Oxidative stress in β-thalassaemia and sickle cell disease

S Voskou, M Aslan, P Fanis, M Phylactides, M Kleanthous, S Voskou, M Aslan, P Fanis, M Phylactides, M Kleanthous

Abstract

Sickle cell disease and β-thalassaemia are inherited haemoglobinopathies resulting in structural and quantitative changes in the β-globin chain. These changes lead to instability of the generated haemoglobin or to globin chain imbalance, which in turn impact the oxidative environment both intracellularly and extracellularly. The ensuing oxidative stress and the inability of the body to adequately overcome it are, to a large extent, responsible for the pathophysiology of these diseases. This article provides an overview of the main players and control mechanisms involved in the establishment of oxidative stress in these haemoglobinopathies.

Copyright © 2015 The Authors. Published by Elsevier B.V. All rights reserved.

Figures

Graphical abstract
Graphical abstract
Fig. 2
Fig. 2
Oxidative events in the bone marrow and circulation. (1) Erythroid expansion. (2) Ineffective erythropoiesis. (3) Endocytosis of RBCs by macrophages through two different mechanisms: eryptosis and senescence. (4) Haemolysis. It leads to Hb release in the plasma, which subsequently autoxidises producing ROS, free haeme and iron. (5) ROS, free haeme and iron intercalate into plasma membranes producing oxidative damage to the endothelium and haematopoietic cells. ROS and haeme activate NF-κB and AP-1, which increase the production of pro-inflammatory cytokines (IL-1, IL-6, TNFα) and adhesion molecules on the endothelium. This increases the adhesion of RBCs, leucocytes and platelets to the endothelium. Activated leucocytes generate more ROS by their NADPH oxidase, creating a vicious cycle of oxidative stress and inflammation. (6) Reduction of NO bioavailability due to free Hb, ROS, and neutrophil activation, (7) Increased expression of plasma and endothelial enzymes (e.g. xanthine oxidase, NADPH oxidase and iNOS) following I/R gives rise to more ROS.
Fig. 1
Fig. 1
(A) Intracellular oxidative events. (1) Oxidative denaturation of Hb. It results in the production of ROS, free haeme and iron. Iron acts as a Fenton reagent in the Haber–Weiss reaction for the generation of hydroxyl radical. Haeme promotes oxidation reactions and a proinflammatory effect by activating NF-κB. (2) Enzymatic generation of superoxide by NADPH oxidase. NADPH oxidase is regulated by intra- (e.g. increased Ca2+) and extra-cellular (pro-inflammatory cytokines) signals that are activated by ROS. (3) ROS and ROS-induced increase of intra-cellular Ca2+ activate caspase-3, which partially degrades band-3, affecting its interaction with the cytoskeleton. ROS and increased Ca2+ also activate the KCC and Gardos channel respectively, resulting in increased exit of K+ from the cell. (4) Haemichromes mediate the oxidative crosslinking and phosphorylation of band-3 leading to band-3 clusterisation and dissociation from cytoskeletal proteins. This results in membrane blebbing and microparticle generation. Band-3/haemachrome clusters are recognised by anti-band-3-NAbs. (5) ROS promote oxidation of protein 4.1, actin and spectrin resulting in impaired interaction. (6) PS exposure results from ROS-induced disruption of normal membrane organisation. (B) Protective mechanisms in RBCs and the circulation. (1) Haptoglobin/Hb and haemopexin/haeme complexes are endocytosed by macrophages. Haeme is then degraded by HO-1 releasing biliverdin, CO and iron. Iron is then taken up by ferritin. (2) Antioxidant enzymes and molecules in RBCs. (3) Stress-response mechanisms in RBCs. FOXO3, HRI/eIF2α/ATF4 and NRF2 are oxidant-response pathways that regulate the expression of antioxidant genes. FOXO3 and HRI/eIF2α/ATF4 are important for terminal differentiation. PRDX2 is a chaperone of Hb and binds free haeme to prevent its oxidative actions. AHSP binds α-globin in the absence of β-globin and in the presence of oxidant insults. (4) Protein quality control pathways. Unpaired α-globins are selectively degraded via ubiquidin-mediated proteasomal degradation (UPS). When UPS becomes overwhelmed, α-globin aggregates are degraded via aggresome-mediated macroautophagy.
Fig. 3
Fig. 3
Scheme of intravascular nitric oxide consumption in sickle cell disease. Superoxide (O2•-) generated by uncoupled eNOS, xanthine oxidase, and NADPH oxidase reacts with NO to form peroxynitrite (ONOO−). Nitric oxide is also consumed by plasma free haemoglobin, released by intravascular haemolysis and myeloperoxidase (MPO). The reaction between MPO and hydrogen peroxide (H2O2) results in the formation of compound I (I), the two electron oxidised form of the enzyme. Compound I can either oxidise halides in a single two-electron step or can oxidise multiple substrates through sequential one electron steps, forming compound II. Tyrosine (Tyr) and ascorbate (Asc) are abundant substrates that can undergo one-electron oxidation by MPO-compound I, forming tyrosyl (•Tyr) and ascorbyl (•Asc) radicals. The reaction of nitric (•NO) with both tyrosyl and ascorbyl radicals is diffusion limited and leads to the catalytic consumption of •NO.

References

    1. Abraham E. NF-kappaB activation. Crit. Care Med. 2000;28:N100–104.
    1. Adamsky K., Weizer O., Amariglio N., Breda L., Harmelin A., Rivella S., Rachmilewitz E., Rechavi G. Decreased hepcidin mRNA expression in thalassemic mice. Br. J. Haematol. 2004;124:123–124.
    1. Aidoo M., Terlouw D.J., Kolczak M.S., McElroy P.D., ter Kuile F.O., Kariuki S., Nahlen B.L., Lal A.A., Udhayakumar V. Protective effects of the sickle cell gene against malaria morbidity and mortality. Lancet. 2002;359:1311–1312.
    1. Alam J., Killeen E., Gong P., Naquin R., Hu B., Stewart D., Ingelfinger J.R., Nath K.A. Heme activates the heme oxygenase-1 gene in renal epithelial cells by stabilizing Nrf2. Am. J. Physiol. Ren. Physiol. 2003;284:F743–752.
    1. Alderton W.K., Cooper C.E., Knowles R.G. Nitric oxide synthases: structure, function and inhibition. Biochem. J. 2001;357:593–615.
    1. Aljurf M., Ma L., Angelucci E., Lucarelli G., Snyder L.M., Kiefer C.R., Yuan J., Schrier S.L. Abnormal assembly of membrane proteins in erythroid progenitors of patients with beta-thalassemia major. Blood. 1996;87:2049–2056.
    1. Allan D., Limbrick A.R., Thomas P., Westerman M.P. Release of spectrin-free spicules on reoxygenation of sickled erythrocytes. Nature. 1982;295:612–613.
    1. Amer J., Fibach E. Chronic oxidative stress reduces the respiratory burst response of neutrophils from beta-thalassaemia patients. Br. J. Haematol. 2005;129:435–441.
    1. Amer J., Ghoti H., Rachmilewitz E., Koren A., Levin C., Fibach E. Red blood cells, platelets and polymorphonuclear neutrophils of patients with sickle cell disease exhibit oxidative stress that can be ameliorated by antioxidants. Br. J. Haematol. 2006;132:108–113.
    1. Angelucci E., Bai H., Centis F., Bafti M.S., Lucarelli G., Ma L., Schrier S. Enhanced macrophagic attack on beta-thalassemia major erythroid precursors. Haematologica. 2002;87:578–583.
    1. Arner E.S., Holmgren A. Physiological functions of thioredoxin and thioredoxin reductase. Eur. J. Biochem./FEBS. 2000;267:6102–6109.
    1. Aslan M., Canatan D. Modulation of redox pathways in neutrophils from sickle cell disease patients. Exp. Hematol. 2008;36:1535–1544.
    1. Aslan M., Freeman B.A. Oxidant-mediated impairment of nitric oxide signaling in sickle cell disease-mechanisms and consequences. Cell. Mol. Biol. 2004;50:95–105.
    1. Aslan M., Freeman B.A. Redox-dependent impairment of vascular function in sickle cell disease. Free Rad. Biol. Med. 2007;43:1469–1483.
    1. Aslan M., Ryan T.M., Adler B., Townes T.M., Parks D.A., Thompson J.A., Tousson A., Gladwin M.T., Patel R.P., Tarpey M.M. Oxygen radical inhibition of nitric oxide-dependent vascular function in sickle cell disease. Proc. Natl. Acad. Sci. USA. 2001;98:15215–15220.
    1. Aslan M., Ryan T.M., Townes T.M., Coward L., Kirk M.C., Barnes S., Alexander C.B., Rosenfeld S.S., Freeman B.A. Nitric oxide-dependent generation of reactive species in sickle cell disease. Actin tyrosine induces defective cytoskeletal polymerization. J. Biol. Chem. 2003;278:4194–4204.
    1. Aslan M., Thornley-Brown D., Freeman B.A. Reactive species in sickle cell disease. Ann. N.Y. Acad. Sci. 2000;899:375–391.
    1. Balagopalakrishna C., Manoharan P.T., Abugo O.O., Rifkind J.M. Production of superoxide from hemoglobin-bound oxygen under hypoxic conditions. Biochemistry. 1996;35:6393–6398.
    1. Baldus S., Eiserich J.P., Brennan M.L., Jackson R.M., Alexander C.B., Freeman B.A. Spatial mapping of pulmonary and vascular nitrotyrosine reveals the pivotal role of myeloperoxidase as a catalyst for tyrosine nitration in inflammatory diseases. Free Rad. Biol. Med. 2002;33:1010.
    1. Baldus S., Eiserich J.P., Mani A., Castro L., Figueroa M., Chumley P., Ma W., Tousson A., White C.R., Bullard D.C. Endothelial transcytosis of myeloperoxidase confers specificity to vascular ECM proteins as targets of tyrosine nitration. J. Clin. Investig. 2001;108:1759–1770.
    1. Balla J., Jacob H.S., Balla G., Nath K., Eaton J.W., Vercellotti G.M. Endothelial-cell heme uptake from heme proteins: induction of sensitization and desensitization to oxidant damage. Proc. Natl. Acad. Sci. USA. 1993;90:9285–9289.
    1. Barodka V., Mohanty J.G., Mustafa A.K., Santhanam L., Nyhan A., Bhunia A.K., Sikka G., Nyhan D., Berkowitz D.E., Rifkind J.M. Nitroprusside inhibits calcium-induced impairment of red blood cell deformability. Transfusion. 2014;54:434–444.
    1. Barodka V.M., Nagababu E., Mohanty J.G., Nyhan D., Berkowitz D.E., Rifkind J.M., Strouse J.J. New insights provided by a comparison of impaired deformability with erythrocyte oxidative stress for sickle cell disease. Blood Cells, Mol. Dis. 2014;52:230–235.
    1. Basu A., Saha S., Karmakar S., Chakravarty S., Banerjee D., Dash B.P., Chakrabarti A. 2D DIGE based proteomics study of erythrocyte cytosol in sickle cell disease: altered proteostasis and oxidative stress. Proteomics. 2013;13:3233–3242.
    1. Bayraktutan U., Draper N., Lang D., Shah A.M. Expression of functional neutrophil-type NADPH oxidase in cultured rat coronary microvascular endothelial cells. Cardiovas. Res. 1998;38:256–262.
    1. Bec N., Gorren A.C., Voelker C., Mayer B., Lange R. Reaction of neuronal nitric-oxide synthase with oxygen at low temperature. Evidence for reductive activation of the oxy-ferrous complex by tetrahydrobiopterin. J. Biol. Chem. 1998;273:13502–13508.
    1. Beckman J.S., Koppenol W.H. Nitric oxide, superoxide, and peroxynitrite: the good, the bad, and ugly. Am. J. Physiol. 1996;271:C1424–1437.
    1. Beetsch J.W., Park T.S., Dugan L.L., Shah A.R., Gidday J.M. Xanthine oxidase-derived superoxide causes reoxygenation injury of ischemic cerebral endothelial cells. Brain Res. 1998;786:89–95.
    1. Beutler E. Red cell metabolism. A. Defects not causing hemolytic disease. B. Environmental modification. Biochimie. 1972;54:759–764.
    1. Blouin M.J., De Paepe M.E., Trudel M. Altered hematopoiesis in murine sickle cell disease. Blood. 1999;94:1451–1459.
    1. Bordin L., Brunati A.M., Donella-Deana A., Baggio B., Toninello A., Clari G. Band 3 is an anchor protein and a target for SHP-2 tyrosine phosphatase in human erythrocytes. Blood. 2002;100:276–282.
    1. Braverman A.S., Lester D. Evidence for increased proteolysis in intact beta thalassemia erythroid cells. Hemoglobin. 1981;5:549–564.
    1. Buehler P.W., Abraham B., Vallelian F., Linnemayr C., Pereira C.P., Cipollo J.F., Jia Y., Mikolajczyk M., Boretti F.S., Schoedon G. Haptoglobin preserves the CD163 hemoglobin scavenger pathway by shielding hemoglobin from peroxidative modification. Blood. 2009;113:2578–2586.
    1. Bunn H.F., Jandl J.H. Exchange of heme among hemoglobins and between hemoglobin and albumin. J. Biol. Chem. 1968;243:465–475.
    1. Burger P., Kostova E., Bloem E., Hilarius-Stokman P., Meijer A.B., van den Berg T.K., Verhoeven A.J., de Korte D., van Bruggen R. Potassium leakage primes stored erythrocytes for phosphatidylserine exposure and shedding of pro-coagulant vesicles. Br. J. Haematol. 2013;160:377–386.
    1. Cao Z., Bell J.B., Mohanty J.G., Nagababu E., Rifkind J.M. Nitrite enhances RBC hypoxic ATP synthesis and the release of ATP into the vasculature: a new mechanism for nitrite-induced vasodilation. Am. J. Physiol. Heart Circ. Physiol. 2009;297:H1494–1503.
    1. Centis F., Tabellini L., Lucarelli G., Buffi O., Tonucci P., Persini B., Annibali M., Emiliani R., Iliescu A., Rapa S. The importance of erythroid expansion in determining the extent of apoptosis in erythroid precursors in patients with beta-thalassemia major. Blood. 2000;96:3624–3629.
    1. Chaves M.A., Leonart M.S., do Nascimento A.J. Oxidative process in erythrocytes of individuals with hemoglobin S. Hematology. 2008;13:187–192.
    1. Chen J.J. Regulation of protein synthesis by the heme-regulated eIF2alpha kinase: relevance to anemias. Blood. 2007;109:2693–2699.
    1. Choi A.M., Alam J. Heme oxygenase-1: function, regulation, and implication of a novel stress-inducible protein in oxidant-induced lung injury. Am. J. Respir. Cell Mol. Biol. 1996;15:9–19.
    1. Clemens M.R., Waller H.D. Lipid peroxidation in erythrocytes. Chem. Phys. Lipids. 1987;45:251–268.
    1. Clementi M.E., Giardina B., Colucci D., Galtieri A., Misiti F. Amyloid-beta peptide affects the oxygen dependence of erythrocyte metabolism: a role for caspase 3. Int. J. Biochem. Cell Biol. 2007;39:727–735.
    1. Comporti M., Signorini C., Buonocore G., Ciccoli L. Iron release, oxidative stress and erythrocyte ageing. Free Rad. Biol. Med. 2002;32:568–576.
    1. Conran N., Costa F.F. Hemoglobin disorders and endothelial cell interactions. Clin. Biochem. 2009;42:1824–1838.
    1. Cytlak U.M., Hannemann A., Rees D.C., Gibson J.S. Identification of the Ca(2)(+) entry pathway involved in deoxygenation-induced phosphatidylserine exposure in red blood cells from patients with sickle cell disease. Pflug. Arch.: Eur. J. Physiol. 2013;465:1651–1660.
    1. De Caterina R., Libby P., Peng H.B., Thannickal V.J., Rajavashisth T.B., Gimbrone M.A., Jr., Shin W.S., Liao J.K. Nitric oxide decreases cytokine-induced endothelial activation. Nitric oxide selectively reduces endothelial expression of adhesion molecules and proinflammatory cytokines. J. Clin. Investig. 1995;96:60–68.
    1. De Franceschi L., Bertoldi M., De Falco L., Santos Franco S., Ronzoni L., Turrini F., Colancecco A., Camaschella C., Cappellini M.D., Iolascon A. Oxidative stress modulates heme synthesis and induces peroxiredoxin-2 as a novel cytoprotective response in beta-thalassemic erythropoiesis. Haematologica. 2011;96:1595–1604.
    1. De Franceschi L., Bertoldi M., Matte A., Santos Franco S., Pantaleo A., Ferru E., Turrini F. Oxidative stress and beta-thalassemic erythroid cells behind the molecular defect. Oxid. Med. Cell. Longev. 2013;2013:985210.
    1. De Franceschi L., Ronzoni L., Cappellini M.D., Cimmino F., Siciliano A., Alper S.L., Servedio V., Pozzobon C., Iolascon A. K-CL co-transport plays an important role in normal and beta thalassemic erythropoiesis. Haematologica. 2007;92:1319–1326.
    1. De Maria R., Zeuner A., Eramo A., Domenichelli C., Bonci D., Grignani F., Srinivasula S.M., Alnemri E.S., Testa U., Peschle C. Negative regulation of erythropoiesis by caspase-mediated cleavage of GATA-1. Nature. 1999;401:489–493.
    1. Dijkers P.F., Medema R.H., Pals C., Banerji L., Thomas N.S., Lam E.W., Burgering B.M., Raaijmakers J.A., Lammers J.W., Koenderman L. Forkhead transcription factor FKHR-L1 modulates cytokine-dependent transcriptional regulation of p27(KIP1) Mol. Cell. Biol. 2000;20:9138–9148.
    1. Dussiot M., Maciel T.T., Fricot A., Chartier C., Negre O., Veiga J., Grapton D., Paubelle E., Payen E., Beuzard Y. An activin receptor IIA ligand trap corrects ineffective erythropoiesis in beta-thalassemia. Nat. Med. 2014;20:398–407.
    1. Eiserich J.P., Baldus S., Brennan M.L., Ma W., Zhang C., Tousson A., Castro L., Lusis A.J., Nauseef W.M., White C.R. Myeloperoxidase, a leukocyte-derived vascular NO oxidase. Science. 2002;296:2391–2394.
    1. Etlinger J.D., Goldberg A.L. A soluble ATP-dependent proteolytic system responsible for the degradation of abnormal proteins in reticulocytes. Proc. Natl Acad. Sci. USA. 1977;74:54–58.
    1. Feng L., Zhou S., Gu L., Gell D.A., Mackay J.P., Weiss M.J., Gow A.J., Shi Y. Structure of oxidized alpha-haemoglobin bound to AHSP reveals a protective mechanism for haem. Nature. 2005;435:697–701.
    1. Ferrali M., Signorini C., Ciccoli L., Comporti M. Iron released from an erythrocyte lysate by oxidative stress is diffusible and in redox active form. FEBS Lett. 1993;319:40–44.
    1. Ferroni P., Vazzana N., Riondino S., Cuccurullo C., Guadagni F., Davi G. Platelet function in health and disease: from molecular mechanisms, redox considerations to novel therapeutic opportunities. Antioxid. Redox Signal. 2012;17:1447–1485.
    1. Ferru E., Giger K., Pantaleo A., Campanella E., Grey J., Ritchie K., Vono R., Turrini F., Low P.S. Regulation of membrane-cytoskeletal interactions by tyrosine phosphorylation of erythrocyte band 3. Blood. 2011;117:5998–6006.
    1. Foller M., Harris I.S., Elia A., John R., Lang F., Kavanagh T.J., Mak T.W. Functional significance of glutamate-cysteine ligase modifier for erythrocyte survival in vitro and in vivo. Cell Death Differ. 2013;20:1350–1358.
    1. Forstermann U., Munzel T. Endothelial nitric oxide synthase in vascular disease: from marvel to menace. Circulation. 2006;113:1708–1714.
    1. Frenette P.S. Sickle cell vaso-occlusion: multistep and multicellular paradigm. Curr. Opin. Hematol. 2002;9:101–106.
    1. George A., Pushkaran S., Konstantinidis D.G., Koochaki S., Malik P., Mohandas N., Zheng Y., Joiner C.H., Kalfa T.A. Erythrocyte NADPH oxidase activity modulated by Rac GTPases, PKC, and plasma cytokines contributes to oxidative stress in sickle cell disease. Blood. 2013;121:2099–2107.
    1. George A., Pushkaran S., Li L., An X., Zheng Y., Mohandas N., Joiner C.H., Kalfa T.A. Altered phosphorylation of cytoskeleton proteins in sickle red blood cells: the role of protein kinase C, Rac GTPases, and reactive oxygen species. Blood Cells Mol. Dis. 2010;45:41–45.
    1. Ghaffari S. Oxidative stress in the regulation of normal and neoplastic hematopoiesis. Antioxid. Redox Signal. 2008;10:1923–1940.
    1. Ghaffari S., Jagani Z., Kitidis C., Lodish H.F., Khosravi-Far R. Cytokines and BCR-ABL mediate suppression of TRAIL-induced apoptosis through inhibition of forkhead FOXO3a transcription factor. Proc. Natl. Acad. Sci. USA. 2003;100:6523–6528.
    1. Goodman S.R. The irreversibly sickled cell: a perspective. Cell. Mol. Biol. (Noisy-le-grand) 2004;50:53–58.
    1. Hamer I., Wattiaux R., Wattiaux-De Coninck S. Deleterious effects of xanthine oxidase on rat liver endothelial cells after ischemia/reperfusion. Biochim. Biophys. Acta. 1995;1269:145–152.
    1. Han A.P., Fleming M.D., Chen J.J. Heme-regulated eIF2alpha kinase modifies the phenotypic severity of murine models of erythropoietic protoporphyria and beta-thalassemia. J. Clin. Investig. 2005;115:1562–1570.
    1. Harding H.P., Zhang Y., Zeng H., Novoa I., Lu P.D., Calfon M., Sadri N., Yun C., Popko B., Paules R. An integrated stress response regulates amino acid metabolism and resistance to oxidative stress. Molecular Cell. 2003;11:619–633.
    1. Harrison M.L., Rathinavelu P., Arese P., Geahlen R.L., Low P.S. Role of band 3 tyrosine phosphorylation in the regulation of erythrocyte glycolysis. J. Biol. Chem. 1991;266:4106–4111.
    1. He C.H., Gong P., Hu B., Stewart D., Choi M.E., Choi A.M., Alam J. Identification of activating transcription factor 4 (ATF4) as an Nrf2-interacting protein. Implication for heme oxygenase-1 gene regulation. J. Biol. Chem. 2001;276:20858–20865.
    1. Hebbel R.P. The systems biology-based argument for taking a bold step in chemoprophylaxis of sickle vasculopathy. Am. J. Hematol. 2009;84:543–545.
    1. Herold S., Exner M., Nauser T. Kinetic and mechanistic studies of the NO*-mediated oxidation of oxymyoglobin and oxyhemoglob. Biochemistry. 2001;40:3385–3395.
    1. Hershko C. Pathogenesis and management of iron toxicity in thalassemia. Ann. N.Y. Acad. Sci. 2010;1202:1–9.
    1. Ibrahim H.A., Fouda M.I., Yahya R.S., Abousamra N.K., Abd Elazim R.A. Erythrocyte phosphatidylserine exposure in beta-thalassemia. Lab. Hematol. 2014;20:9–14.
    1. Iuliano L., Colavita A.R., Leo R., Pratico D., Violi F. Oxygen free radicals and platelet activation. Free Rad. Biol. Med. 1997;22:999–1006.
    1. Jison M.L., Munson P.J., Barb J.J., Suffredini A.F., Talwar S., Logun C., Raghavachari N., Beigel J.H., Shelhamer J.H., Danner R.L. Blood. 2004;104:270–280.
    1. Kanczler J.M., Millar T.M., Bodamyali T., Blake D.R., Stevens C.R. Xanthine oxidase mediates cytokine-induced, but not hormone-induced bone resorption. Free Rad. Res. 2003;37:179–187.
    1. Kang Y.A., Sanalkumar R., O’Geen H., Linnemann A.K., Chang C.J., Bouhassira E.E., Farnham P.J., Keles S., Bresnick E.H. Autophagy driven by a master regulator of hematopoiesis. Mol. Cell. Biol. 2012;32:226–239.
    1. Kanias T., Acker J.P. Biopreservation of red blood cells – the struggle with hemoglobin oxidation. FEBS J. 2010;277:343–356.
    1. Kannan R., Labotka R., Low P.S. Isolation and characterization of the hemichrome-stabilized membrane protein aggregates from sickle erythrocytes. Major site of autologous antibody binding. J. Biolo. Chem. 1988;263:13766–13773.
    1. Kaul D.K., Liu X.D., Choong S., Belcher J.D., Vercellotti G.M., Hebbel R.P. Anti-inflammatory therapy ameliorates leukocyte adhesion and microvascular flow abnormalities in transgenic sickle mice. Am. J. Physiol. Heart Circ. Physiol. 2004;287:H293–301.
    1. Kawatani Y., Suzuki T., Shimizu R., Kelly V.P., Yamamoto M. Nrf2 and selenoproteins are essential for maintaining oxidative homeostasis in erythrocytes and protecting against hemolytic anemia. Blood. 2011;117:986–996.
    1. Kean L.S., Brown L.E., Nichols J.W., Mohandas N., Archer D.R., Hsu L.L. Comparison of mechanisms of anemia in mice with sickle cell disease and beta-thalassemia: peripheral destruction, ineffective erythropoiesis, and phospholipid scramblase-mediated phosphatidylserine exposure. Exp. Hematol. 2002;30:394–402.
    1. Khandros E., Thom C.S., D’Souza J., Weiss M.J. Integrated protein quality-control pathways regulate free alpha-globin in murine beta-thalassemia. Blood. 2012;119:5265–5275.
    1. Khandros E., Weiss M.J. Protein quality control during erythropoiesis and hemoglobin synthesis. Hematol./Oncol Clin. N. Am. 2010;24:1071–1088.
    1. Kiefer C.R., Snyder L.M. Oxidation and erythrocyte senescence. Curr. Opin. Hematol. 2000;7:113–116.
    1. Kiefmann R., Rifkind J.M., Nagababu E., Bhattacharya J. Red blood cells induce hypoxic lung inflammation. Blood. 2008;111:5205–5214.
    1. Kitamuro T., Takahashi K., Ogawa K., Udono-Fujimori R., Takeda K., Furuyama K., Nakayama M., Sun J., Fujita H., Hida W. Bach1 functions as a hypoxia-inducible repressor for the heme oxygenase-1 gene in human cells. J. Biol. Chem. 2003;278:9125–9133.
    1. Kong Y., Zhou S., Kihm A.J., Katein A.M., Yu X., Gell D.A., Mackay J.P., Adachi K., Foster-Brown L., Louden C.S. Loss of alpha-hemoglobin-stabilizing protein impairs erythropoiesis and exacerbates beta-thalassemia. J. Clin. Investig. 2004;114:1457–1466.
    1. Kopito R.R. Aggresomes, inclusion bodies and protein aggregation. Trends Cell Biol. 2000;10:524–530.
    1. Kops G.J., Dansen T.B., Polderman P.E., Saarloos I., Wirtz K.W., Coffer P.J., Huang T.T., Bos J.L., Medema R.H., Burgering B.M. Forkhead transcription factor FOXO3a protects quiescent cells from oxidative stress. Nature. 2002;419:316–321.
    1. Kurantsin-Mills J., Ofosu F.A., Safa T.K., Siegel R.S., Lessin L.S. Plasma factor VII and thrombin-antithrombin III levels indicate increased tissue factor activity in sickle cell patients. Br. J. Haematol. 1992;81:539–544.
    1. Kuypers F.A., de Jong K. The role of phosphatidylserine in recognition and removal of erythrocytes. Cell. Mol. Biol. 2004;50:147–158.
    1. Lang E., Lang F. Mechanisms and pathophysiological significance of eryptosis, the suicidal erythrocyte death. Semin. Cell Dev. Biol. 2015;39:35–42.
    1. Lang F., Busch G.L., Ritter M., Volkl H., Waldegger S., Gulbins E., Haussinger D. Functional significance of cell volume regulatory mechanisms. Physiol. Rev. 1998;78:247–306.
    1. Lappas M., Mitton A., Permezel M. In response to oxidative stress, the expression of inflammatory cytokines and antioxidant enzymes are impaired in placenta, but not adipose tissue, of women with gestational diabetes. J. Endocrinol. 2010;204:75–84.
    1. Lard L.R., Mul F.P., de Haas M., Roos D., Duits A.J. Neutrophil activation in sickle cell disease. J. Leukoc. Biol. 1999;66:411–415.
    1. Lee M.T., Rosenzweig E.B., Cairo M.S. Pulmonary hypertension in sickle cell disease. Clin. Adv. Hematol. Oncol.: H&O. 2007;5:645–653. 585.
    1. Lee S.P., Ataga K.I., Orringer E.P., Phillips D.R., Parise L.V. Biologically active CD40 ligand is elevated in sickle cell anemia: potential role for platelet-mediated inflammation. Arterioscler. Thromb. Vasc. Biol. 2006;26:1626–1631.
    1. Lee V.G., Johnson M.L., Baust J., Laubach V.E., Watkins S.C., Billiar T.R. The roles of iNOS in liver ischemia-reperfusion injury. Shock. 2001;16:355–360.
    1. Leecharoenkiat A., Wannatung T., Lithanatudom P., Svasti S., Fucharoen S., Chokchaichamnankit D., Srisomsap C., Smith D.R. Increased oxidative metabolism is associated with erythroid precursor expansion in beta0-thalassaemia/Hb E disease. Blood Cells, Mol. Dis. 47. 2011:143–157.
    1. Leonard S.S., Harris G.K., Shi X. Metal-induced oxidative stress and signal transduction. Free Rad. Biol. Med. 2004;37:1921–1942.
    1. Li H., Rybicki A.C., Suzuka S.M., von Bonsdorff L., Breuer W., Hall C.B., Cabantchik Z.I., Bouhassira E.E., Fabry M.E., Ginzburg Y.Z. Transferrin therapy ameliorates disease in beta-thalassemic mice. Nat. Med. 2010;16:177–182.
    1. Lincoln T.M., Cornwell T.L. Intracellular cyclic GMP receptor proteins. FASEB J. 1993;7:328–338.
    1. Ling H., Gengaro P.E., Edelstein C.L., Martin P.Y., Wangsiripaisan A., Nemenoff R., Schrier R.W. Effect of hypoxia on proximal tubules isolated from nitric oxide synthase knockout mice. Kidney Int. 1998;53:1642–1646.
    1. Lithanatudom P., Wannatung T., Leecharoenkiat A., Svasti S., Fucharoen S., Smith D.R. Enhanced activation of autophagy in beta-thalassemia/Hb E erythroblasts during erythropoiesis. Ann. Hematol. 2011;90:747–758.
    1. Low F.M., Hampton M.B., Winterbourn C.C. Peroxiredoxin 2 and peroxide metabolism in the erythrocyte. Antioxid. Redox Signal. 2008;10:1621–1630.
    1. Lu L., Han A.P., Chen J.J. Translation initiation control by heme-regulated eukaryotic initiation factor 2alpha kinase in erythroid cells under cytoplasmic stresses. Mol. Cell. Biol. 2001;21:7971–7980.
    1. Lutz H.U., Nater M., Stammler P. Naturally occurring anti-band 3 antibodies have a unique affinity for C3. Immunology. 1993;80:191–196.
    1. Mahaseth H., Vercellotti G.M., Welch T.E., Bowlin P.R., Sonbol K.M., Hsia C.J., Li M., Bischof J.C., Hebbel R.P., Belcher J.D. Polynitroxyl albumin inhibits inflammation and vasoocclusion in transgenic sickle mice. J. Lab. Clin. Med. 2005;145:204–211.
    1. Mandal D., Baudin-Creuza V., Bhattacharyya A., Pathak S., Delaunay J., Kundu M., Basu J. Caspase 3-mediated proteolysis of the N-terminal cytoplasmic domain of the human erythroid anion exchanger 1 (band 3) J. Biol. Chem. 2003;278:52551–52558.
    1. Mandal D., Mazumder A., Das P., Kundu M., Basu J. Fas-, caspase 8-, and caspase 3-dependent signaling regulates the activity of the aminophospholipid translocase and phosphatidylserine externalization in human erythrocytes. J. Biol. Chem. 2005;280:39460–39467.
    1. Mannervik B. The enzymes of glutathione metabolism: an overview. Biochem. Soc. Transact. 1987;15:717–718.
    1. Mannu F., Arese P., Cappellini M.D., Fiorelli G., Cappadoro M., Giribaldi G., Turrini F. Role of hemichrome binding to erythrocyte membrane in the generation of band-3 alterations in beta-thalassemia intermedia erythrocytes. Blood. 1995;86:2014–2020.
    1. Marinkovic D., Zhang X., Yalcin S., Luciano J.P., Brugnara C., Huber T., Ghaffari S. Foxo3 is required for the regulation of oxidative stress in erythropoiesis. J. Clin. Investig. 2007;117:2133–2144.
    1. Martinez-Gac L., Marques M., Garcia Z., Campanero M.R., Carrera A.C. Control of cyclin G2 mRNA expression by forkhead transcription factors: novel mechanism for cell cycle control by phosphoinositide 3-kinase and forkhead. Mol. Cell. Biol. 2004;24:2181–2189.
    1. Matarrese P., Straface E., Pietraforte D., Gambardella L., Vona R., Maccaglia A., Minetti M., Malorni W. Peroxynitrite induces senescence and apoptosis of red blood cells through the activation of aspartyl and cysteinyl proteases. FASEB J. 2005;19:416–418.
    1. Mathias L.A., Fisher T.C., Zeng L., Meiselman H.J., Weinberg K.I., Hiti A.L., Malik P. Ineffective erythropoiesis in beta-thalassemia major is due to apoptosis at the polychromatophilic normoblast stage. Exp. Hematol. 2000;28:1343–1353.
    1. Matte A., Bertoldi M., Mohandas N., An X., Bugatti A., Brunati A.M., Rusnati M., Tibaldi E., Siciliano A., Turrini F. Membrane association of peroxiredoxin-2 in red cells is mediated by the N-terminal cytoplasmic domain of band 3. Free Rad. Biol. Med. 2013;55:27–35.
    1. May J.M. Ascorbate function and metabolism in the human erythrocyte. Front. Biosc.: J. Virtual Libr. 1998;3:1–10. d.
    1. Mayer B., Hemmens B. Biosynthesis and action of nitric oxide in mammalian cells. Trends Biochem. Sci. 1997;22:477–481.
    1. Miller Y.I., Shaklai N. Kinetics of hemin distribution in plasma reveals its role in lipoprotein oxidation. Biochim. Biophys. Acta. 1999;1454:153–164.
    1. Mohanty J.G., Nagababu E., Rifkind J.M. Red blood cell oxidative stress impairs oxygen delivery and induces red blood cell aging. Front. Physiol. 2014;5:84.
    1. Moncada S., Higgs A. The l-arginine-nitric oxide pathway. N. Engl. J. Med. 1993;329:2002–2012.
    1. Morris C.R. Mechanisms of vasculopathy in sickle cell disease and thalassemia. Hematol. Am. Soc. Hematol. Educ. Program. 2008;2008:177–185.
    1. Nagababu E., Chrest F.J., Rifkind J.M. Hydrogen-peroxide-induced heme degradation in red blood cells: the protective roles of catalase and glutathione peroxidase. Biochim. Biophys. Acta. 2003;1620:211–217.
    1. Nagababu E., Mohanty J.G., Friedman J.S., Rifkind J.M. Role of peroxiredoxin-2 in protecting RBCs from hydrogen peroxide-induced oxidative stress. Free Rad. Res. 2013;47:164–171.
    1. Nagababu E., Rifkind J.M. Formation of fluorescent heme degradation products during the oxidation of hemoglobin by hydrogen peroxide. Biochem. Biophys. Res. Commun. 1998;247:592–596.
    1. Nagy E., Eaton J.W., Jeney V., Soares M.P., Varga Z., Galajda Z., Szentmiklosi J., Mehes G., Csonka T., Smith A. Red cells, hemoglobin, heme, iron, and atherogenesis. Arterioscler. Thromb. Vasc. Biol. 2010;30:1347–1353.
    1. Nath K.A., Grande J.P., Croatt A.J., Likely S., Hebbel R.P., Enright H. Kidney Int. 1998;53:100–111.
    1. Nath K.A., Grande J.P., Haggard J.J., Croatt A.J., Katusic Z.S., Solovey A., Hebbel R.P. Oxidative stress and induction of heme oxygenase-1 in the kidney in sickle cell disease. Am. J. Pathol. 2001;158:893–903.
    1. Nemerson Y. The tissue factor pathway of blood coagulation. Semin. Hematol. 1992;29:170–176.
    1. Nemeth E., Tuttle M.S., Powelson J., Vaughn M.B., Donovan A., Ward D.M., Ganz T., Kaplan J. Hepcidin regulates cellular iron efflux by binding to ferroportin and inducing its internalization. Science. 2004;306:2090–2093.
    1. Nemoto S., Finkel T. Redox regulation of forkhead proteins through a p66shc-dependent signaling pathway. Science. 2002;295:2450–2452.
    1. Ney P.A., Christopher M.M., Hebbel R.P. Synergistic effects of oxidation and deformation on erythrocyte monovalent cation leak. Blood. 1990;75:1192–1198.
    1. Nicolas G., Chauvet C., Viatte L., Danan J.L., Bigard X., Devaux I., Beaumont C., Kahn A., Vaulont S. ). The gene encoding the iron regulatory peptide hepcidin is regulated by anemia, hypoxia, and inflammation. J. Clin. Investig. 2002;110:1037–1044.
    1. Nicolas G., Viatte L., Bennoun M., Beaumont C., Kahn A., Vaulont S. Hepcidin, a new iron regulatory peptide. Blood Cells Mol. Dis. 2002;29:327–335.
    1. Ogasawara Y., Funakoshi M., Ishii K. Glucose metabolism is accelerated by exposure to t-butylhydroperoxide during NADH consumption in human erythrocytes. Blood Cells Mol. Dis. 2008;41:237–243.
    1. Ogasawara Y., Ohminato T., Nakamura Y., Ishii K. Structural and functional analysis of native peroxiredoxin 2 in human red blood cells. Int. J. Biochem. Cell Biol. 2012;44:1072–1077.
    1. Olivieri O., De Franceschi L., Capellini M.D., Girelli D., Corrocher R., Brugnara C. Oxidative damage and erythrocyte membrane transport abnormalities in thalassemias. Blood. 1994;84:315–320.
    1. Omura T. Heme-thiolate proteins. Biochem. Biophys. Res. Commun. 2005;338:404–409.
    1. Osarogiagbon U.R., Choong S., Belcher J.D., Vercellotti G.M., Paller M.S., Hebbel R.P. Reperfusion injury pathophysiology in sickle transgenic mice. Blood. 2000;96:314–320.
    1. Oudit G.Y., Sun H., Trivieri M.G., Koch S.E., Dawood F., Ackerley C., Yazdanpanah M., Wilson G.J., Schwartz A., Liu P.P. l-type Ca2+ channels provide a major pathway for iron entry into cardiomyocytes in iron-overload cardiomyopathy. Nat. Med. 2003;9:1187–1194.
    1. Owusu-Ansah A., Choi S.H., Petrosiute A., Letterio J.J., Huang A.Y. Triterpenoid inducers of Nrf2 signaling as potential therapeutic agents in sickle cell disease: a review. Front. Med. 2015;9:46–56.
    1. Paglialunga F., Fico A., Iaccarino I., Notaro R., Luzzatto L., Martini G., Filosa S. G6PD is indispensable for erythropoiesis after the embryonic-adult hemoglobin switch. Blood. 2004;104:3148–3152.
    1. Pak M., Lopez M.A., Gabayan V., Ganz T., Rivera S. Suppression of hepcidin during anemia requires erythropoietic activity. Blood. 2006;108:3730–3735.
    1. Pantaleo A., Giribaldi G., Mannu F., Arese P., Turrini F. Naturally occurring anti-band 3 antibodies and red blood cell removal under physiological and pathological conditions. Autoimmun. Rev. 2008;7:457–462.
    1. Park C.H., Valore E.V., Waring A.J., Ganz T. Hepcidin, a urinary antimicrobial peptide synthesized in the liver. J. Biol. Chem. 2001;276:7806–7810.
    1. Parks D.A., Granger D.N. Ischemia-induced vascular changes: role of xanthine oxidase and hydroxyl radicals. Am. J. Physiol. 1983;245:G285–289.
    1. Peng H.B., Libby P., Liao J.K. Induction and stabilization of I kappa B alpha by nitric oxide mediates inhibition of NF-kappa B. J. Biol. Chem. 1995;270:14214–14219.
    1. Percy M.J., Lappin T.R. Recessive congenital methaemoglobinaemia: cytochrome b(5) reductase deficiency. Br. J. Haematol. 2008;141:298–308.
    1. Perrotta S., Borriello A., Scaloni A., De Franceschi L., Brunati A.M., Turrini F., Nigro V., del Giudice E.M., Nobili B., Conte M.L. The N-terminal 11 amino acids of human erythrocyte band 3 are critical for aldolase binding and protein phosphorylation: implications for band 3 function. Blood. 2005;106:4359–4366.
    1. Pigeon C., Ilyin G., Courselaud B., Leroyer P., Turlin B., Brissot P., Loreal O. A new mouse liver-specific gene, encoding a protein homologous to human antimicrobial peptide hepcidin, is overexpressed during iron overload. J. Biol. Chem. 2001;276:7811–7819.
    1. Porter J.B., Walter P.B., Neumayr L.D., Evans P., Bansal S., Garbowski M., Weyhmiller M.G., Harmatz P.R., Wood J.C., Miller J.L. Mechanisms of plasma non-transferrin bound iron generation: insights from comparing transfused diamond blackfan anaemia with sickle cell and thalassaemia patients. Br. J. Haematol. 2014;167:692–696.
    1. Presta A., Siddhanta U., Wu C., Sennequier N., Huang L., Abu-Soud H.M., Erzurum S., Stuehr D.J. Comparative functioning of dihydro- and tetrahydropterins in supporting electron transfer, catalysis, and subunit dimerization in inducible nitric oxide synthase. Biochemistry. 1998;37:298–310.
    1. Pries A.R., Secomb T.W., Gaehtgens P. Biophysical aspects of blood flow in the microvasculature. Cardiovas. Res. 1996;32:654–667.
    1. Qian Q., Nath K.A., Wu Y., Daoud T.M., Sethi S. Hemolysis and acute kidney failure. Am. J. Kidney Dis. 2010;56:780–784.
    1. Rachmilewitz E.A., Weizer-Stern O., Adamsky K., Amariglio N., Rechavi G., Breda L., Rivella S., Cabantchik Z.I. Role of iron in inducing oxidative stress in thalassemia: Can it be prevented by inhibition of absorption and by antioxidants? Ann. N.Y. Acad. Sci. 2005;1054:118–123.
    1. Radomski M.W., Palmer R.M., Moncada S. The role of nitric oxide and cGMP in platelet adhesion to vascular endothelium. Biochem. Biophys. Res. Commun. 1987;148:1482–1489.
    1. Ramirez J.M., Schaad O., Durual S., Cossali D., Docquier M., Beris P., Descombes P., Matthes T. Growth differentiation factor 15 production is necessary for normal erythroid differentiation and is increased in refractory anaemia with ring-sideroblasts. Br. J. Haematol. 2009;144:251–262.
    1. Ramos C.L., Pou S., Britigan B.E., Cohen M.S., Rosen G.M. Spin trapping evidence for myeloperoxidase-dependent hydroxyl radical formation by human neutrophils and monocytes. J. Biol. Chem. 1992;267:8307–8312.
    1. Redding G.S., Record D.M., Raess B.U. Calcium-stressed erythrocyte membrane structure and function for assessing glipizide effects on transglutaminase activation. Proc. Soc. Exp. Biol. Med. Soc. Exp. Biol. Med. 1991;196:76–82.
    1. Reiter C.D., Wang X., Tanus-Santos J.E., Hogg N., Cannon R.O., 3rd, Schechter A.N., Gladwin M.T. . Cell-free hemoglobin limits nitric oxide bioavailability in sickle-cell disease. Nat. Med. 2002;8:1383–1389.
    1. Reiter R.J., Melchiorri D., Sewerynek E., Poeggeler B., Barlow-Walden L., Chuang J., Ortiz G.G., Acuna-Castroviejo D. A review of the evidence supporting melatonin's role as an antioxidant. J. Pineal Res. 1995;18:1–11.
    1. Ren H., Ghebremeskel K., Okpala I., Lee A., Ibegbulam O., Crawford M. Patients with sickle cell disease have reduced blood antioxidant protection. Int. J. Vitam. Nutr. Res. 2008;78:139–147.
    1. Ribeil J.A., Arlet J.B., Dussiot M., Moura I.C., Courtois G., Hermine O. Ineffective erythropoiesis in beta -thalassemia. Sci. World J. 2013;2013:394295.
    1. Rifkind J.M., Mohanty J.G., Nagababu E. The pathophysiology of extracellular hemoglobin associated with enhanced oxidative reactions. Front. Physiol. 2014;5:500.
    1. Rund D., Rachmilewitz E. Beta-thalassemia. N. Engl. J. Med. 2005;353:1135–1146.
    1. Sadrzadeh S.M., Graf E., Panter S.S., Hallaway P.E., Eaton J.W. Hemoglobin. A biologic fenton reagent. J. Biol. Chem. 1984;259:14354–14356.
    1. Sankaran V.G., Orkin S.H. The switch from fetal to adult hemoglobin. Cold Spring Harbor Perspect. Med. 2013;3:a011643.
    1. Schaer D.J., Buehler P.W., Alayash A.I., Belcher J.D., Vercellotti G.M. Hemolysis and free hemoglobin revisited: exploring hemoglobin and hemin scavengers as a novel class of therapeutic proteins. Blood. 2013;121:1276–1284.
    1. Schnog J.B., Teerlink T., van der Dijs F.P., Duits A.J., Muskiet F.A., Group C.S. Plasma levels of asymmetric dimethylarginine (ADMA), an endogenous nitric oxide synthase inhibitor, are elevated in sickle cell disease. Ann. Hematol. 2005;84:282–286.
    1. Schwartz R.S., Rybicki A.C., Heath R.H., Lubin B.H. Protein 4.1 in sickle erythrocytes. Evidence for oxidative damage. J. Biol. Chem. 1987;262:15666–15672.
    1. Scott M.D., van den Berg J.J., Repka T., Rouyer-Fessard P., Hebbel R.P., Beuzard Y., Lubin B.H. Effect of excess alpha-hemoglobin chains on cellular and membrane oxidation in model beta-thalassemic erythrocytes. J. Clin. Investig. 1993;91:1706–1712.
    1. Shaeffer J.R. ATP-dependent proteolysis of hemoglobin alpha chains in beta-thalassemic hemolysates is ubiquitin-dependent. J. Biol. Chem. 1988;263:13663–13669.
    1. Shalev O., Repka T., Goldfarb A., Grinberg L., Abrahamov A., Olivieri N.F., Rachmilewitz E.A., Hebbel R.P. Deferiprone (L1) chelates pathologic iron deposits from membranes of intact thalassemic and sickle red blood cells both in vitro and in vivo. Blood. 1995;86:2008–2013.
    1. Shinar E., Rachmilewitz E.A. Differences in the pathophysiology of hemolysis of alpha- and beta-thalassemic red blood cells. Ann. N.Y. Acad. Sci. 1990;612:118–126.
    1. Shinar E., Rachmilewitz E.A., Lux S.E. Differing erythrocyte membrane skeletal protein defects in alpha and beta thalassemia. J. Clin. Investig. 1989;83:404–410.
    1. Singer S.T., Ataga K.I. Hypercoagulability in sickle cell disease and beta-thalassemia. Curr. Mol. Med. 2008;8:639–645.
    1. Smith A., Morgan W.T. Haem transport to the liver by haemopexin. Receptor-mediated uptake with recycling of the protein. Biochem. J. 1979;182:47–54.
    1. Solovey A., Gui L., Key N.S., Hebbel R.P. Tissue factor expression by endothelial cells in sickle cell anemia. J. Clin. Investig. 1998;101:1899–1904.
    1. Stuhlmeier K.M., Kao J.J., Wallbrandt P., Lindberg M., Hammarstrom B., Broell H., Paigen B. Antioxidant protein 2 prevents methemoglobin formation in erythrocyte hemolysates. Eur. J. Biochem./FEBS. 2003;270:334–341.
    1. Sudano I., Roas S., Noll G. Vascular abnormalities in essential hypertension. Curr. Pharm. Des. 2011;17:3039–3044.
    1. Suragani R.N., Zachariah R.S., Velazquez J.G., Liu S., Sun C.W., Townes T.M., Chen J.J. Heme-regulated eIF2alpha kinase activated Atf4 signaling pathway in oxidative stress and erythropoiesis. Blood. 2012;119:5276–5284.
    1. Szocs K. Endothelial dysfunction and reactive oxygen species production in ischemia/reperfusion and nitrate tolerance. Gen. Physiol. Biophys. 2004;23:265–295.
    1. Taksande A., Prabhu S., Venkatesh S. Cardiovascular aspect of Beta-thalassaemia. Cardiovas. Hematol. Agents Med. Chem. 2012;10:25–30.
    1. Tamary H., Shalev H., Perez-Avraham G., Zoldan M., Levi I., Swinkels D.W., Tanno T., Miller J.L. Elevated growth differentiation factor 15 expression in patients with congenital dyserythropoietic anemia type I. Blood. 2008;112:5241–5244.
    1. Tanno T., Porayette P., Sripichai O., Noh S.J., Byrnes C., Bhupatiraju A., Lee Y.T., Goodnough J.B., Harandi O., Ganz T. Identification of TWSG1 as a second novel erythroid regulator of hepcidin expression in murine and human cells. Blood. 2009;114:181–186.
    1. Tracz M.J., Alam J., Nath K.A. Physiology and pathophysiology of heme: implications for kidney disease. J. Am. Soc. Nephrol. 2007;18:414–420.
    1. Tran H., Brunet A., Grenier J.M., Datta S.R., Fornace A.J., Jr., DiStefano P.S., Chiang L.W., Greenberg M.E. DNA repair pathway stimulated by the forkhead transcription factor FOXO3a through the Gadd45 protien. Science. 2002;296:530–534.
    1. Umbreit J. Methemoglobin – it's not just blue: a concise review. Am. J. Hematol. 2007;82:134–144.
    1. van Beers E.J., van Tuijn C.F., Mac Gillavry M.R., van der Giessen A., Schnog J.J., Biemond B.J. Haematologica. 2008;93:757–760.
    1. van Zwieten R., Verhoeven A.J., Roos D. Inborn defects in the antioxidant systems of human red blood cells. Free Rad. Biol. Med. 2014;67:377–386.
    1. Vettore L., De Matteis M.C., Di Iorio E.E., Winterhalter K.H. Erythrocytic proteases: preferential degradation of alpha hemoglobin chains. Acta Haematol. 1983;70:35–42.
    1. Vichinsky E., Butensky E., Fung E., Hudes M., Theil E., Ferrell L., Williams R., Louie L., Lee P.D., Harmatz P. Comparison of organ dysfunction in transfused patients with SCD or beta thalassemia. Am. J. Hematol. 2005;80:70–74.
    1. Vokurka M., Krijt J., Sulc K., Necas E. Hepcidin mRNA levels in mouse liver respond to inhibition of erythropoiesis. Physiol. Res./Acad. Sci. Bohemoslov. 2006;55:667–674.
    1. Wagener F.A., Feldman E., de Witte T., Abraham N.G. Heme induces the expression of adhesion molecules ICAM-1, VCAM-1, and E selectin in vascular endothelial cells. Proc. Soc. Exp. Biol. Med. 1997;216:456–463.
    1. Weatherall D., Akinyanju O., Fucharoen S., Olivieri N., Musgrove P. In disease control priorities in developing countries. In: Jamison D.T., Breman J.G., Measham A.R., Alleyne G., Claeson M., Evans D.B., Jha P., Mills A., Musgrove P., editors. Inherited Disorders of Hemoglobin. World Bank; Washington(DC): 2006.
    1. Westerman M., Pizzey A., Hirschman J., Cerino M., Weil-Weiner Y., Ramotar P., Eze A., Lawrie A., Purdy G., Mackie I. Microvesicles in haemoglobinopathies offer insights into mechanisms of hypercoagulability, haemolysis and the effects of therapy. Br. J. Haematol. 2008;142:126–135.
    1. Westwick J., Watson-Williams E.J., Krishnamurthi S., Marks G., Ellis V., Scully M.F., White J.M., Kakkar V.V. Platelet activation during steady state sickle cell disease. J. Med. 1983;14:17–36.
    1. Wickramasinghe S.N., Bush V. Observations on the ultrastructure of erythropoietic cells and reticulum cells in the bone marrow of patients with homozygous beta-thalassaemia. Br. J. Haematol. 1975;30:395–399.
    1. Willekens F.L., Werre J.M., Groenen-Dopp Y.A., Roerdinkholder-Stoelwinder B., de Pauw B., Bosman G.J. Erythrocyte vesiculation: a self-protective mechanism? Br. J. Haematol. 2008;141:549–556.
    1. Winichagoon P., Fucharoen S., Wasi P. Increased circulating platelet aggregates in thalassaemia. Southeast Asian J. Trop. Med. Public Health. 1981;12:556–560.
    1. Wood K.C., Hebbel R.P., Granger D.N. Endothelial cell NADPH oxidase mediates the cerebral microvascular dysfunction in sickle cell transgenic mice. FASEB J. 2005;19:989–991.
    1. Wood K.C., Hebbel R.P., Lefer D.J., Granger D.N. Critical role of endothelial cell-derived nitric oxide synthase in sickle cell disease-induced microvascular dysfunction. Free Rad. Biol. Med. 2006;40:1443–1453.
    1. Wu C.J., Krishnamurti L., Kutok J.L., Biernacki M., Rogers S., Zhang W., Antin J.H., Ritz J. Evidence for ineffective erythropoiesis in severe sickle cell disease. Blood. 2005;106:3639–3645.
    1. Wun T. The role of inflammation and leukocytes in the pathogenesis of sickle cell disease; haemoglobinopathy. Hematology. 2001;5:403–412.
    1. Xia Y., Dawson V.L., Dawson T.M., Snyder S.H., Zweier J.L. Nitric oxide synthase generates superoxide and nitric oxide in arginine-depleted cells leading to peroxynitrite-mediated cellular injury. Proc. Natl. Acad. Sci. USA. 1996;93:6770–6774.
    1. Yu X., Kong Y., Dore L.C., Abdulmalik O., Katein A.M., Zhou S., Choi J.K., Gell D., Mackay J.P., Gow A.J. An erythroid chaperone that facilitates folding of alpha-globin subunits for hemoglobin synthesis. J. Clin. Investig. 2007;117:1856–1865.
    1. Zeuner A., Eramo A., Testa U., Felli N., Pelosi E., Mariani G., Srinivasula S.M., Alnemri E.S., Condorelli G., Peschle C. Control of erythroid cell production via caspase-mediated cleavage of transcription factor SCL/Tal-1. Cell Death Differ. 2003;10:905–913.
    1. Zhang J., Ney P.A. Autophagy-dependent and -independent mechanisms of mitochondrial clearance during reticulocyte maturation. Autophagy. 2009;5:1064–1065.
    1. Zhou S., Olson J.S., Fabian M., Weiss M.J., Gow A.J. Biochemical fates of alpha hemoglobin bound to alpha hemoglobin-stabilizing protein AHSP. J. Biol. Chem. 2006;281:32611–32618.
    1. Zipser Y., Piade A., Barbul A., Korenstein R., Kosower N.S. Ca2+ promotes erythrocyte band 3 tyrosine phosphorylation via dissociation of phosphotyrosine phosphatase from band 3. Biochem. J. 2002;368:137–144.
    1. Zwaal R.F., Schroit A.J. Pathophysiologic implications of membrane phospholipid asymmetry in blood cells. Blood. 1997;89:1121–1132.

Source: PubMed

3
Subscribe