Int J Pharm Pharm Sci, Vol 7, Issue 8, 35-44Review Article


THERAPEUTIC ADVANCEMENTS IN MANAGEMENT OF IRON OVERLOAD–A REVIEW

PRIYANKA TYAGIa, YOUGESH KUMARc, DIKSHI GUPTAa, HARPAL SINGHa AMIT KUMARb*

aCentre for Biomedical Engineering, Indian Institute of Technology, New Delhi, P. O. Box 110016, India, bInstitute of Nuclear Medicine and Allied Sciences, DRDO, New Delhi, P. O. Box 110054, India, cDepartment of Zoology, DAV (PG) College, Muzaffarnagar, C. C. S. University, Meerut, U. P
Email: amityagi@gmail.com

Received: 21 May 2015 Revised and Accepted: 11 Jul 2015


ABSTRACT

Removal of excess of free iron, produced under certain physiological conditions in human body, has always been of great concern for therapists worldwide. A number of naturally occurring and synthetic ligands are being explored for maximal iron decorporation, enhanced patient compliance and minimal side effects. Siderophores are the most important class of currently used chelators and few of them are approved for treating iron overload. This review focuses on the developments taking place in the field of iron overload treatment to find the most novel, efficient and safest iron decorporating agent. The importance of many natural and synthetic iron scavengers, having potential for clinical application in the treatment of iron overload and their associated side effects, are discussed along with the currently followed methods. Special emphasis is given to review the status of new ligands, combination therapy and methodologies adopted for overcoming limitations of existing therapeutic agents for treatment of different iron overload disorders.

Keywords: Iron chelation, Iron overload, Phlebotomy, Siderophores, Deferasirox.


INTRODUCTION

Iron is an essential and most abundant transition metal in human body. Iron overload is a condition of excess iron stores in different parts of body. Since very little iron is excreted from the body normally, it becomes a potentially toxic heavy metal in case of accidental ingestion, transfusional siderosis (e. g. β-thalassemia, sickle cell disease) or hereditary conditions (e. g. hereditary hemochromatosis) leading to the production of destructive free radicals [1, 2] and life threatening complications such as cirrhosis, hepatocellular cancer, diabetes and heart diseases [3]. According to the Institute of Medicine of the National Academy of Sciences, Washington, the tolerable upper level of iron intake for healthy adult males and females is 45 mg/day [4].

Transferrin (Tf) (fig. 1) is a freely circulating natural iron chelator in human blood responsible for uptake of free iron after intestinal absorption and storing it as intracellular ferritin or hemosiderin. In normal humans, Tf is only 25-30% saturated with iron while in iron-overload it becomes completely saturated, leading to the appearance of non transferrin bound iron (NTBI) in blood and tissues.

Fig. 1: Structure of protein transferrin (http://en.wikipedia.org/wiki/Transferrin)

Efforts are being made since long time for effective diagnosis and management of free iron in human body. Phlebotomy and chelation therapy, either alone or in combination, are currently the most widely used clinical practices to reduce iron overload. Lot of work is being done all over the world to find the most suitable therapeutic agent having enhanced efficacy and reduced toxicity so as to overcome the limitations of commercially available iron decorporating agents.

Iron toxicity

The overburden of iron in body leading to the production of NTBI and tissue iron are commonly called as ‘free iron’. Free iron (Fe+3) in body acts as antioxidant because of its ability to gain and loose electrons. The highly reactive free radicals generated by free iron could damage vital organs and tissues of body (fig. 2), by reacting with surrounding biomolecules [5-7], as indicated by increase in biomarkers in iron overloaded thalassemia and sickle cell patients [8].

Fig. 2: Effect of oxidative stress on vital organs of human due to free blood circulating and tissue iron

These oxygen free radicals may be utilized by cancer cell for growth and indefinite division [9]. Iron toxicity is also associated to lipid peroxidation [10], increased plasma histamine and serotonin in patients suffering from hemochromatosis, sickle cell disease, thalassemia, myelodysplastic syndrome, African iron overload, Atransferrinemia, sideroblastic anemia etc.

Classical method of iron decorporation-phlebotomy

Decorporation of any toxic substance from human body has been an important practice since the ancient times. The ancient medical practitioners used to remove the blood borne toxins by removal of excess blood from patient’s body [11]. This clinical process of drawing large volume of blood from patient’s body is called phlebotomy. It is the simplest, safest and most inexpensive treatment for excess iron decorporation [12]. The procedure involves venipuncture in which blood is removed from veins of patient’s arm, through a narrow bore needle, for a period of approximately 30-45 minutes. Before initiating the procedure, the practitioner must be aware of the patient’s serum ferritin levels (men ≥ 300 µg/l & women ≥ 200 µg/l).

Studies have suggested that decrease in serum ferritin level is the most reliable indicator for monitoring the progress of therapeutic phlebotomy instead of transferrin saturation or serum iron [13]. Initially the procedure is done till the serum ferritin levels reach 20% of the initial value after which the therapy should be done at regular intervals [14]. Therapy should be discontinued when serum ferritin level drops below 20-50μg/l. Twice weekly phlebotomies may be suggested for heavily iron-loaded patients. Phlebotomy procedure was found to mobilize iron from body stores when patients have iron level>1000 µg/ml were treated every 1-2 week with 500 ml of blood removed during each session depending on body weight [15]. As compared to women, men usually require twice the phlebotomy units to induce iron depletion [14].

Phlebotomy has been found beneficial in case of iron overload problems occurring due to periodic blood transfusions such as β-thalassemia and sickle cell anemia [16, 17]. Prolonged phlebotomy at a rate of 6-ml/kg-body weight, for several years in β-thalassemia patients, was found to be highly effective to mobilize excess iron from 48 patients under study [18]. Russell and coworkers used a combination of hydroxyurea treatment and phlebotomy to increase fetal hemoglobin and reduce the iron overload and observed decreased serum ferritin from average level of 3134 ng/ml to 617 ng/ml as well as non recurrence of strokes in sickle-cell affected children [19]. They suggested the combined treatment to be a better option rather than commonly used siderophores that suffers numerous side effects. A team of Chinese workers also found considerable decrease in serum ferritin levels among phlebotomy treated patient groups [20]. Bone marrow transplantation is adopted as a treatment procedure to cure thalassemia [21] and phlebotomy performed in such cases has helped patients to get relieved from the iron overload that persisted in their bodies for many years after treatment. Erythropoetin-assisted phlebotomy was found beneficial for thalassemia patients suffering from iron overload due to allogenic hematopoetic stem cell transplantation [22-24].

Therapeutic phlebotomy is the preferred treatment for hereditary hemochromatosis patients. Iron overload in such patients was also found to be associated with hyperlipidemia [25] as a result of mutation in HFE gene [26] and phlebotomy procedure was found to be highly effective for such patients [27]. Cardiac function was greatly improved in patients with primary hemochromatosis by undergoing aggressive phlebotomy [28-31]. Repeated phlebotomy has shown a drastic decrease in spleen ferritin level in iron overloaded rat models [32]. Eleven patients with chronic liver disease and associated iron overload showed an improvement in cytolysis when 300 ml of blood was removed from their bodies for five consecutive weeks [33]. Therapeutic phlebotomy showed resolution in some observed complications of iron overload such as elevated hepatic serum aminotransferases [34], hepatomegaly [35], cutaneous hyperpigmentation [36], hyperferritinemia [37] and rarely improves certain others such as diabetes mellitus [38, 39], cardiomyopathy [40]and refractory arrhythmia [41]. Regular therapeutic phlebotomy has resulted in an increase in longevity as well as quality of life [42]. Regular phlebotomy is suggested as the ultimate treatment for patients with end-organ damage due to iron overload [43].

However iron overload resulted hepatic cirrhosis cannot be resolved by therapeutic phlebotomy leaving the patients at higher risk for liver cancer [36]. Disadvantages associated with the procedure involve exacerbation of joint pains, difficulty to operate on young children, difficult methodology, excessive pain and non-compliance (Fig.3). Several physiological and psychological therapies are given for pain relief to patients undergoing therapeutic phlebotomy such as hot and cold packs, distraction and relaxation [44]. According to the French Superior Health Security, at-home phlebotomy should be recommended for hereditary hemochromatosis patients to reduce health care expenditures and improve patient compliance [45].

Current methods of iron decorporation

Repeated phlebotomy is often required to treat iron overload in patients with thalassemia [46] that leads to incompliance and extreme pain. Also the therapy cannot be given to patients who are hemodynamically unstable and suffer congestion problems related to heart [47]. Therefore drug therapy is commonly prescribed to such patients. Currently drug therapy involving chelation of free iron from blood has become a routine practice for treating iron overloaded patients. The best way to mobilize a toxic metal out of the body is to chelate it with a specific ligand and wash off the stable ligand-metal complex through excretory system of body. Such ligands are mostly organic molecules called as chelating agents, chelators or sequestering agents that have a very high specificity for certain metal ions. Till now researchers have suggested a number of natural and commercially prepared ligands that can be used as iron chelators (fig. 4).

Fig. 3: Advantages and side effects of Phlebotomy procedure for iron decorporation


Fig. 4: Classification of chelators commonly used in iron chelation therapy

Natural chelators

The naturally occurring chemicals in plants are called phytochemicals. Many of these phytochemicals have been used over a long time and found to bind iron. Therefore workers are trying to explore certain natural agents as iron chelators in humans.

Phytochemicals

Phytic acid (InsP6, myo-Inositol hexaphosphate) which is the storage component of most cereal grains and seeds [48] has been found to inhibit free radical formation in a ferric ion standard assay and was effective similar to the commonly used chelator Deferoxamine mesylate [49]. Spinach, beans and nuts have significant levels of oxalic acid that inhibits iron absorption. An administration of wheatgrass juice of a week old leaves daily for six months reduced the iron burden in myelodysplastic syndrome patients [50]. The catechins (flavonoids) such as epigallocatechin gallate and epicatechin gallate are pharmacologically active components of green tea (suggested by many natural medicine practitioners) that form a complex with ferric ions and aid in the chelation of NTBI [51] as well as penetrate the brain barrier and remove the abnormally accumulated iron on dying neurons in certain neurodegenerative disorders such as Parkinson’s disease [52]. The flavonoids and phenolic compounds being the active components of plant extracts have been suggested to possess iron-chelating properties [53, 54] by acting as the free radical terminator [55]. Mohammad Ali and coworkers prepared extract from a series of plant parts and determined the amount and chelation properties of phenols and flavonoids present in them [56]. They found that plants showing highest content of these compounds also showed the highest iron chelating properties. Silybin (200 mg/kg), a flavonoid derived from the herb milk thistle (Silybum marianum), has been found to reduce elevated iron levels in kidney [57-59]. The iron chelating ability of curcumin and its diacetyl derivatives was found to lie in the enolic form of β-diketo moiety [60]. Later the orally available glycosyl derivatives of curcumin were prepared by reducing the molecular weight and hydrophilicity for enhanced absorption through small intestine [61]. Though a number of Ayurveda preparations are prescribed for treating iron overload yet none of them is currently approved for routine clinical application.

Siderophores

Siderophores are low molecular weight (500-1000Da) iron sequestering agents produced by certain microbes in response to the utilization of insoluble form of iron present in soil [62, 63]. They exhibit strong specificity towards ferric iron. The research in this field began around six decades ago. The first siderophores to be chemically identified were amino acid conjugates of 2,3-dihydroxybenzoic acid i. e itoic acid (2,3-dihydroxybenzoylglycine) [64]. Since then more than 500 different types of siderophores have been isolated and characterized [65, 66]. Siderophores play an important role not only to treat iron overload [67,68] but also to keep balance in the dietary iron intake and excretion in body fluids [69, 70]. Siderophores used in human medicine are classified according to the chemical nature of their metal binding functionalities as under:

Hydoxamate based siderophores

These hydroxamic acid containing siderophores are highly hydrophilic and can permeate membranes by forming neutral tris complex with ferric ion [71, 72]. More and more new compounds are being explored under this category due to their good iron chelating properties. Hydroxamates are poorly absorbed through the gastrointestinal tract and are therefore delivered through the parenteral route [73, 74]. Typical tri-hydroxamate siderophores are ferrichromes, fusarinines and ferrioxamines out of which only ferrioxamines are widely used for treating iron overload associated with a number of diseases in humans. It comprises of a group of trihydroxamate siderophores produced by actinomycetes [75]. Since 60 years the most common and effective therapeutic agent belonging to this group is desferrioxamine (DFO), which is an FDA approved hexadentate chelator for treatment of iron toxicity in human subjects [76-78] and is commercially marketed in the form of its methane-sulfonate salt. Before DFO, DTPA and dimerceprol were used [79, 80] and DTPA was found less effective as compared to DFO [81].

DFO is a colourless crystalline substance produced by the fungi Streptomyces pilosus [82]. It is impermeable through cell membrane and is taken up by cells through endocytosis for targeting primarily the excess iron present in liver and spleen. DFO has always been the drug of choice for iron decorporation due to its capacity to chelate iron effectively from ferritin and slowly from transferrin also [83-85]. A combination of DFO with pyrophosphate was suggested to mobilize iron from transferrin which otherwise was not possible with DFO alone [86,87]. DFO is effective not only in treating NTBI but also chelates hepatocellular iron for removal through urine and bile [88]. Free DFO has a very limited stability in plasma and gets metabolized even at very low temperature storage. In-vitro radioactive studies have confirmed that ferrioxamine complex formed after binding of DFO to iron stabilizes the unbound form of the drug in patient’s plasma samples [89].

In the late 1960s it was realized that intramuscular DFO removed only limited amount of iron and hence ascorbic acid should be administered along with the chelator [90, 91]. Smaller doses of ascorbate (100-200 mg/day) given along with DFO chelated iron through urine as well as faeces [92]. DFO therapy was also found beneficial in improving cardiac function of patients suffering from iron overload disorders due to severe congestive cardiomyopathy [93]. Long-term DFO therapy is capable to prevent the severe complications of transfusional iron overload and to improve significantly the life expectancy of thalassemia patients [94-96].

However DFO is not orally available and has a short blood circulation time (5-10 minutes) [97] due to which it is administered as painful continuous infusion for 8-12 h atleast 5 days/week [98, 99] and hence is poorly patient compliant. DFO therapy is expensive [100] with less than 10 mg iron removal in a single session. According to a recent FDA report, DFO treatment in 196 patients in age group of 10-59 years showed serious side-effects such as renal failure, gastritis, anaemia, decreased folate concentration and 6% of them even suffered the persisting iron overload (http://www. ehealthme. com). The chelator has been found unsuccessful in removing small increases in iron in rheumatoid arthritis patients [101]. Still DFO therapy is the choice of prescription for iron-overloaded patients today in view of it being FDA approved and currently most worked ligand.

Catecholate and phenolate based siderophores

Catechol-based ligands are the most powerful known siderophores [102]. Enterobactin is a cyclic tricatecholate siderophore capable of rapid removal of iron from human transferrin [103] as well as possess greater in-vitro affinity for iron than transferrin [104]. But catechol subunits are somewhat air-sensitive [105] and therefore require protection by methyl, benzyl or acetyl groups. Its trimester backbone gets easily hydrolyzed at physiological pH and also has very low aqueous solubility [106]. Phenolic groups play a crucial role in iron chelation [107]. Cranberries are the richest source of total phenolic content among fruits [108] and quercitin present in cranberries is found to possess strong iron-binding properties in physiological conditions due to the presence of three iron-binding motifs [109]. Desferriothicin, a tridentate chelator of ferriothicin discovered in 1980s, is a subgroup of the ortho substituted phenolate class of iron chelators. It was the first orally available chelator capable of efficiently complexing with FeIII in 2:1 ratio in both iron overloaded rodent and primate models. Desferriothicin was however found to be highly nephrotoxic and therefore research is being done to prepare its analogs with similar oral activity but reduced toxicity. A few analogs prepared include Deferitrin, which underwent phase II clinical trials but was found to show severe renal toxicity [110]. FBS0701, an orally available member of the desazadesferrithiocin class of siderophore related tridentate chelators has demonstrated equal iron chelation ability to deferasirox and appears less toxic and is currently under phase II trial [111]. R-apomorphine used in late stage Parkinson disease therapy shows strong iron chelation and antioxidant properties due to its catechol structure [112].

Carboxylates and salicylates

This class of sideophores contains hydroxyl and carboxyl functional groups as donor for iron chelation. Rhizoferrin was found to form 1:1 iron complex as DFO and parabactin [113]. A series of aminocarboxylate chelators designed with a pendant aromatic group resulted in strong iron chelation as well as potent protective effect against oxidative damage [114].

Peptidic siderophores

In view of the oral inactivity of DFO, pyoverdin type of peptidic siderophores isolated from Pseudomonas species were reported to be insensitive to proteases and thus suggested as oral chelators for iron overload disorders [115,116]. The two pyoverdins-PyA (from Pseudomonas aeruginosa ATCC15692) and PyF (from Pseudomonas fluorescens CCM2798) were found to be as effective as DFO in iron-loaded rat hepatocyte cultures due to inhibited release of aspartate aminotransferase and lactate dehydrogenase enzymes. Both chelators decreased the intracellular iron level and increased the concentration of the metal in the culture medium. Later two bidentate chelators-hydroxypyridin-4-ones (CP20 and CP94) were also demonstrated as orally active chelators in iron overloaded animals and were found to be equally effective as the subcutaneous DFO [117]. However a recent comparison of efficacy of pyoverdin, pyochelin and DFO to decorporate iron from ferritin showed DFO to be most effective in iron mobilization [118] and inhibition of lipid peroxidation [119] followed by pyoverdin and pyochelins.

Synthetic chelators

Owing to the expense and inconvenience associated with commonly used chelators such as deferoxamine or ferrichromes, a search for new and effective oral iron chelators has given rise to many synthetically produced compounds that possess improved pharmacological properties. Synthetic siderophores play an important role not only in our understanding of how natural siderophores function but also in the development of new improved therapeutic agents [120, 121]. Such chelators require easy and cheap production and must possess high specificity and high binding constant [122, 123] for specific metal ions. In 1960s, ferrioxamineB was the first siderophore to be synthesized to elucidate its chemical structure. Baret and coworkers proposed a series of tripodal iron sequestering agents based on o, o’-dihydroxybiaryl that were water soluble (due to sulfonation) and possessed well hydrolytic and oxidation stability [124]. The complexation of these agents with Fe+3in water showed similar results as the hydroxamate siderophores, but were poorer than the catecholate siderophores proposed by Raymond and coworkers.

Hydroxypyridinones

Hydroxypyridinones are bidentate, cyclic hydroxamic acid containing oral iron chelating ligands and currently the most important class of chelators being explored. In the late 20th century it was suggested that highly hydrophilic polydentate compounds are mostly orally bioavailable and can act as effective chelators [125]. In this respect a series of hydrophilic ligands based on kojic acid and substituted linkers were prepared (L4-L8) and strongest iron chelating activity was observed with L8 expected to remove iron from Tf in further studies [126]. On the other hand lipophilic hydroxypyridinone ‘CP094’ was prepared that readily suppressed NTBI as well as process of lipid peroxidation due to iron overload in sickle cells as compared to DFO [127]. Clinical studies on tridentate oral ligand pyridoxal isonicotinoyl hydrazone (PIH) suggested its easy permeation through cell membrane [128, 129] and highest selectivity for Fe III at lower doses [130]. A novel group of hybrid and less toxic oral chelators, the 2-pyridylcarboxaldehyde isonicotinoyl hydrazone (PCIH) analogues and pyridylcarboxaldehyde 2-thiophenecarboxyl hydrazone (PCTH) were found two times more effective than DFO or PIH in Tf iron removal from primary myocardiocyte cultures and mice [131, 132]. A range of 3-hydroxypyridin-4-ones (CP102, CP117, CP41) were designed to target the hepatocellular low molecular weight iron pool and were found to mobilize completely the intrahepatic chelatable iron pool in rats [133]. Two lipo-hydrophilic hexadentate 3,4-hydroxypyridinone based compounds were prepared by attaching variable length alkyl-amino arms to tricarboxylic molecular scaffold of KEMP through amide linkage [134] so as to completely wrap free iron and form a stable complex [135]. Many researchers have reviewed the clinical applications and design of hydroxypyridinones for use in iron overload [136, 137]. N, N-bis(2-hydroxybenzyl) ethylenediamine-N,N-diacetic acid (HBED) is a hexadentate synthetic aminocarboxylate ligand that binds to iron in 1:1 ratio [138, 139]. The small size and lipophilicity of phenolic groups maintained ferric iron in redox inert ferrous form thus reducing the free radical toxicity [140]. Galey and coworkers synthesized pro-drugs which got activated by reactive oxygen species into ones with high iron chelation ability and found them effective against in-vitro oxidative injury [141]. The monoethyl and dimethyl ester [142] containing pro-drugs of HBED have been reported to possess great oral bioavailability. However the toxicity of the molecules was observed due to presence of two coordinating N-atoms rendering the molecules to exhibit high affinity for ZnII ions. In order to increase iron chelation ability of 1-hydroxy-2-pyridinones, their tetra-dentate derivatives with amino-bisphosphonate moiety were prepared [143]. A Canadian researcher recently claimed to produce a new novel compound similar to deferiprone for treatment of iron storage in brain tissues of Parkinson affected patients [144]. The low molecular weight bi-and tri-ligands have been prepared for greater efficacy and higher oral bioavailability but are found potentially more toxic than hexadentate ligand [145] and therefore more research is required to prepare safer ligands.

Deferiprone (Ferriprox®)

Alternatively deferiprone is also known as CP2/l1/Kelfer. It is a hydrophilic small molecular weight hydroxypyridone (3-hydroxy-1, 2-dimethylpyridin-4-one) and the first oral iron chelator introduced in Europe in early 1980s that eliminates iron through faeces [146]. Among various hydroxypyridinones discovered, the 3-hydroxy-4-pyridinones are found most effective at physiological pH [147]. Large volume of work was done on this oral chelator between 1990-2000 till deferasirox was found more suitable alternative. Deferiprone is a neutral molecule that readily enters into cell membrane and forms 1:3 iron-chelator complexes. It has a half-life of 160 minutes [148]. The drug could efficiently remove myocardial iron compared to hepatic iron [149,150]. Many groups indicated in their study, greater reduction in iron overload induced cardiac failure due to deferiprone treatment [151, 152]. In a recent study, sixty patients who received thrice daily oral deferiprone (total dose 75 mg/kg/day), showed great enhancement in left-ventricular ejection function as compared to other 180 patients who received deferoxamine alone [153]. No deaths were seen in fifty-five thalassemia patients treated with deferiprone alone or eighteen patients with deferiprone-DFO combination, as compared to eleven cardiac arrests in DFO alone treated patients [154]. The drug has been found to cross blood brain barrier efficiently and is under phase II clinical trials for its safe and efficient use in treating neurodegenerative disorders such as Parkinson’s disease.

However few case reports of deferiprone therapy indicated death due to fatal congestive heart failure with no other known etiology [155, 156]. In a recent case study of a 30-year-old thalassemia patient, the switching from DFO to deferiprone resulted in decrease in serum ferritin levels but also produced seizures after five months of treatment leading to permanent discontinuation of deferiprone treatment [157]. Therefore seizures may be one of the potential adverse effects of deferiprone therapy. Though the drug proved effective in treating iron overload of thalassemia patients [158-160], it also exhibited severe side effects of neutropenia and agranulocytosis and therefore remained as second-line defense only for which deferoxamine treatment remained inadequate [161-163]. Severe side effects such as arthropathy have been observed at dose of 100 mg/kg/day [164]. Deferiprone is found effective in MDS patients but is not licensed in USA and European countries on account of its associated toxicity. Usual doses of treatment are very high (75 mg/kg/day) as compared to DFO (20-40 mg/kg/d) and deferasirox (20-30 mg/kg/d).

Deferasirox (ICL670)

Commercially called as Exjade, deferasirox is a tridentate, synthetic oral iron chelator prepared from salicylic acid, salicylamide and 4-hydrazinobenzoic acid and is proposed to serve the patients who are either incompliant or show intolerance to DFO. It was approved by FDA in USA in 2005 and Europe in 2006 and is under phase II/III clinical trials for iron decorporation in humans [165]. Deferasirox binds to iron in 2:1 ratio and is lipophilic with half life of 8-12 hours so that it has to be administered only once in a day. An affordable single dose of deferasirox per day was found highly tolerable as well as efficient by iron-overloaded patients of sickle cell anaemia [166] and other iron overload disorders [167, 168]. More than 80% of drug is excreted from fecal route while renal excretion was only eight percent [169]. Deferasirox was effective in reducing serum ferritin level in 166 patients suffering from non-transfusion dependent thalassemia with only minor side effects such as nausea, body rashes and diarrhea [170] and increased the quality of life by reducing morbidity and mortality among thalassemia patients [171]. A study on 175 iron loaded myelodysplastic syndrome (MDS) patients was done with dose range 20-30 mg/kg/day, depending on the frequency of blood transfusions, and found that deferasirox showed well-defined and manageable safety profile [172]. In their later study and others [173-174], it was observed that one-third of the patient volunteers did not achieve satisfactory iron balance at doses ≥30 mg/kg per day. A group of workers found increase in toxicity of deferasirox if dose was increased from 30 to 40 mg/kg/d [175]. Deaths in older patients with MDS have been reported who were administered higher doses of deferasirox [176]. More recently it is suggested that instead of a single high dose, twice-daily doses of deferasirox should be administered to patients with chronic iron overload [177].

Several authors have reviewed the development and current status of oral deferasirox [178, 179]. A one-year long study in about six hundred thalassemia patients with iron overload, who contraindicated DFO treatment, was done by European Medicines Agency and observed encouraging results with deferasirox [180]. The drug was found effective in treating iron overload in patients suffering from MDS [181-183], β-thalassemia [184, 185], hereditary hemochromatosis [186], sickle cell disease [187, 188] and allogeneic hematopoietic cell transplantation [189]. However deferasirox was found inefficient in few pediatric patients who underwent frequent blood transfusions during and after high-dose chemotherapy with autologous stem cell transplantation [190]. A phase II clinical study on the same confirmed the poor tolerability and GI toxicity of deferasirox [191].

Many workers have suggested that deferasirox will soon replace the conventional chelation therapy with DFO but the fact that it is contraindicated in patients in advanced stage of hematologic disorders is also unavoidable [192]. Of more importance is that safety data on deferasirox prescribed doses is quite limited and need to be worked upon extensively before it is recommended as first-line treatment against DFO [193].

Hydroxyquinolines

Hydroxyqunoline is an oral bidentate chelator that forms five-membered ring structure with iron [194]. It is lipophilic and readily enters into cell membrane and BBB. VK-28 and its derivatives such as M-10, M-30 are hydroxyquinoline based iron chelators having potency similar to DFO [195]. Their ability to cross BBB has been used for treating neurodegenerative disorders. Hydroxyquinoline based antibiotic such as clioquinol (5-chloro-7-iodo-hydroxyquinoline) chelates iron [196], though not very specifically, but is being explored for possible use in iron reduction in Parkinson disorder [197].

Chelator combination therapy

Many practitioners suggest sequential or combined regimen with two or more chelators to effectively overcome the iron overload associated with many diseases instead of single drug therapy. There is a great deal of work on the combination therapy with deferiprone and DFO since 1990s based on same or different days therapy and many of them are under clinical trials [198-200]. A 7-15 day study on five transfusion-dependent patients showed that instead of increasing deferiprone dose from 75 mg/kg to higher values, a combination of subcutaneous DFO and oral deferiprone could be more effective with reduced toxicity [201]. Few studies have suggested that the combination of deferiprone and deferoxamine is effective if given on same day [202, 203] while others indicated dose dependent chelator combinations to be given on different period of days [204, 205]. The synergistic effect of chelator combinations is expected to minimize the toxicity and inconvenience associated with monotherapy.

High molecular weight iron chelators

Though DFO therapy is currently the most commonly practiced treatment, yet it is costly and unaffordable by a large percentage of patients living in under-developed and developing countries. Short circulation time, patient incompliance and toxicity issues associated with DFO treatment has led to the development of high molecular weight chelators. Iron binding polymers have a great potential at therapeutic level to bind iron irreversibly and form stable and non-toxic complexes. The first of these attempts included inclusion of hydroxamic acids onto polymers to obtain a high stability constant for iron. Hydroxamic acids are long known for their powerful ability to selectively bind iron. In this regard a series of water-soluble acrylic polymers with hydroxamic acid bearing side chains (PHAs) were prepared which proved highly selective for iron with reduced circulation time [206]. But these polymers were non-biodegradable and could lead to intoxication followed by deposition. Ten years later Meshchanov & coworkers suggested that 30-35 hydroxamic acids per hundred monomer units of polymer increase the ability of PHAs to eliminate body iron as well as to undergo biodegradation [207]. To enhance patient compliance, the development of effective and highly selective oral chelators is must. Therefore orally active drugs based on chitosan [208-209] and pectin derivatives [210] were prepared, that got approval by FDA for iron chelation. The adhesive property of chitosan based microspheres functionalized with catechol or hydroxyl-carboxylic acid was used to prepare orally active and therapeutically efficient iron chelators [211]. A team of workers claimed to produce hydrogels based on conjugate of dihydroxybenzoic acid on polyamine vehicles such as polyacrylamide and polyvinyl alcohol that were ten times more effective than any of the current therapeutics [212]. Various other chelators such as desferrithiocin and hydroxypyridinones were also covalently linked to polyamine backbone such as spermine and showed increased permeation of the chelator into cells. Taking into account the selectivity and specificity of DFO for iron, the drug was bound to high molecular weight polymeric vehicles through its amino group. This approach led to the development of high molecular weight chelators having enhanced blood retention times and reduced toxicity. Dextran was proposed as a biodegradable and biocompatible matrix to which DFO was covalently bonded [213]. Hydroxyl ethyl starch-conjugated deferoxamine (HES-DFO) is currently under Phase Ib clinical trial (40SD02) as a novel long-lasting formulation [214].

CONCLUSION

Iron overload is a common problem all over the world. Workers from all over the world are involved in finding the best suitable therapeutic agent that can treat the iron overload and its associated trauma. The active involvement of a number of researchers to find the best suitable iron overload treatment may increase the availability of therapeutic options for patients. A number of iron chelators are approved for treatment of iron overload associated to thalassemia, cancer, asthma etc and others are being improved to get the patients relieved from iron overload with reduced toxicity and maximum efficacy. Though a number of agents, natural or synthetic, are currently available or are being screened to reduce the iron overload, yet the development of a best ligand is challenge in the area of drug discovery and development. The treatment of chronic iron overload is challenge to the modern practitioners, which exposes them to great dilemma regarding management of the problem.

ACKNOWLEDGEMENT

The authors are highly thankful to Department of Science and Technology (DST), New Delhi, India for providing funds to carry out the study.

CONFLICTS OF INTERESTS

The authors declare that they have no conflicts of interest

REFERENCES

  1. Gutteridge JMC, Richmond R, Halliwell B. Inhibition of the iron-catalyzed formation of hydroxyl radicals from superoxide and of lipid peroxidation by desferrioxamine. Biochem J 1979;184:469-72.
  2. Halliwell B. Protection against tissue damage in vivo by desferrioxamine: what is its mechanism of action? Free Radic Biol Med 1989;7:645-51.
  3. Corbett JV. Accidental poisoning with iron supplements. MCN: Am J Maternal/Child Nursing 1995;20:234.
  4. Institute of Medicine. Food and Nutrition Board. Dietary Reference Intakes for Vitamin A, Vitamin K, Arsenic, Boron, Chromium, Copper, Iodine, Iron, Manganese, Molybdenum, Nickel, Silicon, Vanadium and Zinc. Washington DC, National Academy Press; 2001.
  5. Rowley DA, Gutteridge JMC, Halliwell B. Superoxide dependent formation of hydroxyl radicals in the presence of iron salts. Biochem J 1981;199:263-5.
  6. Rank BH, Carlsson J, Hebbel RP. Abnormal redox status of membrane protein thiols in sickle erythrocytes. J Clin Invest 1985;75:1531-7.
  7. Golan DE, Corbett JD. Band 3 and glycophorin are progressively aggregated in density-fractionated sickle and normal red blood cells: Evidence from rotational and lateral mobility studies. J Clin Invest 1993;91:208-17.
  8. Repka T, Hebbel RP. Hydroxyl radical formation by sickle erythrocyte membranes: role of pathologic iron deposits and cytoplasmic reducing agents. Blood 1991;78:2753–8.
  9. Okada S. Iron-induced tissue damage and cancer: the role of reactive oxygen species-free radicals. Pathol Int 1996;46:311–32.
  10. Horton AA, Fairhurst S. Lipid peroxidation and mechanisms of toxicity. Crit Rev Toxicol 1987;18:27-79.
  11. Gilbert RS. Bloodletting over the centuries. N Y State J Med 1980;80:2022-8.
  12. Antle EA. Who needs a therapeutic phlebotomy? Clin J Oncol Nurs 2010:14:694-6.
  13. Barton JC, McDonnel SM, Adams PC, Brissot P, Powell LW, Edwards CQ, et al. Management of hemochromatosis. Hemochromatosis management working group. Ann Intern Med 1998;129:932–9.
  14. Moirand R, Adams PC, Bicheler V, Brissot P, Deugnier Y. Clinical features of genetic hemochromatosis in women compared with men. Ann Intern Med 1997;127:105-10.
  15. Alessandro B, Paola V, Filomena D, Paolo P, Enrico G, Sonia C, et al. Hepatic expression of hemochromatosis genes in two mouse strains after phlebotomy and iron overload. Haematologica 2005;90:1161-7.
  16. Powars D, Wilson B, Imbus C, Pegelow C, Allen J. The natural history of stroke in sickle cell disease. Am J Med 1978;65:461–71.
  17. Balkaran B, Char G, Morris JS, Serjeant BE, Serjeant GR. Stroke in a cohort of patients with homozygous sickle cell disease. J Pediatr 1992;120:360–6.
  18. Emanuele A, Pietro M, Guido L, Marta R, Donatella B, Buket E, et al. Phlebotomy to reduce iron overload in patients cured of thalassemia by bone marrow transplantation. Italian cooperative group for phlebotomy treatment of transplanted thalassemia patients. Blood 1997;90:994-8.
  19. Russell WE, Zimmerman SA, Sylvestre PB, Mortier NA, Davis JS, Treem WR, et al. Prevention of secondary stroke and resolution of transfusional iron overload in children with sickle cell anemia using hydroxyurea and phlebotomy. J Pediatr 2004;145:346-52.
  20. Li CK, Lai DH, Shing MM, Chik KW, Lee V, Yuen PM. Early iron reduction programme for thalassaemia patients after bone marrow transplantation. Bone Marrow Transplant 2000;25:653-6.
  21. Lucarelli G, Galimberti M, Polchi P, Angelucci E, Baronciani D, Giardini C, et al. Bone marrow transplantation in patients with thalassemia. N Engl J Med 1990;322:417-21.
  22. Tomas JF, Pinilla I, Garc!ıa-Buey ML, García A, Figuera A, Gómez-García de Soria VGG, et al. Long-term liver dysfunction after allogeneic bone marrow transplantation: clinical features and course in 61 patients. Bone Marrow Transplant 2000;26:649-55.
  23. Rose C, Ernst O, Hecquet B, Patrice M, Pascale R, Marie PN, et al. Quantification by magnetic resonance imaging and liver consequences of post-transfusional iron overload alone in long-term survivors after allogeneic hematopoietic stem cell transplantation (HSCT). Haematologica 2007;92:850-3.
  24. Butt NM, Clark RE. Autografting as a risk factor for persistingiron overload in long-term survivors of acute myeloid leukaemia. Bone Marrow Transplant 2003;32:909-13.
  25. Mateo-Gallego R, Calmarza P, Jarauta E, Elena B, Ana C, Fernando C. Serum ferritin is a major determinant of lipid phenotype in familial combined hyperlipidemia and familial hypertriglyceridemia. Metabolism 2010;59:154-8.
  26. Solanas-Barca M, Mateo-Gallego R, Calmarza P, Jarauta E, Bea AM, Cenarro A, et al. Mutations in HFE causing hemochromatosis are associated with primary hypertriglyceridemia. J Clin Endocrinol Metab 2009;94:4391-7.
  27. Paola CE, Nuria G, Eva A, Carmen G, Rocio MG, Pilar G, et al. Effect of phlebotomy on lipid metabolism in subjects with hereditary hemochromatosis. Metab Clin Exp 2011;60:830–4.
  28. Bomford A, Williams R. Long term results for venesection therapy in idiopathic hemochromatosis. Q J Med 1976;45:611-23.
  29. McAllen PM, Coghill NF, Lubran M. The treatment of hemochromatosis. Q1 Med 1957;26:251-76.
  30. Skinner C, Kenmure ACF. Hemochromatosis presenting as congestive cardiomyopathy and responding to venesection. Br Heart J 1973;35:466-8.
  31. Williams R, Smith PM, Spicer ElF, Barry M, Sherlock S. Venesection therapy in idiopathic hemochromatosis: an analysis of 40 treated and 18 untreated patients. Q1 Med 1969;38:1-16.
  32. Mostert LJ, Cleton MI, Bruijn WC, Koster JF, Vaneijk HG. Studies on ferritin in rat liver and spleen during repeated phlebotomy. Int J Biochem 1989;21:39-47.
  33. Farinato M, Di Pierro G, Amalfi G, Caravelli V, D'Angelo F, Morace M, et al. Iron overload in chronic liver diseases and clinical usefulness of therapeutic phlebotomy: our experience. Dig Liver Dis 2006;38:S185.
  34. Hayashi H, Takikawa T, Nishimura N, Yano M, Isomura T, Sakamoto N. Improvement of serum aminotransferase levels after phlebotomy in patients with chronic active hepatitis C and excess hepatic iron. Am J Gastroenterol 1994;89:986-8.
  35. Limdi JK, Crampton JR. Review hereditary haemochromatosis. Q J Med 2013;97:315–24.
  36. Adams PC, Barton C. How I treat hemochromatosis? Blood 2010;116:317-25.
  37. Valenti L, Fracanzani AL, Dongiovanni P, Bugianesi E, Marchesini G, Manzini P, et al. Iron depletion by phlebotomy improves insulin resistance in patients with nonalcoholic fatty liver disease and hyperferritinemia: evidence from a case-control study. Am J Gastroenterol 2007;102:1251-8.
  38. Minamiyama Y, Takemura S, Kodai S, Shinkawa H, Tsukioka T, Ichikawa H, et al. Iron restriction improves type 2 diabetes mellitus in otsuka long-evans tokushima fatty rats. Am J Physiol: Endocrinol Metab 2010;298:1140-9.
  39. Niederau C, Fischer R, Purschel A, Stremmel W, Haussinger D, Strohmeyer G. Long-term survival in patients with hereditary hemochromatosis. Gastroenterology 1996;110:1107-19.
  40. Elizabeth MS, Roger AW, Margaret EB. Myocardial involvement in idiopathic hemochromatosis: Morphologic and clinical improvement following venesection. Am J Med 1981;70:1275–9.
  41. Dabestani A, John SC, Eberhard H, Joseph KP, Hans S, William GF, et al. Primary hemochromatosis: Anatomic and physiologic characteristics of the cardiac ventricles and their response to phlebotomy. Am J Cardiol 1984;54;153–9.
  42. Bacon BR, Adams PC, Kowdley KV, Lawrie WP, Anthony ST. Diagnosis and management of hemochromatosis: 2011 practice guideline by the american association for the study of liver diseases. Hepatology 2011;54:328-43.
  43. Koliantzaki G, Sorras KD, Dimitrakopoulos SK, Kathopoulis NI, Bonas A, Saltamavros A, et al. Phlebotomy a longitudinal operating techniques. 9th European Congress on Menopause and Andropause/Maturitas; 2012;71:S1–S82.
  44. McCaffery M, Beebe A. Pain: clinical manual for nursing practice. Baltimore: VV Mosby Company St. Louis; 1994.
  45. Manea P, Loustaud-Ratti V, Mondary D, Arnold V, Ferley JP, Souris S, et al. Evaluation of at-home phlebotomy for iron overload: Feasibility and satisfaction of patients and healthcare workers. Gastroenterol Clin Biol 2008;32:172-9.
  46. Kontoghiorghes GJ, Pattichi K, Hadjigavriel M, Kolnagou A. Transfusional iron overload and chelation therapy with deferoxamine and deferiprone (L1). Transfus Sci 2000;23:211–33.
  47. Fabio G, Minonzio F, Delbini P, Bianchi A, Cappellini MD. Reversal of cardiac complications by deferiprone and deferoxamine combination therapy in a patient affected by a severe type of juvenile hemochromatosis. Blood 2007;109:362-4.
  48. Cosgrove DJ. The determination of myo-inositol hexakisphosphate (phytate). J Sci Food Agric 1980 31;1253–6.
  49. Hawkins PT, Poyner DR, Jackson TR, Letcher AJ, Lander DA, Irvine RF. Inhibition of iron-catalysed hydroxyl radical formation by inositol polyphosphates: a possible physiological function for myo-inositol hexakisphosphate. Biochem J 1993;294:929–34.
  50. Mukhopadhyay S, Basak J, Kar M, Mandal S, Mukhopadhyay A. Wheatgrass as a natural iron chelator in myelodysplastic syndrome, the role of iron chelation activity of wheat grass juice in patients with myelodysplastic syndrome. J Clin Oncol 2009;27:7007-12.
  51. Thephinlap C, Ounjaijean S, Khansuwan U, Fucharoen S, Porter JB. Epigallocatechin-3-gallate and epicatechin-3-gallate from green tea decrease plasma non-transferrin bound iron and erythrocyte oxidative stress. Med Chem 2007;3:289-96.
  52. Mandel S, Amit T, Reznichenko L, Weinreb O, Youdim MB. Green tea catechins as brain-permeable, natural iron chelators-antioxidants for the treatment of neurodegenerative disorders. Mol Nutr Food Res 2006;50:229-34.
  53. Kessler M, Ubeau G, Jung L. Anti-and pro-oxidant activity of rutin and quercetin derivatives. J Pharm Pharmacol 2003;55:131-42.
  54. Cook NC, Samman S. Flavonoids-chemistry, metabolism, cardioprotective effects, and dietary sources. Nutr Biochem 1996;7:66-76.
  55. Shahidi F, Wanasundara PKJ. Phenolic antioxidants. Crit Rev Food Sci Nutr 1992;32:67-103.
  56. Ali ME, Fereshteh P, Ahmad RB. Iron chelating activity, phenol and flavonoid content of some medicinal plants from Iran. Afr J Biotechnol 2008;7:3188-92.
  57. Hutchinson C, Bomford A, Geissler CA. The iron-chelating potential of silybin in patients with hereditary haemochromatosis. Eur J Clin Nutr 2010;64:1239–41.
  58. Gharagozloo M, Khoshdel Z, Amirghofran Z. The effect of an iron (III) chelator, silybin, on the proliferation and cell cycle of Jurkat cells: A comparison with desferrioxamine. Eur J Pharmacol 2008;589:1–7.
  59. Borsari M, Gabbi C, Ghelfi F, Grandi R, Saladini M, Severi S. Silybin, a new iron-chelating agent. J Inorg Biochem 2001;85:123–9.
  60. Borsari M, Ferrari E, Grandi R, Saladini M. Curcuminoids as potential new iron-chelating agents: spectroscopic, polarographic and potentiometric study on their Fe(III) complexing ability. Inorg Chim Acta 2002;328:61–8.
  61. Benassi R, Ferrari E, Grandi R, Lazzari S, Saladini M. Synthesis and characterization of new b-diketo derivatives with iron chelating ability. J Inorg Biochem 2007;101:203–13.
  62. Jurkevitch E, Hadar Y, Chen Y, Libman J, Shanzer A. Iron uptake and molecular recognition in Pseudomonas putida: receptor mapping with ferrichrome and its biomimetic analogs. J Bacterial 1992;174:78-83.
  63. Reid RT, Live DH, Faulkner DJ, Butler A. A siderophore from a marine bacterium with an exceptional ferric ion affinity constant. Nature 1993;366:455-8.
  64. Ito T, Neilands JB. Products of low iron fermentation with Bacillus subtilis: isolation, characterization and synthesis of 2,3-dihydroxybenzoylglycine. J Am Chem Soc 1958;80:4645-7.
  65. Winkelmann G, Drechsel H. Microbial Siderophores. In: Rehm J, Reed G. editors. Biotechnology Set. 2nd ed. Germany: Wiley-VCH Verlag GmbH, Weinheim; 2008.
  66. Sigel A, Sigel H. Metal ions in biological systems, Volume 35: Iron transport and storage microorganisms, plants, and animals. Met Based Drugs 1998;5:262.
  67. Dionis JB, Jenny HB, Peter HH. Handbook of Microbial Iron Chelates. In: Winkelmann G Editor. CRC Press; 1991. p. 339-56.
  68. Porter JB, Huehns ER, Hider RC. The development of iron chelating drugs. Baillieres Clin Haematol 1989;2:257-92.
  69. Bulman RA. The chemistry of chelating agents in medical sciences. Struct Bonding 1987;67:91-141.
  70. Zevin S, Link G, Grady RW, Hider RC, Peter HH, Hershko C. Origin and fate of iron mobilized by the 3-hydroxypyridin-4-one oral chelators: studies in hypertransfused rats by selective radioiron probes of reticuloendothelial and hepatocellular iron stores. Blood 1992;79:248-53.
  71. Miller MJ. Syntheses and therapeutic potential of hydroxamic acid based siderophores and analogues. Chem Rev 1989;89:1563–79.
  72. Hider RC, Hall AD. Clinically useful chelators of tripositive elements. Prog Med Chem 1991;28:41–173.
  73. Keberle H. The biochemistry of desferrioxamine and its relation to iron metabolism. Ann New York Acad Sci 1964;119:758-68.
  74. Summers JB, Gunn BP, Martin JG, Mazdiyasni H, Stewart AO, Young PR, et al. Orally active hydroxamic acid inhibitors of leukotriene biosynthesis. J Med Chem 1988;31:3–5.
  75. Dhungana S, White PS, Crumbliss AL. Crystal structure of ferrioxamine b: a comparative analysis and implications for molecular recognition. J Biol Inorg Chem 2001;6:810.
  76. Westlin WF. Deferoxamine as a chelating agent. Clin Toxicol 1971;4:597-02.
  77. Robotham JL, Lietman PS. Acute iron poisoning. a review. Am J Dis Child 1980;34:875-9.
  78. Cohen A, Schwartz E. Iron chelation therapy with deferoxamine in Cooley anemia. J Pediatr 1978;92:643-7.
  79. Ohlsson WTL, Kullendorff GT, Ljungberg KL. Transfusion hemosiderosis; report of a case treated with BAL. Acta Med Scand 1953;145:410–8.
  80. Seven MJ, Gottlieb H, Israel HL, Reinhold J0, Rubin M. N-hydroxyethyl-ethylenedianine triacetic acid, versenol, in the treatment of hemochromatosis. Am J Med Sci Direct 1954;228:646–51.
  81. Bannerman RM, Callender ST, Williams DL. Effect of Desferrioxamine and DTPA in iron overload. Br Med J 1962;2:1573–7.
  82. Bickel H, Bosshardt R, Gaumann E, Reusser P, Vischer E, Voser W. Stoffwechsel-produkte von actinomyceten, über die isolierung und charakterisierung der ferrioxamine A-F, neuer wuchsstoff der sideramin-gruppe. Helv Chim Acta 1960;43:2105-18.
  83. Tufano TP, Pecoraro VL, Raymond KN. Ferric ion sequestering agents: kinetics of iron release from ferritin to catechoylamides. Biochim Biophys Acta 1981;668:420-8.
  84. Kontoghiorghes GJ, Evans RW. Site specificity of iron removal from transferrin by ketohydroxypyridine chelators. FEBS Lett 1985;189:141–4.
  85. Kretchmar SA, Raymond KN. Biphasic kinetics and temperature dependence of iron removal from transferrin by 3,4-LICAMS. J Am Chem Soc 1986;1086:6212-8.
  86. Simeon P, Grace V, Fred L. Iron removal from transferrin: an experimental study. Biochim Biophys Acta 1977;497:481-7.
  87. Sawas DC, Soulpi C. Alterations of the [59Fe] ferric citrate biodistribution in hyperferremic mice after the administration of pyrophosphate and desferrioxamine. J Pharmacol Exp Ther 1983;224:415-8.
  88. Hershko C, Grady RW, Cerami A. Mechanism of iron chelation in the hypertransfused rat: definition of two alternative pathways of iron mobilization. J Lab Clin Med 1978;92:144-51.
  89. Singh S, Mohammed N, Ackerman R, Porter JB, Hider RC. Quantificationof desferrioxamine and its iron chelating metabolites by high performance chromatography and simultaneous radioactive detection. Anal Biochem 1992;203:116-20.
  90. Wapnick AA, Lynch SR, Charlton RW, Seftel HC, Bothwell TH. The effect of ascorbic acid deficiency on desferrioxamine induced urinary iron excretion. Br J Haematol 1969;17:563-8.
  91. O’Brien RT. Ascorbic acid enhancement of desferrioxamine induced urinary iron excretion in thalassemia major. Ann NY Acad Sci 1974;232:221-5.
  92. Pippard MJ, Callender ST, Finch CA. Ferrioxamine excretion in iron-loaded man. Blood 1982;60:288-94.
  93. Rahko PS, Salerni R, Uretsky BF. Successful reversal by chelation therapy of congestive cardiomyopathy due to iron overload. J Am Coll Cardiol 1986;8:436-40.
  94. Barry M, Flynn DM, Letsky EA, Risdon RA. Long-term chelation therapy in thalassaemia major: effect on liver iron concentration, liver histology and clinical progress. Br Med J 1974;2:16-20.
  95. Pignatti BC, Rugolotto S, Stefano PD, Zhao H, Cappellini MD, Vecchio GCD, et al. Survival and complications in patients with thalassemia major treated with transfusion and deferoxamine. Haematol 2004;89:1187-93.
  96. Zurlo MG, Stefano PD, Borgna-Pignatti C, Palma AD, Piga A, Melevendi C, et al. Survival and causes of death in thalassemia major. Lancet 1989;334:27-30.
  97. Summers MR, Jacobs A, Tudway D, Perera P, Ricketts C. Studies in desferrioxamine and ferrioxamine metabolism in normal and iron-loaded subjects. Br J Haematol 1979;42:547–55.
  98. Olivieri NF, Brittenham GM. Iron chelation therapy and the treatment of thalassemia. Blood 1997;89:739–61.
  99. Wong B, Richardson DR. Beta-thalassemia: emergence of new and improved iron chelators for treatment. Int J Biochem Cell Biol 2003;35:1144–9.
  100. Catsch A, Hartmuth-Hoene AE. Pharmacology and therapeutic applications of agents used in heavy metal poisoning. Pharmacol Ther 1976;1:1–118.
  101. Polson RJ, Jawed A, Bomford A. Treatment of rheumatoid arthritis with desferrioxamine: relation between stores of iron before treatment and side effects. BMJ 1985;291:448.
  102. Raymond KN, Muller G, Matzanke BF. Complexation of iron by siderophores. A review of their solution and structural chemistry and biological functions. Top Curr Chem 1984;123:49-102.
  103. Carrano CJ, Raymond KN. Ferric ion sequestering agents. Kinetics and mechanism of iron removal from transferrin by enterobactin and synthetic tricatechols. J Am Chem Soc 1979;101:5401–4.
  104. Guterman SK, Morris PM, Tannenberg WJK. Feasibility of enterochelin as an iron-chelating drug: Studies with human serum and a mouse model system. General Pharmacol Vascular System 1978;9:123–7.
  105. Smithgall TE, Penning TM. Indomethacin-sensitive 3a-hydroxysteroid dehydrogenase in rat tissues. Biochem Pharmacol 1985;34:831-5.
  106. O'Brien IG, Cox GB, Gibson F. Enterochelin hydrolysis and iron metabolism in Escherichia coli. Biochim Biophys Acta 1971;237:537-49.
  107. Wei Y, Guo M. Hydrogen peroxide triggered prochelator activation, subsequent metal chelation and attenuation of the Fenton reaction. Angew Chem Int Ed 2007;46:4722–5.
  108. Sun J, Chu Y, Wu X, Liu RH. Antioxidant and antiproliferative activities of common fruits. J Agric Food Chem 2002;50:7449–54.
  109. Guo M, Perez C, Wei Y, Rapoza E, Su G, Bou-Abdallah F, Chasteen ND. Iron-binding properties of plant phenolics and cranberry’s bio-effects. Dalton Trans 2007;43:4951–61.
  110. Bergeron RJ, Wiegand J, Brittenham GM. HBED: a potential alternative to deferoxamine for iron-chelating therapy. Blood 1998;91:1446–52.
  111. Neufeld EJ, Galanello R, Viprakasit V, Aydinok Y, Piga A, Harmatz P, et al. A phase 2 study of the safety, tolerability, and pharmacodynamics of FBS0701, a novel oral iron chelator, in transfusional iron overload. Blood 2012;119:3263-8.
  112. Grünblatt E, Mandel S, Berkuzki T, Youdim MB. Apomorphine protects against MPTP-induced neurotoxicity in mice. Mov Disord 1999;14:612–8.
  113. Drechsel H, Jung G, Winkelmann. Stereochemical characterization of rhizoferrin and identification of its dehydration products. G Biol Met 1992;5:141-8.
  114. Galey JB, Destree O, Dumats J, Pichaud P, March J, Genard S, et al. Protection of U937 cells against oxidative injury by a novel series of iron chelators. Free Radical Biol Med 1998;25:881–90.
  115. Cox CD, Adams P. Siderophore activity of pyoverdin for Pseudomonas aeruginosa. Infect Immun 1985;48:130-8.
  116. Jego P, Lescoat G, Hubert N, Pasdeloup N, Brissot P, Ocaktan AZ, et al. Inhibition of iron overload toxicity by pyoverdins in adult rat hepatocyte cultures. J Inorg Biochem 1991;43:639.
  117. Chenoufi N, Hubert N, Loréal O, Morel I, Pasdeloup N, Cillard J, et al. Inhibition of iron toxicity in rat and human hepatocyte cultures by the hydroxypyridin-4-ones CP20 and CP94. J Hepatol 1995;23:166-73.
  118. Somporn S, Susan A, Malulee T, Sumonrat M. Comparison between Pseudomonas aeruginosa siderophores and desferrioxamine for iron acquisition from ferritin. Asian Biomed 2010;4:631-35.
  119. Morel I, Cillard J, Lescoat G, Sergent O, Pasdeloup N, Ocaktan AZ, et al. Antioxidant and free radical scavenging activities of the iron chelators pyoverdin and hydroxypyrid-4-ones in iron-loaded hepatocyte cultures: comparison of their mechanism of protection with that of desferrioxamine. Free Radic Biol Med 1992;13:499-08.
  120. Richardson DR. Iron chelators as therapeutic agents for the treatment of cancer. Crit Rev Oncol Hematol 2002;42:267-81.
  121. Liu ZD, Hider RC. Design of iron chelators with therapeutic application. Coord Chem Rev 2002;232:151-71.
  122. Dean RT, Nicholson P. The action of nine chelators on iron-dependent radical damage. Free Radic Res 1994;20:83–101.
  123. Singh S, Khodr H, Taylor MI, Hider RC. Therapeutic iron chelators and their potential side-effects. Biochem Soc Symp 1995;61:127–37.
  124. Baret P, Claude B, Gaude D, Gellon G, Mourral C, Pierre JL, et al. Tripodal ligands possessing six convergent hydroxyl groups a novel family of iron sequestering agents based on o, o’-dihydroxybiphenyl subunits tetrahedron. 1994;50:2077-94.
  125. Fredenburg AM, Wedlund PJ, Skinner TL, Damani LA, Hider RC, Yokel RA. Pharmacokinetics of representative 3-hydroxypyridin-4-ones in rabbits: CP20 and CP94. Drug Metab Dispos 1993;21:255-8.
  126. Toso L, Crisponi G, Nurchi VM, Crespo-Alonso M, Lachowicz JI, Santos MA, et al. A family of hydroxypyrone ligands designed and synthesized as iron chelator. J Inorg Biochem 2013;127:220–31.
  127. Hartley A, Rice-Evans C. The chelation of nonheme iron within sickle erythrocytes by the hydroxypyridinone chelator CP094. Arch Biochem Biophys 1992;297:377-82.
  128. Ponka P, Borova J, Neuwirt J, Fuchs O. Mobilization of iron from reticulocytes. Identification of pyridoxal isonicotinoyl hydrazone as a new iron chelating agent. FEBS Lett 1979;97:317–21.
  129. Hoy T, Humphreys J, Jacobs A, Williams A, Ponka P. Effective iron chelation following oral administration of an isoniazid pyridoxalhydrazone. Br J Haematol 1979;43:443–9.
  130. Brittenham GM. Pyridoxal isonicotinoyl hydrazone: an effective chelator after oral administration. Semin Hematol 1990;27:112–6.
  131. Wong CSM, Kwok JC, Richardson DR. PCTH: a novel orally active chelator of the aroylhydrazone class that induces iron excretion from mice. Biochim Biophys Acta 2004;1739:70–80.
  132. Becker E, Richardson DR. Development of novel aroylhydrazone ligands for iron chelation therapy: 2-pyridylcarboxaldehydeisonicotinoyl hydrazine analogs. J Lab Clin Med 1999;134:510–21.
  133. Zanninelli G, Choudury R, Loreal O, Guyader D, Lescoat G, Arnaud J, et al. Novel orally active iron chelators (3-hydroxypyridin-l-ones) enhance the biliary excretion of plasma non-transferrin-bound iron in rats. J Hepat 1991;27:116-84.
  134. Grazina R, Gano L, Sebestialk J, Santos MA. New tripodal hydroxypyridinone based chelating agents for Fe(III), Al(III)and Ga(III): Synthesis, physicochemical properties and bioevaluation. J Inorg Biochem 2009;103:262–73.
  135. Santos MA, Gama S, Gano L, Cantinho G, Farkas E. A new bis(3-hydroxy-4-pyridinone)-IDA derivative as a potential therapeutic chelating agent. Synthesis, metal-complexation and biological assays. Dalton Trans 2004;7:3772-81.
  136. Liu ZD, Hider RC. Design of clinically useful iron (III)-selective chelators. Med Res Rev 2002;22:26-64.
  137. Santos MA. Review: Recent developments on 3-hydroxy-4-pyridinones with respect to their clinical applications-Mono and combined ligand approaches. Coord Chem Rev 2008;252:1213–24.
  138. Hershko C, Grady RW, Link G. Phenolic ethylenediamine derivatives: a study of orally effective iron chelators. J Lab Clin Med 1984;103:337–46.
  139. Grady RW, Salbe AD, Hilgartner MW, Giardina PJ. Results from a phase I clinical trial of HBED. Adv Exp Med Biol 1994;356:351–9.
  140. Samuni AM, Afeworki M, Stein W, Yordanov AT, DeGraff W, Krishna MC, et al. Multifunctional antioxidant activity of HBED iron chelator. Free Radical Biol Med 2001;30:170–7.
  141. Galey JB, Dumats J, Beck I, Fernandez B, Hocquaux M. N, N-9-Bis dibenzyl ethylenediaminediacetic acid (DBED): a site-specific hydroxyl radical scavenger acting as an oxidative stress activatable iron chelator in vitro. Free Radic Res 1995;22:67–86.
  142. Maxton DG, Bjarnason I, Reynolds AP, Catt SD, Peters TJ, Menzies IS. Lactulose [15]Cr-labelled ethylenediaminetetra-acetate, L-rhamnose and polyethyleneglycol 500 as probe marker for assessment in vivo of human intestinal permeability. Clin Sci 1986;71:71-80.
  143. Bailly T, Burgada R, Prange T, Lecouvey M. Synthesis of tetradentate mixed bisphosphonates-new hydroxypyridinonate ligands for metal chelation therapy. Tetrahedron Lett 2003;44:189–92.
  144. Wodzinska J. Proof of concept of treatment with novel hydroxypyridinone iron chelators in a non-clinical model of parkinson's disease. Therapeutics Development Initiative. The Micheal J. Fox Foundation; 2011.
  145. Hider RC, Liu ZD, Piyamongkol S. The design and properties of 3-hydroxypyridin-4-one iron chelators with high pFe3 values. Transfus Sci 2000;23:201-9.
  146. Hider RC, Singh S, Porter JB, Huehns ER. The development of hydroxypyridin-4-ones as orally active iron chelators. Ann N Y Acad Sci 1990;612:327–8.
  147. Scarrow RC, Riley PE, Abu-Dari K, White DL, Raymond KN. Ferric ion sequestering agents. Synthesis, structures, and thermodynamics of complexation of cobalt(III) and iron(III) tris complexes of several chelating hydroxypyridinones. Inorg Chem 1985;24:954–67.
  148. Matsui D, Klein J, Hermann C, Grunau V, McClelland R, Chung D, et al. Relationship between the pharmacokinetics and iron excretion pharmacodynamics of the new oral iron chelator 1,2-dimethyl-3-hydroxypyrid-4-one in patients with thalassemia. Clin Pharmacol Ther 1991;50:294–8.
  149. Richardson R. Deferiprone: greater efficacy at depleting myocardial than hepatic iron? Lancet 2002;360:501-2.
  150. Anderson LJ, Wonke B, Prescott E, Holden S, Walker JM, Pennell DJ. Comparison of effects of oral deferiprone and subcutaneous desferrioxamine on myocardial iron concentrations and ventricularfunction in beta-thalassaemia. Lancet 2002;360:516–20.
  151. Piga A, Gagliotic C, Fogliacco E, Tricta F. Comparative effects of deferioprone and deferoxamine on survival and cardiac disease in patients with thalassemia major: a retrospective analysis. Haematologica 2003;88:489–96.
  152. Peng CT, Tsai CH, Wu KH, Hsu CC, Sheng YT. Improvement of cardiac function in thalassemia patients using deferiprone-Review article. Tzu Chi Med J 2007;19:192–9.
  153. Filosa A, Vitrano A, Rigano P, Calvaruso G, Barone R, Capra M, et al. Long-term treatment with deferiprone enhances left ventricular ejection function when compared to deferoxamine in patients with thalassemia major. Blood Cells Mol Dis 2013;51:85–8.
  154. Maggio A, Vitrano A, Capra M, Cuccia L, Gagliardotto F, Filosa A, et al. Improving survival with deferiprone treatment in patients with thalassemia major: A prospective multicenter randomised clinical trial under the auspices of the italian society for thalassemia and Hemoglobinopathies. Blood Cells Mol Dis 2009;42:247–51.
  155. Aqodad N, Loréal O, Erdtman L, Brissot P, Guyader D. Fatal congestive heart failure with deferiprone. Gastroenterol Clin Biol 2008;32:656-9.
  156. Ladis V, Chouliaras G, Berdoukas V. Fatal congestive heart failure with deferiprone. Gastroenterol Clin Biol 2009;33:593-4.
  157. Mallat NS, Beydoun A, Musallam KM, Koussa S. Deferiprone-induced seizures in a patient withβ-thalassemia major. Blood Cells Mol Dis 2013;51:94–5.
  158. Olivieri NF, Brittenham GM, Matsui D, Berkovitch M, Blendis LM, Cameron RG, et al. Iron chelation therapy with oral deferiprone in patients with thalassemia major. N Engl J Med 1995;332:918-22.
  159. Tondury P, Kontoghiorghes GJ, Rdolfi-Luthy A, Hirt A, Hoffbrand AV, Lottenbach AM, et al. L1(1,2-dimethyl-3-hydroxypyrid-4-one) for oral iron chelation in patients with beta-thalassemia major. Br J Haematol 1990;76:550-3.
  160. Farmaki K, Tzoumari I, Pappa C. Oral chelators in transfusion-dependent thalassemia major patients may prevent or reverse iron overload complications. Blood Cells Mol Dis 2011;47:33–40.
  161. Hoffbrand V. Deferiprone therapy for transfusional iron overload. Best Pract Res Clin Haematol 2005;18:299–317.
  162. Cohen AR, Galanello R, Piga A, De Sanctis V, Tricta F. Safety and effectiveness of long-term therapy with the oral iron chelator deferiprone. Blood 2003;102:1583–7.
  163. Ceci A, Baiardi P, Felisi M, Cappellini MD, Carnelli V, De Sanctis V, et al. The safety and effectiveness of deferiprone in a large-scale, 3-year study in Italian patients. Br J Haematol 2002;118:330-6.
  164. Agarwal MB, Gupta SS, Viswanathan C, Vasandani D, Ramanathan J, Desai N, et al. Long-term assessment of efficacy and safety of L1, an oral iron chelator, in transfusion dependent thalassemia: Indian trial. Br J Haematol 1992;82:460–6.
  165. Pennell DJ, Porter JB, Piga A, Lai Y, El-Beshlawy A, Belhoul KM, et al. A 1-year randomized controlled trial of deferasirox vs deferoxamine for myocardial iron removal in β-thalassemia major. Blood 2014;123:1447-54.
  166. Vichinsky E, Onyekwere O, Porter J, Swerdlow P, Eckman J, Lane P, et al. A randomised comparison of deferasirox versus deferoxamine for the treatment of transfusional iron overload in sickle cell disease. Br J Haematol 2007;136:501-8.
  167. Kim J, Kim Y. A time-cost augmented economic evaluation of oral deferasirox versus infusional deferoxamine for patients with iron overload in South Korea. Value Health 2009;12:78-81.
  168. Cappellini MD, Bejaoui M, Agaoglu L, Canatan D, Capra M, Cohen A, et al. Iron chelation with deferasirox in adult and pediatric patients with thalassemia major: efficacy and safety during 5 years follow-up. Blood 2011;118:884–93.
  169. Waldmeier F, Bruin GJ, Glaenzel U, Hazell K, Sechaud R, Warrington S, et al. Pharmacokinetics, metabolism, and disposition of deferasirox in beta-thalassemic patients with transfusion-dependent iron overload who are at pharmacokinetic steady state. Drug Metab Dispos 2010;38:808-16.
  170. Taher AT, Porter J, Viprakasit V, Kattamis A, Chuncharunee S, Sutcharitchan P, et al. Deferasirox reduces iron overload significantly in nontransfusion-dependent thalassemia: 1-year results from a prospective, randomized, double blind, placebo-controlled study. Blood 2012;120:970-7.
  171. Ho WL, Chung KP, Yang SS, Lu MY, Jou ST, Chang HH, et al. A pharmaco-economic evaluation of deferasirox for treating patients with iron overload caused by transfusion-dependent thalassemia in Taiwan. J Formosan Med Assoc 2013;112:221-9.
  172. Fenauxd P, Gansere A, Guerci-Breslerf A, Schmidg M, Taylorh K, Vassilieff D, et al. Deferasirox in iron-overloaded patients with transfusion-dependent myelodysplastic syndromes. Leukemia Res 2010;34:1143–50.
  173. Chang HH, Lu MY, Liao YM, Lin PC, Yang YL, Lin DT, et al. Improved efficacy and tolerability of oral deferasirox by twice-daily dosing for patients with transfusion dependent beta-thalassemia. Pediatr Blood Cancer 2011;56:420–4.
  174. Chirnomas D, Smith AL, Braunstein J, Finkelstein Y, Pereira L, Bergmann AK, et al. Deferasirox pharmacokinetics in patients with adequate versus inadequate response. Blood 2009;114:4009–13.
  175. Chen J. The youth team. Nature 2001;411:13–4.
  176. Kuhnigk O, Bothern AM, Reimer J, Schäfer I, Biegler A, Jueptner M, et al. Benefits and pitfalls of scientific research during undergraduate medical education. GMS Z Med Ausbild 2010;27:72.
  177. Pongtanakul B, Viprakasit V. Letter to the Editor. Blood Cells Mol Dis 2013;51:96–7.
  178. Meerpohl JJ, Antes G, Rücker G, Fleeman N, Motschall E, Niemeyer CM, et al. Deferasirox for managing iron overload in people with thalassemia. Cochrane Database Syst Rev 2012;2:CD007476.  doi: 10.1002/14651858.CD007476. [Article in Press]
  179. Neufeld EJ. Oral chelators deferasirox and deferiprone for transfusional iron overload in thalassemia major: new data, new questions. Blood 2006;107:3436-41.
  180. European Medicines Agency (EMA), Committee for Medicinal Products for Human Use plenary meeting monthly report. London, EMEA/CHMP/674356/2009; 2009.
  181. Stein J. Deferasirox beneficial in patients with transfusion-dependent myelodysplastic syndrome and iron overload; 2013. Available from: URL: http://www.docguide.com/deferasirox-beneficial-patients-transfusion-dependent-myelodysplastic-syndrome-and-iron-overload?tsid=6. [Last accessed on 20 Apr 2015]
  182. Guarigliaa R, Martorellia MC, Villani O, Pietrantuonoa G, Mansuetoa G, Auriaa FD, et al. Positive effects on hematopoiesis in patients with myelodysplastic syndrome receiving deferasirox as oral iron chelation therapy: A brief review. Leukemia Res 2011;35:566–70.
  183. Breccia M, Finsinger P, Latagliata R, Cannella L, Loglisci G, Federico V, et al. Deferasirox treatment in myelodysplastic syndromes: real-life efficacy and safety in a single institution patient population. Leukemia Res 2011;35:S27–S142.
  184. Cappellini MD, Bejaoui M, Agaoglu L, Canatan D, Capra M, Cohen A, et al. Iron chelation with deferasirox in adult and pediatric patients with thalassemia major: efficacy and safety during 5 years follow-up. Blood 2011;118:884–93.
  185. Lindsey WT, Olin BR. Deferasirox for transfusion-related iron overload: a clinical review. Clin Ther 2007;29:2154-66.
  186. Phatak P, Brissot P, Wurster M, Adams PC, Bonkovsky HL, Gross J, et al. A phase 1/2, dose-escalation trial of deferasirox for the treatment of iron overload in HFE-related hereditary hemochromatosis. Hepatology 2010;52:1671-779.
  187. Adlette I. Recent advances in improving the management of sickle cell disease. Blood Rev 2009;23:S9–S13.
  188. Thuret I. Post-transfusional iron overload in the haemoglobinopathies. CR Biologies 2013;336:164–72.
  189. Navneet SM, Lazarus HM, Burns LJ. A prospective study of iron overload management in allogeneic hematopoietic cell transplantation survivors. Biol Blood Marrow Transplant 2010;16:832-37.
  190. Bae SJ, Kang C, Sung KW, Chueh HW, Son MH, Lee SH. Iron overload during follow-up after tandem high-dose chemotherapy and autologous stem cell transplantation in patients with high-risk neuroblastoma. J Korean Med Sci 2012;27:363–9.
  191. Kennedy GA, Morris KL, Subramonpillai E, Curley C, Butler J, Durrant S. A prospective phase II randomized study of deferasirox to prevent iatrogenic iron overload in patients undertaking induction/consolidation chemotherapy for acute myeloid leukaemia. Br J Haematol 2013;161:794-801.
  192. McDevitt P. A viable alternative to deferoxamine. Community Oncol 2007;4:722-3.
  193. Kontoghiorghes G J. Turning a blind eye to deferasirox's toxicity? Lancet 2013;381:1183–4.
  194. Pierre JL, Baret P, Serratrice G. Hydroxyquinolines as iron chelators. Curr Med Chem 2003;10:1077–84.
  195. Zheng H, Weiner LM, Epsztejn OS, Cabantchik ZI, Warshawsky A, Youdim MB, et al. Design, synthesis, and evaluation of novel bifunctional iron-chelators as potential agents for neuroprotection in Alzheimer's, Parkinson's, and other neurodegenerative diseases. Bioorg Med Chem 2005;13:773–83.
  196. Kidani Y, Naga S, Koike H. Mass spectrometry of 5-chloro-7-iodo-8-quinol metal chelates. JPN Analyst 1974;23:1375–8.
  197. Mounsey RB, Teismann P. Chelators in the treatment of iron accumulation in Parkinson's disease. Int J Cell Biol 2012;1-12. doi.org/10.1155/2012/983245. [Article in Press]
  198. Aydinok Y, Ulger Z, Nart D, Terzi A, Cetiner N, Ellis G, et al. A randomized controlled 1-year study of daily deferiprone plus twice weekly desferrioxamine compared with daily deferiprone monotherapy in patients with thalassemia major. Haematol 2007;92:1599-606.
  199. Porcu M, Landis N, Salis S, Corda M, Orrùa P, Serraa E, et al. Effects of combined deferiprone and desferrioxamine iron chelating therapy in {beta}-thalassemia major end-stage heart failure. Eur J Heart Failure 2009;9:320-2.
  200. Telfer PT, Warburton F, Christou S, Hadjigavriel M, Sitarou M, Kolnagou A. Improved survival in thalassemia major patients on switching from desferrioxamine to combined chelation therapy with desferrioxamine and deferiprone. Haematol 2009;94:1777-8.
  201. Wonke B, Wright C, Hoffbrand AV. Combined therapy with deferiprone and desferrioxamine. Br J Haematol 1998;103:3614.
  202. Balveer K, Pyar K, Wonke B. Combined oral and parenteral iron chelation in beta thalassaemia major. Med J Malaysia 2000;55:4937.
  203. Mourad FH, Hoffbrand AV, Sheikh-Taha M, Koussa S, Khoriaty AI, Taher A. Comparison between desferrioxamine and combined therapy with desferrioxamine and deferiprone in iron overloaded thalassaemia patients. Br J Haematol 2003;121:187–9.
  204. Origa R, Bina P, Agus A, Crobu G, Defraia E, Dessì C, et al. Combined therapy with deferiprone and desferrioxamine in thalassemia major. Haematologica 2005;90:1309-14.
  205. Zareifar S, Jabbari A, Cohan N, Haghpanah S. Efficacy of combined desferrioxamine and deferiprone versus single desferrioxamine therapy in patients with major thalassemia. Arch Iran Med 2009;12:488-91.
  206. Winston A. Iron complexing bioactive polymers. Handbook of bioactive polymeric systems. In: Gebelein CG, Carraher CE. editors. Chemistry/Food Science, general. Springer US; 1985. p. 621-49.
  207. Meshchanov AY, Kol'tsova GN, Minina LT. Biodegradable polymeric hydroxamic acids for the elimination of iron from the body. Pharm Chem J 1994;28:173-6.
  208. Onsoyen E, Skaugrud OJ. Metal recovery using chitosan. Chem Technol Biotechnol 1990;49:395–404.
  209. Burke A, Yilmaz E, Hasirci N. Evaluation of chitosan as a potential medical iron (III) ion adsorbent. Turk J Med Sci 2000;30:341-8.
  210. Phillips SF, Fernandez R. Pectin and cellulose binding of iron in vitro. Am Soc Clin Nutr 1981;34:2322–3.
  211. Brzonova I, Steiner W, Zankel A, Nyanhongo GS, Guebitz GM. Enzymatic synthesis of catechol and hydroxyl-carboxylic acid functionalized chitosan microspheres for iron overload therapy. Eur J Pharm Biopharm 2011;79:294–303.
  212. Berkland C, Mohammadi Z. Polyamine-dihydroxybenzoic acid conjugate hydrogels as iron chelators. Patent publication number EP2542248A1; 2011.
  213. Hallaway PE, Eaton JW, Panter SS, Hedlund BE. Modulation of deferoxamine toxicity and clearance by covalent attachment to biocompatible polymers. Proc Natl Acad Sci 1989;86:10108-12.
  214. Harmatz P, Grady RW, Dragsten P, Vichinsky E, Giardina P, Madden J, et al. Phase Ib clinical trial of starch-conjugated deferoxamine (40SD02): a novel long-acting iron chelator. Br J Haematol 2007;138:374-8.