The pharmacokinetics and pharmacodynamics of iron preparations

Peter Geisser, Susanna Burckhardt, Peter Geisser, Susanna Burckhardt

Abstract

Standard approaches are not appropriate when assessing pharmacokinetics of iron supplements due to the ubiquity of endogenous iron, its compartmentalized sites of action, and the complexity of the iron metabolism. The primary site of action of iron is the erythrocyte, and, in contrast to conventional drugs, no drug-receptor interaction takes place. Notably, the process of erythropoiesis, i.e., formation of new erythrocytes, takes 3-4 weeks. Accordingly, serum iron concentration and area under the curve (AUC) are clinically irrelevant for assessing iron utilization. Iron can be administered intravenously in the form of polynuclear iron(III)-hydroxide complexes with carbohydrate ligands or orally as iron(II) (ferrous) salts or iron(III) (ferric) complexes. Several approaches have been employed to study the pharmacodynamics of iron after oral administration. Quantification of iron uptake from radiolabeled preparations by the whole body or the erythrocytes is optimal, but alternatively total iron transfer can be calculated based on known elimination rates and the intrinsic reactivity of individual preparations. Degradation kinetics, and thus the safety, of parenteral iron preparations are directly related to the molecular weight and the stability of the complex. High oral iron doses or rapid release of iron from intravenous iron preparations can saturate the iron transport system, resulting in oxidative stress with adverse clinical and subclinical consequences. Appropriate pharmacokinetics and pharmacodynamics analyses will greatly assist our understanding of the likely contribution of novel preparations to the management of anemia.

Figures

Figure 1.
Figure 1.
Schematic representation of iron metabolism. Under normal conditions, the iron in the body is in a dynamic equilibrium between different compartments (solid arrows). From approximately 10 mg of iron ingested with food, 1–2 mg are absorbed by duodenal enterocytes and the same amount is lost, e.g., via skin exfoliation. In the circulation, iron is bound to transferrin (ca. 3 mg), which safely transports it e.g., to the bone marrow for hemoglobin synthesis. Approximately two-thirds of the iron in the body is found in the form of hemoglobin, in red blood cells (1800 mg) and in erythroid precursors in the bone marrow (300 mg), whereas 10–15% is present in myoglobin and in a variety of different essential enzymes. Iron is stored in parenchymal cells of the liver (ca. 1000 mg). Reticuloendothelial macrophages temporarily store the iron recycled from senescent red blood cells (600 mg) in a readily available form. Erythropoetin, produced in the kidneys, regulates duodenal iron absorption and erythropoiesis (dashed lines). Adapted from Crichton, 2008 [7].
Figure 2.
Figure 2.
In vitro reactivity of Ferinject®, Venofer® and Ferrlecit® with apotransferrin. Urea polyacrylamide gel electrophoresis (PAGE) of transferrin incubated with different amounts of various intravenous iron preparations. Apo-Tf, transferrin with no iron; Fe-Tf, transferrin with one iron-binding site occupied; Fe2-Tf, transferrin with both iron-binding sites occupied [holotransferrin]. The reactivity towards apotransferrin was the lowest with the most stable complex, i.e. Ferinject®. At concentrations equivalent to those expected in the serum of an adult after a therapeutic dose of ∼200 or ∼2,000 mg of iron, transferrin saturation was observed with Ferrlecit® and Venofer® but not with Ferinject® (Technical communication, Vifor Pharma – Vifor International Inc).
Figure 3.
Figure 3.
Normalized simulated single first-order elimination kinetics for different intravenous iron preparations, depicted as fraction of total serum iron over time. Values of the terminal elimination rates given in Table 2 were used to calculate an overall first-order kinetics and t1/2 values. The figure clearly shows that the AUC is negatively correlated to the elimination rate constants.
Figure 4.
Figure 4.
Serum concentration of non-transferrin bound iron (NTBI) and percentage transferrin saturation (TSAT) following administration of a single oral dose of 100 mg iron in the form of three different ferrous salts to healthy adult volunteers. Broken blue lines indicate the percentage transferrin saturation (right-hand axis). Solid red lines indicate NTBI concentration (left-hand axis). Values shown are mean ± SD. Modified from Dresow et al. 2008 [34].
Figure 5.
Figure 5.
Increase in serum iron concentration after administration of 25, 50 and 100 mg ferrous iron in 6 healthy subjects [37]. Data are shown as mean ± SEM. The data clearly show that there is no linear relationship between serum iron increase (Cmax and AUC) and dose.
Figure 6.
Figure 6.
Utilization of iron following a single intravenous administration of radiolabeled iron sucrose (Venofer®) in a patient with iron deficiency anemia (modified from Beshara et al. 1999 [10]).
Figure 7.
Figure 7.
Illustration of the mean measured serum iron concentration (red lines) and the calculated curve (green lines) based on the following equation: C(t) = a (1−e−kin*t) − k0t, where C(t) is the serum iron concentration at time t, a is a constant, kin is the rate constant for iron absorption, and k0 is the rate constant for elimination. Data are from an open-label, single-dose, randomized, crossover bioequivalence study in 20 healthy female volunteers given standard oral ferrous fumarate or slow-release ferrous fumarate at a dose equivalent to 100 mg iron per intake (Modified from Geisser et al., 2009 [13]).

References

    1. Gotloib L., Silverberg D., Fudin R., Shostak A. Iron deficiency is a common cause of anemia in chronic kidney disease and can often be corrected with intravenous iron. J. Nephrol. 2006;19:161–167.
    1. Nanas P.N., Matsouka C., Karageorgopoulos D., Leonti A., Tsolakis E., Drakos S.G., Tsagalou E.P., Maroulidis G.D., Alexopoulos G.P., Kanakakis J.E., et al. Etiology of anemia in patients with advanced heart failure. J. Amer. Coll. Cardiol. 2006;48:2485–2489.
    1. Gasche C., Lomer M.C.E., Cavill I., Weiss G. Iron, anaemia, and inflammatory bowel diseases. Gut. 2004;53:1190–1197.
    1. Kalantar-Zadeh K., Streja E., Miller J.E., Nissenson A.R. Intravenous iron versus erythropoiesis-stimulating agents: friends or foes in treating chronic kidney disease anemia? Adv. Chronic Kidney Dis. 2009;16:143–151.
    1. Schümann K., Classen H.G., Hages M., Prinz-Langenohl R., Pietrzik K., Biesalski H.K. Bioavailability of oral vitamins, minerals, and trace elements in perspective. Drug Res. 1997;47:369–380.
    1. Andrews N.C. Disorders of iron metabolism. N. Engl. J. Med. 1999;341:1986–1995.
    1. Crichton R.R., Danielson B.G., Geisser P. Iron Therapy with Special Emphasis on Intravenous Administration. 4th ed. UNI-MED Verlag AG; Bremen, Germany: 2008.
    1. European Agency for the Evaluation of Medicinal Products Committee for Proprietary Medicinal Products (CPMP) CPMP/EWP/QWP/1401/98 . Accessed on 7 July 2010.
    1. Wienk K.J., Marx J.J., Beynen A.C. The concept of iron and its bioavailability. Eur. J. Nutr. 1999;38:51–75.
    1. Beshara S., Lundqvist H., Sundin J., Lubberink M., Tolmachev V., Valind S., Antoni G., Långström B., Danielson B.G. Pharmacokinetics and red cell utilization of iron(III)-hydroxide-sucrose complex in anaemic patients: a study using positron emission tomography. Br. J. Haematol. 1999;104:296–302.
    1. Forth W. Iron: Bioavailability, Absorption, Utilization. BI Wissenschaftsverlag; Mannheim, Germany: 1992. p. 36.
    1. Ekenved G. Iron Absorption Studies. Munksgaard; Copenhagen, Denmark: 1976.
    1. Geisser P., Philipp E. True iron bioavailability, iron pharmacokinetics and clinically silent side effects. Nutr. Immun. Health. 2009;1:3–11.
    1. Danielson J. Structure, chemistry, and pharmacokinetics of intravenous iron agents. Am. Soc. Nephrol. 2004;15:S93–S98.
    1. Evans R.W., Rafique R., Zarea A., Rapisarda C., Cammack R., Evans P.J., Porter J.B., Hider R.C. Nature of non-transferrin-bound iron: studies on iron citrate complexes and the thalassemic sera. J. Biol. Inorg. Chem. 2008;13:57–74.
    1. Geisser P. Reactivity of iron compounds (letter) Am. J. Kidney Dis. 2003;42:1103–1104.
    1. Geisser P., Baer M., Schaub E. Structure/histotoxicity relationship of parenteral iron preparations. Drug Res. 1992;42:1439–1452.
    1. Van Wyck D., Anderson J., Johnson K. Labile iron in parenteral iron formulations: a quantitative and comparative study. Nephrol. Dialysis Transplant. 2004;19:561–565.
    1. Chandler G., Harchowal J., Macdougall I.C. Intravenous iron sucrose: establishing a safe dose. Am. J. Kidney Dis. 2001;38:988–991.
    1. Danielson B.G., Salmonson T., Derendorf H., Geisser P. Pharmacokinetics of iron(III)-hydroxide sucrose complex after a single intravenous dose in healthy volunteers. Drug Res. 1996;46:615–621.
    1. Seligman P.A., Dahl N.V., Strobos J., Kimko H.C., Schleicher R.B., Jones M., Ducharme M.P. Single-dose pharmacokinetics of sodium ferric gluconate complex in iron-deficient subjects. Pharmacotherapy. 2004;24:574–583.
    1. Geisser P., Banké-Bochita J. Pharmacokinetics, safety and tolerability of intravenous ferric carboxymaltose: a dose-escalation study in volunteers with mild iron-deficiency anaemia. Drug Res. 2010;60:362–372.
    1. Balakrishnan V.S., Rao M., Kausz A.T., Brenner L., Pereira B.J., Frigo T.B., Lewis J.M. Physico chemical properties of ferumoxytol, a new intravenous iron preparation. Eur. J. Clin. Invest. 2009;39:489–496.
    1. Henderson P.A., Hillman R.S. Characteristics of iron dextran utilization in man. Blood. 1969;34:357–375.
    1. Landry R., Jacobs P.M., Davis R., Shenouda M., Bolton W.K. Pharmacokinetics study of ferumoxytol: a new iron replacement therapy in normal subjects and hemodialysis patients. Am. J. Nephrol. 2005;25:400–410.
    1. Geisser P., Rumyantsev V. Pharmacodynamics and safety of ferric carboxymaltose: a multiple-dose study in patients with iron-deficiency anaemia secondary to a gastrointestinal disorder. Drug Res. 2010;60:373–385.
    1. Hentze M.W., Muckenthaler M.U., Galy B., Camaschella C. Two to tango: regulation of Mammalian iron metabolism. Cell. 2010;142:24–38.
    1. Latunde-Dada G.O., Takeuchi K., Simpson R.J., McKie A.T. Haem carrier protein 1 (HCP1): Expression and functional studies in cultured cells. FEBS Lett. 2006;580:6865–6870.
    1. Qiu A., Jansen M., Sakaris A., Min S.H., Chattopadhyay S., Tsai E., Sandoval C., Zhao R., Akabas M.H., Goldman I.D. Identification of an intestinal folate transporter and the molecular basis for hereditary folate malabsorption. Cell. 2006;127:917–928.
    1. Nakai Y., Inoue K., Abe N., Hatakeyama M., Ohta K.Y., Otagiri M., Hayashi Y., Yuasa H. Functional characterization of human proton-coupled folate transporter/heme carrier protein 1 heterologously expressed in mammalian cells as a folate transporter. J. Pharmacol. Exp. Ther. 2007;322:469–476.
    1. Andrew N.C. When is a heme transporter not a heme transporter? When it's a folate transporter. Cell Metab. 2007;5:5–6.
    1. Ferris C.D., Jaffrey S.R., Sawa A., Takahashi M., Brady S.D., Barrow R.K., Tysoe S.A., Wolosker H., Barañano D.E., Doré S., et al. Haem oxygenase-1 prevents cell death by regulating cellular iron. Nat. Cell. Biol. 1999;1:152–157.
    1. Heinrich H.C., Gabbe E.E., Whang D.H. Dose relationship of intestinal iron absorption in men with normal iron stores and persons with prelatent/latent iron deficiency. Z. Naturforshg. 1969;24b:1301–1310.
    1. Dresow B., Petersen D., Fischer R., Nielsen P. Non-transferrin-bound iron in plasma following administration of oral iron drugs. Biometals. 2008;21:273–276.
    1. Hutchinson C., Al-Ashgar W., Liu D.Y., Hider R.C., Powell J.J., Geissler C.A. Oral ferrous sulphate leads to a marked increase in pro-oxidant nontransferrin-bound iron. Eur. J. Clin. Invest. 2004;34:782–784.
    1. Macdougall I.C. Strategies for iron supplementation: oral versus intravenous. Kidney Int. Suppl. 1999;69:S61–S66.
    1. Ekenved G., Norrby A., Sölvell L. Serum iron increase as a measure of iron absorption -studies on the correlation with total absorption. Scand. J. Haematol. 1976;17(Suppl. 28):31–49.
    1. Beutler E., Buttenwieser E. The regulation of iron absorption I. a search for humoral factors. J. Lab. Clin. Med. 1960;55:274–280.
    1. Werner E., Kaltwasser J.P. Judgement of measured values of intestinal iron absorption. Drug Res. 1987;37:116–121.
    1. Heinrich H.C., Fischer R. Correlation of postaborptive serum iron increase and erythrocyte-59Fe-incorporation with the whole body retention of absorbed 59Fe. Klin. Wochenschr. 1982;60:1493–1496.
    1. Kaltwasser J.P., Werner E., Niechzial M. Bioavailability and therapeutic efficacy of bivalent and trivalent iron preparations. Drug Res. 1987;37:122–129.
    1. Potgieter M.A., Potgieter J.H., Venter C., Venter J.L., Geisser P. Effect of oral aluminium hydroxide on iron absorption from iron(III)-hydroxide polymaltose complex in patients with iron deficiency anemia: A single-centre randomized controlled isotope study. Drug Res. 2007;57:392–400.
    1. Kaltwasser J.P., Hansen C., Oebike Y., Werner E. Assessment of iron availability using stable 54Fe. Eur. J. Clin. Invest. 1991;21:436–442.
    1. Geisser P., Müller A. Pharmacokinetics of iron salts and ferric hydroxide-carbohydrate complexes. Drug Res. 1987;37:100–104.
    1. Jacobs P., Johnson G., Wood L. Oral iron therapy in human subjects, comparative absorption between ferrous salts and iron polymaltose. J. Med. 1984;15:367–377.
    1. Jacobs P., Wood L., Bird A.R. Erythrocytes: better tolerance of iron polymaltose complex compared with ferrous sulphate in the treatment of anaemia. Hematology. 2000;5:77–83.
    1. Beshara S., Sorensen J., Lubberink M., Tolmachev V., Långström B., Antoni G., Danielson B.G., Lundqvist H. Pharmacokinetics and red cell utilization of 52Fe/59Fe-labelled iron polymaltose in anaemic patients using positron emission tomography. Br. J. Haematol. 2003;120:853–859.
    1. Jacobs P., Wormald L.A., Gregory M.C. Absorption of iron polymaltose and ferrous sulphate in rats and humans. S. Afr. Med. J. 1979;55:1065–1072.
    1. Walczyk T., Davidsson L., Zavaleta N., Hurrell R.F. Stable isotope labels as a tool to determine the iron absorption by Peruvian school children from a breakfast meal. Fresenius J. Anal. Chem. 1997;359:445–449.
    1. Widness J.A., Serfass R.E., Haiden N., Nelson S.E., Lombard K.A., Pollak A. Erythrocyte iron incorporation but not absorption is increased by intravenous iron administration in erythropoietin-treated premature infants. J. Nutr. 2006;136:1868–1873.
    1. Hallberg L., Sölvell L. Iron absorption studies. Acta Med. Scand. 1960;168:3–108.
    1. Cantone M.C., Molho N., Pirola L., Gambarini G., Hansen C., Roth P., Werner E. An iron metabolism study in humans by means of stable tracers. Med. Phys. 1988;15:862–866.
    1. Oudit G.Y., Trivieri M.G., Khaper N., Liu P.P., Backx P.H. Role of L-type Ca2+ channels in iron transport and iron-overload cardiomyopathy. J. Mol. Med. 2006;84:349–364.
    1. Fishbane S. Iron management in nondialysis-dependent CKD. Am. J. Kidney Dis. 2007;49:736–743.
    1. Santos M., De Sousa M. In vitro modulation of T-cell surface molecules by iron. Cell. Immunol. 1994;154:498–506.
    1. Rudd M.F., Good M.F., Chapman D.E., Powell L.W., Halliday J.W. Clonal analysis of the effect of iron on human cytotoxic and proliferation T lymphocytes. Immunol. Cell. Biol. 1990;68:317–324.
    1. Lyseng-Williamson K.A., Keating G.M. Ferric carboxymaltose: a review of its use in iron-deficiency anaemia. Drugs. 2009;69:739–756.
    1. Anker S.D., Comin Colet J.C., Filippatos G., Willenheimer R., Dickstein K., Drexler H., Lüscher T.F., Bart B., Banasiak W., Niegowska J., et al. Ferric carboxymaltose in patients with heart failure and iron deficiency. N. Engl. J. Med. 2009;361:2436–2448.
    1. Macdougall I.C., Ashenden M. Current and upcoming erythropoiesis-stimulating agents, iron products, and other novel anemia medication. Adv. Chronic Kidney Dis. 2009;6:117–130.
    1. Bailie G.R., Clark J.A., Lane C.E., Lane P.L. Hypersensitivity reactions and deaths associated with intravenous iron preparations. Nephrol. Dialysis Transplant. 2005;20:1443–1449.
    1. Chertow G.M., Mason P.D., Vaage-Nilsen O., Ahlmén J. Update on adverse drug events associated with parenteral iron. Nephrol. Dialysis Transplant. 2006;21:378–382.
    1. Lu M., Cohen M.H., Rieves D., Pazdur R. FDA report: ferumoxytol for intravenous iron therapy in adult patients with chronic kidney disease. Am. J. Hematol. 2010;85:315–319.
    1. Santosh S., Podaralla P., Miller B. Anaphylaxis with elevated serum tryptase after administration of intravenous ferumoxytol. Nephrol. Dialysis Transplant. 2010;3:341–342.
    1. Qunibi W.Y. The efficacy and safety of current intravenous iron preparations for the management of iron-deficiency anaemia: a review. Drug Res. 2010;60:399–412.
    1. Olsson K.S., Weinfeld A. Availability of iron dextran for hemoglobin synthesis. Acta Med. Scand. 1972;192:543–549.
    1. Tuomainen T.P., Nyyssönen K., Porkkala-Sarataho E., Salonen R., Baumgartner J.A., Geisser P., Salonen J.T. Oral supplementation with ferrous sulphate but not with non-ionic iron polymaltose complex increases the susceptibility of plasma lipoproteins to oxidation. Nutr. Res. 1999;19:1121–1132.
    1. Geisser P. Safety and efficacy of iron(III)-hydroxide polymaltose complex / a review of over 25 years experience. Drug Res. 2007;57:439–452.

Source: PubMed

3
Subskrybuj