Increasingly, perturbations in cellular iron and ferritin are emerging as an important element in the pathogenesis of disease. These changes in ferritin are important not only in the classic diseases of iron acquisition, transport, and storage, such as primary hemochromatosis, but also in diseases characterized by inflammation, infection, injury, and repair. Among these are some of the most common diseases that afflict mankind: neurodegenerative diseases such as Parkinson disease1 and Alzheimer disease,2vascular diseases such as cardiac and neuronal ischemia-reperfusion injury,3,4 atherosclerosis itself,5 pulmonary inflammatory states,6 rheumatoid arthritis,7,8 and a variety of premalignant conditions and frank cancers.9 We are just beginning to learn the mechanisms and implications of alterations in ferritin and iron homeostasis from these natural “experiments.” However, what is increasingly apparent is that ferritin appears to be a key molecule that limits the extent, character, and location of the pro-oxidant stress that typifies inflammatory diseases, cancer, and conditions of altered oxygenation. The link between alteration in ferritin regulation and these diseases is forged through a diverse set of cellular stress pathways that alter ferritin subunit composition and/or content within cells. The 2 broad themes developed in this review are that understanding the signals and pathways that regulate ferritin may lead to insights into the pathophysiology of these diseases, and that attention to how ferritin responds to stress will in turn teach us more about the normal functions of this complex protein.

Ferritin is a ubiquitous and highly conserved iron-binding protein. In vertebrates, the cytosolic form consists of 2 subunits, termed H and L. Twenty-four ferritin subunits assemble to form the apoferritin shell. Each apoferritin molecule of 450 000 d can sequester up to approximately 4500 iron atoms.10 Depending on the tissue type and physiologic status of the cell, the ratio of H to L subunits in ferritin can vary widely, from predominantly L in such tissues as liver and spleen, to predominantly H in heart and kidney.11 The H-to-L ratio is not fixed, but is rather quite plastic: it is readily modified in many inflammatory and infectious conditions, and in response to xenobiotic stress, differentiation, and developmental transitions, as well as other stimuli. Ferritin H and L subunits are encoded by separate genes.12,13 Although a single functional H and L gene was thought to be expressed in all vertebrate species, a functional mitochondrial ferritin gene has recently been described.14Multiple pseudogenes are also present.15-17 Ferritin also has enzymatic properties, converting Fe(II) to Fe(III) as iron is internalized and sequestered in the ferritin mineral core. Use of recombinant ferritins has demonstrated that this function is an inherent feature of the H subunit of ferritin, which has a ferroxidase activity.18 The ferroxidase center is evolutionarily conserved,10 and ferroxidase activity is dramatically reduced following mutation of residues His65 and Glu62 in both human and mouse.18,19 

Small quantities of ferritin are also present in human serum, and are elevated in conditions of iron overload and inflammation.20-22 Serum ferritin is iron-poor, resembles ferritin L immunologically, and may contain a novel “G” (glycosylated) subunit.23 Despite widespread use of serum ferritin as a clinical indicator of body iron stores, little is known of the source of this ferritin. However, the increase in serum ferritin in patients with mutations in ferritin L has led to the suggestion that serum ferritin and ferritin L derive from the same gene product24 (see below).

The critical role of ferritin in cellular and organismal iron homeostasis is intimately linked to its primary and best-studied function—iron sequestration. Iron in heme is necessary for the transport, binding, and release of oxygen; the ready availability of iron for incorporation to heme is essential to organismal survival. Iron is also essential for the function of enzymes that participate in numerous critical cellular processes, including the cell cycle, the reductive conversion of ribonucleotides to deoxyribonucleotides, electron transport, and others. However, iron also donates electrons for the generation of the superoxide radical, and can participate in the generation of hydroxyl radicals via the Fenton reaction (Fe (II) + H2O2 → Fe (III) + OH + OH.).25 Thus, iron status dramatically affects the generation of oxygen as well as ferryl and nitrogen radicals. The toxicity of iron in cellular systems is attributable in large part to its capacity to participate in the generation of such reactive species, which can directly damage DNA, lipids, and proteins, leading to profound cellular toxicity. At an organismal level, iron balance is maintained with exquisite care.26,27 Ferritin, by capturing and “buffering” the intracellular labile iron pool,28-31 plays a key role in maintaining iron homeostasis. It is not surprising, then, that homozygous murine knockouts of ferritin H are lethal.32 

Recently, it has become evident that regulatory factors, in addition to those that regulate iron flux, have an important impact on cellular ferritin (Table 1). In fact, ferritin can be viewed not only as part of a group of iron regulatory proteins that include transferrin and the transferrin receptor, but also as a member of the protein family that orchestrates the cellular defense against stress and inflammation.33 This review focuses on the molecular mechanisms and biologic implications of ferritin regulation by cytokines, oxidants, oncogenes, growth factors, and other stimuli, as well as their relevance to the complex and still poorly understood events that perturb ferritin and iron homeostasis in a number of disease states.

Table 1.

Mechanisms and effectors of ferritin regulation

MechanismEffector
Transcriptional regulation TNF, cAMP, hematopoetic differentiation, iron, oxidative stress, chemopreventive agents,c-myc, E1A  
Posttranscriptional regulation via modulation of IRP proteins Iron, phorbol ester, nitric oxide, superoxide and hydroxyl radicals, hypoxia-reoxygenation  
Posttranscriptional regulation independent of IRP, including mRNA stability and protein stability IL-1β, phorbol ester, hemin  
Release of iron from sequestered intracellular compartments (including ferritin itself)—“secondary” iron regulation Oxidant stress (GSH depletion, menadione) 
MechanismEffector
Transcriptional regulation TNF, cAMP, hematopoetic differentiation, iron, oxidative stress, chemopreventive agents,c-myc, E1A  
Posttranscriptional regulation via modulation of IRP proteins Iron, phorbol ester, nitric oxide, superoxide and hydroxyl radicals, hypoxia-reoxygenation  
Posttranscriptional regulation independent of IRP, including mRNA stability and protein stability IL-1β, phorbol ester, hemin  
Release of iron from sequestered intracellular compartments (including ferritin itself)—“secondary” iron regulation Oxidant stress (GSH depletion, menadione) 

Mechanisms of ferritin regulation described in the literature. These are not mutually exclusive; examples are illustrative and not exhaustive. See text for references.

TNF indicates tumor necrosis factor; IRP, iron regulatory protein; IL-1β, interleukin 1 beta; GSH, glutathione.

Not only does ferritin sequester iron in a nontoxic form, but levels of “labile” iron regulate cellular ferritin levels, protecting cells from damage triggered by excess iron. Iron-mediated, largely posttranscriptional pathways of ferritin regulation have been identified through a series of elegant experiments over the last 15 years.34-42 The events that coordinate ferritin regulation are described briefly below and are illustrated in Figure1. The reader is also referred to detailed reviews of this subject.38,43 Although there is widespread agreement that these regulatory mechanisms are utilized in many cell types, additional regulatory pathways may be operant in erythroid cells, due to the specialized function of these cells in hemoglobin synthesis.44 

Fig. 1.

Foundations of ferritin biology: IRE/IRP and iron-mediated regulation.

Note model of 5′ IRE repression of ferritin translation in low-iron conditions is similar to mechanism of iron-mediated regulation of erythroid ALA synthase (e-ALAS) gene. Similarly, destabilization of TfR mRNA under high-iron conditions by binding of IRP to 3′ IRE is similar to proposed regulation of divalent metal transporter-1 (DMT-1). C-acon indicates cytosolic cis-aconitase; IRP1, iron regulatory protein 1; IRP2, iron regulatory protein 2; IRE, iron responsive element.

Fig. 1.

Foundations of ferritin biology: IRE/IRP and iron-mediated regulation.

Note model of 5′ IRE repression of ferritin translation in low-iron conditions is similar to mechanism of iron-mediated regulation of erythroid ALA synthase (e-ALAS) gene. Similarly, destabilization of TfR mRNA under high-iron conditions by binding of IRP to 3′ IRE is similar to proposed regulation of divalent metal transporter-1 (DMT-1). C-acon indicates cytosolic cis-aconitase; IRP1, iron regulatory protein 1; IRP2, iron regulatory protein 2; IRE, iron responsive element.

Close modal

The content of cytoplasmic ferritin is regulated by the translation of ferritin H and L mRNAs in response to an intracellular pool of “chelatable” or “labile” iron.45,46 Thus, when iron levels are low, ferritin synthesis is decreased; conversely, when iron levels are high, ferritin synthesis increases. Although in certain circumstances there is an increase in ferritin mRNA in response to iron,47 the regulatory response of ferritin to iron is largely posttranscriptional,48 and is due to the recruitment of stored mRNA from monosomes to polysomes in the presence of iron.46 This process is mediated by interaction between RNA binding proteins and a region in the 5′ untranslated region of ferritin H and L mRNA termed the iron responsive element (IRE) that has a “stem-loop” secondary structure. There are 2 RNA binding proteins, iron regulatory proteins 1 and 2 (IRP1 and IRP2), that bind to this stem loop structure and inhibit mRNA translation. However, the proteins are regulated differently: IRP1 is an iron-sulfur cluster protein that exists in 2 forms. When iron is abundant, it exists as a cytosolic aconitase. When iron is scarce, it assumes an open configuration associated with the loss of iron atoms in the iron-sulfur cluster, and can bind the IRE stem loop, acting as a repressor of ferritin translation. In contrast, IRP2 is regulated by degradation: IRP2 protein is abundant in iron scarcity, but is degraded rapidly in iron excess through targeting of a unique 73 amino acid sequence.49 Although both IRP1 and IRP2 bind the IRE and exert an inhibitory effect on ferritin synthesis, there is evidence that IRP1 and IRP2 may have distinct tissue-specific roles.50-52 Thus, IRP2 knockout mice exhibit a pronounced misregulation of iron metabolism in the intestinal mucosa and central nervous system,53 suggesting that the function of IRP2 in these tissues cannot be complemented by IRP1. Further, relative ratios of IRP1/IRP2 differ in a tissue-specific fashion, with IRP1 being more abundant than IRP2 in liver, kidney, intestine, and brain, and less abundant in pituitary and a pro–B-lymphocytic cell line.51 Action of IRP proteins can be further modulated through the activation of signal transduction cascades. For example, activation of protein kinase C (PKC) by phorbol esters phosphorylates IRP1 and increases its binding to the IRE.54 Similarly, PKC can activate IRP2, but through phosphorylation of different serine residues.54 Recently, a protein distinct from IRP1 and IRP2 that binds to a 5′ stem loop structure of mitochondrial complex 1 has been identified.55 

Perhaps the most interesting feature of the IRE-IRP interaction is the conservation of the IRE sequence in other genes that regulate iron homeostasis. For example, the 3′ untranslated region (UTR) of the transferrin receptor gene contains 5 tandem IRE sequences. IRE-IRP binding lengthens transferrin receptor mRNA half-life, ultimately leading to increased transferrin receptor display on the cell surface in situations of iron depletion. Thus, similar RNA-protein binding motifs can have strikingly different biologic effects when located in different positions on different genes. In the case of ferritin, IRP binding results in inhibition of translation, whereas in the transferrin receptor, IRP binding increases transferrin receptor mRNA half-life.56 In addition to ferritin H and L and transferrin receptor, inducible eALA synthase (the enzyme catalyzing the rate-limiting step in heme biosynthesis), mitochondrial aconitase, and DMT-1 (the recently discovered divalent metal transporter, also termed DCT-1 and Nramp2), have functional IRE sequences.57IRE sequences have been identified in a number of other genes including ferroportin1/IREG1/MTP127,58,59; however, whether they function in IRP binding has not yet been determined.

Before one can approach the regulation of ferritin by signals other than cellular iron, it is essential to have an understanding of the overall gene structure of ferritin. This topic, as well as ferritin protein structure, has been extensively reviewed, and the reader is referred to excellent and detailed reviews of the protein, its crystal structure, and the implications for iron catalysis and storage.10,60,61 

The structure of ferritin genes and proteins are highly conserved, likely due to the critical role of ferritin in the maintenance of iron homeostasis in species ranging from plants to humans. In vertebrates, the structure of cytoplasmic ferritin genes in all species studied thus far shows 3 introns and 4 exons, with the intron-exon boundaries occurring at similar locations.62 Sequences that code for a stem loop structure—the IRE—in the 5′ untranslated region of the ferritin mRNA are particularly conserved among species. A recently reported mitochondrial ferritin gene is the exception to this rule: it is intronless, and contains a domain with only weak homology to the classical IRE.14 

The importance of the IRE in ferritin regulation is highlighted by the discovery of an autosomal dominant disorder resulting in hyperferritinaemia and cataracts, which can be attributed to point mutations in the IRE of ferritin L mRNA, leading to the constitutive activation of ferritin L translation and high serum ferritin in the absence of iron excess (Table 2). A recently discovered autosomal dominant mutation in the IRE of ferritin H leads to increased affinity of the IRE for IRP, reduced ferritin H protein, and iron overload.63 A dominantly inherited mutation in the C-terminal domain of ferritin L with an associated decrease in serum ferritin and abnormal deposition of ferritin and iron in the brain has also been described: this mutation has been suggested to underlie a new syndrome termed “neuroferritinopathy.”64 

Table 2.

Human ferritin genes and ferritin gene mutations

Gene or mutationChromosomal locationMutation effectReference no.
Cytoplasmic ferritin L 19q13.3 —  
 Ferritin L IRE mutations (13 different mutations known)  Hyperferritinemia/cataract 24, 193, 194, 195, 196, 197, 198, 199, 200, 201, 202, 203  
 Ferritin L non-IRE mutations  Neuroferritinopathy 64  
Cytoplasmic ferritin H 11q12-q13 —   
 Ferritin H IRE mutations  Iron overload 63  
Mitochondrial ferritin 5q23.1 —  14  
Gene or mutationChromosomal locationMutation effectReference no.
Cytoplasmic ferritin L 19q13.3 —  
 Ferritin L IRE mutations (13 different mutations known)  Hyperferritinemia/cataract 24, 193, 194, 195, 196, 197, 198, 199, 200, 201, 202, 203  
 Ferritin L non-IRE mutations  Neuroferritinopathy 64  
Cytoplasmic ferritin H 11q12-q13 —   
 Ferritin H IRE mutations  Iron overload 63  
Mitochondrial ferritin 5q23.1 —  14  

IRE indicates iron responsive element.

In nonvertebrate animals and plants, structural analysis and functional studies in many species have either failed to reveal structures compatible with posttranscriptional regulation by iron, or have revealed sequences that are not functional IRP binding sites. Instead, transcriptional regulation of ferritin is most frequently (although not universally) identified as a primary mode of response to exogenous iron as well as stress-related signals.65-72 

The cytokine tumor necrosis factor alpha (TNFα; cachectin) is synthesized by stimulated macrophages and other cell types. Specific binding of TNFα to cell surface receptors triggers apoptotic pathways in susceptible cells; its pleiotrophic effects also mediate, in concert with other inflammatory cytokines, many of the events in inflammation and septic shock (for review see Beutler and Cerami,73Tsuji and Torti,74 and Baud and Karin75). TNFα is also an important contributor to the syndrome of cancer cachexia and other chronic inflammatory conditions,76diseases in which ferritin levels are frequently altered.5,8,77 

TNFα and interleukin 1α (IL-1α), another proinflammatory cytokine, transcriptionally induce the H chain of ferritin, suggesting that pathways related to inflammation and stress can impact on ferritin regulation33,78 (Figure 2). In mesenchymal cells (normal human skeletal myoblasts and myocytes, adipocytes, human and mouse fibroblasts), these TNFα and IL-1α effects were striking, selective for ferritin H mRNA, and resulted in the accumulation of a population of H-rich ferritin proteins, substantially altering the cellular subunit composition and content of ferritin.33 This observation provided a plausible physiologic explanation for the differential regulation of ferritin subunits in inflammation. Since cytokines markedly affect aspects of iron homeostasis in malignancy and inflammatory diseases,79 the observation that ferritin H could be selectively transcriptionally regulated provided a molecular model to explain the linkage between inflammation and the modulation of subunit composition and content of ferritin, largely inexplicable based on posttranscriptional regulation alone.

Fig. 2.

Mouse and human ferritin H genes.

In the murine gene, regulatory regions reside between 4 kb and 5 kb distal to the transcriptional start site; in the human gene, regulatory regions are just 5′ of the TATA box. However, similar sequence motifs and binding proteins suggest similar functional regulation of these genes. IF1 has not been definitively identified as binding to the G-fer region of the human ferritin H gene.115 See text for further explanation.

Fig. 2.

Mouse and human ferritin H genes.

In the murine gene, regulatory regions reside between 4 kb and 5 kb distal to the transcriptional start site; in the human gene, regulatory regions are just 5′ of the TATA box. However, similar sequence motifs and binding proteins suggest similar functional regulation of these genes. IF1 has not been definitively identified as binding to the G-fer region of the human ferritin H gene.115 See text for further explanation.

Close modal

The regulatory elements in the murine gene that respond to cytokines have been mapped in detail, and the DNA binding proteins that mediate this response have been identified (Figure 2). Ferritin H is regulated by TNFα through a cis-acting region (FER2) located 4.8 kb upstream of the transcriptional start site that binds the transcription factor NFκB.80 Two of the multiple NFκB subunits are specifically involved in this binding, p50 and p65. They bind to an NFκB full consensus sequence and to an adjacent NFκB core sequence. Both sequences are necessary for maximal ferritin H induction by TNFα.

Cytokines also have transcriptional effects on ferritin in other cell types. Ferritin induction in macrophages may be particularly important, given their central role in iron homeostasis as scavengers of old and damaged red blood cells, a critical and quantitatively important element in whole body iron turnover. In the U937 macrophage cell line, the proinflammatory cytokines TNFα and interferonγ increased mRNA for ferritin H, but not ferritin L. IL-1β had no effect on ferritin mRNA levels. TNFα uniquely resulted in a greater proportion of iron incorporated into ferritin.81 In A549 cells (a human cell line with properties of type 2 alveolar pneumocytes), TNFα and IL-1β in the presence or absence of exogenous iron, induced predominately H ferritins and increased iron uptake. In contrast, treatment with iron alone (transferrin-bound or non–transferrin-bound iron) induced the synthesis of L ferritin.82Overexpression of myotonic dystrophy protein kinase in C2C12 myocytes led to an increase in ferritin H mRNA, presumably by inducing IL-1.83 

Cytokines also regulate ferritin posttranscriptionally. Early experimental models involving induction of inflammation in rats with turpentine demonstrated increased ferritin synthesis in the liver84; an increase in ferritin synthesis but not ferritin mRNA was also seen in liver slices from turpentine-treated rats.85 A subsequent study revealed an increase in ferritin mRNA in the livers of turpentine-treated rats; in this series of experiments, concomitant nitric oxide (NO)–mediated induction of IRP activity prevented a coordinate increase in ferritin protein (see below).86 The response of ferritin in cultured hepatocytes treated with defined cytokines has also been investigated. In the HepG2 hepatic cell line, induction of ferritin synthesis was observed with a number of cytokines: IL-1β, IL-6, TNFα. In each case, desferrioxamine inhibited the increase in ferritin, suggesting a posttranscriptional induction in liver cells.87 The mechanism of this posttranscriptional regulation of ferritin by IL-1β has been evaluated in detail88; it is not mediated by protein binding to the IRE, but rather to a distinct G+C–rich region of the mRNA distinct from the IRE.89 A similar sequence is present on genes for other acute phase proteins. IL-1β also appears to affect ferritin accumulation posttranscriptionally in human astrocytoma cells and as a consequence reduces the labile iron pool.90 

Cytokines may also affect ferritin translation indirectly through their ability to induce iNOS (nitric oxide synthase) and hence increase NO.50,91,92 NO in turn causes the activation of both IRP1 and IRP2 (although effects on IRP2 appear to exhibit some cell type specificity), effects that may be particularly important in inflammatory conditions. Mechanisms hypothesized to underlie NO-mediated induction of IRP binding activity include cluster disassembly (IRP1), intracellular iron chelation (IRP1 and IRP2), or increased de novo synthesis (IRP2).93 

Secretion of ferritin is stimulated by cytokines. In primary cultured human hepatocytes, IL-1α and IL-6 induced a transient secretion of ferritin at 24 hours followed by a decline to baseline, whereas TNF treatment resulted in a sustained increase in ferritin secretion, reaching a level 10 times that found in untreated cells.94Similar effects on secreted ferritin were shown in HepG2 liver cells. In this study, both TNFα and IL-1β induced the secretion of ferritin, and the combination was at least additive. Iron in the form of either iron-dextran or ferric nitrilotriacetate (FeNTA) also induced the secretion of ferritin to about the same level as the combination of cytokines. Secretion was inhibited by brefeldin A, an inhibitor of Golgi function, and by an inhibitor of transcription.95 

Cytokines play a pivotal role in the cellular response to infection, and ferritin plays a prominent role in the cytokine response. Lipopolysaccharide (LPS; endotoxin), a component of the outer membrane of gram-negative bacteria, elicits a variety of reactions that involve ferritin. Although stimulation of a number of inflammatory cytokines is associated with LPS, the cellular reaction is complex, and in animals can involve vascular leak syndrome, the coagulation cascade, activation of complement, and prostaglandin synthesis. LPS administered endotracheally to rats induced ferritin protein but not mRNA.96 Similarly, tail vein injection of rats with LPS increased immunoreactive ferritin in the spleen.97 Viral infection with mengo picornavirus has been reported to lead to an increase in ferritin.98 Cyclopentenone prostaglandins (A-type prostaglandins A1, A2, J2, etc), which are involved in inflammatory and febrile responses as well as viral replication, induced L chain ferritin, heme oxygenase, and HSP 70 in human monocytes.99 

Ferritin is also involved in the inflammatory processes of atherosclerosis. In a study to determine the genes regulated in atherogenesis, cDNA libraries were constructed from atherosclerotic aorta and screened for genes differentially expressed in normal and atherosclerotic plaques.5 Ferritin H and L mRNAs were markedly induced in the aortas of rabbits fed an atherogenic diet for 6 weeks. In situ hybridization revealed that both H and L ferritin were induced in endothelial cells and in macrophages. Cells in culture were then used to model elements of the atherosclerotic process. In the THP-1 monocytic cell line and in aortic smooth muscle, ferritin was up-regulated by IL-1 and TNFα, but not TGF, platelet-derived growth factor (PDGF), or oxidized low-density lipoprotein (LDL).5 

Transcription of the human ferritin H gene is induced in response to both hormones and second messengers, including cAMP. Thecis-acting elements mediating these responses have mapped to a relatively small region in the proximal promoter of the human ferritin H gene (Figure 2).

There were 2 groups that identified ferritin H as a gene differentially expressed in response to thyrotropin in rodent cells.100,101 Subsequent work revealed that dibutyryl-cAMP recapitulated the effect of thyrotropin on ferritin H transcripts, albeit with different kinetics.102 Short fragments of the rat 5′ flanking region (up to 400 bp) but not longer fragments were responsive to dibutyryl-cAMP and thyrotropin in murine 3T3 cells and FRTL5 thyroid cells.103,104 Nuclear run-on assays confirmed the transcriptional effect of thyrotropin on ferritin H.105 The cAMP-dependent induction of ferritin was inhibited by ras in a rat thyroid cell line.106 

Collectively, these experiments demonstrated that thyrotropin increased ferritin H transcription, probably by elevating cAMP. cAMP-mediated induction of ferritin H transcription was further defined in human HeLa cells.107 The human cAMP-responsive region (the B-box) binds a protein complex termed B-box binding factors (Bbf), comprised of the transcription factor NFY, the coactivator p300, and the histone acetylase p300/CBP associated factor (PCAF).108,109 The adenoviral oncogene E1A reduces the formation of this complex. Overexpression of p300 in HeLa cells reverses the E1A-mediated inhibition of the ferritin promoter driven by Bbf.110 Okadaic acid, a phosphatase inhibitor, stimulates H ferritin transcription in HeLa cells by increasing the interaction between the p300 coactivator molecule and other components of Bbf.111 In cells with low expression of human ferritin H, overexpression of the histone acetylase PCAF activates transcription from the B-box of ferritin H.112 The B-box may also mediate the increase in ferritin H mRNA that occurs during spontaneous differentiation of Caco-2 colon carcinoma cells108 and vascular smooth muscle.113 Other important regulatory elements in the human ferritin H gene include a region called the A-box at position −132, which contains an SP-1 consensus sequence.114 

Since evaluation of the rodent cAMP regulatory region showed that longer promoter fragments exhibited a reduced rather than enhanced response to thyrotropin, these experiments also suggested the presence of negative cis-acting elements that may counteract the effect of cAMP and thyrotropin. Additional evidence for a negative regulator(s) of ferritin H transcription was obtained by Barresi et al.115 This group demonstrated that there is a stretch of 10 G's, which they termed “G-fer” between −272 and −291 of the human ferritin H gene. A 3-bp substitution mutation in this region increased promoter activity in HeLa cells, suggesting an inhibitory effect of this sequence on ferritin transcription. Inhibitory factor 1 (IF-1), which binds ubiquitously to G-rich sequences, was suggested to bind to this region.

Thryoid hormone may also regulate ferritin posttranscriptionally: T3 modulates the activity of IRP1, affecting its ability to bind to the ferritin IRE, possibly through induction of signal transduction cascades that result in phosphorylation of IRP1.116 T3 and TRH also induce the phosphorylation of IRP2.51 

Similarities between the murine and human ferritin H gene highlight the conservation not only of ferritin function, but of ferritin regulation across species. The murine ferritin H gene contains similar elements to those described above; however, they are located almost 5 kb 5′ to the corresponding regulatory elements identified in the human ferritin H gene (see Figure 2). The murine ferritin H gene contains a basal enhancer FER1 that also binds p300 and is inhibited by E1A.117 Contained within FER1 is a region of dyad symmetry that binds SP1, like the A-box of the human ferritin H gene. However, to date the FER1 region has not been shown to respond to cAMP.

In addition to thyroid hormone, insulin and IGF-1 have also been implicated in regulation of ferritin at the mRNA level. Insulin and IGF-1 both induced mRNA for H and L ferritin in C6 glioma cells.118 There was no additive effect on ferritin induction when both hormones were combined at the optimal concentration of each, suggesting that insulin might be acting through the IGF-1 receptor.118 In contrast to the equal induction of ferritin H and L by insulin and IGF-1, in pancreatic cells high glucose caused selective induction of ferritin H mRNA, with a 4-fold to 8-fold increase in ferritin H mRNA, a 75% to 90% decrease in ferritin L, and an overall 3-fold increase in ferritin as assayed by immunostaining.119 

Among the most carefully studied areas of ferritin biology have been the changes in ferritin and other proteins of iron metabolism that occur during hematopoietic differentiation. This emphasis is appropriate for many reasons: the availability of iron during erythropoiesis is a critical aspect of mammalian homeostasis; the unique role of macrophages and monocytes in iron handling is essential to the understanding of iron recycling; and finally, the experimental systems, derived for the most part from malignant cells, provide an excellent model for studying ferritin regulation in proliferating malignant cells and their differentiated (and less malignant) counterparts. In many of these model systems ferritin H transcription is selectively induced, leading to H-rich ferritin protein over the time course of differentiation. However, some of the reports infer transcriptional activity from the changes in steady-state levels of mRNA, without either assessment of transcription rates, mRNA stability, or evaluation of transcription from heterologous promoters. It should also be noted that inducers of cellular differentiation trigger a complex process often spanning many days, and the proximal regulators of ferritin alterations observed in many of the experiments described below have often not been identified.

HL60 promyelocytic cells reproducibly demonstrate a shift toward the accumulation of H-rich ferritin protein and mRNA with differentiation.120,121 In HL60 cells induced to differentiate into macrophages with the phorbol ester PMA, ferritin H mRNA levels increased up to 16-fold in 3 days; in cells induced to differentiate into neutrophils with dimethylsulfoxide (DMSO) there was a more modest 3-fold increase. In an HL60 cell line variant which produced phenotypes of intermediate differentiation, ferritin H and L were expressed at different stages of differentiation: H ferritin mRNA was expressed in the most differentiated promyelocytic cells, whereas early in the differentiation process L subunit induction was observed.122 

Erythroleukemia cell lines have also been investigated extensively for their ability to express ferritin during differentiation and on exposure to hemin. Much of this work has been performed using Friend erythroleukemia cells, which are mouse erythroleukemia cells that differentiate no further along the erythropoietic lineage than proerythroblasts. In response to the differentiation inducer DMSO, a biphasic transcriptional induction of ferritin H and L mRNA was observed; however, no corresponding increase in ferritin protein synthesis was detected.123,124 Hemin increased the concentration of ferritin H mRNA 10-fold and ferritin protein 20-fold in Friend leukemia cells. Protoporphyrin IX increased ferritin H mRNA but not ferritin protein, possibly due to its iron chelation effect. Ferric ammonium citrate was a less potent inducer of both ferritin H mRNA and ferritin protein than hemin, in both Friend erythroleukemia cells125 and in fibroblasts.126 Although the differentiation inducers DMSO and hexamethylenbisacetamide (HMBA) had only a 2-fold effect on ferritin mRNA or protein synthesis when used alone as differentiation inducers, when hemin was added to these inducers, a synergistic effect on ferritin was seen. The effect was both transcriptional and translational, with ferritin H and L mRNA induction of approximately 15-fold as well as a 20-fold to 25-fold increase in ferritin protein. Desferrioxamine had no effect on ferritin mRNA accumulation. Interestingly, the induction of ferritin was not associated with a decrease in transferrin receptor expression, as might be predicted through IRP-mediated posttranscriptional mechanisms. Rather, hemin and protoporphyrin IX transcriptionally induced both ferritin H and L and transferrin receptor genes. A similar finding of induction of ferritin genes associated with an increase, not decrease, in transferrin receptor was seen in erythropoietin-induced differentiation of J2E erythroid cells.127 In murine and human erythroleukemic cells, erythropoietin treatment modulated IRP, resulting in IRP activation,128 possibly via induction of a signal transduction cascade and phosphorylation of IRP.51 Taken in aggregate, these experiments with erythroleukemia model systems show that either differentiation inducers or hemin result in a transcriptional induction of both H and L ferritin genes. It is interesting that a wide range of inducers, which can direct differentiation along different hematopoietic lineages, regulate the coordinate induction of ferritin genes.

The molecular details of transcriptional regulation of human ferritin in the mouse Friend erythroleukemia system have been investigated.129 The minimum region of the ferritin H promoter that was able to confer transcriptional regulation by heme was 77 bp upstream of the TATA box. This region binds a protein complex referred to as the heme responsive factor, which was identified as NF-Y, an ubiquitous transcription factor. The CCAAT element in this region is critical, since a point mutation abolished binding to the heme responsive factor and transcriptional activation. The pathway of hemin activation was not defined, but the induction of ferritin H and L transcription proceeded in both normal and cAMP protein kinase–deficient murine erythroleukemia (MEL) cells, suggesting the cAMP pathway is not involved in this induction.130 

In addition to transcriptional regulation and posttranscriptional IRP-mediated regulation, altered ferritin mRNA stability has also been documented in hematopoetic cells. Using K562 cells in which hemin was added, the level of ferritin H and L mRNA increased 2-fold to 5-fold or 2-fold to 3-fold over 24 hours, respectively, whereas the protein increased 10-fold to 30-fold. This mRNA increase was not inhibited by desferrioxamine, suggesting that it was not mediated through chelatable iron. Further, transcription assays for ferritin H and L genes were unchanged. Although mRNA stability was not directly measured, these results led to the suggestion that changes in mRNA stability explained the ferritin mRNA rise with hemin treatment.131,132Similar conclusions were reached in the human monocytic cell line THP-1, which can be induced to differentiate into macrophages by treatment with the phorbol ester PMA; PMA treatment elicited an induction of ferritin H.133 Subsequent studies that directly investigated the stability of ferritin H mRNA found evidence that PMA stabilized ferritin H mRNA. This effect was mediated by pyrimidine-rich sequences within the 3′ UTR of the ferritin gene.134 

Recently, a novel mitochondrial ferritin gene has been reported.14 This intronless gene contains a mitochondrial localization signal and is expressed in the mitochondrial matrix. It exhibits more than 75% sequence identity to the ferritin H gene, and appears to sequester iron more avidly than cytosolic H-rich ferritins. Northern blot analysis revealed that expression of mitochondrial ferritin is normally restricted to testicular tissue. However, use of a specific antibody demonstrated that mitochondrial ferritin can also be expressed in erythroblasts from patients with X-linked sideroblastic anemia. Although work on this new form of ferritin is still in its early stages, these results suggest that mitochondrial ferritin may be induced under conditions of pathologic iron accumulation in heme-synthesizing cells.

In another area of intense recent investigation, a novel role for ferritin H in hemoglobin switching has also been proposed. Using gel mobility shift assays in K562 erythroleukemic cells, Broyles and coworkers demonstrated specific binding of a protein with properties of ferritin H to a conserved CAGTGC motif in the beta globin promoter.135 Transient transfection assays revealed that ferritin H repressed synthesis of beta globin, suggesting that ferritin may play a role in hemoglobin switching. Although a nuclear distribution and function for ferritin has not been unequivocally documented, the accumulation of reports of nuclear ferritin localization needs careful attention,136,137 particularly given the report of an E coli protein structurally related to ferritin that binds to and protects DNA from oxidative damage.138 

One of the major functions of ferritin is to limit Fe(II) available to participate in the generation of oxygen free radicals (ROS). Oxidant stress is an ever-present threat to organismal survival, both from exogenous and endogenous cellular sources; it is therefore not surprising that oxidant stress activates multiple pathways of ferritin regulation. How these pathways interact is just beginning to be understood, as is the role of ferritin in the substantial cassette of gene and protein alterations that coordinately limit oxidant toxicity.

There is strong experimental support for ferritin as a protectant against oxidant stress. Early studies demonstrated that exposure to heme induced ferritin synthesis in endothelial cells and concordantly reduced their cytotoxic response to hydrogen peroxide.139In tumor cell lines, sensitivity to oxidants was inversely correlated with ferritin protein levels; modulation of ferritin levels with hemin could alter oxidant sensitivity.140 These results are consistent with more recent observations that increased ferritin levels reduce the low molecular weight (“labile” or “regulatory”) iron pool.30 They are also consistent with observations that a reduction in ferritin sensitizes cells to pro-oxidant cytotoxicity,141 that overexpression of ferritin reduces oxidant species in cells challenged with oxidants142,143 and reduces oxidant toxicity,144 as well as the importance of ferritin H ferroxidase activity142 in limiting oxidant toxicity.

Both transcriptional and posttranscriptional mechanisms have been implicated in ferritin induction by oxidants. Oxidants induce ferritin transcription by directly targeting conserved regions of ferritin genes.145 Transcriptional induction of ferritin H and L genes was also observed in rat livers after injection with phorone, which reduces glutathione concentration and therefore limits free radical defense mechanisms.146 Oxidative stress can also contribute to ferritin induction by inactivating IRP1 through reversible oxidation of critical cysteine residues.147 However, oxidant-mediated inactivation of IRP1 is not always seen. In fact, in other experimental systems, oxidants had the opposite effect: hydrogen peroxide activated the iron responsive protein (IRP1), possibly through induction of a signaling pathway that mobilizes iron from the 4Fe-4S cubane cluster.148 This results in reduced ferritin synthesis posttranscriptionally, potentially leaving the cell more susceptible to oxidative injury.149 These observations are difficult to reconcile with a postulated role for ferritin in the protection against oxidative stress.

Recent work has offered a model that permits observations regarding IRP activation, ferritin induction, and the protection from oxidative stress to be resolved. In cultured BNLCL2 mouse liver cells following acute oxidant challenge, IRP was activated only transiently, and thus ferritin translation was only transiently inhibited. At the same time, a sustained increase in ferritin transcription was induced. The ultimate result was an increase in ferritin protein in oxidant-treated cells, a condition that permits reduction in oxidant-induced cell injury.143,145 These results emphasize that transcriptional and translational controls collaborate in the determination of ultimate cellular ferritin content (see Figure3).

Fig. 3.

Complexities in ferritin regulation.

The schematic highlights temporal changes in ferritin transcription and translation in response to oxidants and cytokines in cultured cells. Note that hydrogen peroxide induces both H and L ferritin; TNF induces H ferritin but not L ferritin in cells of mesenchymal lineage. IRPs have not been investigated temporally after TNFα treatment.

Fig. 3.

Complexities in ferritin regulation.

The schematic highlights temporal changes in ferritin transcription and translation in response to oxidants and cytokines in cultured cells. Note that hydrogen peroxide induces both H and L ferritin; TNF induces H ferritin but not L ferritin in cells of mesenchymal lineage. IRPs have not been investigated temporally after TNFα treatment.

Close modal

Oxidants may also alter ferritin transcription and translation through release of iron from cellular proteins. Oxidants, including ROS86,146,150 and nitric oxide,151 may release iron from ferritin, IRP1, or hemoglobin,152 either directly or through heme oxygenase (in the case of heme-containing proteins).39 This can lead to ferritin induction through IRP inhibition (above), and perhaps through direct iron-mediated transcriptional regulation of ferritin.153,154 Assessment of alterations in cellular iron due to breakdown of ferritin, aconitase iron clusters, heme-containing proteins, and other proteins will be immeasurably aided by recently described methods to directly measure a “labile” iron pool.31 

Although much experimental work has centered on hydrogen peroxide–mediated generation of ROS, other forms of oxidant stress also alter ferritin. For example, UV irradiation, which produces oxygen-free radicals and damages DNA, induced ferritin H mRNA, and protein.155,156 Menadione, a synthetic vitamin K derivative, and its water soluble form, menadione sodium bisulfite, were shown to induce ferritin in the rat liver, an effect which was preceded by heme oxygenase induction, as well as hydrogen peroxide generation and changes in the intrahepatic GSH pool. Menadione sodium bisulfite was also shown to decrease both IRP1 and aconitase activity in B6 fibroblasts.157 Another oxidant, oxidized lipoproteins, dramatically stimulated ferritin L in the THP-1 macrophage line.158 Various components of oxidized LDL were able to recapitulate this response, which may be mediated by peroxisome proliferator–activated receptor γ (PPARγ).159 However, other studies using oxidized LDL in THP-1 cells failed to show either ferritin H or ferritin L induction.5 

A major stress faced by hematopoietic and other cells and organisms is exposure to oxidant radicals induced by xenobiotic pro-oxidants. These include a broad range of agents, including environmental toxicants such as cigarette smoke, herbicides, pesticides, ozone, and many others. Exposure to these agents leads to the induction of a group of antioxidant protective enzymes (“phase II” enzymes), including glutathione S transferase, MnSOD, heme oxygenase, NAD(P)H quinone oxidoreductase (NQO1), gamma glutamyl cysteine synthetase, and others, that collectively act to mitigate cell damage mediated by these toxicants. Chemopreventives are synthetic or naturally occurring compounds that are intended to deliberately elicit this protective response without exerting toxic effects of their own (reviewed in Kensler et al160). Intriguingly, they are frequently mild oxidants, and as such, might be expected to induce ferritin. Indeed, Primiano et al161 showed that rats treated with the chemopreventive dithiolethione increased levels of ferritin in the liver. Further, the induction of phase II enzymes by chemopreventives is transcriptionally mediated via an “EpRE” (electrophilic response element, also termed antioxidant response element, [ARE] or oxidative stress response element [OSRE]). A sequence search revealed the presence of a putative EpRE in the murine ferritin L gene; when ligated to a reporter construct, the sequence was shown to be inducible by tert butyl hydroquinone (tBHQ), a model phase II enzyme inducer, in HEPG2 cells.162 Recent evidence indicates that ferritin H and L are induced by oltipraz, a chemopreventive agent currently in clinical trials, in murine and human liver cells.163 An induction of ferritin has also been observed in animals treated with oltipraz by gavage.164 

Tissue ischemia and cellular hypoxia have been modeled in a number of conditions and changes in ferritin documented. Hypoxia in neonatal rat oligodendrocytes and human oligodendrogliomas induced the synthesis of ferritin. This effect was not inhibited by actinomycin D, nor did mRNA levels of ferritin H mRNA change. The effect in oligodendrocytes was recapitulated by exogenous iron and blocked by desferrioxamine.165-167 Similar effects on ferritin induction were observed in a rat model of acute hypoxic/ischemic insult. Shortly after the hypoxic-ischemic insult, ferritin-positive microglia accumulate in subcortical white matter. In addition, the ratio of H to L ferritins shifts toward H-rich ferritins, especially in the hemisphere in which both hypoxic and ischemic insult was applied.3 In another ischemic-reperfusion model of the rat brain, induction of both ferritin H and L mRNAs occurred in the ischemic hemisphere, beginning 12 hours after 60 minutes of ischemia and lasting 14 days. Protein levels, as determined by immunohistochemistry, paralleled rises in mRNA; surprisingly, although overall induction of H and L mRNA was equal, the distribution of H and L ferritin mRNA as determined by in situ hybridization of rat brains was completely different.4 

Ferritin changes in hypoxia are at least in part mediated by altered regulation of the IRP proteins. IRP1 binding activity decreased under hypoxic conditions in rat hepatoma cells168; IRP1 decreased and ferritin levels increased in hypoxic mouse macrophages.169 In contrast, IRP2 activity was found to increase under similar conditions.170However, in 2 human cell lines, one group found an induction of IRP activity and a decrease in ferritin protein.171 

The period of reperfusion after ischemia is thought to be a critical period during which oxidant damage is maximal in many tissues, including heart, brain, and other organs. During postischemic reoxygenation of rat liver, early ferritin degradation was counteracted by enhanced ferritin transcription and concomitant IRP down-regulation. It was suggested that this might act to re-establish ferritin levels and limit reperfusion damage.153 Similarly, in a model of transient surgically-induced segmental intestinal ischemia reperfusion in rats, cytosolic ferritin mRNA and protein (in addition to brush border enzymes) decreased after 3 hours and 6 hours of reperfusion. By 12 hours, ferritin mRNA but not protein had increased to higher than normal levels. Ferritin appeared to be regulated both pretranslationally and translationally in response to ischemia reperfusion.172 Consistent with these findings, reoxygenation was found to induce IRP1 in a hepatoma cell line.168 

Raising levels of inspired oxygen (hyperoxia) is an effective treatment of lung diseases such as the adult respiratory distress syndrome (ARDS) as well as treatment of preterm infants with inadequate pulmonary maturation. Unfortunately, lung damage on exposure to high concentrations of oxygen is a major side effect of such treatment. Exposure of the lung to high oxygen concentration can cause the production of partially reduced oxygen species such as O2 and H2O2. When rats were exposed to hyperoxic conditions (95% O2), a selective increase in mRNA for ferritin L was seen at 24 hours and 48 hours, but not earlier. This result is reminiscent of acute iron overload in the rat.154 Indeed, hypotransferrinemic mice, which have high levels of ferritin and lactoferrin, are resistant to hyperoxia-induced lung injury.173 In contrast, at early time points (4 hours) in murine peritoneal macrophages treated with 80% O2, IRP activity was increased and ferritin synthesis, both H and L, was repressed.169 

Early views of the relationship between ferritin and cancer stem from work demonstrating an increase in total ferritin as well as a shift toward acidic (H-rich) ferritins in the serum of patients with various malignancies.174 However, subsequent evaluations of ferritin levels in tumor tissue itself have revealed a complex, perhaps disease-specific picture: for example, in some cases such as colon cancer,175 testicular seminoma,176 and breast cancer,177,178 increases in ferritin in tumor tissue versus comparable normal tissue have been reported; in other cases, including liver cancer, a decrease in ferritin is seen.179 New forms of ferritin may be involved in certain situations: Moroz and coworkers have described a novel isoform of ferritin that is elevated in the serum of patients with neoplastic breast disease as well as during pregnancy and HIV infection.180-182 

Molecular explanations for the perturbations of ferritin in cancer have been slow to advance, with often conflicting conclusions. For example, Modjtahedi et al183 studied subclones of an SW 613-W human colon carcinoma cell line that differ in their ability to develop tumors in nude mice. Differential screening revealed markedly higher expression of ferritin H but not ferritin L mRNA in the most tumorigenic cell lines. These highly tumorigenic clones had a high copy number of the c-myc gene and expressed high levels ofc-myc mRNA. Similarly, in immortalized (MCF-10F) breast cancer cells, ferritin H mRNA was elevated relative to the mortal HBEC line S-130.184 A study of breast cancer tissue also showed high levels of ferritin H mRNA in malignant tissues: in an examination of human breast tissue removed at mastectomy for breast cancer, cells with ductal hyperplasia, carcinoma-in-situ, and infiltrating ductal carcinoma showed ferritin H expression by in situ hybridization, whereas normal breast tissue had the lowest levels of expression measured.184 Conversely, cancers with the highest metastatic potential showed the lowest levels of ferritin expression in rat transitional cell carcinoma of the urothelium.185Unfortunately, ferritin protein levels were not reported in any of these studies.183-185 

A study published in 1993 was the first to demonstrate that a defined oncogene modulated ferritin H mRNA and protein.186 This experiment demonstrated that the E1A oncogene, an immortalizing nuclear oncogene of adenovirus that falls within the same complementation group as c-myc in transformation assays,187selectively repressed ferritin H.186 Subsequent experiments demonstrated that this was the result of E1A-dependent transcriptional repression of the ferritin H gene.117 

Consistent with these findings, down-regulation of ferritin H in highly tumorigenic, clonogenic c-myc–transformed, Epstein-Barr virus–immortalized B cells was seen.188 Transcriptional induction of IRP2 was also observed, leading to the suggestion that c-myctransformation maximizes iron available for proliferation through coordinate transcriptional and translational effects on proteins of iron metabolism. A key finding in this paper was that reexpression of high levels of ferritin H reverted c-myc–induced transformation, indicating that repression of ferritin H was critical to c-myc–mediated transformation.188 

Immortalization and transformation are linked to changes in the control of cellular proliferation. In some cell types, ferritin has been observed to increase in growth arrest. Thus, downregulation ofc-myc and coordinate upregulation of ferritin H was seen in U937 cells induced to differentiate with TPA (12-O-tetradecanoylphorbol 13-acetate).189 Upregulation of ferritin was also associated with induction of differentiation and growth arrest in hematopoetic systems (see above), premature replicative senescence of fibroblasts,190 and differentiation of preadipocytes to adipocytes.33 Growth suppression associated with overexpression of ferritin H has also been reported.142 However, in rat-1 fibroblast cells, microarray analysis of c-myc–responsive genes did not identify ferritin H in either the upregulated or downregulated gene profile.191 

In part, conflicting findings relating to the relationship between ferritin and cancer reflect the relative paucity of experiments in this area; additional experiments focused on defined oncogenes and the pathways they elicit will no doubt clarify these issues. Perhaps some contradictions are inevitable (and accurate), given the complexity and variability of pathophysiologic processes that underlie the collection of diseases grouped together as “cancer.”

Observations over the last few years have placed the regulation of ferritin within the broad context of cell injury and stress as well as altered growth regulation. This has led to the discovery of new hormonal and growth factors that regulate ferritin, many through transcriptional targeting of ferritin genes. The pathways that link these factors to changes in ferritin gene expression and protein levels are just beginning to be understood, but clearly involve multiple pathways, some of which are cell-type– and disease-specific: they include NFκB, cellular oxidant pathways that target the electrophilic response element, cAMP signaling pathways, as well as signals important for growth and cell cycle progression, such as the proto-oncogenec-myc and the adenovirus E1A oncoprotein. These observations are inseparably linked to classical posttranscriptional ferritin regulation, since cellular mRNA translation of ferritin genes remains dependent on cellular iron status through IRE-IRP–mediated regulation.192 How these pathways are perturbed in diseases in which ferritin dysregulation has been described, and whether changes in ferritin have a role in causation or cellular protective response, are questions that can now be reasonably posed.

Supported in part by grants DK42412 and DK57781 from the National Institutes of Health.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 U.S.C. section 1734.

1
Linert
W
Jameson
GNL
Redox reactions of neurotransmitters possibly involved in the progression of Parkinson's Disease.
J Inorg Biochem.
79
2000
319
326
2
Kondo
T
Shirasawa
T
Itoyama
Y
Mori
H
Embryonic genes expressed in Alzheimer's disease brains.
Neurosci Lett.
209
1996
157
160
3
Cheepsunthorn
P
Palmer
C
Menzies
S
Roberts
RL
Connor
JR
Hypoxic/ischemic insult alters ferritin expression and myelination in neonatal rat brains.
J Comp Neurol.
431
2001
382
396
4
Chi
SI
Wang
CK
Chen
JJ
Chau
LY
Lin
TN
Differential regulation of H- and L-ferritin messenger RNA subunits, ferritin protein and iron following focal cerebral ischemia-reperfusion.
Neuroscience.
100
2000
475
484
5
Pang
JH
Jiang
MJ
Chen
YL
et al
Increased ferritin gene expression in atherosclerotic lesions.
J Clin Invest.
97
1996
2204
2212
6
Ryan
TP
Krzesicki
RF
Blakeman
DP
et al
Pulmonary ferritin: differential effects of hyperoxic lung injury on subunit mRNA levels.
Free Radic Biol Med.
22
1997
901
908
7
Biemond
P
Swaak
AIG
Vaneijk
HG
Koster
JF
Intraarticular ferritin-bound iron in rheumatoid arthritis—a factor that increases oxygen free radical-induced tissue destruction.
Arthritis Rheum.
29
1986
1187
1193
8
Ahmadzadeh
N
Shingu
M
Nobunaga
M
Yasuda
M
Correlation of metal-binding proteins and proteinase inhibitors with immunological parameters in rheumatoid synovial fluids.
Clin Exp Rheumatol.
8
1990
547
551
9
Wu
CG
Groenink
M
Bosma
A
Reitsma
PH
vanDeventer
SJH
Chamuleau
RAFM
Rat ferritin-H: CDNA cloning, differential expression and localization during hepatocarcinogenesis.
Carcinogenesis.
18
1997
47
52
10
Harrison
PM
Arosio
P
The ferritins: molecular properties, iron storage function and cellular regulation.
Biochim Biophys Acta.
1275
1996
161
203
11
Arosio
P
Yokota
M
Drysdale
JW
Structural and immunological relationships of isoferritins in normal and malignant cells.
Cancer Res.
36
1976
1735
1739
12
Caskey
JH
Jones
C
Mills
KHG
Seligman
PA
Human ferritin gene is assigned to chromosome 19.
Proc Natl Acad Sci U S A.
80
1983
482
486
13
Worwood
M
Brook
JD
Cragg
SJ
et al
Assignment of human ferritin genes to chromosomes 11 and 19q13.3–-19qter.
Hum Genet.
69
1985
371
374
14
Levi
S
Corsi
B
Bosisio
M
et al
A human mitochondrial ferritin encoded by an intronless gene.
J Biol Chem.
270
2001
24437
24440
15
Jain
SK
Barrett
KJ
Boyd
D
Favreau
MF
Crampton
J
Drysdale
JW
Ferritin H and L chains are derived from different multigene families.
J Biol Chem.
260
1985
11762
11768
16
Lebo
RV
Kan
YW
Cheung
MC
Jain
SK
Drysdale
J
Human ferritin light chain gene sequences mapped to several sorted chromosomes.
Hum Genet.
71
1985
325
328
17
Cragg
SJ
Drysdale
J
Worwood
M
Genes for the ‘H’ subunit of human ferritin are present on a number of human chromosomes.
Hum Genet.
71
1985
108
112
18
Lawson
DM
Treffry
A
Artymiuk
PJ
et al
Identification of the ferroxidase centre in ferritin.
FEBS Lett.
254
1989
207
210
19
Rucker
P
Torti
FM
Torti
SV
Role of H and L subunits in mouse ferritin.
J Biol Chem.
271
1996
33352
33357
20
Lipschitz
DA
Cook
JD
Finch
CA
A clinical evaluation of serum ferritin as an index of iron stores.
N Eng J Med.
290
1974
1213
1216
21
Koziol
JA
Ho
NJ
Felitti
VJ
Beutler
E
Reference centiles for serum ferritin and percentage of transferrin saturation, with application to mutations of the HFE gene.
Clin Chem.
47
2001
1804
1810
22
Torti
SV
Torti
FM
Iron and ferritin in inflammation and cancer.
Adv Inorg Biochem.
10
1994
119
137
23
Santambrogio
P
Cozzi
A
Levi
S
Arosio
P
Human serum ferritin G-peptide is recognized by anti-L ferritin subunit antibodies and concanavalin-A.
Br J Haematol.
65
1987
235
237
24
Beaumont
C
Leneuve
P
Devaux
I
et al
Mutation in the iron responsive element of the L ferritin mRNA in a family with dominant hyperferritinaemia and cataract.
Nat Genet.
11
1995
444
446
25
Halliwell
B
Gutteridge
JM
Protection against oxidants in biological systems: the superoxide theory of oxygen toxicity
Free radicals in biology and medicine.
1989
86
179
Oxford University Press
New York, NY
26
Conrad
ME
Umbreit
JN
Moore
EG
Iron absorption and transport.
Am J Med Sci.
318
1999
213
229
27
Gunshin
H
Mackenzie
B
Berger
UV
et al
Cloning and characterization of a mammalian proton-coupled metal-ion transporter.
Nature.
388
1997
482
488
28
Kakhlon
O
Gruenbaum
Y
Cabantchik
ZL
Repression of the heavy ferritin chain increases the labile iron pool of human K562 cells.
Biochem J.
356
2001
311
316
29
Picard
V
Renaudie
F
Porcher
C
Hentze
MW
Grandchamp
B
Beaumont
C
Overexpression of the ferritin H subunit in cultured erythroid cells changes the intracellular iron distribution.
Blood.
87
1996
2057
2064
30
Picard
V
Epsztejn
S
Santambrogio
P
Cabantchik
ZI
Beaumont
C
Role of ferritin in the control of the labile iron pool in murine erythroleukemia cells.
J Biol Chem.
273
1998
15382
15386
31
Konijn
AM
Glickstein
H
Vaisman
B
Meyron-Holtz
EG
Slotki
IN
Cabantchik
ZI
The cellular labile iron pool and intracellular ferritin in K562 cells.
Blood.
94
1999
2128
2134
32
Ferreira
C
Bucchini
D
Martin
ME
et al
Early embryonic lethality of H ferritin gene deletion in mice.
J Biol Chem.
275
2000
3021
3024
33
Torti
SV
Kwak
EL
Miller
SC
et al
The molecular cloning and characterization of murine ferritin heavy chain, a tumor necrosis factor-inducible gene.
J Biol Chem.
263
1988
12638
12644
34
Hentze
MW
Rouault
TA
Caughman
SW
Dancis
A
Harford
JB
Klausner
RD
A cis-acting element is necessary and sufficient for translational regulation of human ferritin expression in response to iron.
Proc Natl Acad Sci U S A.
84
1987
6730
6734
35
Walden
WE
Daniels-McQueen
S
Brown
PH
et al
Translational repression in eukaryotes: partial purification and characterization of a repressor of ferritin mRNA translation.
Proc Natl Acad Sci U S A.
85
1988
9503
9507
36
Casey
JL
Hentze
MW
Koeller
DM
et al
Iron-responsive elements: regulatory RNA sequences that control mRNA levels and translation.
Science.
240
1988
924
928
37
Leibold
EA
Munro
HN
Cytoplasmic protein binds in vitro to a highly conserved sequence in the 5′ untranslated region of ferritin heavy- and light-subunit mRNAs.
Proc Natl Acad Sci U S A.
85
1988
2171
2175
38
Theil
EC
The ferritin family of iron storage proteins.
Adv Enzymol Relat Areas Mol Biol.
63
1990
421
449
39
Eisenstein
RS
Garcia-Mayol
D
Pettingell
W
Munro
HN
Regulation of ferritin and heme oxygenase synthesis in rat fibroblasts by different forms of iron.
Proc Natl Acad Sci U S A.
88
1991
688
692
40
Harrell
CM
McKenzie
AR
Patino
MM
Walden
WE
Theil
EC
Ferritin mRNA: interactions of iron regulatory element with translational regulator protein P-90 and the effect on base-paired flanking regions.
Proc Natl Acad Sci U S A.
88
1991
4166
4170
41
Gray
NK
Quick
S
Goossen
B
et al
Recombinant iron-regulatory factor functions as an iron-responsive element-binding protein, a translational repressor and an aconitase: a functional assay for translational repression and direct demonstration of the iron switch.
Eur J Biochem.
218
1993
657
667
42
Mascotti
DP
Goessling
LS
Rup
D
Thach
RE
Effects of the ferritin open reading frame on translational induction by iron.
Prog Nucleic Acid Res Mol Biol.
55
1996
121
134
43
Hentze
MW
Kuhn
LC
Molecular control of vertebrate iron metabolism: mRNA-based regulatory circuits operated by iron, nitric oxide, and oxidative stress.
Proc Natl Acad Sci U S A.
93
1996
8175
8182
44
Ponka
P
Tissue-specific regulation of iron metabolism and heme synthesis: distinct control mechanisms in erythroid cells.
Blood.
89
1997
1
25
45
Aziz
N
Munro
HN
Both subunits of rat liver ferritin are regulated at a translational level by iron induction.
Nucleic Acids Res.
14
1986
915
927
46
Rogers
J
Munro
H
Translation of ferritin light and heavy subunit mRNAs is regulated by intracellular chelatable iron levels in rat hepatoma cells.
Proc Natl Acad Sci U S A.
84
1987
2277
2281
47
Cairo
G
Bardella
L
Schiaffonati
L
Arosio
P
Levi
S
Bernelli-Zazzera
A
Multiple mechanisms of iron-induced ferritin synthesis in HeLa cells.
Biochem Biophys Res Commun.
133
1985
314
321
48
Zahringer
J
Baliga
BS
Munro
HN
Novel mechanism for translational control in regulation of ferritin synthesis by iron.
Proc Natl Acad Sci U S A.
73
1976
857
861
49
Iwai
K
Drake
SK
Wehr
NB
et al
Iron-dependent oxidation, ubiquitination, and degradation of iron regulatory protein 2: implications for degradation of oxidized proteins.
Proc Natl Acad Sci U S A.
95
1998
4924
4928
50
Eisenstein
RS
Iron regulatory proteins and the molecular control of mammalian iron metabolism.
Annu Rev Nutr.
20
2000
627
662
51
Thomson
AM
Rogers
JT
Leedman
PJ
Iron-regulatory proteins, iron-responsive elements and ferritin mRNA translation.
Int J Biochem Cell Biol.
31
1999
1139
1152
52
Ke
YH
Wu
JY
Leibold
EA
Walden
WE
Theil
EC
Loops and bulge/loops in iron-responsive element isoforms influence iron regulatory protein binding—fine-tuning of mRNA regulation?
J Biol Chem.
273
1998
23637
23640
53
LaVaute
T
Smith
S
Cooperman
S
et al
Targeted deletion of the gene encoding iron regulatory protein-2 causes misregulation of iron metabolism and neurodegenerative disease in mice.
Nat Genet.
27
2001
209
214
54
Schalinske
KL
Eisenstein
RS
Phosphorylation and activation of both iron regulatory proteins 1 and 2 in HL-60 cells.
J Biol Chem.
271
1996
7168
7176
55
Lin
E
Graziano
JH
Freyer
GA
Regulation of the 75-kDa subunit of mitochondrial complex I by iron.
J Biol Chem.
276
2001
27685
27692
56
Harford
JB
Klausner
RD
Coordinate post-transcriptional regulation of ferritin and transferrin receptor expression: the role of regulated RNA-protein interaction.
Enzyme.
44
1990
28
41
57
Sheth
S
Brittenham
GM
Genetic disorders affecting proteins of iron metabolism: clinical implications.
Annu Rev Med.
51
2000
443
464
58
Donovan
A
Brownlie
A
Zhou
Y
et al
Positional cloning of zebrafish ferroportin1 identifies a conserved vertebrate iron exporter.
Nature.
403
2000
776
781
59
McKie
AT
Marciani
P
Rolfs
A
et al
A novel duodenal iron-regulated transporter, IREG1, implicated in the basolateral transfer of iron to the circulation.
Mol Cell.
5
2000
299
309
60
Chasteen
ND
Ferritin: uptake, storage, and release of iron.
Met Ions Biol Syst.
35
1998
479
514
61
Massover
WH
Ultrastructure of ferritin and apoferritin—a review.
Micron.
24
1993
389
437
62
Harrison
PM
Ford
GC
Smith
JM
White
JL
The location of exon boundaries in the multimeric iron-storage protein ferritin.
Biol Met.
4
1991
95
99
63
Kato
J
Fujikawa
K
Kanda
M
et al
A mutation in the iron-responsive element of H ferritin mRNA causing autosomal dominant iron overload.
Am J Hum Genet.
69
2001
191
197
64
Curtis
ARJ
Fey
C
Morris
CM
et al
Mutation in the gene encoding ferritin light polypeptide causes dominant adult-onset basal ganglia disease.
Nat Genet.
28
2001
350
354
65
Schussler
P
Potters
E
Winnen
R
Michel
A
Bottke
W
Kunz
W
Ferritin mRNAs in Schistosoma mansoni do not have iron-responsive elements for post-transcriptional regulation.
Eur J Biochem.
241
1996
64
69
66
Lind
MI
Ekengren
S
Melefors
O
Soderhall
K
Drosophila ferritin mRNA: alternative RNA splicing regulates the presence of the iron-responsive element.
FEBS Lett.
436
1998
476
482
67
Georgieva
T
Dunkov
BC
Harizanova
N
Ralchev
K
Law
JH
Iron availability dramatically alters the distribution of ferritin subunit messages in Drosophila melanogaster.
Proc Natl Acad Sci U S A.
96
1999
2716
2721
68
Nichol
H
Locke
M
Secreted ferritin subunits are of two kinds in insects—molecular cloning of cDNAs encoding two major subunits of secreted ferritin from Calpodes ethlius.
Insect Biochem Mol Biol.
29
1999
999
1013
69
Proudhon
D
Wei
J
Briat
J
Theil
EC
Ferritin gene organization: differences between plants and animals suggest possible kingdom-specific selective constraints.
J Mol Evol.
42
1996
325
336
70
Wei
JZ
Theil
EC
Identification and characterization of the iron regulatory element in the ferritin gene of a plant (soybean).
J Biol Chem.
275
2000
17488
17493
71
Petit
JM
Van Wuytswinkel
O
Briat
JF
Lobréaux
S
Characterization of an iron-dependent regulatory sequence involved in the transcriptional control of AtFer1 and ZmFer1 plant ferritin genes by iron.
J Biol Chem.
276
2001
5584
5590
72
Fobis-Loisy
I
Aussel
L
Briat
JF
Post-transcriptional regulation of plant ferritin accumulation in response to iron as observed in the maize mutant ys1.
FEBS Lett.
397
1996
149
154
73
Beutler
B
Cerami
A
The biology of cachectin/TNF—a primary mediator of the host response.
Ann Rev Immunol.
7
1989
625
655
74
Tsuji
Y
Torti
FM
Tumor necrosis factor structure and function.
Cytokines in Health and Disease.
Kunkel
SL
Remick
DG
1992
131
150
Marcel Dekker
New York, NY
75
Baud
V
Karin
M
Signal transduction by tumor necrosis factor and its relatives.
Trends Cell Biol.
11
2001
372
377
76
Torti
FM
Dieckmann
B
Beutler
B
Cerami
A
Ringold
GM
A macrophage factor inhibits adipocyte gene expression—an in vitro model of cachexia.
Science.
229
1985
867
869
77
Biemond
P
Swaak
AJ
van Eijk
HG
Koster
JF
Intraarticular ferritin-bound iron in rheumatoid arthritis: a factor that increases oxygen free radical-induced tissue destruction.
Arthritis Rheum.
29
1986
1187
1193
78
Wei
Y
Miller
SC
Tsuji
Y
Torti
SV
Torti
FM
Interleukin 1 induces ferritin heavy chain in human muscle cells.
Biochem Biophys Res Commun.
169
1990
289
296
79
Konijn
AM
Hershko
C
The anemia of inflammation and chronic disease. In Desousa M, Brock JH, eds. Iron In Immunity, Cancer and Inflammation.
1988
111
144
John Wiley & Sons
New York, NY
80
Kwak
EL
Larochelle
DA
Beaumont
C
Torti
SV
Torti
FM
Role for NF-kappa B in the regulation of ferritin H by tumor necrosis factor-alpha.
J Biol Chem.
270
1995
15285
15293
81
Fahmy
M
Young
SP
Modulation of iron metabolism in monocyte cell line U937 by inflammatory cytokines: changes in transferrin uptake, iron handling and ferritin mRNA.
Biochem J.
296
1993
175
181
82
Smirnov
IM
Bailey
K
Flowers
CH
Garrigues
NW
Wesselius
LJ
Effects of TNF-alpha and IL-1 beta on iron metabolism by A549 cells and influence on cytotoxicity.
Am J Physiol Lung Cell Mol Physiol.
277
1999
L257
L263
83
Watanabe
T
Sasagawa
N
Usuki
F
et al
Overexpression of myotonic dystrophy protein kinase in C2C12 myogenic culture involved in the expression of ferritin heavy chain and interleukin-1 alpha mRNAs.
J Neurol Sci.
167
1999
26
33
84
Konijn
AM
Carmel
N
Levy
R
Hershko
C
Ferritin synthesis in inflammation. II. Mechanism of increased ferritin synthesis.
Br J Haematol.
49
1981
361
370
85
Schiaffonati
L
Rappocciolo
E
Tacchini
L
et al
Mechanisms of regulation of ferritin synthesis in rat liver during experimental inflammation.
Exp Mol Pathol.
48
1988
174
181
86
Pietrangelo
A
Casalgrandi
G
Quaglino
D
et al
Duodenal ferritin synthesis in genetic hemochromatosis [published erratum appears in Gastroenterology. 1995;108:1963].
Gastroenterology.
108
1995
208
217
87
Hirayama
M
Kohgo
Y
Kondo
H
et al
Regulation of iron metabolism in HepG2 cells—a possible role for cytokines in the hepatic deposition of iron.
Hepatology.
18
1993
874
880
88
Rogers
JT
Bridges
KR
Durmowicz
GP
Glass
J
Auron
PE
Munro
HN
Translational control during the acute phase response: ferritin synthesis in response to interleukin-1.
J Biol Chem.
265
1990
14572
14578
89
Rogers
JT
Andriotakis
JL
Lacroix
L
Durmowicz
GP
Kasschau
KD
Bridges
KR
Translational enhancement of H-ferritin mRNA by interleukin-1 beta acts through 5′ leader sequences distinct from the iron responsive element.
Nucleic Acids Res.
22
1994
2678
2686
90
Pinero
DJ
Hu
J
Cook
BM
Scaduto
RC
Connor
JR
Interleukin-1 beta increases binding of the iron regulatory protein and the synthesis of ferritin by increasing the labile iron pool.
Biochim Biophys Acta.
1497
2000
279
288
91
Weiss
G
Goossen
B
Doppler
W
et al
Translational regulation via iron-responsive elements by the nitric oxide NO-synthase pathway.
EMBO J.
12
1993
3651
3657
92
Drapier
JC
Hirling
H
Wietzerbin
J
Kaldy
P
Kuhn
L
Biosynthesis of nitric oxide activates iron regulatory factor in macrophages.
EMBO J.
12
1993
3643
3649
93
Domachowske
JB
Rafferty
SP
Singhania
N
Mardiney
M
Malech
HL
Nitric oxide alters the expression of gamma-globin, H-ferritin, and transferrin receptor in human K562 cells at the posttranscriptional level.
Blood.
88
1996
2980
2988
94
Muntanerelat
J
Ourlin
JC
Domergue
J
Maurel
P
Differential effects of cytokines on the inducible expresion of CYP1A1, CYP1A2 and CYP3A4 in human hepatocytes in primary culture.
Hepatology.
22
1995
1143
1153
95
Tran
TN
Eubanks
SK
Schaffer
KJ
Zhou
CY
Linder
MC
Secretion of ferritin by rat hepatoma cells and its regulation by inflammatory cytokines and iron.
Blood.
90
1997
4979
4986
96
Carraway
MS
Ghio
AJ
Taylor
JL
Piantadosi
CA
Induction of ferritin and heme oxygenase-1 by endotoxin in the lung.
Am J Physiol Lung Cell Mol Physiol.
275
1998
L583
L592
97
Kumagai
T
Awai
M
Okada
S
Mobilization of iron and iron-related proteins in rat spleen after intravenous injection of lipopolysaccharides (LPS).
Pathol Res Pract.
188
1992
931
941
98
Mulvey
MR
Kuhn
LC
Scraba
DG
Induction of ferritin synthesis in cells infected with mengo virus.
J Biol Chem.
271
1996
9851
9857
99
Elia
G
Polla
B
Rossi
A
Santoro
MG
Induction of ferritin and heat shock proteins by prostaglandin A1 in human monocytes—evidence for transcriptional and post-transcriptional regulation.
Eur J Biochem.
264
1999
736
745
100
Ursini
MV
de Franciscis
V
TSH regulation of ferritin H chain messenger RNA levels in the rat thyroids.
Biochem Biophys Res Commun.
150
1988
287
295
101
Cox
F
Gestautas
J
Rapoport
B
Molecular cloning of cDNA corresponding to mRNA species whose steady state levels in the thyroid are enhanced by thyrotropin: homology of one of these sequences with ferritin H.
J Biol Chem.
263
1988
7060
7067
102
Chazenbalk
GD
Wadsworth
HL
Rapoport
B
Transcriptional regulation of ferritin H messenger RNA levels in FRTL5 rat thyroid cells by thyrotropin.
J Biol Chem.
265
1990
666
670
103
Chazenbalk
GD
Wadsworth
HL
Foti
D
Rapoport
B
Thyrotropin and adenosine 3′,5′-monophosphate stimulate the activity of the ferritin-H promoter.
Mol Endocrinol.
4
1990
1117
1124
104
Coluccidamato
LG
Ursini
MV
Colletta
G
Cirafici
A
Defranciscis
V
Thyrotropin stimulates transacription from the ferritin heavy chain promoter.
Biochem Biophys Res Commun.
165
1989
506
511
105
Iwasa
Y
Aida
K
Yokomori
N
Inoue
M
Onaya
T
Transcriptional regulation of ferritin heavy chain messenger RNA expression by thyroid hormone.
Biochem Biophys Res Commun.
167
1990
1279
1285
106
Gallo
A
Feliciello
A
Varrone
A
Cerillo
R
Gottesman
ME
Avvedimento
VE
Ki-ras oncogene interferes with the expression of cyclic AMP-dependent promoters.
Cell Growth Differ.
6
1995
91
95
107
Bevilacqua
MA
Faniello
MC
Russo
T
Cimino
F
Costanzo
F
Transcriptional regulation of the human H ferritin-encoding gene (FERH) in G418-treated cells: role of the B-box-binding factor.
Gene.
141
1994
287
291
108
Bevilacqua
MA
Faniello
MC
D'Agostino
P
et al
Transcriptional activation of the H-ferritin gene in differentiated Caco-2 cells parallels a change in the activity of the nuclear factor Bbf.
Biochem J.
311
1995
769
773
109
Faniello
MC
Bevilacqua
MA
Condorelli
G
et al
The B subunit of the CAAT-binding factor NFY binds the central segment of the co-activator p300.
J Biol Chem.
274
1999
7623
7626
110
Bevilacqua
MA
Faniello
MC
Quaresima
B
et al
A common mechanism underlying the E1A repression and the cAMP stimulation of the H ferritin transcription.
J Biol Chem.
272
1997
20736
20741
111
Bevilacqua
MA
Faniello
MC
Cimino
F
Costanzo
F
Okadaic acid stimulates H ferritin transcription in HeLa cells by increasing the interaction between the p300 Co-activator molecule and the transcription factor Bbf.
Biochem Biophys Res Commun.
240
1997
179
182
112
Bevilacqua
MA
Faniello
MC
Russo
T
Cimino
F
Costanzo
F
P/CAF/p300 complex binds the promoter for the heavy subunit of ferritin and contributes to its tissue-specific expression.
Biochem J.
335
1998
521
525
113
Liau
G
Chan
LM
Feng
P
Increased ferritin gene expression is both promoted by cAMP and a marker of growth arrest in rabbit vascular smooth muscle cells.
J Biol Chem.
266
1991
18819
18826
114
Bevilacqua
MA
Giordano
M
D'Agostino
P
Santoro
C
Cimino
F
Costanzo
F
Promoter for the human ferritin heavy chain-encoding gene (FERH): structural and functional characterization.
Gene.
111
1992
255
260
115
Barresi
R
Sirito
M
Karsenty
G
Ravazzolo
R
A negative cis-acting G-fer element participates in the regulation of expression of the human H-ferritin-encoding gene (FERH).
Gene.
140
1994
195
201
116
Leedman
PJ
Stein
AR
Chin
WW
Rogers
JT
Thyroid hormone modulates the interaction between iron regulatory proteins and the ferritin mRNA iron-responsive element.
J Biol Chem.
271
1996
12017
12023
117
Tsuji
Y
Akebi
N
Lam
TK
Nakabeppu
Y
Torti
SV
Torti
FM
FER-1, an enhancer of the ferritin H gene and a target of E1A-mediated transcriptional repression.
Mol Cell Biol.
15
1995
5152
5164
118
Yokomori
N
Iwasa
Y
Aida
K
Inoue
M
Tawata
M
Onaya
T
Transcriptional regulation of ferritin messenger ribonucleic acid levels by insulin in cultured rat glioma cells.
Endocrinology.
128
1991
1474
1480
119
Macdonald
MJ
Cook
JD
Epstein
Ml
Flowers
CH
Large amount of (apo)ferritin in the pancreatic insulin cell and its stimulation by glucose.
FASEB J.
8
1994
777
781
120
Dorner
MH
Silverstone
AE
Desostoa
A
Munn
G
Desousa
M
Relative subunit composition of the ferritin synthesized by selected human lymphomyeloid cell populations.
Exp Hematol.
11
1983
866
872
121
Chou
CC
Gatti
RA
Fuller
ML
et al
Structure and expression of ferritin genes in a human promyelocytic cell line that differentiates in vitro.
Mol Cell Biol.
6
1986
566
573
122
Cayre
Y
Raynal
MC
Darzynkiewicz
Z
Dorner
MH
Model for intermediate steps in monocytic differentiation: c-myc, c-fms, and ferritin as markers.
Proc Natl Acad Sci U S A.
84
1987
7619
7623
123
Beaumont
C
Jain
SK
Bogard
M
Nordmann
Y
Drysdale
J
Ferritin synthesis in differentiating Friend erythroleukemic cells.
J Biol Chem.
262
1987
10619
10623
124
Beaumont
C
Dugast
I
Renaudie
F
Souroujon
M
Grandchamp
B
Transcriptional regulation of ferritin H and L subunits in adult erythroid and liver cells from the mouse: unambiguous identification of mouse ferritin subunits and in vitro formation of the ferritin shells.
J Biol Chem.
264
1989
7498
7504
125
Coccia
EM
Profita
V
Fiorucci
G
et al
Modulation of ferritin H-chain expression in Friend erythroleukemia cells: transcriptional and translational regulation by hemin.
Mol Cell Biol.
12
1992
3015
3022
126
Coccia
EM
Stellacci
E
Orsatti
R
Testa
U
Battistini
A
Regulation of ferritin H-chain expression in differentiating Friend leukemia cells.
Blood.
86
1995
1570
1579
127
Busfield
SJ
Tilbrook
PA
Callus
BA
Spadaccini
A
Kuhn
L
Klinken
SP
Complex regulation of transferrin receptors during erythropoietin-induced differentiation of J2E erythroid cells—elevated transcription and mRNA stabilisation produce only a modest rise in protein content.
Euro J Biochem.
249
1997
77
84
128
Weiss
G
Houston
T
Kastner
S
Johrer
K
Grunewald
K
Brock
JH
Regulation of cellular iron metabolism by erythropoietin: activation of iron-regulatory protein and upregulation of transferrin receptor expression in erythroid cells.
Blood.
89
1997
680
687
129
Marziali
G
Perrotti
E
Ilari
R
Testa
U
Coccia
EM
Battistini
A
Transcriptional regulation of the ferritin heavy-chain gene: the activity of the CCAAT binding factor NF-Y is modulated in heme-treated Friend leukemia cells and during monocyte-to-macrophage differentiation.
Mol Cell Biol.
17
1997
1387
1395
130
Pilz
RB
Impaired erythroid-specific gene expression in cAMP-dependent protein kinase-deficient murine erythroleukemia cells.
J Biol Chem.
268
1993
20252
20258
131
Mattia
E
den Blaauwen
J
Ashwell
G
van Renswoude
J
Multiple post-transcriptional regulatory mechanisms in ferritin gene expression.
Proc Natl Acad Sci U S A.
86
1989
1801
1805
132
Mattia
E
den Blaauwen
J
van Renswoude
J
Role of protein synthesis in the accumulation of ferritin mRNA during exposure of cells to iron.
Biochem J.
267
1990
553
555
133
Scaccabarozzi
A
Arosio
P
Weiss
G
et al
Relationship between TNF-α and iron metabolism in differentiating human monocytic THP-1 cells.
Br J Haematol.
110
2000
978
984
134
Ai
LS
Chau
LY
Post-transcriptional regulation of H-ferritin mRNA—identification of a pyrimidine-rich sequence in the 3′-untranslated region associated with message stability in human monocytic THP-1 cells.
J Biol Chem.
274
1999
30209
30214
135
Broyles
RH
Belegu
V
DeWitt
CR
et al
Specific repression of β-globin promoter activity by nuclear ferritin.
Proc Natl Acad Sci U S A.
98
2001
9145
9150
136
Cai
CX
Birk
DE
Linsenmayer
TF
Ferritin is a developmentally regulated nuclear protein of avian corneal epithelial cells.
J Biol Chem.
272
1997
12831
12839
137
Cai
CX
Birk
DE
Linsenmayer
TF
Nuclear ferritin protects DNA from UV damage in corneal epithelial cells.
Mol Biol Cell.
9
1998
1037
1051
138
Grant
RA
Filman
DJ
Finkel
SE
Kolter
R
Hogle
JM
The crystal structure of Dps, a ferritin homolog that binds and protects DNA.
Nat Struct Biol.
5
1998
294
303
139
Balla
G
Jacob
HS
Balla
J
et al
Ferritin: a cytoprotective antioxidant strategem of endothelium.
J Biol Chem.
267
1992
18148
18153
140
Cermak
J
Balla
J
Jacob
HS
et al
Tumor cell heme uptake induces ferritin synthesis resulting in altered oxidant sensitivity: possible role in chemotherapy efficacy.
Cancer Res.
53
1993
5308
5313
141
Orino
K
Tsuji
Y
Torti
FM
Torti
SV
Adenovirus E1A blocks oxidant-dependent ferritin induction and sensitizes cells to pro-oxidant cytotoxicity.
FEBS Lett.
461
1999
334
338
142
Cozzi
A
Corsi
B
Levi
S
Santambrogio
P
Albertini
A
Arosio
P
Overexpression of wild type and mutated human ferritin H-chain in HeLa cells—in vivo role of ferritin ferroxidase activity.
J Biol Chem.
275
2000
25122
25129
143
Orino
K
Lehman
L
Tsuji
Y
Ayaki
H
Torti
SV
Torti
FM
Ferritin and the response to oxidative stress.
Biochem J.
357
2001
241
247
144
Epsztejn
S
Glickstein
H
Picard
V
et al
H-ferritin subunit overexpression in erythroid cells reduces the oxidative stress response and induces multidrug resistance properties.
Blood.
94
1999
3593
3603
145
Tsuji
Y
Ayaki
H
Whitman
SP
Morrow
CS
Torti
SV
Torti
FM
Coordinate transcriptional and translational regulation of ferritin in response to oxidative stress.
Mol Cell Biol.
20
2000
5818
5827
146
Cairo
G
Tacchini
L
Pogliaghi
G
Anzon
E
Tomasi
A
Bernelli-Zazzera
A
Induction of ferritin synthesis by oxidative stress: transcriptional and post-transcriptional regulation by expansion of the “free” iron pool.
J Biol Chem.
270
1995
700
703
147
Cairo
G
Castrusini
E
Minotti
G
Bernelli-Zazzera
A
Superoxide and hydrogen peroxide-dependent inhibition of iron regulatory protein activity: a protective stratagem against oxidative injury.
FASEB J.
10
1996
1326
1335
148
Pantopoulos
K
Hentze
MW
Activation of iron regulatory protein-1 by oxidative stress in vitro.
Proc Natl Acad Sci U S A.
95
1998
10559
10563
149
Pantopoulos
K
Hentze
MW
Rapid responses to oxidative stress mediated by iron regulatory protein.
EMBO J.
14
1995
2917
2924
150
Minotti
G
Recalcati
S
Mordente
A
et al
The secondary alcohol metabolite of doxorubicin irreversibly inactivates aconitase iron regulatory protein-1 in cytosolic fractions from human myocardium.
FASEB J.
12
1998
541
552
151
Kennedy
MC
Antholine
W
Beinert
H
An EPR investigation of the products of the reaction of cytosolic and mitochodrial aconitases with nitric oxide.
J Biol Chem.
272
1997
20340
20347
152
Balla
J
Jacob
HS
Balla
G
Nath
K
Vercellotti
GM
Endothelial cell heme oxygenase and ferritin induction by heme proteins: a possible mechanism limiting shock damage.
Trans Assoc Am Physicians.
105
1992
1
6
153
Tacchini
L
Recalcati
S
Bernelli-Zazzera
A
Cairo
G
Induction of ferritin synthesis in ischemic-reperfused rat liver: analysis of the molecular mechanisms.
Gastroenterology.
113
1997
946
953
154
White
K
Munro
HN
Induction of ferritin subunit synthesis by iron is regulated at both the transcriptional and translational levels.
J Biol Chem.
263
1988
8938
8942
155
Rosen
CF
Poon
R
Drucker
DJ
UVB radiation-activated genes induced by transcriptional and posttranscriptional mechanisms in rat keratinocytes.
Am J Physiol Cell Physiol.
37
1995
C846
C855
156
Applegate
LA
Scaletta
C
Panizzon
R
Frenk
E
Evidence that ferritin is UV inducible in human skin: part of a putative defense mechanism.
J Invest Dermatol.
111
1998
159
163
157
Gehring
NH
Hentze
MW
Pantopoulos
K
Inactivation of both RNA binding and aconitase activities of iron regulatory protein-1 by quinone-induced oxidative stress.
J Biol Chem.
274
1999
6219
6225
158
Ossola
JO
Kristoff
G
Tomaro
ML
Heme oxygenase induction by menadione bisulfite adduct-generated oxidative stress in rat liver.
Comp Biochem Physiol.
127
2000
91
99
159
Jang
MK
Choi
MS
Park
YB
Regulation of ferritin light chain gene expression by oxidized low-density lipoproteins in human monocytic THP-1 cells.
Biochem Biophys Res Commun.
265
1999
577
583
160
Kensler
TW
Groopman
JD
Sutter
TR
Curphey
TJ
Roebuck
BD
Development of cancer chemopreventive agents: oltipraz as a paradigm.
Chem Res Toxicol.
12
1999
113
126
161
Primiano
T
Kensler
TW
Kuppusamy
P
Zweier
JL
Sutter
TR
Induction of hepatic heme oxygenase-1 and ferritin in rats by cancer chemopreventive dithiolethiones.
Carcinogenesis.
17
1996
2291
2296
162
Wasserman
WW
Fahl
WE
Functional antioxidant responsive elements.
Proc Natl Acad Sci U S A.
94
1997
5361
5366
163
Pietsch
EC
Morrow
C
Torti
FM
Torti
SV
Oltipraz-mediated induction of ferritin expression.
Proc Amer Assoc Cancer Res.
40
1999
258
164
Wilkinson
J
Everley
LC
Pfeiffer
GR
et al
Effect of cfos genotype and oltipraz treatment on phase II protein expression in mice.
Proc Amer Assoc Cancer Res.
42
2001
208
165
Qi
Y
Dawson
G
Hypoxia induces synthesis of a novel 22 kDa protein in neonatal rat oligodendrocytes.
J Neurochem.
59
1992
1709
1716
166
Qi
Y
Dawson
G
Hypoxia specifically and reversibly induces the synthesis of ferritin in oligodendrocytes and human oligodendrogliomas.
J Neurochem.
63
1994
1485
1490
167
Qi
Y
Jamindar
TM
Dawson
G
Hypoxia alters iron homeostasis and induces ferritin synthesis in oligodendrocytes.
J Neurochem.
64
1995
2458
2464
168
Hanson
ES
Leibold
EA
Regulation of iron regulatory protein 1 during hypoxia and hypoxia/reoxygenation.
J Biol Chem.
273
1998
7588
7593
169
Kuriyama-Matsumura
K
Sato
H
Suzuki
M
Bannai
S
Effects of hyperoxia and iron on iron regulatory protein-1 activity and the ferritin synthesis in mouse peritoneal macrophages.
Biochim Biophys Acta Protein Struct Mol Enzymol.
1544
2001
370
377
170
Hanson
ES
Foot
LM
Leibold
EA
Hypoxia post-translationally activates iron-regulatory protein 2.
J Biol Chem.
274
1999
5047
5052
171
Toth
I
Yuan
LP
Rogers
JT
Boyce
H
Bridges
KR
Hypoxia alters iron-regulatory protein-1 binding capacity and modulates cellular iron homeostasis in human hepatoma and erythroleukemia cells.
J Biol Chem.
274
1999
4467
4473
172
Yeh
KY
Yeh
M
Glass
J
Expression of intestinal brush-border membrane hydrolases and ferritin after segmental ischemia-reperfusion in rats.
Am J Physiol Gastrointest Liver Physiol.
38
1998
G572
G583
173
Yang
FM
Coalson
JJ
Bobb
HH
Carter
JD
Banu
J
Ghio
AJ
Resistance of hypotransferrinemic mice to hyperoxia-induced lung injury.
Am J Physiol Lung Cell Mol Physiol.
277
1999
L1214
L1223
174
Hazard
JT
Drysdale
JW
Ferritinemia in cancer.
Nature.
265
1977
755
756
175
Vaughn
CB
Weinstein
R
Bond
B
et al
Ferritin content in human cancerous and noncancerous colonic tissue.
Cancer Invest.
5
1987
7
10
176
Cohen
C
Shulman
G
Budgeon
LR
Immunohistochemical ferritin in testicular seminoma.
Cancer.
54
1984
2190
2194
177
Guner
G
Kirkali
G
Yenisey
C
Tore
IR
Cytosol and serum ferritin in breast carcinoma.
Cancer Lett.
67
1992
103
112
178
Weinstein
RE
Bond
BH
Silberberg
BK
Tissue ferritin concentration in carcinoma of the breast.
Cancer.
50
1982
2406
2409
179
Zhou
XD
DeTolla
L
Custer
RP
London
WT
Iron, ferritin, hepatitis B surface and core antigens in the livers of Chinese patients with hepatocellular carcinoma.
Cancer.
59
1987
1430
1437
180
Shterman
N
Kupfer
B
Moroz
C
Expression of messenger RNA species coding for a Mr 43 000 peptide associated with ferritin in human leukemia-K562 cells and its down regulation during differentiation.
Cancer Res.
49
1989
5033
5036
181
Rosen
HR
Moroz
C
Reiner
A
et al
Expression of p43 in breast cancer tissue, correlation with prognostic parameters.
Cancer Lett.
67
1992
35
45
182
Maymon
R
Jauniaux
E
Greenwold
N
Moroz
C
Localization of p43 placental isoferritin in human maternal-fetal tissue interface.
Am J Obstet Gynecol.
182
2000
670
674
183
Modjtahedi
N
Frebourg
T
Fossar
N
Lavialle
C
Cremisi
C
Brison
O
Increased expression of cytokeratin and ferritin-H genes in tumorigenic clones of the SW 613-S human colon carcinoma cell line.
Exp Cell Res.
201
1992
74
82
184
Higgy
NA
Salicioni
AM
Russo
IH
Zhang
PL
Russo
J
Differential expression of human ferritin H chain gene in immortal human breast epithelial MCF-10F cells.
Mol Carcinog.
20
1997
332
339
185
Vet
JA
van Moorselaar
RJ
Debruyne
FM
Schalken
JA
Differential expression of ferritin heavy chain in a rat transitional cell carcinoma progression model.
Biochim Biophys Acta.
1360
1997
39
44
186
Tsuji
Y
Kwak
E
Saika
T
Torti
SV
Torti
FM
Preferential repression of the H subunit of ferritin by adenovirus E1A in NIH-3T3 mouse fibroblasts.
J Biol Chem.
268
1993
7270
7275
187
Land
H
Parada
LF
Weinberg
RA
Tumorigenic conversion of primary embryo fibroblasts requires at least two cooperating oncogenes.
Nature.
304
1983
596
602
188
Wu
KJ
Polack
A
Dalla-Favera
R
Coordinated regulation of iron-controlling genes, H-ferritin and IRP2, by c-MYC.
Science.
283
1999
676
679
189
Larsson
A
Flodin
M
Kollberg
H
Increased serum concentrations of carbohydrate-deficient transferrin (CDT) in patients with cystic fibrosis.
Ups J Med Sci.
103
1998
231
236
190
Thweatt
R
Murano
S
Fleischmann
RD
Goldstein
S
Isolation and characterization of gene-sequences overexpressed in Werner syndrome fibroblasts during premature replicative senescence.
Exp Gerontol.
27
1992
433
440
191
Guo
QM
Malek
RL
Kim
S
et al
Identification of c-Myc responsive genes using rat cDNA microarray.
Cancer Res.
60
2000
5922
5928
192
Miller
LL
Miller
SC
Torti
SV
Tsuji
Y
Torti
FM
Iron-independent induction of ferritin H chain by tumor necrosis factor.
Proc Natl Acad Sci U S A.
88
1991
4946
4950
193
Cremonesi
L
Fumagalli
A
Soriani
N
et al
Double-gradient denaturing gradient gel electrophoresis assay for identification of L-ferritin iron-responsive element mutations responsible for hereditary hyperferritinemia-cataract syndrome: identification of the new mutation C14G.
Clin Chem.
47
2001
491
497
194
Cazzola
M
Bergamaschi
G
Tonon
L
et al
Hereditary hyperferritinemia-cataract syndrome: relationship between phenotypes and specific mutations in the iron-responsive element of ferritin light-chain mRNA.
Blood.
90
1997
814
821
195
Arosio
C
Fossati
L
Viganò
M
Trombini
P
Cazzaniga
G
Piperno
A
Hereditary hyperferritinemia cataract syndrome: a de novo mutation in the iron responsive element of the L-ferritin.
Haematologica.
84
1999
560
561
196
Mumford
AD
Vulliamy
T
Lindsay
J
Watson
A
Hereditary hyperferritinemia-cataract syndrome: two novel mutations in the L-ferritin iron-responsive element.
Blood.
91
1998
367
368
197
Girelli
D
Corrocher
R
Bisceglia
L
et al
Molecular basis for the recently described hereditary hyperferritinemia-cataract syndrome: a mutation in the iron-responsive element of ferritin L-subunit gene (the “Verona mutation”).
Blood.
86
1995
4050
4053
198
Martin
ME
Fargion
S
Brissot
P
Pellat
B
Beaumont
C
A point mutation in the bulge of the iron-responsive element of the L ferritin gene in two families with the hereditary hyperferritinemia-cataract syndrome.
Blood.
91
1998
319
323
199
Cicilano
M
Zecchina
G
Roetto
A
et al
Recurrent mutations in the iron regulatory element of L-ferritin in hereditary hyperferritinemia-cataract syndrome.
Haematologica.
84
1999
489
492
200
Balas
A
Aviles
MJ
Garcia-Sanchez
F
Vicario
JL
Cervera
A
Description of a new mutation in the L-ferritin iron-responsive element associated with hereditary hyperferritinemia-cataract syndrome in a Spanish family.
Blood.
93
1999
4020
4021
201
Aguilar-Martinez
P
Biron
C
Masmejean
C
Jeanjean
P
Schved
JF
A novel mutation in the iron responsive element of ferritin L-subunit gene as a cause for hereditary hyperferritinemia-cataract syndrome.
Blood.
88
1996
1895
202
Barton
JC
Beutler
E
Gelbart
T
Coinheritance of alleles associated with hemochromatosis and hereditary hyperferritinemia-cataract syndrome.
Blood.
92
1998
4480
203
Camaschella
C
Zecchina
G
Lockitch
G
et al
A new mutation (G51C) in the iron-responsive element (IRE) of L-ferritin associated with hyperferritinaemia-cataract syndrome decreases the binding affinity of the mutated IRE for iron-regulatory proteins.
Br J Haematol.
108
2000
480
482

Author notes

Frank M. Torti, Comprehensive Cancer Center, Wake Forest University School of Medicine, Medical Center Blvd, Winston-Salem, NC 27157; e-mail: ftorti@wfubmc.edu.

Sign in via your Institution