throbber
ADIS DRUG EVALUATION
`
`Drugs 1999 Sep; 58 (3): 553-578
`0012-6667/99/0009-0553/$26.00/0
`7298
`© Adis International Limited. All rights reserved.
`
`Deferiprone
`A Review of its Clinical Potential in Iron Overload
`in β-Thalassaemia Major and Other
`Transfusion-Dependent Diseases
`
`Julia A. Barman Balfour and Rachel H. Foster
`Adis International Limited, Auckland, New Zealand
`
`Various sections of the manuscript reviewed by:
`M.B. Agarwal, Department of Haematology, LTMG Hospital and LTM Medical College, Bombay, India; C.
`Giardini, Divisione di Ematologia, Centro Trapianti Midollo Osseo di Muraglia, Pesaro, Italy; A.V. Hoffbrand,
`Department of Haematology, Royal Free Hospital, London, England; G. Koren, Division of Clinical
`Pharmacology and Toxicology, Hospital for Sick Children, Toronto, Ontario, Canada; A. Piga, University of
`Turin, Turin, Italy; M.H. van Weel-Sipman, Department of Paediatrics/BMT Unit, Leiden University Medical
`Centre, Leiden, The Netherlands; B. Wonke, Department of Haematology, Whittington Hospital, London,
`England.
`
`Data Selection
`Sources: Medical literature published in any language since 1966 on deferiprone, identified using AdisBase (a proprietary database of Adis
`International, Auckland, New Zealand) and Medline. Additional references were identified from the reference lists of published articles.
`Bibliographical information, including contributory unpublished data, was also requested from the company developing the drug.
`Search strategy: AdisBase search terms were ‘deferiprone’ or ‘1-2-Dimethyl-3-hydroxypyrid-4-one’ or ‘CGP-37391’ or ‘CP-20’ or ‘L-1’.
`Medline search terms were ‘deferiprone’ or ‘CGP 37391’ or ‘CP 020’ or ‘CP 20’. Searches were last updated 28 Jul 1999.
`Selection: Studies in patients with iron overload who received deferiprone. Inclusion of studies was based mainly on the methods section
`of the trials. When available, large, well controlled trials with appropriate statistical methodology were preferred. Relevant pharmacodynamic
`and pharmacokinetic data are also included.
`Index terms: deferiprone, thalassaemia, iron-overload, pharmacodynamics, pharmacokinetics, therapeutic use.
`
`Contents
`
` . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 554
`Summary
`1.
`Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
`2. Pharmacodynamic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558
`2.1 Binding of Iron and Other Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558
`2.1.1 Promotion of Iron Excretion in Vivo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559
`2.1.2 Source of Iron Chelated by Deferiprone . . . . . . . . . . . . . . . . . . . . . . . . . . . 559
`2.2 Oxidant/Antioxidant Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560
`2.3 Cytotoxic and Antiproliferative Effects
` . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 561
`2.4 Other Effects
` . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 561
`3. Pharmacokinetic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 561
`3.1 Absorption and Plasma Concentrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 561
`3.2 Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 562
`3.3 Metabolism and Elimination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 563
`4. Therapeutic Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 563
`4.1 Urinary Iron Excretion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 565
`4.2 Serum Ferritin Levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566
`
`
`1 of 26
`
`Taro Pharmaceuticals, Ltd.
`Exhibit 1017
`
`

`

`554
`
`Barman Balfour & Foster
`
` . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
`4.3 Hepatic Iron Content
`4.4 Other Effects
` . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
`4.5 Comparisons with Deferoxamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
`4.6 Use in Combination with Deferoxamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569
`4.7 Compliance with Therapy and Quality of Life . . . . . . . . . . . . . . . . . . . . . . . . . . . 569
`5. Tolerability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 570
`5.1 Arthropathy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 570
`5.2 Neutropenia/Agranulocytosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 570
`5.3 Gastrointestinal and Liver Enzyme Disturbances . . . . . . . . . . . . . . . . . . . . . . . . . . 571
`5.4 Immunological Abnormalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
`5.5 Zinc Deficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
`5.6 Hepatic Fibrosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 572
`5.7 Other Effects
` . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 572
`6. Dosage and Administration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 572
`7. Potential Place of Deferiprone in the Management of Iron Overload in
`β-Thalassaemia Major and Other Transfusion-Dependent Diseases . . . . . . . . . . . . . . . . . . 573
`
`Summary
`Abstract
`
`Patients with β-thalassaemia and other transfusion-dependent diseases develop
`iron overload from chronic blood transfusions and require regular iron chelation
`to prevent potentially fatal iron-related complications. The only iron chelator
`currently widely available is deferoxamine, which is expensive and requires
`prolonged subcutaneous infusion 3 to 7 times per week or daily intramuscular
`injections. Moreover, some patients are unable to tolerate deferoxamine and com-
`pliance with the drug is poor in many patients.
`Deferiprone is the most extensively studied oral iron chelator to date. Non-
`comparative clinical studies mostly in patients with β-thalassaemia have demon-
`strated that deferiprone 75 to 100 mg/kg/day can reduce iron burden in regularly
`transfused iron-overloaded patients. Serum ferritin levels are generally reduced
`in patients with very high pretreatment levels and are frequently maintained
`within an acceptable range in those who are already adequately chelated.
`Deferiprone is not effective in all patients (some of whom show increases in serum
`ferritin and/or liver iron content, particularly during long term therapy). This may
`reflect factors such as suboptimal dosage and/or severe degree of iron overload
`at baseline in some instances.
`Although few long term comparative data are available, deferiprone at the
`recommended dosage of 75 mg/kg/day appears to be less effective than defer-
`oxamine; however, compliance is superior with deferiprone, which may partly
`compensate for this. Deferiprone has additive, or possibly synergistic, effects on
`iron excretion when combined with deferoxamine.
`The optimum dosage and long term efficacy of deferiprone, and its effects on
`survival and progression of iron-related organ damage, remain to be established.
`The most important adverse effects in deferiprone-treated patients are arthrop-
`athy and neutropenia/agranulocytosis. Other adverse events include gastrointes-
`tinal disturbances, ALT elevation, development of antinuclear antibodies and zinc
`deficiency. With deferiprone, adverse effects occur mostly in heavily iron-loaded
`patients, whereas with deferoxamine adverse effects occur predominantly when
`body iron burden is lower.
`Conclusion: Deferiprone is the most promising oral iron chelator under de-
`
`© Adis International Limited. All rights reserved.
`
`Drugs 1999 Sep; 58 (3)
`
`
`2 of 26
`
`Taro Pharmaceuticals, Ltd.
`Exhibit 1017
`
`

`

`Deferiprone in β-Thalassaemia: A Review
`
`555
`
`Pharmacodynamic
`Properties
`
`Pharmacokinetic
`Properties
`
`velopment at present. Further studies are required to determine the best way to
`use this new drug. Although it appears to be less effective than deferoxamine at
`the recommended dosage and there are concerns regarding its tolerability, it may
`nevertheless offer a therapeutic alternative in the management of patients unable
`or unwilling to receive the latter drug. Deferiprone also shows promise as an
`adjunct to deferoxamine therapy in patients with insufficient response and may
`prove useful as a maintenance treatment to interpose between treatments.
`Deferiprone is an oral bidentate iron chelator which binds to iron in a 3 : 1 ratio.
`It also binds other metals including aluminium, gallium, copper and zinc, but not
`calcium or magnesium.
`Deferiprone reduces body iron content in iron-overloaded animals and hu-
`mans. Iron excretion is related to dosage and the degree of iron overload, and
`occurs largely by the renal route. Deferiprone appears to mobilise iron from both
`reticuloendothelial and hepatocellular pools, from transferrin, ferritin and
`haemosiderin and from pathological iron deposits in intact red blood cells from
`patients with thalassaemia or sickle-cell anaemia.
`Depending on concentration, deferiprone has been reported to promote (at low
`concentrations, in vitro), and conversely to protect against (at high concentra-
`tions), oxidative damage caused by oxygen free radicals.
`As with deferoxamine, deferiprone inhibits proliferation of several cell lines
`in vitro and may induce apoptosis. It has also shown myelosuppressive effects in
`animals and humans. Although in vitro data suggest that deferiprone is markedly
`less toxic than deferoxamine to bone marrow myeloid progenitors, the clinical
`relevance of this is unclear, as deferiprone-induced myelosuppression may occur
`via a reactive metabolite-induced event mediated by the immune system.
`Peak plasma concentrations (Cmax) are reached within approximately 1 hour after
`oral administration of deferiprone. Food intake reduces the rate, but not the extent,
`of absorption of the drug. Administration of deferiprone 75 mg/kg/day at 12-
`hourly intervals produced a Cmax of 34.6 mg/L and area under the plasma con-
`centration-time curve (AUC) of 137.5 mg/L • h in patients with β-thalassaemia.
`Coadministration of iron (ferrous sulfate 600mg) reduced the AUC by about 20%
`in healthy volunteers.
`It is not clear whether deferiprone induces its own metabolism in vivo. This
`has been demonstrated in vitro. Trough plasma concentrations of deferiprone
`decreased during long term treatment with the drug in 1 study, but this was not
`corroborated by other studies.
`The volume of distribution after administration of deferiprone 75 mg/kg/day
`was 1.55 or 1.73 L/kg at steady state (depending on the dosage schedule) in
`patients with β-thalassaemia. Deferiprone was found to accumulate (≈3-fold) in
`thalassaemic, but not normal or sickle, red blood cells in vitro.
`Deferiprone is metabolised predominantly (>85%) to a glucuronide conjugate
`that lacks chelating properties. The drug, its conjugate and the deferiprone-iron
`complex are mainly excreted by the kidney and approximately 80% of a dose is
`recovered in the urine. Deferiprone is rapidly eliminated, with an elimination
`half-life (t1⁄2β) of approximately 1 to 2.5 hours in patients with β-thalassaemia.
`The t1⁄2β of deferiprone glucuronide was significantly correlated with creatinine
`clearance and this metabolite was found to accumulate in a patient with renal
`dysfunction. Although deferiprone is metabolised by the liver, the effects of he-
`patic impairment on the pharmacokinetics of the drug are yet to be determined.
`
`© Adis International Limited. All rights reserved.
`
`Drugs 1999 Sep; 58 (3)
`
`
`3 of 26
`
`Taro Pharmaceuticals, Ltd.
`Exhibit 1017
`
`

`

`556
`
`Barman Balfour & Foster
`
`Therapeutic Potential
`
`Tolerability
`
`Clinical studies, mostly in patients with β-thalassaemia, have demonstrated that
`deferiprone 75 to 100 mg/kg/day is capable of reducing iron burden in regularly
`transfused iron-overloaded patients. Factors affecting response to deferiprone
`appear to include the degree of iron overload and duration, dosage and degree of
`compliance with therapy.
`Serum ferritin levels (an indirect indicator of body iron load) are generally
`decreased in patients with very high pretreatment levels. In patients who are
`already adequately chelated at baseline, serum ferritin levels frequently remain
`stable. However, increases or inadequate decreases in serum ferritin and/or he-
`patic iron content were seen in some patients, especially after long term treatment.
`In some instances, this may reflect suboptimal dosage and/or severe degree of
`iron overload at baseline. Beneficial effects noted in some long term studies
`include lightening of the skin and decreased serum ALT and non-transferrin-
`bound iron levels.
`It should be noted that long term clinical trials reported to date have generally
`been noncomparative and conducted in small numbers of patients, who differed
`greatly with regard to baseline chelation status and underlying disease. Moreover,
`in many studies, the proportion of patients who were adequately chelated on
`deferiprone was not reported.
`Short term comparative studies have demonstrated that deferiprone ≤75
`mg/kg/day is less effective than deferoxamine in increasing iron excretion. How-
`ever, compliance during clinical use is superior with deferiprone, which may
`compensate for this to some degree. Few data from long term prospective ran-
`domised studies comparing deferiprone with deferoxamine have been reported.
`In these studies, deferiprone appeared to be slightly less effective than deferoxam-
`ine in reducing serum ferritin and less effective in controlling hepatic iron levels.
`Preliminary data suggest that deferiprone can be used successfully in combi-
`nation with deferoxamine, with additive or synergistic effects on urinary iron
`excretion and substantial reductions in serum ferritin levels being achieved.
`
`The most common adverse events in deferiprone-treated patients have been ar-
`thropathy (musculoskeletal stiffness and pain, accompanied by effusion in severe
`cases) and gastrointestinal disturbances (anorexia, nausea, vomiting). Arthropa-
`thy occurred in up to 39% of patients in clinical trials and generally resolves on
`dosage reduction or drug withdrawal.
`The most serious adverse effect associated with deferiprone is severe neu-
`tropenia/agranulocytosis (approximately 2% of patients each). This appears to be
`reversible.
`Other adverse events include elevated ALT and immunological abnormalities
`(development of antinuclear and antihistone antibodies). Deferiprone also pro-
`motes increased urinary excretion of zinc, particularly in patients with diabetes
`mellitus. This may occasionally lead to clinical signs of zinc deficiency (e.g.
`dry/itchy skin), which respond to zinc supplementation.
`Progression of existing liver fibrosis in 5 of a series of 14 patients treated with
`deferiprone was attributed to the drug, but this conclusion was subsequently
`questioned on the basis of methodological flaws in the study concerned. Long
`term follow-up of deferiprone-treated patients by other investigators implicates
`chronic hepatitis C infection and iron overload, rather than deferiprone, in pro-
`gression of hepatic fibrosis in transfusional iron-overloaded patients.
`
`© Adis International Limited. All rights reserved.
`
`Drugs 1999 Sep; 58 (3)
`
`
`4 of 26
`
`Taro Pharmaceuticals, Ltd.
`Exhibit 1017
`
`

`

`Deferiprone in β-Thalassaemia: A Review
`
`557
`
`Dosage and
`Administration
`
`The recommended dosage of deferiprone is 25 mg/kg 3 times daily, although
`some investigators recommend use of dosages up to 100 mg/kg/day and/or twice
`daily administration. Special monitoring is required in all patients. Particular
`caution is recommended (with monitoring of renal or hepatic function) when
`treating patients with impaired renal or hepatic function. Deferiprone is contra-
`indicated in patients with neutropenia or a history of agranulocytosis or recurrent
`episodes of neutropenia, those taking drugs known to cause neutropenia, and in
`pregnant or lactating women. Women of childbearing potential should use con-
`traceptives while taking deferiprone. Weekly monitoring of neutrophil count is
`recommended and patients should be advised to report immediately any symp-
`toms of infection, such as fever, sore throat or flu-like symptoms.
`Deferiprone may interact with concomitantly administered medications con-
`taining metallic cations, including aluminium-based antacids.
`
`1. Introduction
`
`Thalassaemia major is one of the most common
`worldwide causes of iron overload. The thalas-
`saemias are a heterogeneous group of hereditary
`haemolytic anaemias characterised by inadequate
`synthesis of one or more haemoglobin polypeptide
`chains. They are classified according to which chain
`is involved (α, β or δ). In β-thalassaemia, de-
`creased β-chain production leads to a relative ex-
`cess of α-chains, which are unstable.
`Patients with β-thalassaemia major (or Cooley’s
`anaemia; the most severe form of the disease) usu-
`ally manifest symptoms at 4 to 6 months of age,
`developing severe anaemia with a haematocrit of
`<20%. They require regular blood transfusions in
`order to avoid the complications of anaemia and
`ineffective erythropoiesis, allow normal growth
`and development and prevent early death. How-
`ever, iron from the transfused blood accumulates
`in the reticuloendothelial system and parenchymal
`cells.
`Iron toxicity begins when the load of iron in a
`particular tissue exceeds the binding capacity of
`ferritin in the cell and of transferrin in the plasma.
`This results in accumulation of free or non-trans-
`ferrin-bound iron (NTBI), which can participate in
`free radical formation, leading to peroxidation and
`damage in various tissues. As reviewed by Olivieri
`and Brittenham[1] the degree of clinical iron toxic-
`ity in iron-overloaded patients is determined by:
`• the amount of excess iron
`
`• the rate of iron accumulation
`• the duration of exposure to elevated iron levels
`• the distribution of iron between highly hazard-
`ous and less hazardous body sites
`• ascorbate status (which influences body distri-
`bution of iron)
`• presence of viral hepatitis
`• alcohol intake.
`The sequelae of iron overload include hepatic
`fibrosis and cirrhosis, multiple endocrinopathies
`(diabetes mellitus, hypogonadism, hypopara-
`thyroidism, hypothyroidism), immunological dys-
`function, growth and bone abnormalities, short
`stature, cardiac disease (congestive heart failure,
`arrhythmias), pulmonary dysfunction and hyper-
`pigmentation of the skin. Progressive organ dys-
`function, affecting the heart, liver and endocrine
`system in particular, ultimately leads to death in the
`second or third decade of life if left untreated.
`A high proportion of regularly transfused tha-
`lassaemic patients have hepatitis C acquired via
`transfused blood, whereas vaccination has drasti-
`cally decreased the incidence of hepatitis B in this
`population.
`The management of thalassaemia and iron over-
`load has been reviewed by other authors.[1-6]
`The iron-chelating agent deferoxamine has been
`available for over 2 decades. Effective chelation
`with this agent maintains body iron stores in pa-
`tients with thalassaemia major at 5 to 10 times the
`levels found in healthy individuals.[7] Neverthe-
`less, with this level of control of body iron, patients
`
`© Adis International Limited. All rights reserved.
`
`Drugs 1999 Sep; 58 (3)
`
`
`5 of 26
`
`Taro Pharmaceuticals, Ltd.
`Exhibit 1017
`
`

`

`558
`
`Barman Balfour & Foster
`
`are protected against impaired glucose tolerance,
`diabetes mellitus, cardiac disease and early death.
`However, the high cost of deferoxamine pre-
`cludes its use in many parts of the world where
`β-thalassaemia is endemic (e.g. Asia). Addition-
`ally, many patients, particularly adolescents, fail to
`comply adequately with deferoxamine therapy be-
`cause of the onerous administration schedule (sub-
`cutaneous infusion over 8 to 24 hours 3 to 7 times
`weekly or intramuscular injection daily plus intra-
`venous infusion at the time of blood transfu-
`sion[8,9]). This noncompliance can allow iron-re-
`lated complications to develop. Serious adverse
`effects of deferoxamine include sensorineural
`hearing loss, retinal/visual damage, abnormal car-
`tilage metabolism and stunted linear growth. Other
`adverse effects include increased susceptibility to
`Yersinia, and more rarely other, infections. The risk
`of adverse effects increases when serum ferritin
`levels fall below 1000 μg/L (i.e. when body iron
`burden is only very modestly elevated or becomes
`normalised).[1] Deferoxamine-associated toxicity
`is less likely to occur at dosages ≤40 mg/kg/day in
`patients with thalassaemia (the recommended sub-
`cutaneous dose is 20 to 40 mg/kg/day in the US[8]
`and 20 to 50 mg/kg in the UK[9]).
`Diethylenetetra-penta-acetic acid (DTPA), an-
`other hexadentate chelator, is also given by slow
`subcutaneous infusion. However, it also chelates
`zinc avidly, which causes difficulties in maintain-
`ing adequate levels of zinc during long term ther-
`apy. Nevertheless, it provides a possible alternative
`to deferoxamine in those who cannot tolerate the
`latter drug.[3,10]
`Bone marrow transplantation provides the pos-
`sibility of radical cure in the minority of eligible
`patients for whom a suitable donor is available and
`facilities and funds allow.
`Thus, there is a considerable need to develop an
`oral iron chelator. Deferiprone is the best studied
`oral chelator to date. It is a bidentate iron ligand
`(i.e. it has 2 binding sites per molecule for iron,
`whereas deferoxamine has 6; fig. 1). The remainder
`of this review focuses on its clinical potential in the
`management of iron overload in β-thalassaemia.
`
`O
`
`N
`
`CH3
`
`OH
`
`CH3
`
`Deferiprone (molecular weight 139)
`
`CONH
`(CH2)2
`
`H2N
`
`(CH2)5
`N C
`
`OH O
`
`(CH2)5
`N C
`
`OH O
`
`CONH
`(CH2)2
`
`CH3
`
`(CH2)5
`N C
`
`OH O
`
`Deferoxamine (molecular weight 656)
`
`Fig. 1. Comparison of structural formulae of deferiprone and
`deferoxamine.
`
`Deferiprone has also been studied in small numbers
`of iron-overloaded patients with other transfusion-
`dependent diseases,
`including myelodysplastic
`syndrome, and sickle cell or Blackfan-Diamond
`anaemia and relevant data are included in this re-
`view.
`
`2. Pharmacodynamic Properties
`
`2.1 Binding of Iron and Other Metals
`
`Deferiprone forms strong, water-soluble 3:1 com-
`plexes with the Fe+++ ion, with a binding constant
`of 37, which is markedly higher than that of
`deferoxamine.[11] It also binds aluminium,[12-16]
`gallium,[17] copper[16] and zinc, but not calcium or
`magnesium.[11] Although deferiprone appears to
`have a low affinity for zinc in vitro,[16] zinc defi-
`ciency has been reported in a small number of iron-
`overloaded patients receiving long term deferi-
`prone treatment (see section 5.5). Lead, aluminium
`and copper, but not magnesium, calcium or man-
`ganese, compete with iron for binding to deferi-
`prone in vitro.[16]
`
`© Adis International Limited. All rights reserved.
`
`Drugs 1999 Sep; 58 (3)
`
`
`6 of 26
`
`Taro Pharmaceuticals, Ltd.
`Exhibit 1017
`
`

`

`Deferiprone in β-Thalassaemia: A Review
`
`559
`
`patients treated immediately post-transfusion,
`when haemoglobin levels are relatively high.[29]
`Other investigators found that urinary iron excre-
`tion in patients with thalassaemia was indepen-
`dently affected both by steady-state trough plasma
`concentrations[30] and AUC[31] of deferiprone.
`The amount of iron excretion promoted by
`deferiprone is also related to iron load,[20,26,32] and
`the drug does not appear to chelate iron apprecia-
`bly in healthy volunteers who are not iron-over-
`loaded.[20,32]
`The therapeutic potential of deferiprone in in-
`ducing iron excretion in iron-overloaded patients
`is discussed in section 4.
`
`2.1.2 Source of Iron Chelated by Deferiprone
`Studies using radiolabelled iron in hyper-
`transfused rats indicated that deferiprone mobi-
`lises iron from both reticuloendothelial and
`hepatocellular pools. Iron from the reticuloendo-
`thelial pool was excreted in urine and bile, whereas
`all of the iron originating from the hepatic pool was
`excreted in the bile.[33] Similarly, when rats with
`normal iron stores were fed radiolabelled 3,5,5-
`trimethylhexanoyl-ferrocene (which delivers iron
`predominantly to hepatocytes), most of the iron re-
`
`0
`
`75
`50
`25
`Deferiprone dosage (mg/kg/day)
`
`100
`
`50
`
`40
`
`30
`
`20
`
`10
`
`0
`
`Mean urinary iron excretion (mg/24h)
`
`Fig. 2. Dose-related urinary excretion of iron in deferiprone-
`treated patients. All patients (n = 52) in this crossover study had
`thalassaemia major and iron overload and received deferiprone
`in 3 divided daily doses (duration of treatment was not
`stated).[27]
`
`2.1.1 Promotion of Iron Excretion In Vivo
`Orally administered deferiprone lowers body
`iron levels in animals and humans. Excretion of
`iron occurs mainly by the renal route.[18]
`When the 24-hour urinary iron excretion was
`divided by the amount of iron that a given oral dose
`was capable of binding, the chelation efficiency of
`deferiprone in iron-overloaded patients was calcu-
`lated to be 3.8[19] or 6.8%.[20] The iron-binding ef-
`ficiency of deferiprone was lower in animal stud-
`ies, ranging from 1.3 to 2.1% in iron-overloaded
`monkeys and rats.[21,22]
`intra-
`In rats with acute iron intoxication,
`peritoneal administration of deferiprone (400 mg/
`kg 15 minutes after iron administration, followed
`by 1 dose of 200 mg/kg and 2 doses of 100 mg/kg
`at hourly intervals) decreased mortality (from 59
`to 15% of animals at 14 hours; p = 0.013).[23] How-
`ever, a similar study found that, although concom-
`itant administration of deferiprone with a lethal
`dose of iron protected rats against iron toxicity,
`administration of deferiprone (≤9 mmol/kg) 10 to
`15 minutes after the iron dose appeared to increase
`mortality (p = 0.009).[24] For a review of preclinical
`studies with deferiprone see Al-Refaie and Hoff-
`brand.[18]
`In iron-overloaded patients, deferiprone in-
`creases urinary excretion of iron in a dose-depend-
`ent manner,[25-27] although increases in urinary iron
`excretion with increased dosage are more than pro-
`portional. Mean daily excretion ranged from 6.2
`mg/day at a dosage of 25 mg/day to 42.3 mg/day
`at a dosage of 100 mg/day (fig. 2).[27]
`More frequent administration of deferiprone,
`resulting in more sustained plasma drug concentra-
`tions but very similar 24-hour area under the
`plasma concentration-time curve (AUC24), achie-
`ved better chelation with the same total daily dos-
`age. Urinary iron excretion in 10 patients with β-
`thalassaemia was 0.59 mg/kg/day during 6-hourly,
`compared with 0.40 mg/kg/day (p = 0.0129) during
`12-hourly, administration of deferiprone 75
`mg/kg/day (for 3 days).[28] It was subsequently
`suggested that sustained plasma concentrations of
`deferiprone may achieve better iron excretion in
`
`© Adis International Limited. All rights reserved.
`
`Drugs 1999 Sep; 58 (3)
`
`
`7 of 26
`
`Taro Pharmaceuticals, Ltd.
`Exhibit 1017
`
`

`

`560
`
`Barman Balfour & Foster
`
`moved by deferiprone (200 mg/kg/day) was found
`in the faeces.[34]
`Deferiprone chelates iron bound by transferrin
`(iron transport protein),[20,35] ferritin (iron storage
`protein)[36] and haemosiderin (partly degraded fer-
`ritin),[36] as well as free iron. It has been calculated
`that up to 20% of iron excreted in the urine after a
`single dose of deferiprone may originate from iron
`bound to transferrin. Deferiprone was more effec-
`tive than deferoxamine in mobilising iron from
`transferrin and ferritin in vitro.[35]
`Deferiprone, but not deferoxamine, was able to
`chelate pathological iron deposits from intact red
`blood cells (RBCs) from patients with thalassae-
`mia or sickle-cell anaemia, both in vitro and in vivo.
`No free iron was detected in thalassaemic RBCs
`after 24 hours’ incubation with deferiprone 0.5
`mmol/L. After administration of deferiprone 50 or
`70 mg/kg/day for 2 weeks to thalassaemic patients,
`66 and 78% of free iron was removed from RBC
`membranes.[37] In 11 patients on long term therapy,
`deferiprone removed membrane free iron from β-
`thalassaemic erythrocytes with a reduction in KCl
`cotransport activity.[38] These findings may reflect
`the enhanced ability of deferiprone (compared with
`deferoxamine) to permeate cell membranes. Bi-
`dentate ligands are able to penetrate cells more ef-
`ficiently than hexadentate chelators on account of
`their lower molecular weight.[39]
`
`2.2 Oxidant/Antioxidant Effects
`
`Liberation of iron in vivo can catalyse formation
`of hydroxyl radicals, which are highly reactive and
`toxic to tissues.[40] Depending on concentration,
`deferiprone has been reported to promote (at low
`concentrations in vitro), and to protect against (at
`high concentrations), oxidative damage caused by
`oxygen free radicals.
`Binding of 1 atom of iron (which has 6 coordi-
`nation sites) requires 3 molecules of deferiprone
`but only 1 molecule of deferoxamine. Moreover,
`the chelates formed by deferiprone have relatively
`low stability.[41] At low deferiprone concentrations
`(relative to available iron concentrations), partially
`bound iron species (bound to only 1 or 2 deferi-
`
`prone molecules) can be formed. In some in vitro
`systems, devoid of substrates capable of ‘mopping
`up’ free radicals, the unoccupied coordination sites
`of these complexes can catalyse formation of hy-
`droxyl radical or other reactive oxygen species.[41]
`Exposure to deferiprone 1 mmol/L for 1 hour
`before exposure to hydrogen peroxide markedly
`potentiated hydrogen peroxide-mediated oxidative
`DNA damage in iron-loaded hepatic (HepG2) cells
`in vitro.[42] Deferoxamine 0.33 mmol/L had no ef-
`fect. However, when deferiprone exposure was
`maintained throughout hydrogen peroxide expo-
`sure, deferiprone protected against oxidative dam-
`age. Thus, the influence of the drug on iron-cata-
`lysed free radical formation may depend on the
`intracellular deferiprone : iron concentration ratio.
`In vitro studies investigating the effects of deferi-
`prone concentrations on iron-mediated ascorbate
`oxidation and deoxyribose degradation indicate
`that the deferiprone:iron ratio must be at least 3 : 1 to
`inhibit free radical generation. At lower deferi-
`prone concentrations, free radical generation is in-
`creased.[42]
`These findings may serve as an indication of the
`theoretical potential for generating reactive oxygen
`species in the presence of low tissue concentrations
`of deferiprone, but the nonphysiological condi-
`tions and concentrations employed in these studies
`should be taken into account.
`Deferiprone prevented oxidation of human low-
`density lipoproteins (LDL) in vitro, with complete
`inhibition at a concentration of 50 μmol/L, and at
`a concentration of >50 μmol/L protected human
`umbilical vein endothelial cells against the cyto-
`toxic action of oxidised LDL (p < 0.001 vs un-
`treated control). These effects were concentration-
`dependent.[43] It has been suggested that oxidation
`of lipoproteins may play an important role in ath-
`erogenesis. Administration of deferiprone 100 mg/
`kg/day to rabbits fed a high-cholesterol diet signif-
`icantly decreased aortic cholesterol accumulation
`(by 72% vs untreated controls; p < 0.0001).[43]
`Deferiprone 50 μmol/L attenuated methaemo-
`globin formation by 25% (p < 0.0001 vs untreated
`control) in β-thala

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket