosis and iron deficiency. Haematologica. 2006;91:303–310.
    1. Kennedy M.C., Mende-Mueller L., Blondin G.A., Beinert H. Purification and characterization of cytosolic aconitase from beef liver and its relationship to the iron-responsive element binding protein. Proc. Natl. Acad. Sci. USA. 1992;89:11730–11734. doi: 10.1073/pnas.89.24.11730.
    1. Elford H.L., Freese M., Passamani E., Morris H.P. Ribonucleotide reductase and cell proliferation. I. Variations of ribonucleotide reductase activity with tumor growth rate in a series of rat hepatomas. J. Biol. Chem. 1970;245:5228–5233.
    1. Trinder D., Zak O., Aisen P. Transferrin receptor-independent uptake of differic transferrin by human hepatoma cells with antisense inhibition of receptor expression. Hepatology. 1996;23:1512–1520. doi: 10.1002/hep.510230631.
    1. Richardson D.R., Baker E. The uptake of iron and transferrin by the human malignant melanoma cell. Biochim. Biophys. Acta. 1990;1053:1–12. doi: 10.1016/0167-4889(90)90018-9.
    1. Dang C.V. Myc on the path to cancer. Cell. 2012;149:22–35. doi: 10.1016/j.cell.2012.03.003.
    1. Adhikary S., Eilers M. Transcriptional regulation and transformation by myc proteins. Nat. Rev. Mol. Cell. Biol. 2005;6:635–645. doi: 10.1038/nrm1703.
    1. Dang C.V. C-myc target genes involved in cell growth, apoptosis, and metabolism. Mol. Cell. Biol. 1999;19:1–11.
    1. Wu K.J., Polack A., Dalla-Favera R. Coordinated regulation of iron-controlling genes, h-ferritin and irp2, by c-myc. Science. 1999;283:676–679. doi: 10.1126/science.283.5402.676.
    1. Graslund A., Ehrenberg A., Thelander L. Characterization of the free radical of mammalian ribonucleotide reductase. J. Biol. Chem. 1982;257:5711–5715.
    1. Green D.A., Antholine W.E., Wong S.J., Richardson D.R., Chitambar C.R. Inhibition of malignant cell growth by 311, a novel iron chelator of the pyridoxal isonicotinoyl hydrazone class: Effect on the r2 subunit of ribonucleotide reductase. Clin. Cancer Res. 2001;7:3574–3579.
    1. Shang H., Li Q., Feng G., Cui Z. Molecular analysis and functions of p53r2 in zebrafish. Gene. 2011;475:30–38. doi: 10.1016/j.gene.2010.12.008.
    1. Wei P.P., Tomter A.B., Rohr A.K., Andersson K.K., Solomon E.I. Circular dichroism and magnetic circular dichroism studies of the active site of p53r2 from human and mouse: Iron binding and nature of the biferrous site relative to other ribonucleotide reductases. Biochemistry. 2006;45:14043–14051. doi: 10.1021/bi061127p.
    1. Chen D., Li M., Luo J., Gu W. Direct interactions between hif-1 αand mdm2 modulate p53 function. J. Biol. Chem. 2003;278:13595–13598. doi: 10.1074/jbc.C200694200.
    1. Chen K., Albano A., Ho A., Keaney J.F., Jr. Activation of p53 by oxidative stress involves platelet-derived growth factor-βreceptor-mediated ataxia telangiectasia mutated (atm) kinase activation. J. Biol. Chem. 2003;278:39527–39533. doi: 10.1074/jbc.M304423200.
    1. Fukuchi K., Tomoyasu S., Watanabe H., Kaetsu S., Tsuruoka N., Gomi K. Iron deprivation results in an increase in p53 expression. Biol. Chem. Hoppe-Seyler. 1995;376:627–630.
    1. Liang S.X., Richardson D.R. The effect of potent iron chelators on the regulation of p53: Examination of the expression, localization and DNA-binding activity of p53 and the transactivation of waf1. Carcinogenesis. 2003;24:1601–1614. doi: 10.1093/carcin/bgg116.
    1. Saletta F., Suryo Rahmanto Y., Noulsri E., Richardson D.R. Iron chelator-mediated alterations in gene expression: Identification of novel iron-regulated molecules that are molecular targets of hypoxia-inducible factor-1 alpha and p53. Mol. Pharmacol. 2010;77:443–458. doi: 10.1124/mol.109.061028.
    1. Dongiovanni P., Fracanzani A.L., Cairo G., Megazzini C.P., Gatti S., Rametta R., Fargion S., Valenti L. Iron-dependent regulation of mdm2 influences p53 activity and hepatic carcinogenesis. Am. J. Pathol. 2010;176:1006–1017. doi: 10.2353/ajpath.2010.090249.
    1. Zhang F., Wang W., Tsuji Y., Torti S.V., Torti F.M. Post-transcriptional modulation of iron homeostasis during p53-dependent growth arrest. J. Biol. Chem. 2008;283:33911–33918. doi: 10.1074/jbc.M806432200.
    1. Weizer-Stern O., Adamsky K., Margalit O., Ashur-Fabian O., Givol D., Amariglio N., Rechavi G. Hepcidin, a key regulator of iron metabolism, is transcriptionally activated by p53. Br. J. Haematol. 2007;138:253–262. doi: 10.1111/j.1365-2141.2007.06638.x.
    1. Campisi J. Cellular senescence as a tumor-suppressor mechanism. Trends Cell Biol. 2001;11:S27–S31.
    1. Lagger G., Doetzlhofer A., Schuettengruber B., Haidweger E., Simboeck E., Tischler J., Chiocca S., Suske G., Rotheneder H., Wintersberger E., et al. The tumor suppressor p53 and histone deacetylase 1 are antagonistic regulators of the cyclin-dependent kinase inhibitor p21/waf1/cip1 gene. Mol. Cell. Biol. 2003;23:2669–2679. doi: 10.1128/MCB.23.8.2669-2679.2003.
    1. Abbas T., Dutta A. P21 in cancer: Intricate networks and multiple activities. Nat. Rev. Cancer. 2009;9:400–414. doi: 10.1038/nrc2657.
    1. Mitra J., Dai C.Y., Somasundaram K., El-Deiry W.S., Satyamoorthy K., Herlyn M., Enders G.H. Induction of p21(waf1/cip1) and inhibition of cdk2 mediated by the tumor suppressor p16(ink4a) Mol. Cell. Biol. 1999;19:3916–3928.
    1. Cayrol C., Knibiehler M., Ducommun B. P21 binding to pcna causes g1 and g2 cell cycle arrest in p53-deficient cells. Oncogene. 1998;16:311–320.
    1. Stein G.H., Drullinger L.F., Soulard A., Dulic V. Differential roles for cyclin-dependent kinase inhibitors p21 and p16 in the mechanisms of senescence and differentiation in human fibroblasts. Mol. Cell. Biol. 1999;19:2109–2117.
    1. Zhang Y., Fujita N., Tsuruo T. Caspase-mediated cleavage of p21waf1/cip1 converts cancer cells from growth arrest to undergoing apoptosis. Oncogene. 1999;18:1131–1138. doi: 10.1038/sj.onc.1202426.
    1. Kawata S., Ariumi Y., Shimotohno K. P21(waf1/cip1/sdi1) prevents apoptosis as well as stimulates growth in cells transformed or immortalized by human t-cell leukemia virus type 1-encoded tax. J. Virol. 2003;77:7291–7299. doi: 10.1128/JVI.77.13.7291-7299.2003.
    1. Biankin A.V., Kench J.G., Dijkman F.P., Biankin S.A., Henshall S.M. Molecular pathogenesis of precursor lesions of pancreatic ductal adenocarcinoma. Pathology. 2003;35:14–24.
    1. Weiss R.H. P21waf1/cip1 as a therapeutic target in breast and other cancers. Cancer Cell. 2003;4:425–429. doi: 10.1016/S1535-6108(03)00308-8.
    1. Gazitt Y., Reddy S.V., Alcantara O., Yang J., Boldt D.H. A new molecular role for iron in regulation of cell cycling and differentiation of hl-60 human leukemia cells: Iron is required for transcription of p21(waf1/cip1) in cells induced by phorbol myristate acetate. J. Cell. Physiol. 2001;187:124–135. doi: 10.1002/1097-4652(2001)9999:9999<::AID-JCP1061>;2-E.
    1. Kramer J.L., Baltathakis I., Alcantara O.S., Boldt D.H. Differentiation of functional dendritic cells and macrophages from human peripheral blood monocyte precursors is dependent on expression of p21 (waf1/cip1) and requires iron. Br. J. Haematol. 2002;117:727–734. doi: 10.1046/j.1365-2141.2002.03498.x.
    1. Fu D., Richardson D.R. Iron chelation and regulation of the cell cycle: 2 Mechanisms of posttranscriptional regulation of the universal cyclin-dependent kinase inhibitor p21cip1/waf1 by iron depletion. Blood. 2007;110:752–761. doi: 10.1182/blood-2007-03-076737.
    1. Deshpande A., Sicinski P., Hinds P.W. Cyclins and cdks in development and cancer: A perspective. Oncogene. 2005;24:2909–2915. doi: 10.1038/sj.onc.1208618.
    1. Gao J., Richardson D.R. The potential of iron chelators of the pyridoxal isonicotinoyl hydrazone class as effective antiproliferative agents, iv: The mechanisms involved in inhibiting cell-cycle progression. Blood. 2001;98:842–850. doi: 10.1182/blood.V98.3.842.
    1. Kulp K.S., Green S.L., Vulliet P.R. Iron deprivation inhibits cyclin-dependent kinase activity and decreases cyclin d/cdk4 protein levels in asynchronous mda-mb-453 human breast cancer cells. Exp. Cell Res. 1996;229:60–68. doi: 10.1006/excr.1996.0343.
    1. Nurtjahja-Tjendraputra E., Fu D., Phang J.M., Richardson D.R. Iron chelation regulates cyclin d1 expression via the proteasome: A link to iron deficiency-mediated growth suppression. Blood. 2007;109:4045–4054. doi: 10.1182/blood-2006-10-047753.
    1. Masamha C.P., Benbrook D.M. Cyclin d1 degradation is sufficient to induce g1 cell cycle arrest despite constitutive expression of cyclin e2 in ovarian cancer cells. Cancer Res. 2009;69:6565–6572. doi: 10.1158/0008-5472.CAN-09-0913.
    1. Oexle H., Gnaiger E., Weiss G. Iron-dependent changes in cellular energy metabolism: Influence on citric acid cycle and oxidative phosphorylation. Biochim. Biophys. Acta. 1999;1413:99–107.
    1. King A., Selak M.A., Gottlieb E. Succinate dehydrogenase and fumarate hydratase: Linking mitochondrial dysfunction and cancer. Oncogene. 2006;25:4675–4682. doi: 10.1038/sj.onc.1209594.
    1. Wen F.Q., Jabbar A.A., Chen Y.X., Kazarian T., Patel D.A., Valentino L.A. C-myc proto-oncogene expression in hemophilic synovitis: In vitro studies of the effects of iron and ceramide. Blood. 2002;100:912–916. doi: 10.1182/blood-2002-02-0390.
    1. Fan L., Iyer J., Zhu S., Frick K.K., Wada R.K., Eskenazi A.E., Berg P.E., Ikegaki N., Kennett R.H., Frantz C.N. Inhibition of N-myc expression and induction of apoptosis by iron chelation in human neuroblastoma cells. Cancer Res. 2001;61:1073–1079.
    1. Shim H., Dolde C., Lewis B.C., Wu C.S., Dang G., Jungmann R.A., Dalla-Favera R., Dang C.V. C-myc transactivation of ldh-a: Implications for tumor metabolism and growth. Proc. Natl. Acad. Sci. USA. 1997;94:6658–6663. doi: 10.1073/pnas.94.13.6658.
    1. Keith B., Johnson R.S., Simon M.C. Hif1α and hif2α: Sibling rivalry in hypoxic tumour growth and progression. Nat. Rev. Cancer. 2012;12:9–22.
    1. Talks K.L., Turley H., Gatter K.C., Maxwell P.H., Pugh C.W., Ratcliffe P.J., Harris A.L. The expression and distribution of the hypoxia-inducible factors hif-1α and hif-2α in normal human tissues, cancers, and tumor-associated macrophages. Am. J. Pathol. 2000;157:411–421. doi: 10.1016/S0002-9440(10)64554-3.
    1. Lidgren A., Hedberg Y., Grankvist K., Rasmuson T., Vasko J., Ljungberg B. The expression of hypoxia-inducible factor 1α is a favorable independent prognostic factor in renal cell carcinoma. Clin. Cancer Res. 2005;11:1129–1135.
    1. Noguera R., Fredlund E., Piqueras M., Pietras A., Beckman S., Navarro S., Pahlman S. Hif-1α and hif-2α are differentially regulated in vivo in neuroblastoma: High hif-1α correlates negatively to advanced clinical stage and tumor vascularization. Clin. Cancer Res. 2009;15:7130–7136. doi: 10.1158/1078-0432.CCR-09-0223.
    1. Callapina M., Zhou J., Schnitzer S., Metzen E., Lohr C., Deitmer J.W., Brune B. Nitric oxide reverses desferrioxamine- and hypoxia-evoked hif-1α accumulation—Implications for prolyl hydroxylase activity and iron. Exp. Cell Res. 2005;306:274–284. doi: 10.1016/j.yexcr.2005.02.018.
    1. Ellen T.P., Ke Q., Zhang P., Costa M. Ndrg1, a growth and cancer related gene: Regulation of gene expression and function in normal and disease states. Carcinogenesis. 2008;29:2–8. doi: 10.1093/carcin/bgm200.
    1. Le N.T., Richardson D.R. Iron chelators with high antiproliferative activity up-regulate the expression of a growth inhibitory and metastasis suppressor gene: A link between iron metabolism and proliferation. Blood. 2004;104:2967–2975. doi: 10.1182/blood-2004-05-1866.
    1. Guan R.J., Ford H.L., Fu Y., Li Y., Shaw L.M., Pardee A.B. Drg-1 as a differentiation-related, putative metastatic suppressor gene in human colon cancer. Cancer Res. 2000;60:749–755.
    1. Bandyopadhyay S., Pai S.K., Gross S.C., Hirota S., Hosobe S., Miura K., Saito K., Commes T., Hayashi S., Watabe M., et al. The drg-1 gene suppresses tumor metastasis in prostate cancer. Cancer Res. 2003;63:1731–1736.
    1. Akiba J., Ogasawara S., Kawahara A., Nishida N., Sanada S., Moriya F., Kuwano M., Nakashima O., Yano H. N-myc downstream regulated gene 1 (ndrg1)/cap43 enhances portal vein invasion and intrahepatic metastasis in human hepatocellular carcinoma. Oncol. Rep. 2008;20:1329–1335.
    1. Chua M.S., Sun H., Cheung S.T., Mason V., Higgins J., Ross D.T., Fan S.T., So S. Overexpression of ndrg1 is an indicator of poor prognosis in hepatocellular carcinoma. Mod. Pathol. 2007;20:76–83. doi: 10.1038/modpathol.3800711.
    1. Blatt J., Stitely S. Antineuroblastoma activity of desferoxamine in human cell lines. Cancer Res. 1987;47:1749–1750.
    1. Foa P., Maiolo A.T., Lombardi L., Villa L., Polli E.E. Inhibition of proliferation of human leukaemic cell populations by deferoxamine. Scand. J. Haematol. 1986;36:107–110.
    1. Reddel R.R., Hedley D.W., Sutherland R.L. Cell cycle effects of iron depletion on t-47d human breast cancer cells. Exp. Cell Res. 1985;161:277–284. doi: 10.1016/0014-4827(85)90085-0.
    1. Hodges Y.K., Antholine W.E., Horwitz L.D. Effect on ribonucleotide reductase of novel lipophilic iron chelators: The desferri-exochelins. Biochem. Biophys. Res. Commun. 2004;315:595–598. doi: 10.1016/j.bbrc.2004.01.101.
    1. Kalinowski D.S., Richardson D.R. The evolution of iron chelators for the treatment of iron overload disease and cancer. Pharmacol. Rev. 2005;57:547–583. doi: 10.1124/pr.57.4.2.
    1. Keberle H. The biochemistry of desferrioxamine and its relation to iron metabolism. Ann. N. Y. Acad. Sci. 1964;119:758–768. doi: 10.1111/j.1749-6632.1965.tb54077.x.
    1. Pullarkat V., Sehgal A., Li L., Meng Z., Lin A., Forman S., Bhatia R. Deferasirox exposure induces reactive oxygen species and reduces growth and viability of myelodysplastic hematopoietic progenitors. Leuk. Res. 2012;36:966–973. doi: 10.1016/j.leukres.2012.03.018.
    1. Ohyashiki J.H., Kobayashi C., Hamamura R., Okabe S., Tauchi T., Ohyashiki K. The oral iron chelator deferasirox represses signaling through the mtor in myeloid leukemia cells by enhancing expression of redd1. Cancer Sci. 2009;100:970–977. doi: 10.1111/j.1349-7006.2009.01131.x.
    1. Kim J.L., Kang H.N., Kang M.H., Yoo Y.A., Kim J.S., Choi C.W. The oral iron chelator deferasirox induces apoptosis in myeloid leukemia cells by targeting caspase. Acta Haematol. 2011;126:241–245. doi: 10.1159/000330608.
    1. Chantrel-Groussard K., Gaboriau F., Pasdeloup N., Havouis R., Nick H., Pierre J.L., Brissot P., Lescoat G. The new orally active iron chelator icl670a exhibits a higher antiproliferative effect in human hepatocyte cultures than O-trensox. Eur. J. Pharmacol. 2006;541:129–137. doi: 10.1016/j.ejphar.2006.05.001.
    1. Torti S.V., Torti F.M., Whitman S.P., Brechbiel M.W., Park G., Planalp R.P. Tumor cell cytotoxicity of a novel metal chelator. Blood. 1998;92:1384–1389.
    1. Samuni A.M., Krishna M.C., DeGraff W., Russo A., Planalp R.P., Brechbiel M.W., Mitchell J.B. Mechanisms underlying the cytotoxic effects of tachpyr—A novel metal chelator. Biochim. Biophys. Acta. 2002;1571:211–218.
    1. Turner J., Koumenis C., Kute T.E., Planalp R.P., Brechbiel M.W., Beardsley D., Cody B., Brown K.D., Torti F.M., Torti S.V. Tachpyridine, a metal chelator, induces g2 cell-cycle arrest, activates checkpoint kinases, and sensitizes cells to ionizing radiation. Blood. 2005;106:3191–3199. doi: 10.1182/blood-2005-03-1263.
    1. Baker E., Vitolo M.L., Webb J. Iron chelation by pyridoxal isonicotinoyl hydrazone and analogues in hepatocytes in culture. Biochem. Pharmacol. 1985;34:3011–3017. doi: 10.1016/0006-2952(85)90142-X.
    1. Richardson D.R., Hefter G.T., May P.M., Webb J., Baker E. Iron chelators of the pyridoxal isonicotinoyl hydrazone class. III. Formation constants with calcium(II), magnesium(II) and zinc(II) Biol. Metals. 1989;2:161–167.
    1. Kim B.K., Huebers H.A., Finch C.A. Effectiveness of oral iron chelators assayed in the rat. Am. J. Hematol. 1987;24:277–284. doi: 10.1002/ajh.2830240307.
    1. Hershko C., Avramovici-Grisaru S., Link G., Gelfand L., Sarel S. Mechanism of in vivo iron chelation by pyridoxal isonicotinoyl hydrazone and other imino derivatives of pyridoxal. J. Lab. Clin. Med. 1981;98:99–108.
    1. Becker E.M., Lovejoy D.B., Greer J.M., Watts R., Richardson D.R. Identification of the di-pyridyl ketone isonicotinoyl hydrazone (pkih) analogues as potent iron chelators and anti-tumour agents. Br. J. Pharmacol. 2003;138:819–830. doi: 10.1038/sj.bjp.0705089.
    1. Thelander L., Graslund A. Mechanism of inhibition of mammalian ribonucleotide reductase by the iron chelate of 1-formylisoquinoline thiosemicarbazone. Destruction of the tyrosine free radical of the enzyme in an oxygen-requiring reaction. J. Biol. Chem. 1983;258:4063–4066.
    1. DeConti R.C., Toftness B.R., Agrawal K.C., Tomchick R., Mead J.A., Bertino J.R., Sartorelli A.C., Creasey W.A. Clinical and pharmacological studies with 5-hydroxy-2-formylpyridine thiosemicarbazone. Cancer Res. 1972;32:1455–1462.
    1. Gojo I., Tidwell M.L., Greer J., Takebe N., Seiter K., Pochron M.F., Johnson B., Sznol M., Karp J.E. Phase I and pharmacokinetic study of triapine, a potent ribonucleotide reductase inhibitor, in adults with advanced hematologic malignancies. Leuk. Res. 2007;31:1165–1173. doi: 10.1016/j.leukres.2007.01.004.
    1. Yuan J., Lovejoy D.B., Richardson D.R. Novel di-2-pyridyl-derived iron chelators with marked and selective antitumor activity: In vitro and in vivo assessment. Blood. 2004;104:1450–1458. doi: 10.1182/blood-2004-03-0868.
    1. Whitnall M., Howard J., Ponka P., Richardson D.R. A class of iron chelators with a wide spectrum of potent antitumor activity that overcomes resistance to chemotherapeutics. Proc. Natl. Acad. Sci. USA. 2006;103:14901–14906. doi: 10.1073/pnas.0604979103.
    1. Eberhard Y., McDermott S.P., Wang X., Gronda M., Venugopal A., Wood T.E., Hurren R., Datti A., Batey R.A., Wrana J., et al. Chelation of intracellular iron with the antifungal agent ciclopirox olamine induces cell death in leukemia and myeloma cells. Blood. 2009;114:3064–3073. doi: 10.1182/blood-2009-03-209965.
    1. Roth M., Will B., Simkin G., Narayanagari S., Barreyro L., Bartholdy B., Tamari R., Mitsiades C.S., Verma A., Steidl U. Eltrombopag inhibits the proliferation of leukemia cells via reduction of intracellular iron and induction of differentiation. Blood. 2012;120:386–394. doi: 10.1182/blood-2011-12-399667.
    1. Anastas J.N., Moon R.T. Wnt signalling pathways as therapeutic targets in cancer. Nat. Rev. Cancer. 2013;13:11–26. doi: 10.1038/nrc3419.
    1. Nelson R.L. Iron and colorectal cancer risk: Human studies. Nutr. Rev. 2001;59:140–148. doi: 10.1111/j.1753-4887.2001.tb07002.x.
    1. Chua A.C., Klopcic B., Lawrance I.C., Olynyk J.K., Trinder D. Iron: An emerging factor in colorectal carcinogenesis. World J. Gastroenterol. 2010;16:663–672. doi: 10.3748/wjg.v16.i6.663.
    1. Brooks B.E., Buchanan S.K. Signaling mechanisms for activation of extracytoplasmic function (ecf) sigma factors. Biochim. Biophys. Acta. 2008;1778:1930–1945.
    1. Song S., Christova T., Perusini S., Alizadeh S., Bao R.Y., Miller B.W., Hurren R., Jitkova Y., Gronda M., Isaac M., et al. Wnt inhibitor screen reveals iron dependence of β-catenin signaling in cancers. Cancer Res. 2011;71:7628–7639. doi: 10.1158/0008-5472.CAN-11-2745.
    1. Coombs G.S., Schmitt A.A., Canning C.A., Alok A., Low I.C., Banerjee N., Kaur S., Utomo V., Jones C.M., Pervaiz S., et al. Modulation of wnt/beta-catenin signaling and proliferation by a ferrous iron chelator with therapeutic efficacy in genetically engineered mouse models of cancer. Oncogene. 2012;31:213–225. doi: 10.1038/onc.2011.228.
    1. Daniels T.R., Bernabeu E., Rodriguez J.A., Patel S., Kozman M., Chiappetta D.A., Holler E., Ljubimova J.Y., Helguera G., Penichet M.L. The transferrin receptor and the targeted delivery of therapeutic agents against cancer. Biochim. Biophys. Acta. 2012;1820:291–317.
    1. Ohgami R.S., Campagna D.R., McDonald A., Fleming M.D. The steap proteins are metalloreductases. Blood. 2006;108:1388–1394. doi: 10.1182/blood-2006-02-003681.
    1. Gomes I.M., Maia C.J., Santos C.R. Steap proteins: From structure to applications in cancer therapy. Mol. Cancer Res. 2012;10:573–587. doi: 10.1158/1541-7786.MCR-11-0281.
    1. Grunewald T.G., Bach H., Cossarizza A., Matsumoto I. The steap protein family: Versatile oxidoreductases and targets for cancer immunotherapy with overlapping and distinct cellular functions. Biol. Cell. 2012;104:641–657. doi: 10.1111/boc.201200027.
    1. Becton D.L., Roberts B. Antileukemic effects of deferoxamine on human myeloid leukemia cell lines. Cancer Res. 1989;49:4809–4812.
    1. Noulsri E., Richardson D.R., Lerdwana S., Fucharoen S., Yamagishi T., Kalinowski D.S., Pattanapanyasat K. Antitumor activity and mechanism of action of the iron chelator, dp44mt, against leukemic cells. Am. J. Hematol. 2009;84:170–176. doi: 10.1002/ajh.21350.
    1. Callens C., Coulon S., Naudin J., Radford-Weiss I., Boissel N., Raffoux E., Wang P.H., Agarwal S., Tamouza H., Paubelle E., et al. Targeting iron homeostasis induces cellular differentiation and synergizes with differentiating agents in acute myeloid leukemia. J. Exp. Med. 2010;207:731–750. doi: 10.1084/jem.20091488.
    1. Rodriguez J.A., Luria-Perez R., Lopez-Valdes H.E., Casero D., Daniels T.R., Patel S., Avila D., Leuchter R., So S., Ortiz-Sanchez E., et al. Lethal iron deprivation induced by non-neutralizing antibodies targeting transferrin receptor 1 in malignant b cells. Leuk. Lymphoma. 2011;52:2169–2178. doi: 10.3109/10428194.2011.596964.
    1. Fukushima T., Kawabata H., Nakamura T., Iwao H., Nakajima A., Miki M., Sakai T., Sawaki T., Fujita Y., Tanaka M., et al. Iron chelation therapy with deferasirox induced complete remission in a patient with chemotherapy-resistant acute monocytic leukemia. Anticancer Res. 2011;31:1741–1744.
    1. Estrov Z., Tawa A., Wang X.H., Dube I.D., Sulh H., Cohen A., Gelfand E.W., Freedman M.H. In vitro and in vivo effects of deferoxamine in neonatal acute leukemia. Blood. 1987;69:757–761.
    1. Yee K.W., Cortes J., Ferrajoli A., Garcia-Manero G., Verstovsek S., Wierda W., Thomas D., Faderl S., King I., O’Brien S.M., et al. Triapine and cytarabine is an active combination in patients with acute leukemia or myelodysplastic syndrome. Leuk. Res. 2006;30:813–822. doi: 10.1016/j.leukres.2005.12.013.
    1. Choi J.G., Kim J.L., Park J., Lee S., Park S.J., Kim J.S., Choi C.W. Effects of oral iron chelator deferasirox on human malignant lymphoma cells. Korean J. Hematol. 2012;47:194–201. doi: 10.5045/kjh.2012.47.3.194.
    1. Vazana-Barad L., Granot G., Mor-Tzuntz R., Levi I., Dreyling M., Nathan I., Shpilberg O. Mechanism of the antitumoral activity of deferasirox, an iron chelation agent, on mantle cell lymphoma. Leuk. Lymphoma. 2013;54:851–859. doi: 10.3109/10428194.2012.734614.
    1. Maris J.M. Recent advances in neuroblastoma. N. Engl. J. Med. 2010;362:2202–2211. doi: 10.1056/NEJMra0804577.
    1. Blatt J. Deferoxamine in children with recurrent neuroblastoma. Anticancer Res. 1994;14:2109–2112.
    1. Chaston T.B., Lovejoy D.B., Watts R.N., Richardson D.R. Examination of the antiproliferative activity of iron chelators: Multiple cellular targets and the different mechanism of action of triapine compared with desferrioxamine and the potent pyridoxal isonicotinoyl hydrazone analogue 311. Clin. Cancer Res. 2003;9:402–414.
    1. Donfrancesco A., Deb G., Dominici C., Pileggi D., Castello M.A., Helson L. Effects of a single course of deferoxamine in neuroblastoma patients. Cancer Res. 1990;50:4929–4930.
    1. Donfrancesco A., de Bernardi B., Carli M., Mancini A., Nigro M., de Sio L., Casale F., Bagnulo S., Helson L., Deb G. Deferoxamine followed by cyclophosphamide, etoposide, carboplatin, thiotepa, induction regimen in advanced neuroblastoma: Preliminary results. Italian neuroblastoma cooperative group. Eur. J. Cancer. 1995;31A:612–615.
    1. Lui G.Y., Obeidy P., Ford S.J., Tselepis C., Sharp D.M., Jansson P.J., Kalinowski D.S., Kovacevic Z., Lovejoy D.B., Richardson D.R. The iron chelator, deferasirox, as a novel strategy for cancer treatment: Oral activity against human lung tumor xenografts and molecular mechanism of action. Mol. Pharmacol. 2013;83:179–190. doi: 10.1124/mol.112.081893.
    1. Ford S., Obeidy P., Lovejoy D., Bedford M., Nichols L., Chadwick C., Tucker O., Lui G., Kalinowski D., Jansson P., et al. Deferasirox (icl670a) effectively inhibits oesophageal cancer growth in vitro and in vivo. Br. J. Pharmacol. 2013;168:1316–1328. doi: 10.1111/bph.12045.
    1. Dixon K.M., Lui G.Y., Kovacevic Z., Zhang D., Yao M., Chen Z., Dong Q., Assinder S.J., Richardson D.R. Dp44mt targets the akt, tgf-beta and erk pathways via the metastasis suppressor ndrg1 in normal prostate epithelial cells and prostate cancer cells. Br. J. Cancer. 2013;108:409–419. doi: 10.1038/bjc.2012.582.
    1. Rao V.A., Klein S.R., Agama K.K., Toyoda E., Adachi N., Pommier Y., Shacter E.B. The iron chelator dp44mt causes DNA damage and selective inhibition of topoisomerase iiα in breast cancer cells. Cancer Res. 2009;69:948–957. doi: 10.1158/0008-5472.CAN-08-1437.

Source: PubMed

3
Abonnere