ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Regulation of iron balance

Regulation of iron balance
Literature review current through: Jan 2024.
This topic last updated: May 02, 2023.

INTRODUCTION — A tight regulation of iron balance is essential to avoid both iron deficiency and overload. The regulation of iron metabolism involves the interaction of a number of specific proteins as well as the interplay between iron absorption, recycling, and iron loss. This topic review will discuss these factors [1].

Clinical implications are discussed separately:

Iron deficiency – (See "Iron deficiency in infants and children <12 years: Treatment" and "Iron requirements and iron deficiency in adolescents" and "Causes and diagnosis of iron deficiency and iron deficiency anemia in adults".)

Iron overload – (See "Approach to the patient with suspected iron overload" and "Clinical manifestations and diagnosis of hereditary hemochromatosis" and "HFE and other hemochromatosis genes" and "Gene test interpretation: HFE (hereditary hemochromatosis gene)".)

ROLE OF SPECIFIC PROTEINS — Our understanding of iron metabolism and the hallmarks of iron deficiency and iron excess is based upon the biology of a number of critical proteins, including but not limited to the following (figure 1) [2]:

Transferrin (Tf), the plasma iron transport protein.

Transferrin receptor (TfR), the cellular receptor for iron-bound transferrin.

Ferritin (Ft), the cellular iron storage protein.

Iron regulatory protein 1 and 2 (IRP1 and IRP2), the cellular iron sensing proteins.

Divalent metal transporter 1 (DMT1, Nramp2, DCT1, solute carrier family 11, member 2 [SLC11A2]), the duodenal iron transporter.

Ferroportin (Ireg1, SLC40A1, Mtp1), the cellular iron exporter. Pathogenic variants in the SLC40A1 (also called FPN1) gene, which encodes ferroportin, cause rare forms of iron overload.

Hephaestin, which likely cooperates with ferroportin for exporting iron from enterocytes to transferrin.

Ceruloplasmin, a plasma metalloprotein required by ferroportin to export iron from macrophages, hepatocytes, and glial cells.

Poly(rC) binding proteins 1 and 2 (PCBP1 and PCBP2), the intracellular iron chaperones that bind to cytosines in RNA.

Hepcidin, the key negative regulator of intestinal iron absorption as well as macrophage iron release. Mutations of hepcidin cause a rare form of juvenile hemochromatosis.

HFE, mutations of which are responsible for the common form of hereditary hemochromatosis.

TFR2, mutations of which are responsible for a rare form of hereditary hemochromatosis.

Hemojuvelin, a hepcidin regulator, mutations of which are responsible for the common form of juvenile hemochromatosis.

Bone morphogenetic proteins (BMPs; cytokines produced by endothelial cells) such as BMP6 and BMP2, which activate hepcidin by binding to BMP receptors and signaling through SMAD proteins.

Matriptase 2/TMPRSS6, the liver hepcidin inhibitor with a major role in iron deficiency.

Erythroferrone, a hormone produced by erythroblasts stimulated by erythropoietin, a kidney-derived hormone produced in response to hypoxia and iron deficiency that downregulates hepcidin in response to erythropoiesis expansion.

Duodenal cytochrome B (DyctB, also called DcytB), an iron reductase at the luminal site of enterocyte enabling reduction of iron for absorption.

STEAP3, the intracellular ferric reductase essential for transcellular iron transport.

NCOA4, the nuclear receptor co-activator 4, cargo receptor for intracellular ferritin degradation and cellular iron mobilization.

Transferrin — The gene for apotransferrin is on the long arm chromosome 3. It codes for a protein (mol wt 80 kDa) that tightly binds one or two ferric (Fe3+) iron molecules and is the major transporter for iron trafficking through the plasma. Most of the Tf, which has a half-life of eight days, is made in the liver where its synthesis is considerably increased in states of iron deficiency or reduced upon iron loading or inflammation by unknown mechanisms [1,3].

ApoTf acquires iron from ferroportin, becoming monoferric or (when iron is abundant) diferric. A 2019 study reported that Tf binds iron at the N-terminal and C-terminal lobes with different affinities; binding to the C-terminal lobe favors the N-terminal lobe binding and the formation of diferric transferrin [4]. Diferric transferrin is the true ligand of both transferrin receptors (TfR1 and TfR2). (See 'Transferrin receptor' below.)

Tf can be measured in the plasma using an enzyme-linked immunosorbent assay (ELISA) or turbidimetric method to determine the mg of the transferrin protein/dL of plasma.

The total iron binding capacity of Tf (ie, TIBC) can be measured directly using iron binding methodology (ie, mg of iron binding capacity/dL of plasma), or it can be calculated by multiplying the results of the chemical or immunologic method by a conversion factor calculated by the individual laboratory, as follows:

   TIBC (microg Fe/dL) = Tf (mg protein/dL) x (conversion factor)
                (Conversion factor range: 1.40 to 1.49)

Complete lack of transferrin is most likely incompatible with life. Hypotransferrinemia is a rare autosomal recessive disorder associated with low transferrin levels (<10 mg/dL), severe iron deficiency anemia with hypochromic, microcytic red cells, and iron overloading of the liver and other parenchymal organs [5].

Transferrin receptor — The gene for the TfR (TFRC) is located on the long arm of chromosome 3 (as is the gene for transferrin). TFRC codes for a homodimeric transmembrane protein (mol wt 94 kDa) that is found on most cells, most densely on erythroid precursors and placental cells.

There is a binding site for the transcription factor Stat5 in the first intron of the gene that encodes TfR [6]. Expression of constitutively active Stat5A in an erythroid cell line increased TfR levels, while lethally irradiated mice that received a transplant of Stat5a/b-/- liver cells developed microcytic, hypochromic anemia [6].

Each TfR molecule can bind two diferric Tf molecules (four Fe3+ atoms), which it endocytoses after clustering on clathrin coated pits. The iron is offloaded in acidified vacuoles and the apotransferrin-TfR complex is recycled to the cell surface, where apo-Tf is discharged and released back into the circulation. TfR mRNA has five 3' iron-response elements (IREs) and, as occurs in other iron genes, is post-transcriptionally regulated by iron-response proteins (IRPs), being stabilized in iron deficiency and degraded in iron overload. (See 'Systemic iron homeostasis' below.)

Important roles in several tissues have been illustrated in mouse models:

Germline inactivation of TfR in mice is embryonically lethal due to severe anemia and abnormal central nervous system development. Haploinsufficiency of TfR causes microcytosis and a reduction of total body iron [7].

Conditional heart inactivation of TfR in mice causes cardiomegaly, decreased cardiac function, failure of mitochondrial respiration, and early death [8].

Conditional inactivation of TfR in the epithelial cells of the mouse intestine causes disruption of the epithelial barrier and early death. Since the phenotype was unresponsive to parenteral iron treatment, the authors suggested that TfR is implicated in the maintenance of the intestinal epithelium [9].

Conditional inactivation of TfR in the liver indicates that TfR is not essential for basal hepatocellular iron supply but is indispensable for the fine-tuning of hepcidin expression in response to hepatocyte iron loading [10]. Another study on liver conditional inactivation of TfR shows that the most important function of TfR in hepatocytes is to sequester HFE in iron deficiency, to block hepcidin upregulation [11]. (See 'HFE' below.)

The only disease associated with pathogenic variants in TfR causes only mild anemia and leads to a type of combined immunodeficiency. It is due to homozygosity for a variant that disrupts the TfR internalization signal, strongly impairing endocytosis and underscoring the essential role of iron in B and T lymphocyte development and differentiation [12].

Serum (soluble) TfR (sTfR) is a product of membrane TfR that is released into the circulation by membrane proteases when TfR is not associated with its ligand, diferric transferrin, as occurs in iron deficiency [13]. Levels of sTfR measured in serum correlate directly with erythropoietic expansion [14]. sTfR levels can be a measure of iron deficiency and the bone marrow erythroid progenitors' need for iron. (See "Causes and diagnosis of iron deficiency and iron deficiency anemia in adults", section on 'Iron studies (list of available tests)'.)

TfR is the receptor for Arenavirus and Plasmodium vivax, but it ameliorates cell entry by other viruses [15]. The same TfR ectodomain used by these pathogens to enter human cells interacts with the extracellular domain of H ferritin (ferritin heavy chain), as demonstrated by cryo-electron microscopy [16]. (See 'Ferritin' below.)

Ferritin — Ferritin is the cellular storage protein for iron. It is a huge 24 subunit protein (mol wt 440 kDa) consisting of light chains (L ferritin, 20 kd, gene on chromosome 19) and heavy chains (H ferritin, 21 kd, gene on chromosome 11), that can store up to 4500 atoms of iron within its spherical cavity [17]. H Ferritin possesses ferroxidase activity necessary for iron uptake by the ferritin molecule. L ferritin stabilizes the multimeric ferritin shell. A poly (rC) binding protein, (PCBP1), appears to be required as a cytosolic chaperone to deliver iron to ferritin [18]. PCBP1 gene deletion in mice causes microcytic anemia [19]. PCBP1 and PCBP2 form homodimeric and heteromeric complexes with high iron binding affinity and distinct functions in cellular iron trafficking, with PCBP2 thought to affect iron delivery from DMT1 and to FPN1 [20,21].

Ferritin synthesis is subject to different levels of control, including DNA transcription via its promoter, and mRNA translation via interactions with iron regulatory proteins [22]. (See 'Iron regulatory proteins or iron-responsive element binding protein' below.)

Ferritin is an acute phase reactant, and, along with transferrin and the transferrin receptor, is a member of the protein family that orchestrates cellular defense against oxidative stress and inflammation [22,23]. The gene for H ferritin is activated by oxidative stress via an upstream enhancer antioxidant-responsive element (ARE) and by several pro-inflammatory cytokines such as interleukin (IL)-1, IL-6 or tumor necrosis factor (TNF) alpha. (See "Acute phase reactants".)

Mice lacking H ferritin die early in gestation [24]. Mice heterozygous for H ferritin have slightly elevated tissue L ferritin and 7- to 10-fold more serum L ferritin than controls, although they do not have tissue iron overload [25]. These observations suggest that reduced H ferritin expression in humans might be responsible for cases in which high serum ferritin is present in the absence of iron overload [25]. Conditional inactivation of H ferritin in duodenal cells in mice causes iron overload, indicating that ferritin in the gut is important in the control of iron absorption [26].

Much of the iron stored in ferritin is accessible for metabolic needs. Ferritin is degraded by lysosomes through a process called ferritinophagy that requires the cargo protein NCOA4 [27,28]. This process is important for recovering intracellular iron when needed. Inactivation of NCOA4 in mice causes increased ferritin aggregates in spleen and liver macrophages and other organs [27,29]. In vitro downregulation of NCOA4 affects erythroid maturation and hemoglobin synthesis [19]. Ferritin within erythroid precursors may donate iron for heme synthesis, especially at the beginning of hemoglobin accumulation, at a time when the transferrin-transferrin receptor pathway is still insufficient [30]. However, the most important role of NCOA4 is in macrophage ferritinophagy, to recover iron stored in ferritin in iron deficiency [31] and in the gut where it contributes to iron absorption (see below).

When ferritin accumulates, it aggregates and is proteolyzed by lysosomal enzymes; it is then converted to an iron-rich, poorly characterized hemosiderin which releases its iron slowly and is detected by the Prussian blue reaction, which is used in the common Perls staining for iron in bone marrow aspirates.

Ferritin measured clinically in plasma is usually apoferritin, a non-iron containing molecule. The plasma level generally reflects overall iron storage, with 1 ng of ferritin per mL indicating approximately 10 mg of total iron stores. Thus,

An adult male with a plasma ferritin level of 50 to 100 ng/mL has iron stores of approximately 500 to 1000 mg [32].

A serum ferritin less than 15 ng/mL is 99 percent specific for making a diagnosis of iron deficiency. (See "Causes and diagnosis of iron deficiency and iron deficiency anemia in adults", section on 'Iron studies (list of available tests)'.)

An elevated serum ferritin in the absence of infection or inflammation may suggest the presence of an iron overload state. (See "Approach to the patient with suspected iron overload", section on 'CBC, LFTs, and iron studies'.)

Ferritin levels may be extremely high in patients with hemophagocytic lymphohistiocytosis or certain rheumatologic disorders as well as in acute infections with concomitant inflammation such as in severe coronavirus disease 2019 (COVID-19) [33-35]. In such cases, the ferritin tends to be less glycosylated than normal. (See "Clinical features and diagnosis of hemophagocytic lymphohistiocytosis", section on 'Serum ferritin levels'.)

A separate ferritin (m-ferritin) is present within mitochondria and is the product of an intronless nuclear gene [17]. Its expression is increased in tissues with high numbers of mitochondria, rather than in tissues involved in iron storage. M-ferritin likely protects mitochondria from oxidative damage [17].

The role and the origin of circulating ferritin as well as the pathways of cellular ferritin secretion are largely unknown [36]. Scavenger receptor, member 5 (Scara5), is a proposed receptor for circulating L-ferritin and mediates the delivery of iron bound to ferritin to the developing kidney [37].The T cell immunoglobulin-domain and mucin-domain (TIM) proteins Tim-2 on B lymphocytes and liver and kidney cells is proposed as a H-ferritin receptor [38]. TfR binds H ferritin (ferritin heavy chain) independently from transferrin [16]. Whether this interaction may locally provide iron to cells when transferrin saturation is low remains to be demonstrated.

Iron regulatory proteins or iron-responsive element binding protein — The expression of proteins involved in cellular iron uptake and storage is regulated by the iron status of the cell.

Iron-regulatory proteins 1 and 2 (IRP1 and IRP2) are cytosolic RNA-binding proteins that bind to iron-responsive elements (IRE), consisting of a loop configuration of nucleotides, that are located in the 5' or 3' untranslated regions of specific mRNAs encoding for iron and other genes (eg, ferritin, TfR, DMT1, ferroportin, the erythroid specific form of delta-aminolevulinic acid synthase [eALAS], hypoxia inducible factor-2 alpha [HIF-2 alpha; HIF-2α]).

Binding of IRPs to their target sequences occurs when the cell is iron deficient; this has different effects according to whether the IRE position is at the 5' or the 3' UTR:

When IRPs bind to the 5' IRE of ferritin, ferroportin, eALAS, or HIF-2 alpha, the rates of mRNA translation and protein biosynthesis are decreased.

When IRPs bind to the 3' end of transcripts such as TfR or DMT1, the mRNA half-life is prolonged and rates of biosynthesis are increased (figure 2).

IRP1 and IRP2 sense the availability of metabolically active iron in the cell in different ways. When cellular iron levels increase, assembly with iron sulfur clusters changes IRP1 to cytoplasmic aconitase and its binding ability is lost. Under the same conditions of cellular iron increase, IRP2 interacts with the iron stabilized FBXL5 protein, which recruits an E3 ligase. This causes IRP2 ubiquitination and proteasomal degradation [39]. FBXL5 has an iron- and oxygen-sensing domain, representing an example of the iron-oxygen connection.

The net effect of these IRPs is that the iron overloaded state is characterized by increased production of ferritin (to permit adequate storage) and ferroportin (to export excess iron) and eALAS (for iron consumption) as well as decreased production of TfR and DMT1 (to minimize iron uptake). These changes are reversed in iron deficiency, which is characterized by reduced ferritin and ferroportin and elevated TfR synthesis (figure 2).

Targeted deletion of the gene encoding IRP1 in the mouse leads to no evident hematologic change in adulthood, while targeted deletion of the gene encoding IRP2 causes misregulation of iron metabolism, refractory microcytic anemia, and a neurodegenerative disease due to abnormal neuronal accumulation of iron [40].

These observations establish a major role for IRP2 in the regulation of iron uptake in erythroid cells. However, it has been shown that HIF-2 alpha, which has an IRE in the 5' UTR, is specifically controlled by IRP1 [41]. This demonstrates another example of the iron-hypoxia connection. Irp1-/- mice, when young or iron deficient, develop polycythemia, pulmonary hypertension, and sudden death from hemorrhages, due to an inability to suppress the synthesis of HIF-2 alpha that stimulates both erythropoietin in the kidney and endothelin 1 in pulmonary endothelial cells [42,43].

There are differences in tissue-specific expression of the two IRPs, and also specific targets. (See "Molecular pathogenesis of congenital erythrocytoses and polycythemia vera", section on 'Oxygen sensor'.)

HFE — HFE, a product of the HFE gene on the short arm of chromosome 6, is an MHC class I-like protein present ubiquitously at low levels, with high-level expression in hepatocytes. The C282Y variant in the HFE gene is responsible for the vast majority of cases of hereditary hemochromatosis in adults. Iron overload is also noted in mice with HFE gene deletion [44]. Conditional deletion of HFE in the liver produces the phenotype of hemochromatosis [45]. In contrast, conditional deletion of HFE in duodenal cells or macrophages does not alter systemic iron metabolism [46].

The HFE protein is associated in a complex with TfR as a possible iron sensor. Diferric transferrin competes with this binding, releasing HFE from TfR when iron is abundant. Free HFE increases the response of hepcidin to iron, since patients with hemochromatosis due to an HFE mutation have low hepcidin levels and do not respond to oral iron challenge [47]. (See 'Hepcidin' below.)

It has been reported that when HFE is free from TfR, it binds to TfR2, encoded by a gene mutated in type 3 hereditary hemochromatosis that is considered to be a sensor of the level of transferrin saturation [48]. (See "HFE and other hemochromatosis genes".)

However, HFE-TFR2 binding has been disputed [49]. Indeed, the diseases caused by pathogenic variants affecting HFE and TfR2 are different. Simultaneous gene deletion of both HFE and TfR2 in mice results in marked dysregulation of hepcidin and more severe iron overload than loss of the single molecules [50]. These observations suggest distinct functions for HFE and TfR2. According to an in vitro study, HFE stabilizes the BMP receptor ALK3 that is degraded when HFE is inactive [51]. HFE has been shown to interact with low density lipoprotein- receptor (LDL-R) in hepatocytes, thereby affecting systemic lipid-homeostasis and atherosclerosis development [52].

Transferrin receptor 2 — Transferrin receptor 2 (TfR2), encoded by a gene on chromosome 7q22, is a member of the TfR family and is homologous to TfR1, but it has no IRE elements and has a lower affinity for diferric transferrin than TfR1 [53]. It displays a restricted expression pattern, being present at high levels in hepatocytes and erythroid cells. TfR2 may bind diferric transferrin and is considered a sensor of Tf saturation [54].

Mutations of TfR2 cause a form of hereditary hemochromatosis. Mice with either germline or liver conditional inactivation of TfR2 develop iron overload [55,56]. (See "HFE and other hemochromatosis genes", section on 'Transferrin receptor 2 (TFR2)'.)

TfR2 is expressed in immature erythroid cells, where it interacts with the erythropoietin receptor, stabilizing it on the cell surface [57]. In mice, conditional inactivation of TfR2 in the bone marrow causes erythrocytosis with normal erythropoietin levels. It has been shown that TfR2 modulates the erythropoietin sensitivity of erythroblasts by sensing iron deficiency through decreased diferric transferrin. Through its sensor activity, TfR2 may coordinate erythropoiesis with hepcidin synthesis [58]. TfR2 is also expressed in osteoclasts and osteoblasts, where it controls bone structure by modulating the BMP pathway [59].

Hemojuvelin — Hemojuvelin (HJV) is a glycosylphosphatidylinositol (GPI)-anchored protein that regulates hepcidin production in hepatocytes. It is the product of a gene on chromosome 1q21 and is homologous to RGM (Repulsive Guidance Molecule, also known as RGMc) expressed in the central nervous system. HJV is highly expressed in liver, skeletal muscles and heart. Mutations of the gene encoding HJV (HFE2) produce a form of juvenile hemochromatosis with extremely low hepcidin levels [60]. (See "HFE and other hemochromatosis genes", section on 'Hemojuvelin (HJV)'.)

HJV is present in a membrane-associated form, bound to a glycosylphosphatidylinositol (GPI) anchor, and also as a soluble form with opposite effects on hepcidin activation [61,62]. Membrane HJV is a coreceptor for bone morphogenetic proteins (BMPs) and is essential in hepcidin activation. In vitro cleavage of HJV by furin in hypoxia and iron deficiency produces soluble HJV components and serves to downregulate hepcidin [63]. Its role in vivo is uncertain. In iron deficiency, membrane HJV is cleaved by the liver serine protease TMPRSS6 to attenuate the BMP signaling and suppress hepcidin [64]. (See 'Matriptase-2/TMPRSS6' below.)

Divalent metal transporter 1 and duodenal cytochrome b — The duodenal divalent metal transporter 1 (DMT1), encoded by the SLC11A2 gene on chromosome 12p13, is the major route for the uptake of non-heme iron from the intestinal lumen (figure 1). It was identified by positional cloning as gene responsible for a form of microcytic anemia in mice with a missense mutation that had a marked impairment in intestinal iron transport [65].

Parallel functional studies in Xenopus oocytes found a single cDNA that stimulated iron transport. This divalent metal transporter protein (DMT1) was identical to Nramp2 [66]. The transporter also transports other heavy metals such as lead, zinc, and copper by an ATP and proton-dependent process. DMT1 is widely expressed, particularly in the proximal duodenum. Expression of the isoform containing an iron responsive element is specifically upregulated in dietary iron deficiency [65] and hypoxia through the action of HIF-2 alpha [67], with greatest expression at the brush border of the apical pole of the enterocytes in the apical two-thirds of the villi, the major site of iron absorption.

Pathogenic variants in SLC11A2 cause life-long microcytic anemia, increased Tf saturation, and liver iron accumulation, but with low or only moderately high serum ferritin levels [68].

In iron import, DMT1 cooperates with duodenal cytochrome b (DYCTB), a membrane reductase that facilitates iron absorption as ferrous iron from the lumen of the duodenum (figure 1) [69].

Local intestinal hypoxia has an important role in increasing the expression of genes involved in iron transport, especially the luminal proteins DMT1 and DcytB and the iron exporter ferroportin (see 'Ferroportin' below). It has been shown that HIF-2 alpha binds to the DMT1 promoter and regulates DMT1 expression in duodenal cells; tissue-specific deletion of HIF-2 alpha in mouse enterocytes decreases intestinal iron absorption as well as the expression of DMT1 in enterocytes [67].

Four different isoforms of DMT1 have been described in cells and various tissues including intestinal cells, erythroid cells, macrophages, brain cells, or kidney cells, which have different organ-specific and subcellular functions for iron trafficking, an issue that still awaits clarification [70].

HIF-2 alpha — Hypoxia-inducible factor 2 alpha (HIF-2α), encoded by the EPAS1 gene, has an IRE in the 5' UTR and is controlled by IRP1 [41]. (See 'Iron regulatory proteins or iron-responsive element binding protein' above.)

HIF-2α is an essential mediator of iron absorption that cooperates with low hepcidin to increase absorption in iron deficiency, anemia, and hypoxia [71]. In hypoxia, HIF-2α increases the expression of key genes (DMT1, DCYTB, and ferroportin) that contribute to enhanced iron absorption [67]. (See 'Intestinal iron absorption' below.)

A number of families with gain-of-function mutations in EPAS1 have been described with autosomal dominant erythrocytosis. (See "Molecular pathogenesis of congenital erythrocytoses and polycythemia vera", section on 'EPAS1 mutations'.)

HIF-1α is regulated by the von Hippel-Lindau (VHL) gene, and its dysregulation may contribute to some of the findings in VHL disease. (See "Molecular biology and pathogenesis of von Hippel-Lindau disease", section on 'Hypoxia-inducible factor 1 and 2'.)

Importantly, HIFs regulate the production of erythropoietin, which subsequently reduces hepcidin expression by mechanisms detailed below [72]. (See "Regulation of erythropoiesis", section on 'Hypoxia and EPO expression'.)

Ferroportin — Ferroportin-1 (Ireg1, encoded by SLC40A1, formerly called SLC11A3, Mtp1) is an iron export protein found by positional cloning in mutant zebrafish with hypochromic anemia [73]. Human ferroportin is highly expressed in the basal portion of placental syncytiotrophoblasts, the basolateral surface of duodenal enterocytes, macrophages, hepatocytes, cardiomyocytes, erythrocytes, and other cells [74-77]. The ferroportin gene maps to 2q32.

Ferroportin functions as a major exporter of iron, transporting iron from mother to fetus, transferring absorbed iron from enterocytes into the circulation, and allowing macrophages to recycle iron from damaged and senescent red cells back into the circulation (figure 1).

In animal and human in vitro models, ferroportin is posttranscriptionally regulated by the amount of available iron, due to the presence of an IRE in the 5' UTR [78-81]. An isoform of ferroportin without the 5' IRE (ferroportin B) has been identified in duodenal mucosa. It appears to function to evade iron-mediated post-transcriptional regulation in conditions of iron deficiency [80]. The same isoform is expressed in erythroid cells and may function to "fine-tune" systemic iron usage [77,82].

The most important control of ferroportin is post-translational, since ferroportin is downregulated through its interaction with hepcidin [75,83,84]. When hepcidin levels increase, hepcidin binds to ferroportin, occludes the central cavity that exports iron, and induces ferroportin internalization and lysosomal degradation [85]. This reduces the amount of iron that is released into the circulation from duodenal cells as well as macrophages [83].

Besides post-translational control by IRPs, ferroportin expression is increased by heme independently from iron [86]. Both iron and erythrophagocytosis (through heme increase) stimulate ferroportin transcription [87]. Ferroportin transcription is reduced in inflammation via Toll-like receptors and pro-inflammatory cytokine signaling [88,89].

Pathogenic variants in the gene for ferroportin due to loss of function mutations have been found in a number of families with an autosomal dominant form of iron overload, which has a distinct phenotype called "ferroportin disease" [90-92]. In contrast, gain of function mutations make ferroportin resistant to the effect of hepcidin, producing a phenotype that overlaps with hemochromatosis. (See "HFE and other hemochromatosis genes", section on 'Ferroportin (SLC40A1; FPN1)'.)

The hepcidin-ferroportin axis may have important tissue specific roles (eg, in the heart). Mice with conditional inactivation of ferroportin in cardiomyocytes develop cardiac iron overload and dilated cardiomyopathy, and they have decreased survival, strengthening the concept that these cells need to maintain an intact iron export system [87].

Hephaestin and ceruloplasmin — Hephaestin and ceruloplasmin are involved in iron export as multioxidases:

Hephaestin is the product of an X-linked gene mutated in mice with sex-linked anemia, a disorder in which enterocytes are iron-loaded but efflux of iron through the basolateral membrane and into the plasma is inhibited [93]. Hephaestin cooperates with ferroportin in iron export in enterocytes. It has significant homology to the serum protein ceruloplasmin and has ferroxidase activity as well.

Ceruloplasmin is a copper-containing protein encoded by the CP gene on chromosome 3q24-25; it is a ferroxidase required for efficient recycling of iron in the liver, reticuloendothelial system, and glial cells. It is present as a soluble form and as a membrane form bound to a GPI anchor in the central nervous system. Mutations of ceruloplasmin lead to aceruloplasminemia. (See 'Iron release from macrophages' below.)

FLVCR — A mammalian heme export protein has been described (FLVCR, feline leukemia virus, subgroup C, receptor), which has been postulated to protect developing erythroid cells from the toxicity of unbound cytoplasmic heme. Interference with this protein results in a loss of erythroid progenitors and severe anemia in experimental animals, while its inhibition in human erythroleukemia cells decreases heme export, impairs their erythroid maturation, and leads to apoptosis [94].

The heme exporter FLVCR1 regulates expansion and differentiation of committed erythroid progenitors by controlling intracellular heme accumulation in mice [95] and plays a crucial role in maintaining intestinal heme homeostasis [96].

ZIP14 — ZIP14, a metal transporter of zinc and manganese encoded by the SLC39A1 gene, has been shown to be important for non-transferrin bound iron (NTBI) uptake by hepatocytes, pancreatic acinar cells [97], and beta-cells of the islets, while other NTBI transporters such as L-type calcium channels, DMT1, ZIP8 are active in the heart and anterior pituitary [98].

Hepcidin — Hepcidin (also called liver-expressed antimicrobial peptide [LEAP-1] or hepcidin antimicrobial peptide [HAMP]) is an acute phase reactant with intrinsic antimicrobial activity [99-102]. It is encoded as a propeptide by a gene on 19q13. There are two isoforms of the mature peptide; hepcidin-25 has a central role in iron homeostasis, while the function of hepcidin-20, which lacks the five amino acid sequence crucial for iron regulation, is unknown.

Serum hepcidin levels have correlated directly with serum ferritin in healthy people, are highest in inflammation, and are lowest in iron deficiency anemia. Reference ranges for hepcidin levels in healthy controls were noted to be wide in one study, but when rigorous criteria were required to eliminate individuals with inflammation or renal or hepatic disease, variability was considerably less [103].

Initiatives are ongoing for standardization of hepcidin assays [104]. Two large studies have measured serum hepcidin in the general population. In the first study, hepcidin levels were analyzed using a competitive enzyme-linked immunosorbent assay in 2998 well-characterized participants from the Nijmegen Biomedical Study [105]. In the second report, serum hepcidin levels were measured by mass spectrometry in 1545 individuals from an Italian cohort [106].

Both studies observed stable values in males across ages but strong variation in females, with low values in younger and higher levels in older females [106].

There was a trend of increasing hepcidin concentrations during the day, although there was no evidence for a primary or secondary circadian variation in hepcidin levels [105].

Levels were strongly associated with serum ferritin levels and were less strongly associated with C-reactive protein and total iron binding capacity (TIBC) in men and for TIBC, alanine aminotransferase, and glomerular filtration rate in women [105].

Because of the lack of a standardized assay, hepcidin testing is not ready for regular clinical use, although a number of test platforms are under active investigation [103]. Serum hepcidin levels may have diagnostic value in certain iron disorders [103].

Hepcidin is produced in many tissues, but the primary site of synthesis is in the liver [107-109]. Other tissues that produce hepcidin include macrophages in inflammation, adipocytes, and retinal cells [110-112]. An essential cell-autonomous role for cardiomyocyte hepcidin in cardiac iron homeostasis has also been demonstrated [113].

Hepcidin is rapidly excreted by the kidney and reabsorbed in the proximal tubules; its levels increase in chronic kidney disease. It serves as an important mediator in the pathogenesis of the anemia of chronic disease [108,114,115]. It may decrease in chronic liver disease [103]. Its deficiency or inappropriate production explains the pathogenesis of iron overload in hereditary hemochromatosis [116-118] (see "HFE and other hemochromatosis genes"). Its excessive inhibition by ineffective erythropoiesis explains iron overload in iron loading anemias [119]. Hepcidin levels are also influenced by hormones, especially inhibited by testosterone [120].

The following examples demonstrate the importance of hepcidin in iron balance and indicate that hepcidin plays a major role as a negative regulator of intestinal iron absorption and iron release from macrophages:

Hepcidin knockout (KO) mice develop iron overload [100,121]. Liver-specific KO mice fully recapitulate the severe iron overload phenotype observed in the total hepcidin KO mice, demonstrating that the hepatocyte constitutes the predominant reservoir for systemic hepcidin and that the other tissues capable of synthesizing hepcidin are unable to compensate [122].

In humans, mutation in hepcidin causes a rare form of juvenile hemochromatosis [123]. (See "HFE and other hemochromatosis genes", section on 'Hepcidin (HAMP)'.)

Injection of hepcidin inhibited intestinal iron absorption in mice independent of their iron status and did not require the HFE gene product. In other experiments, injection of a synthetic hepcidin in mice was associated with a rapid and prolonged reduction in serum iron, along with accumulation of hepcidin in ferroportin-rich organs (eg, liver, spleen, proximal duodenum) [84].

Constitutive overexpression of hepcidin leads to severe iron deficiency anemia at birth [124] and slows dietary iron absorption and cycling through macrophages, resulting in iron-restricted erythropoiesis and a failure to respond to adequate erythropoietin levels [125].

Overexpression of hepcidin inhibits the iron accumulation normally observed in HFE-deficient mice [116] and in mouse models of beta thalassemia [126].

The BMP SMAD signaling pathway activating hepcidin in response to iron, and the IL-6 signaling pathway increasing hepcidin in inflammation are discussed below. (See 'Iron sensing and signaling pathway' below.)

Hepcidin reduces iron absorption in the intestine and releases iron from macrophages via a mechanism that involves interactions with and inactivation of the iron export protein ferroportin [83,101]. The interaction occurs only with ferroportin bound to iron [127]; this explains the strong effect on cells such as macrophages and enterocytes, which export iron to plasma. Hepcidin also appears to regulate levels of non-transferrin-bound iron (NTBI) in a mouse model of bacterial infection; this may be the key mechanism by which hepcidin exerts its antimicrobial properties against circulating "siderophilic" bacterial strains that use NTBI and thrive in an iron-rich environment [102].

A distinct example of the antimicrobial role played by hepcidin occurs in necrotizing fasciitis caused by certain strains of group A Streptococcus. Hepcidin production is induced in the skin of the infected patients and has a protective role [128]. Hepcidin loss in keratinocytes of mice infected with the same microorganism blocks the production of the CXCL1 chemokine, reducing neutrophil recruitment and enabling infection to disseminate. This outcome can be reverted by the injection of exogenous hepcidin.

Another example of the antimicrobial function of hepcidin is illustrated in the gut, where hepcidin production by dendritic cells can lead to sequestering of iron from the microbiome, allowing mucosal repair of gut inflammation such as would occur in inflammatory bowel disease [129].

Strategies are underway for altering/inhibiting the function of hepcidin and its receptor ferroportin to alleviate some of these disorders of iron metabolism [130]. (See "Anemia of chronic disease/anemia of inflammation", section on 'Pathogenesis'.)

BMP and BMP receptors — Bone morphogenetic proteins (BMPs; cytokines produced by endothelial cells) such as BMP6 and BMP2 activate hepcidin binding to BMP receptors and signaling through SMAD proteins [131].

BMP6-null mice have a phenotype resembling hereditary hemochromatosis, with reduced hepcidin expression and tissue iron overload, indicating the essential role of BMP6 in iron regulation in mammals [131,132]. BMP6 is highly expressed in liver sinusoidal endothelial cells (LSEC), and conditional inactivation of BMP6 in LSEC causes iron overload [133]. Also, BMP2, produced by the same cells, regulates hepcidin; its conditional inactivation in LSEC and in endothelial cells in mice causes iron overload, defining its role in signaling for hepcidin, for which the mechanism still needs to be clarified [134].

The phenotype of BMP6 and BMP2 double knock out mice is not more severe than the individual knockouts, suggesting that the two BMPs cooperate in hepcidin activation, possibly through the formation of BMP heterodimers [135].

Matriptase-2/TMPRSS6 — The most powerful inhibitor of hepcidin expression is matriptase-2, which is encoded by TMPRSS6, the liver transmembrane protease, serine 6 gene. Matriptase-2 exerts its hepcidin regulatory effects by cleaving membrane hemojuvelin, a protein that normally signals to promote hepcidin expression and likely blocks other activators of hepcidin [64,136,137]. (See 'Hemojuvelin' above.)

In agreement with hemojuvelin being the substrate of matriptase-2, mice that are double knockout for both TMPRSS6 and HJV have the same iron overload phenotype as that of HJV-/- mice [138].

The important role of TMPRSS6 in human iron deficiency is indicated by the observation that mutations in TMPRSS6 cause a rare autosomal recessive disorder characterized by iron deficiency anemia unresponsive to treatment with oral iron, but partially responsive to parenteral iron, a condition termed iron-refractory iron deficiency anemia (IRIDA). This condition is discussed in detail separately. (See "Causes and diagnosis of iron deficiency and iron deficiency anemia in adults", section on 'Inherited disorders/IRIDA'.)

Single nucleotide polymorphisms in the TMPRSS6 gene described in several populations appear to affect serum iron concentrations, hemoglobin levels, erythrocyte characteristics (eg, mean corpuscular volume [MCV], mean corpuscular hemoglobin [MCH]), and, indirectly, erythropoiesis [139-143]. These findings suggest a genetic susceptibility to the development of iron deficiency.

Anti-TMPRSS6 compounds increase hepcidin and control iron overload in animal models of iron disorders and are being studied in clinical trials to increase hepcidin to alleviate iron loading.

Erythroferrone — Erythroferrone (ERFE) is encoded by the ERFE gene (also called FAM132B or CTRP15). The protein is a member of the tumor necrosis factor alpha (TNF-alpha) family. Erythroferrone is stimulated by erythropoietin, a circulating hormone essential for the maturation and survival of erythroid progenitor cells to raise the red blood cell count in response to hypoxia or anemia. ERFE downregulates hepcidin to ensure iron supply.

The mechanism is mediated by blocking the binding of a subgroup of BMP proteins (including BMP6 and BMP2) to their receptors [144,145]. ERFE behaves as a ligand trap for BMPs, to attenuate BMP signaling and acquire iron [146]. Antibodies raised against the N-terminal domain of ERFE prevent hepcidin suppression and are proposed as a therapeutic tool for iron loading disorders due to low hepcidin [146]. (See "Regulation of erythropoiesis", section on 'Erythropoietin'.)

Serum ERFE is low/undetectable in the absence of disease. Its levels are increased by erythropoietin in hypoxia or after hemorrhage [147]. ERFE function must be transient to allow acquisition of only the amount of iron needed to compensate erythropoiesis. Chronic production of ERFE induces iron overload and developmental abnormalities in transgenic mice [148]. ERFE is responsible of transfusion-independent iron overload in patients with ineffective erythropoiesis [2,147].

SYSTEMIC IRON HOMEOSTASIS — The normal iron content of the body is approximately 3 to 4 grams. It exists in the following forms (figure 3):

Hemoglobin in circulating red cells and developing erythroblasts – approximately 2.0 to 2.5 g

Iron-containing proteins (eg, myoglobin, cytochromes, enzymes) – 300 to 400 mg

Plasma transferrin-bound iron – 3 to 4 mg

The remainder is storage iron in the form of ferritin or hemosiderin

Males have approximately 0.7 to 1.0 g of storage iron (mostly in liver, spleen, muscle, and bone marrow). Females have less storage iron (0.3 g), depending upon the extent of menses, pregnancies, deliveries, and iron intake. (See "Anemia in pregnancy", section on 'Iron deficiency'.)

Only a small amount of iron (1 to 2 mg) enters and leaves the body via enterocyte or keratinocyte desquamation on a daily basis. Most iron is recycled from the breakdown of old red blood cells by macrophages of the reticuloendothelial system. To avoid iron-mediated free radical toxicity, iron is always bound to proteins in the less reactive ferric iron (Fe3+) form. Essentially all circulating iron is bound to transferrin. This chelation renders the iron soluble and prevents toxicity. Iron homeostasis is regulated strictly at the level of intestinal absorption and release of iron from macrophages.

Intestinal iron absorption — The gastrointestinal mucosa plays a major role in regulating iron absorption, which varies according to the form of iron in the diet. A Western daily diet contains approximately 15 mg of iron. Some of this is heme iron, of which approximately 30 percent is promptly absorbed, likely via its own transport system. A candidate heme iron transporter, named heme carrier protein 1, has been found in the apical brush border membrane of duodenal enterocytes in the mouse [149]. However, the role in heme transport was subsequently dismissed [150].

The remaining non-heme iron, accounting for almost all of the iron in the diet in non-Western countries, is poorly absorbed, with less than 10 to 20 percent being taken into the mucosal cells (figure 3) [151]. Iron absorption is increased in iron deficiency states and may be different from food sources relative to oral iron supplements [152].

The table summarizes various factors that can impact iron absorption (table 1). As examples:

Dietary sources of heme iron (fish, poultry, and meat) have a higher bioavailability than do non-heme (vegetable) sources (30 versus <10 percent).

Intraluminal factors and genetic polymorphisms of iron metabolism genes can affect absorption.

Ascorbic acid and meat sources enhance the absorption of non-animal sources of iron such as cereal, breads, fruits, and vegetables, whereas tannates (teas), bran foods rich in phosphates, and phytates inhibit iron absorption [153-157].

Certain medications can impact iron absorption, as shown for reduced iron uptake in individuals treated with proton pump inhibitors [158].

The composition of the gut microbiota can affect iron absorption, as some specific bacteria can produce metabolites that block enterocyte iron import [159]. (See "Treatment of iron deficiency anemia in adults", section on 'Dosing and administration (oral iron)'.)

Molecular mechanisms of intestinal heme absorption are unclear. The proposed heme carrier protein 1, which is highly expressed in the gut and stimulated by hypoxia [149], functions as a folate transporter [150]. A heme exporter, Feline Leukemia Virus Receptor 5 (FLVR5) is expressed in enterocytes, macrophages and erythroblasts with the likely function of exporting heme excess [94]. (See 'FLVCR' above.)

Iron in food is prominently ferric (Fe3+), which is poorly soluble above a pH of 3 and is therefore poorly absorbed. In comparison, ferrous iron (Fe2+) is more soluble, even at the pH of 7 to 8 seen in the duodenum. As a result, it is more easily absorbed.

Ferrous iron is taken up at the mucosal side by the intestinal transporter, DMT1. In this process, duodenal cytochrome b (DcytB) reduces ferric iron to ferrous iron. Transcription of both DMT1 and DCYTB is stimulated by hypoxia inducible factor-2 alpha (HIF-2 alpha; HIF-2α) in the hypoxic environment of intestinal mucosa [160,161]. The microbiota-derived products 1,3-diaminopropane and reuterin reduce HIF-2a stability and reduce iron absorption, pointing to the importance of the gut microbiota for systemic iron homeostasis [159].

The iron is thought to enter the cell where it binds to cytosolic low molecular weight iron carriers and is transported to the basolateral portion of the cell. To enter the circulation, iron must be transported across the basolateral membrane by the duodenal iron exporter, ferroportin. Upon its release, the ferrous iron is oxidized to the ferric form and loaded onto transferrin. This oxidation process involves hephaestin, a homologue of ceruloplasmin, a known ferroxidase.

Mucosal cells, in addition to the cellular mechanisms noted above, respond to specific physiologic signals (see 'Hepcidin' above). In hypoxia, hepcidin is downregulated to allow increased iron export through ferroportin, while HIF-2α increases the expression of key genes (DMT1, DCYTB, and ferroportin) that enhance iron absorption [67]. The rate of iron absorption is appropriately enhanced when iron stores are reduced or absent (figure 1). HIF-2α is an essential mediator of iron absorption that cooperates with low hepcidin to increase absorption in iron deficiency, anemia and hypoxia [71]. (See 'HIF-2 alpha' above.)

HIF-2 alpha also controls NCOA4 and stimulates ferritinophagy as an additional mechanism to increase iron export to plasma, a mechanism contributing to iron overload in hemochromatosis [162].

The degree of erythropoiesis has an indirect effect, as iron absorption is increased when driven by increased erythropoiesis, a process mediated by hepcidin suppression.

Iron absorption is especially increased in disorders that cause ineffective erythropoiesis, such as beta thalassemia, dyserythropoietic and sideroblastic anemia, or variants of the myelodysplastic syndrome. On the other hand, intestinal cells can hold onto iron in the iron-replete state; this iron is lost when the mucosal cells are sloughed. All of these regulatory processes are mainly mediated by hepcidin through its interaction with ferroportin [1]. This function of hepcidin appears to be especially important when there are competing needs for iron, such as when anemia, iron deficiency, and infection coexist, as shown in African children in the malaria setting [163].

Iron absorption is decreased in conditions of iron excess through hepcidin increase, a process lost in hereditary hemochromatosis. (See "Clinical manifestations and diagnosis of hereditary hemochromatosis", section on 'Pathophysiology'.)

Transferrin saturation — Circulating transferrin normally is approximately one-third saturated with iron (ie, Fe ÷ TIBC = 1/3, when both are expressed as microg of iron per 100 mL of plasma) [164,165]. The formula to change transferrin levels from mg of protein to microg of iron binding capacity (see 'Transferrin' above) can be used if this information has not already been provided by the laboratory.

Conditions in which transferrin saturation (TSAT) is reduced include those in which the supply of iron to the plasma from the macrophage and other storage sites is reduced. These include:

Iron deficiency anemia

The anemia of chronic disease (anemia of inflammation)

Some patients with a ferroportin mutation (see 'Ferroportin' above and "HFE and other hemochromatosis genes", section on 'Ferroportin (SLC40A1; FPN1)')

Conversely, TSAT is increased in those conditions in which the supply of iron is excessive or is greater than the current demand. These include:

Most cases of hereditary and acquired hemochromatosis

Aplastic anemia, bone marrow suppression

Sideroblastic anemias

Ineffective erythropoiesis

Heavily transfused patients

Liver disease with reduced transferrin synthesis

Monoclonal immunoglobulin with anti-transferrin activity (rare) [166]

The use of TSAT to evaluate disease states is discussed separately:

Iron deficiency – (See "Causes and diagnosis of iron deficiency and iron deficiency anemia in adults", section on 'Iron studies (list of available tests)'.)

Iron overload – (See "Approach to the patient with suspected iron overload", section on 'CBC, LFTs, and iron studies' and "Clinical manifestations and diagnosis of hereditary hemochromatosis", section on 'Diagnostic evaluation'.)

Iron loss — There is no normal mechanism of regulated iron excretion. Iron is lost in sweat, shed skin cells, and some gastrointestinal loss at a rate of approximately 1 to 2 mg/day (figure 3). A normal adult Western male has 1 to 2 mg of heme iron and 10 to 15 mg of other iron in his diet. If 30 percent of the heme iron and 10 percent of the other iron is absorbed, then the total rate of iron absorption is 1 to 2 mg/day [164]. Thus, a man can easily stay in iron balance and even build up iron stores. On the other hand, a woman with an additional menstrual iron loss of 1 to 2 mg/day generally has lower iron stores than men, and is always delicately poised to become iron deficient.

There is abundant expression of DMT1 in the proximal tubule and collecting ducts of the kidney [65]; these cells also express TFR1, ZIP14, and ZIP8 [167]. The iron handling by the kidney in terms of uptake and export is not fully elucidated. Iron is filtered in the glomerulus and reabsorbed in the tubules both as transferrin and non-transferrin-bound iron (NTBI). The expression of ferroportin on the apical membrane of proximal tubules suggests a possible contribution of iron excretion in iron overload [168].

Iron release from macrophages — Approximately 20 to 25 mg of iron are released daily from the breakdown of senescent red cells in the macrophages (figure 3). Hemoglobin heme released from phagocytosed red cells is catabolized by microsomal heme oxygenase to biliverdin and carbon monoxide and the resulting iron is released to the circulation through ferroportin or stored in ferritin according to the body needs and to the local concentration of hepcidin [169]. Thus, hepcidin coordinates both duodenal iron absorption as well as macrophage iron release. (See 'Hepcidin' above.)

Upon its release from ferroportin, ferrous iron is oxidized to the ferric form, and loaded onto transferrin. The oxidation process involves ceruloplasmin, a known copper dependent multioxidase. This could explain the iron overload that is seen in aceruloplasminemia, an autosomal recessive disorder of iron metabolism characterized by anemia, diabetes, retinal degeneration, and neurologic symptoms. Affected patients have inherited mutations in the ceruloplasmin gene, along with progressive parenchymal iron accumulation in conjunction with strongly decrease of circulating serum ceruloplasmin [170]. (See "Bradykinetic movement disorders in children", section on 'Neurodegeneration with brain iron accumulation'.)

Iron sensing and signaling pathway — Hepcidin is upregulated in response to increased circulating and body iron levels (figure 1), inflammation, infection, and endotoxin; it is downregulated following hypoxia, anemia, iron deficiency, ineffective erythropoiesis, all conditions characterized by increased levels of erythropoietin but also by steroid hormones or alcohol consumption (figure 4) [171-176]. Hepcidin has also been shown to be regulated by a transferrin-dependent pathway in the zebrafish embryo [177]. In humans hepcidin increases following the absorption of amounts of iron sufficient to acutely increase transferrin saturation [107]. (See 'Transferrin saturation' above.)

Transcription of hepcidin in response to increased plasma or tissue iron is mediated by bone morphogenetic proteins (BMP), requires hemojuvelin (HJV) as a BMP coreceptor, and is SMAD-dependent (figure 4) [62,178]. Increased hepcidin production is seen in the acute and chronic inflammation, mediated by lipopolysaccharide, interleukin (IL)-6 and IL-1-beta (figure 4) [114] and is an essential component of anemia of acute inflammation and anemia of chronic diseases [179,180].

Since most iron is used by maturing erythroid cells, several conditions including iron deficiency anemia, hypoxia, and erythropoietic expansion decrease hepcidin production to favor iron acquisition [119]. Suppression of hepcidin in hypoxia, initially proposed to be mediated by HIF-1 alpha (HIF-1α), occurs indirectly through erythropoietic expansion [181].

Ferrokinetic studies provided earlier information about erythropoiesis. The methods consisted of intravenously injecting a tracer label of Fe-59 bound to plasma transferrin. Three measurements were then made: the disappearance rate of Fe-59 from plasma over a period of minutes provided an index of beginning erythropoiesis; the appearance of Fe-59 by scanning over the spleen, liver, and bone marrow indicated the erythropoietic sites; and the total amount of iron appearing in red blood cells (RBCs) in the circulation 7 to 10 days later provided a measure of effective erythropoiesis. As an example, in severe beta thalassemia, the very rapid disappearance of plasma iron indicated a massive onset of erythropoiesis, and the low incorporation of injected iron in RBCs (20 to 30 percent, versus a normal value of >80 percent) showed that the increased erythropoiesis was severely ineffective. The existence of a regulator of iron absorption produced by erythroblasts was first proposed based on ferrokinetic studies [182]. It was subsequently documented that both effective and ineffective erythropoiesis upregulate iron acquisition by suppressing hepcidin.

Candidates for this hepcidin-suppressing function are cytokines produced by the erythroblasts. Growth differentiation factor 15 (GDF15), which is especially increased in serum of patients with homozygous beta-thalassemia, was first proposed as a major hepcidin inhibitor. However, mice with GDF15 gene deletion treated with phlebotomy are able to suppress hepcidin similarly to wild type animals [183]. Erythroferrone (ERFE) plays a major role in hepcidin inhibition [184]. In vitro, ERFE binds and sequesters different BMPs, including BMP6 and BMP2, attenuating the SMAD signaling [144,145]. However, full hepcidin inhibition by ERFE requires an attenuated BMP pathway, strengthening the complex process of hepcidin inhibitory control [185].

A role for PDGF-BB as a hepcidin inhibitor in hypoxia has been demonstrated in humans [186].

In iron deficiency, the BMP pathway is attenuated. In addition to cleavage of HJV by TMPRSS6, a study has described a role for epigenetic downregulation of the pathway [187].

The iron sensing and signaling pathway involving hepcidin is complex and not fully elucidated [101,188,189]. The proposed model is shown in the figure (figure 4).

BMP6 in iron overload binds and activates its own receptors (BMPRs) in the presence of the coreceptor HJV. In agreement with this model, BMP6 inactivation in mice causes severe iron overload with low hepcidin [131,132].

BMP2 participates to hepcidin activation likely setting up the basal hepcidin levels. It does not respond significantly to iron increase and signals prevalently through ALK3.

Inactivation of BMP2 in mice causes severe iron overload that is not compensated by BMP6.

BMPs bind to BMPRs. Conditional inactivation of the BMPR Alk2 and Alk3 in mouse liver causes iron overload of different severity [190]. Signal transduction of BMPs occurs through SMAD proteins; conditional inactivation of Smad4 in mouse liver causes liver iron accumulation and the inability to upregulate hepcidin, findings that are similar to those seen in hereditary hemochromatosis [191].

Whether and how the HFE TFR2 make a complex in the presence of increased diferric transferrin that cooperate with the BMP-HJV-SMAD pathway for hepcidin activation is not fully defined. In HFE-/- mice, the BMP pathway is indeed less active, and treatment with BMP6 appears to ameliorate iron overload [192].

Inflammatory cytokines, especially IL-6 (and IL1-beta) activate hepcidin transcription through interaction with the IL-6 receptor and signal transduction through STAT3 [47,62,107,193].

There is a crosstalk between the two pathways of hepcidin activation (inflammation- and iron-dependent) as shown by improved hepcidin control by compounds that inhibit the BMP-SMAD pathway in inflammation [194].

In iron deficiency, TMPRSS6 seems to be the major regulator of hepcidin suppression, cleaving the BMP coreceptor HJV. When erythropoiesis is expanded by erythropoietin, ERFE plays a major role in downregulating hepcidin expression, both in effective and ineffective erythropoiesis.

IRON AND THE BRAIN — Iron is essential for neuronal function, neurotransmitter synthesis, and myelinization [195]. Iron is taken up from plasma transferrin by the endothelial cells of the blood-brain barrier, and then it is released through ferroportin to the cerebral fluid, where it is bound to local transferrin.

Iron regulation, distribution, and exchange between neurons and glial cells remain incompletely understood. Neurons express high levels of the transferrin receptor (TfR), while glial cells mainly take up non-transferrin bound iron (NTBI) and express high levels of ferritin. Some areas of the brain, such as the globus pallidus and substantia nigra, accumulate ferritin-iron over time, perhaps as a form of iron storage.

Specific regions of the brain are susceptible to iron accumulation in rare disorders collectively called neurodegeneration with brain iron accumulation (NBIA). (See "Bradykinetic movement disorders in children", section on 'Neurodegeneration with brain iron accumulation'.)

Neurodegenerative disorders of aging, such as Parkinson disease, Alzheimer disease, Huntington disease, amyotrophic lateral sclerosis, and Friedreich ataxia have been associated with excessive intracellular free iron and ferroptosis, a form of cell death due to iron-induced lipid peroxidation [196-198].

In Friedreich ataxia, an autosomal recessive progressive neurodegenerative disorder mainly affecting the spinal cord and cerebellum reduced expression of frataxin results in toxic iron accumulation [199]. This causes mitochondrial oxidative stress and dysfunction along with an abnormality in iron sulfur-cluster formation [200]. (See "Friedreich ataxia".)

In Parkinson disease and Alzheimer disease, it is still debated whether spatio-temporal accumulation of iron is a causative disease trigger or a secondary effect due to cell degradation [201]. The neurotransmitter dopamine has been identified as an iron chaperone that can shuttle iron across membranes and affect cellular iron homeostasis [202]. Dopamine deficiency likely results in iron misdistribution in the brain. (See "Epidemiology, pathogenesis, and genetics of Parkinson disease", section on 'Basal ganglia circuits' and "Epidemiology, pathogenesis, and genetics of Parkinson disease", section on 'Iron metabolism' and "Epidemiology, pathology, and pathogenesis of Alzheimer disease", section on 'Pathogenesis'.)

In Parkinson disease-related restless legs syndrome, a frequent sensorimotor disorder in which iron deficiency has been causally related to disease severity, treatment with L-dopamine results in improvement of symptoms [203,204], in parallel with improvement of mitochondrial iron deficiency and mitochondrial functionality [205]. (See "Clinical manifestations of Parkinson disease", section on 'Sleep disorders'.)  

Iron accumulation has been observed in dopaminergic neurons in individuals with Parkinson disease [206]. Based on this, a 2022 study investigated the efficacy of the iron-chelator deferiprone in patients with newly diagnosed Parkinson disease; however, this found worsening of symptoms, likely due to reduced cellular and mitochondrial iron in neurons [207].

An improved understanding of cellular and spatiotemporal iron fluxes in the brain and their underlying regulatory processes will be essential to estimate the true impact of abnormalities in iron regulation in neurological disorders and may offer an opportunity to develop targeted interventions.

SUMMARY

Proteins involved in iron regulation – The regulation of iron metabolism involves the interaction of a number of specific proteins as well as the interplay between iron absorption from the gastrointestinal tract, recycling of iron from red cells at the end of their life span, release of iron stores from the monocyte-macrophage system, and iron loss from the body (figure 1 and figure 2 and figure 3). (See 'Role of specific proteins' above.)

Intestinal iron absorption – Intestinal iron absorption is tightly controlled, since there is no physiologic means of excreting iron from the body once it is absorbed. Factors that affect absorption are summarized in the table (table 1). HIF-2alpha and hepcidin are the most important regulators of apical and basolateral transporters, respectively (See 'Intestinal iron absorption' above.)

Role of ferroportin – Ferroportin functions as a major exporter of iron, transporting iron from mother to fetus, transferring absorbed iron from enterocytes into the circulation, and allowing macrophages to recycle iron from senescent and damaged red cells back into the circulation. (See 'Ferroportin' above and 'Iron release from macrophages' above.)

Role of hepcidin – Hepcidin downregulates ferroportin post-translationally. This is the most important control of iron cycling through ferroportin (figure 4). (See 'Hepcidin' above.)

Iron loss – Iron is lost in sweat, shed skin cells, and perhaps some gastrointestinal loss at a rate of approximately 1 mg/day. Females with additional menstrual iron loss of 1 to 2 mg/day generally have lower iron stores than males or postmenopausal females and are delicately poised to become iron deficient. (See 'Iron loss' above and "Causes and diagnosis of iron deficiency and iron deficiency anemia in adults", section on 'Search for source of blood and iron loss'.)

Iron in the brain – Iron is essential in the brain, but its regulation and distribution remain incompletely understood. Abnormal iron accumulation has been seen in Parkinson disease, Alzheimer disease, and neurodegeneration with brain iron accumulation (NBIA), but it remains unclear whether iron accumulation is a cause of neuronal dysfunction or a consequence. Greater understanding might offer the opportunity to develop targeted interventions. (See 'Iron and the brain' above.)

ACKNOWLEDGMENTS

UpToDate gratefully acknowledges Stanley L Schrier, MD (deceased), who contributed as Section Editor on earlier versions of this topic review and was a founding Editor-in-Chief for UpToDate in Hematology.

The UpToDate editorial staff also acknowledges the extensive contributions of William C Mentzer, MD, to earlier versions of this and many other topic reviews.

  1. Hentze MW, Muckenthaler MU, Galy B, Camaschella C. Two to tango: regulation of Mammalian iron metabolism. Cell 2010; 142:24.
  2. Camaschella C, Nai A, Silvestri L. Iron metabolism and iron disorders revisited in the hepcidin era. Haematologica 2020; 105:260.
  3. Kawabata H. Transferrin and transferrin receptors update. Free Radic Biol Med 2019; 133:46.
  4. Parrow NL, Li Y, Feola M, et al. Lobe specificity of iron binding to transferrin modulates murine erythropoiesis and iron homeostasis. Blood 2019; 134:1373.
  5. Athiyarath R, Arora N, Fuster F, et al. Two novel missense mutations in iron transport protein transferrin causing hypochromic microcytic anaemia and haemosiderosis: molecular characterization and structural implications. Br J Haematol 2013; 163:404.
  6. Zhu BM, McLaughlin SK, Na R, et al. Hematopoietic-specific Stat5-null mice display microcytic hypochromic anemia associated with reduced transferrin receptor gene expression. Blood 2008; 112:2071.
  7. Levy JE, Jin O, Fujiwara Y, et al. Transferrin receptor is necessary for development of erythrocytes and the nervous system. Nat Genet 1999; 21:396.
  8. Xu W, Barrientos T, Mao L, et al. Lethal Cardiomyopathy in Mice Lacking Transferrin Receptor in the Heart. Cell Rep 2015; 13:533.
  9. Chen AC, Donovan A, Ned-Sykes R, Andrews NC. Noncanonical role of transferrin receptor 1 is essential for intestinal homeostasis. Proc Natl Acad Sci U S A 2015; 112:11714.
  10. Fillebeen C, Charlebois E, Wagner J, et al. Transferrin receptor 1 controls systemic iron homeostasis by fine-tuning hepcidin expression to hepatocellular iron load. Blood 2019; 133:344.
  11. Xiao X, Moschetta GA, Xu Y, et al. Regulation of iron homeostasis by hepatocyte TfR1 requires HFE and contributes to hepcidin suppression in β-thalassemia. Blood 2023; 141:422.
  12. Jabara HH, Boyden SE, Chou J, et al. A missense mutation in TFRC, encoding transferrin receptor 1, causes combined immunodeficiency. Nat Genet 2016; 48:74.
  13. Kaup M, Dassler K, Weise C, Fuchs H. Shedding of the transferrin receptor is mediated constitutively by an integral membrane metalloprotease sensitive to tumor necrosis factor alpha protease inhibitor-2. J Biol Chem 2002; 277:38494.
  14. Beguin Y, Clemons GK, Pootrakul P, Fillet G. Quantitative assessment of erythropoiesis and functional classification of anemia based on measurements of serum transferrin receptor and erythropoietin. Blood 1993; 81:1067.
  15. Wessling-Resnick M. Crossing the Iron Gate: Why and How Transferrin Receptors Mediate Viral Entry. Annu Rev Nutr 2018; 38:431.
  16. Montemiglio LC, Testi C, Ceci P, et al. Cryo-EM structure of the human ferritin-transferrin receptor 1 complex. Nat Commun 2019; 10:1121.
  17. Arosio P, Levi S. Cytosolic and mitochondrial ferritins in the regulation of cellular iron homeostasis and oxidative damage. Biochim Biophys Acta 2010; 1800:783.
  18. Shi H, Bencze KZ, Stemmler TL, Philpott CC. A cytosolic iron chaperone that delivers iron to ferritin. Science 2008; 320:1207.
  19. Ryu MS, Zhang D, Protchenko O, et al. PCBP1 and NCOA4 regulate erythroid iron storage and heme biosynthesis. J Clin Invest 2017; 127:1786.
  20. Yanatori I, Richardson DR, Toyokuni S, Kishi F. The new role of poly (rC)-binding proteins as iron transport chaperones: Proteins that could couple with inter-organelle interactions to safely traffic iron. Biochim Biophys Acta Gen Subj 2020; 1864:129685.
  21. Philpott CC, Patel SJ, Protchenko O. Management versus miscues in the cytosolic labile iron pool: The varied functions of iron chaperones. Biochim Biophys Acta Mol Cell Res 2020; 1867:118830.
  22. Hintze KJ, Theil EC. DNA and mRNA elements with complementary responses to hemin, antioxidant inducers, and iron control ferritin-L expression. Proc Natl Acad Sci U S A 2005; 102:15048.
  23. Wang W, Knovich MA, Coffman LG, et al. Serum ferritin: Past, present and future. Biochim Biophys Acta 2010; 1800:760.
  24. Ferreira C, Bucchini D, Martin ME, et al. Early embryonic lethality of H ferritin gene deletion in mice. J Biol Chem 2000; 275:3021.
  25. Ferreira C, Santambrogio P, Martin ME, et al. H ferritin knockout mice: a model of hyperferritinemia in the absence of iron overload. Blood 2001; 98:525.
  26. Vanoaica L, Darshan D, Richman L, et al. Intestinal ferritin H is required for an accurate control of iron absorption. Cell Metab 2010; 12:273.
  27. Mancias JD, Wang X, Gygi SP, et al. Quantitative proteomics identifies NCOA4 as the cargo receptor mediating ferritinophagy. Nature 2014; 509:105.
  28. Dowdle WE, Nyfeler B, Nagel J, et al. Selective VPS34 inhibitor blocks autophagy and uncovers a role for NCOA4 in ferritin degradation and iron homeostasis in vivo. Nat Cell Biol 2014; 16:1069.
  29. Bellelli R, Federico G, Matte' A, et al. NCOA4 Deficiency Impairs Systemic Iron Homeostasis. Cell Rep 2016; 14:411.
  30. Vaisman B, Meyron-Holtz EG, Fibach E, et al. Ferritin expression in maturing normal human erythroid precursors. Br J Haematol 2000; 110:394.
  31. Nai A, Lidonnici MR, Federico G, et al. NCOA4-mediated ferritinophagy in macrophages is crucial to sustain erythropoiesis in mice. Haematologica 2021; 106:795.
  32. Finch CA, Bellotti V, Stray S, et al. Plasma ferritin determination as a diagnostic tool. West J Med 1986; 145:657.
  33. Wafa A, Hicham H, Naoufal R, et al. Clinical spectrum and therapeutic management of systemic lupus erythematosus-associated macrophage activation syndrome: a study of 20 Moroccan adult patients. Clin Rheumatol 2022; 41:2021.
  34. Bellmann-Weiler R, Lanser L, Barket R, et al. Prevalence and Predictive Value of Anemia and Dysregulated Iron Homeostasis in Patients with COVID-19 Infection. J Clin Med 2020; 9.
  35. Annous Y, Manning S, Khoujah D. Ferritin, fever, and frequent visits: Hyperferritinemic syndromes in the emergency department. Am J Emerg Med 2021; 48:249.
  36. Truman-Rosentsvit M, Berenbaum D, Spektor L, et al. Ferritin is secreted via 2 distinct nonclassical vesicular pathways. Blood 2018; 131:342.
  37. Li JY, Paragas N, Ned RM, et al. Scara5 is a ferritin receptor mediating non-transferrin iron delivery. Dev Cell 2009; 16:35.
  38. Chen TT, Li L, Chung DH, et al. TIM-2 is expressed on B cells and in liver and kidney and is a receptor for H-ferritin endocytosis. J Exp Med 2005; 202:955.
  39. Iwai K. Regulation of cellular iron metabolism: Iron-dependent degradation of IRP by SCFFBXL5 ubiquitin ligase. Free Radic Biol Med 2019; 133:64.
  40. Wilkinson N, Pantopoulos K. The IRP/IRE system in vivo: insights from mouse models. Front Pharmacol 2014; 5:176.
  41. Wilkinson N, Pantopoulos K. IRP1 regulates erythropoiesis and systemic iron homeostasis by controlling HIF2α mRNA translation. Blood 2013; 122:1658.
  42. Anderson SA, Nizzi CP, Chang YI, et al. The IRP1-HIF-2α axis coordinates iron and oxygen sensing with erythropoiesis and iron absorption. Cell Metab 2013; 17:282.
  43. Ghosh MC, Zhang DL, Jeong SY, et al. Deletion of iron regulatory protein 1 causes polycythemia and pulmonary hypertension in mice through translational derepression of HIF2α. Cell Metab 2013; 17:271.
  44. Zhou XY, Tomatsu S, Fleming RE, et al. HFE gene knockout produces mouse model of hereditary hemochromatosis. Proc Natl Acad Sci U S A 1998; 95:2492.
  45. Vujić Spasić M, Kiss J, Herrmann T, et al. Hfe acts in hepatocytes to prevent hemochromatosis. Cell Metab 2008; 7:173.
  46. Vujic Spasic M, Kiss J, Herrmann T, et al. Physiologic systemic iron metabolism in mice deficient for duodenal Hfe. Blood 2007; 109:4511.
  47. Piperno A, Girelli D, Nemeth E, et al. Blunted hepcidin response to oral iron challenge in HFE-related hemochromatosis. Blood 2007; 110:4096.
  48. Goswami T, Andrews NC. Hereditary hemochromatosis protein, HFE, interaction with transferrin receptor 2 suggests a molecular mechanism for mammalian iron sensing. J Biol Chem 2006; 281:28494.
  49. Rishi G, Crampton EM, Wallace DF, Subramaniam VN. In situ proximity ligation assays indicate that hemochromatosis proteins Hfe and transferrin receptor 2 (Tfr2) do not interact. PLoS One 2013; 8:e77267.
  50. Wallace DF, Summerville L, Crampton EM, et al. Combined deletion of Hfe and transferrin receptor 2 in mice leads to marked dysregulation of hepcidin and iron overload. Hepatology 2009; 50:1992.
  51. Wu XG, Wang Y, Wu Q, et al. HFE interacts with the BMP type I receptor ALK3 to regulate hepcidin expression. Blood 2014; 124:1335.
  52. Demetz E, Tymoszuk P, Hilbe R, et al. The haemochromatosis gene Hfe and Kupffer cells control LDL cholesterol homeostasis and impact on atherosclerosis development. Eur Heart J 2020; 41:3949.
  53. Kawabata H, Yang R, Hirama T, et al. Molecular cloning of transferrin receptor 2. A new member of the transferrin receptor-like family. J Biol Chem 1999; 274:20826.
  54. Johnson MB, Chen J, Murchison N, et al. Transferrin receptor 2: evidence for ligand-induced stabilization and redirection to a recycling pathway. Mol Biol Cell 2007; 18:743.
  55. Wallace DF, Summerville L, Subramaniam VN. Targeted disruption of the hepatic transferrin receptor 2 gene in mice leads to iron overload. Gastroenterology 2007; 132:301.
  56. Roetto A, Di Cunto F, Pellegrino RM, et al. Comparison of 3 Tfr2-deficient murine models suggests distinct functions for Tfr2-alpha and Tfr2-beta isoforms in different tissues. Blood 2010; 115:3382.
  57. Forejtnikovà H, Vieillevoye M, Zermati Y, et al. Transferrin receptor 2 is a component of the erythropoietin receptor complex and is required for efficient erythropoiesis. Blood 2010; 116:5357.
  58. Nai A, Lidonnici MR, Rausa M, et al. The second transferrin receptor regulates red blood cell production in mice. Blood 2015; 125:1170.
  59. Rauner M, Baschant U, Roetto A, et al. Transferrin receptor 2 controls bone mass and pathological bone formation via BMP and Wnt signaling. Nat Metab 2019; 1:111.
  60. Papanikolaou G, Samuels ME, Ludwig EH, et al. Mutations in HFE2 cause iron overload in chromosome 1q-linked juvenile hemochromatosis. Nat Genet 2004; 36:77.
  61. Lin L, Goldberg YP, Ganz T. Competitive regulation of hepcidin mRNA by soluble and cell-associated hemojuvelin. Blood 2005; 106:2884.
  62. Babitt JL, Huang FW, Wrighting DM, et al. Bone morphogenetic protein signaling by hemojuvelin regulates hepcidin expression. Nat Genet 2006; 38:531.
  63. Silvestri L, Pagani A, Camaschella C. Furin-mediated release of soluble hemojuvelin: a new link between hypoxia and iron homeostasis. Blood 2008; 111:924.
  64. Silvestri L, Pagani A, Nai A, et al. The serine protease matriptase-2 (TMPRSS6) inhibits hepcidin activation by cleaving membrane hemojuvelin. Cell Metab 2008; 8:502.
  65. Gunshin H, Mackenzie B, Berger UV, et al. Cloning and characterization of a mammalian proton-coupled metal-ion transporter. Nature 1997; 388:482.
  66. Fleming MD, Trenor CC 3rd, Su MA, et al. Microcytic anaemia mice have a mutation in Nramp2, a candidate iron transporter gene. Nat Genet 1997; 16:383.
  67. Mastrogiannaki M, Matak P, Peyssonnaux C. The gut in iron homeostasis: role of HIF-2 under normal and pathological conditions. Blood 2013; 122:885.
  68. Mims MP, Guan Y, Pospisilova D, et al. Identification of a human mutation of DMT1 in a patient with microcytic anemia and iron overload. Blood 2005; 105:1337.
  69. McKie AT, Barrow D, Latunde-Dada GO, et al. An iron-regulated ferric reductase associated with the absorption of dietary iron. Science 2001; 291:1755.
  70. Yanatori I, Kishi F. DMT1 and iron transport. Free Radic Biol Med 2019; 133:55.
  71. Schwartz AJ, Das NK, Ramakrishnan SK, et al. Hepatic hepcidin/intestinal HIF-2α axis maintains iron absorption during iron deficiency and overload. J Clin Invest 2019; 129:336.
  72. Liu Q, Davidoff O, Niss K, Haase VH. Hypoxia-inducible factor regulates hepcidin via erythropoietin-induced erythropoiesis. J Clin Invest 2012; 122:4635.
  73. Donovan A, Brownlie A, Zhou Y, et al. Positional cloning of zebrafish ferroportin1 identifies a conserved vertebrate iron exporter. Nature 2000; 403:776.
  74. Abboud S, Haile DJ. A novel mammalian iron-regulated protein involved in intracellular iron metabolism. J Biol Chem 2000; 275:19906.
  75. Donovan A, Lima CA, Pinkus JL, et al. The iron exporter ferroportin/Slc40a1 is essential for iron homeostasis. Cell Metab 2005; 1:191.
  76. Zhang DL, Wu J, Shah BN, et al. Erythrocytic ferroportin reduces intracellular iron accumulation, hemolysis, and malaria risk. Science 2018; 359:1520.
  77. Zhang DL, Ghosh MC, Ollivierre H, et al. Ferroportin deficiency in erythroid cells causes serum iron deficiency and promotes hemolysis due to oxidative stress. Blood 2018; 132:2078.
  78. Chen H, Su T, Attieh ZK, et al. Systemic regulation of Hephaestin and Ireg1 revealed in studies of genetic and nutritional iron deficiency. Blood 2003; 102:1893.
  79. Delaby C, Pilard N, Gonçalves AS, et al. Presence of the iron exporter ferroportin at the plasma membrane of macrophages is enhanced by iron loading and down-regulated by hepcidin. Blood 2005; 106:3979.
  80. Zhang DL, Hughes RM, Ollivierre-Wilson H, et al. A ferroportin transcript that lacks an iron-responsive element enables duodenal and erythroid precursor cells to evade translational repression. Cell Metab 2009; 9:461.
  81. Keel SB, Abkowitz JL. The microcytic red cell and the anemia of inflammation. N Engl J Med 2009; 361:1904.
  82. Zhang DL, Senecal T, Ghosh MC, et al. Hepcidin regulates ferroportin expression and intracellular iron homeostasis of erythroblasts. Blood 2011; 118:2868.
  83. Nemeth E, Tuttle MS, Powelson J, et al. Hepcidin regulates cellular iron efflux by binding to ferroportin and inducing its internalization. Science 2004; 306:2090.
  84. Rivera S, Nemeth E, Gabayan V, et al. Synthetic hepcidin causes rapid dose-dependent hypoferremia and is concentrated in ferroportin-containing organs. Blood 2005; 106:2196.
  85. Aschemeyer S, Qiao B, Stefanova D, et al. Structure-function analysis of ferroportin defines the binding site and an alternative mechanism of action of hepcidin. Blood 2018; 131:899.
  86. Marro S, Chiabrando D, Messana E, et al. Heme controls ferroportin1 (FPN1) transcription involving Bach1, Nrf2 and a MARE/ARE sequence motif at position -7007 of the FPN1 promoter. Haematologica 2010; 95:1261.
  87. Lakhal-Littleton S, Wolna M, Carr CA, et al. Cardiac ferroportin regulates cellular iron homeostasis and is important for cardiac function. Proc Natl Acad Sci U S A 2015; 112:3164.
  88. Guida C, Altamura S, Klein FA, et al. A novel inflammatory pathway mediating rapid hepcidin-independent hypoferremia. Blood 2015; 125:2265.
  89. Ludwiczek S, Aigner E, Theurl I, Weiss G. Cytokine-mediated regulation of iron transport in human monocytic cells. Blood 2003; 101:4148.
  90. Njajou OT, Vaessen N, Joosse M, et al. A mutation in SLC11A3 is associated with autosomal dominant hemochromatosis. Nat Genet 2001; 28:213.
  91. Pietrangelo A, Montosi G, Totaro A, et al. Hereditary hemochromatosis in adults without pathogenic mutations in the hemochromatosis gene. N Engl J Med 1999; 341:725.
  92. Montosi G, Donovan A, Totaro A, et al. Autosomal-dominant hemochromatosis is associated with a mutation in the ferroportin (SLC11A3) gene. J Clin Invest 2001; 108:619.
  93. Vulpe CD, Kuo YM, Murphy TL, et al. Hephaestin, a ceruloplasmin homologue implicated in intestinal iron transport, is defective in the sla mouse. Nat Genet 1999; 21:195.
  94. Keel SB, Doty RT, Yang Z, et al. A heme export protein is required for red blood cell differentiation and iron homeostasis. Science 2008; 319:825.
  95. Mercurio S, Petrillo S, Chiabrando D, et al. The heme exporter Flvcr1 regulates expansion and differentiation of committed erythroid progenitors by controlling intracellular heme accumulation. Haematologica 2015; 100:720.
  96. Fiorito V, Forni M, Silengo L, et al. Crucial Role of FLVCR1a in the Maintenance of Intestinal Heme Homeostasis. Antioxid Redox Signal 2015; 23:1410.
  97. Jenkitkasemwong S, Wang CY, Coffey R, et al. SLC39A14 Is Required for the Development of Hepatocellular Iron Overload in Murine Models of Hereditary Hemochromatosis. Cell Metab 2015; 22:138.
  98. Knutson MD. Non-transferrin-bound iron transporters. Free Radic Biol Med 2019; 133:101.
  99. Pigeon C, Ilyin G, Courselaud B, et al. A new mouse liver-specific gene, encoding a protein homologous to human antimicrobial peptide hepcidin, is overexpressed during iron overload. J Biol Chem 2001; 276:7811.
  100. Nicolas G, Bennoun M, Devaux I, et al. Lack of hepcidin gene expression and severe tissue iron overload in upstream stimulatory factor 2 (USF2) knockout mice. Proc Natl Acad Sci U S A 2001; 98:8780.
  101. Ganz T. Hepcidin and iron regulation, 10 years later. Blood 2011; 117:4425.
  102. Stefanova D, Raychev A, Arezes J, et al. Endogenous hepcidin and its agonist mediate resistance to selected infections by clearing non-transferrin-bound iron. Blood 2017; 130:245.
  103. Girelli D, Nemeth E, Swinkels DW. Hepcidin in the diagnosis of iron disorders. Blood 2016; 127:2809.
  104. van der Vorm LN, Hendriks JC, Laarakkers CM, et al. Toward Worldwide Hepcidin Assay Harmonization: Identification of a Commutable Secondary Reference Material. Clin Chem 2016; 62:993.
  105. Galesloot TE, Vermeulen SH, Geurts-Moespot AJ, et al. Serum hepcidin: reference ranges and biochemical correlates in the general population. Blood 2011; 117:e218.
  106. Traglia M, Girelli D, Biino G, et al. Association of HFE and TMPRSS6 genetic variants with iron and erythrocyte parameters is only in part dependent on serum hepcidin concentrations. J Med Genet 2011; 48:629.
  107. Lin L, Valore EV, Nemeth E, et al. Iron transferrin regulates hepcidin synthesis in primary hepatocyte culture through hemojuvelin and BMP2/4. Blood 2007; 110:2182.
  108. Muckenthaler M, Roy CN, Custodio AO, et al. Regulatory defects in liver and intestine implicate abnormal hepcidin and Cybrd1 expression in mouse hemochromatosis. Nat Genet 2003; 34:102.
  109. Wang CY, Babitt JL. Liver iron sensing and body iron homeostasis. Blood 2019; 133:18.
  110. Theurl I, Theurl M, Seifert M, et al. Autocrine formation of hepcidin induces iron retention in human monocytes. Blood 2008; 111:2392.
  111. Hintze KJ, Snow D, Nabor D, Timbimboo H. Adipocyte hypoxia increases hepatocyte hepcidin expression. Biol Trace Elem Res 2011; 143:764.
  112. Gnana-Prakasam JP, Martin PM, Mysona BA, et al. Hepcidin expression in mouse retina and its regulation via lipopolysaccharide/Toll-like receptor-4 pathway independent of Hfe. Biochem J 2008; 411:79.
  113. Lakhal-Littleton S, Wolna M, Chung YJ, et al. An essential cell-autonomous role for hepcidin in cardiac iron homeostasis. Elife 2016; 5.
  114. Nicolas G, Chauvet C, Viatte L, et al. The gene encoding the iron regulatory peptide hepcidin is regulated by anemia, hypoxia, and inflammation. J Clin Invest 2002; 110:1037.
  115. Weinstein DA, Roy CN, Fleming MD, et al. Inappropriate expression of hepcidin is associated with iron refractory anemia: implications for the anemia of chronic disease. Blood 2002; 100:3776.
  116. Nicolas G, Viatte L, Lou DQ, et al. Constitutive hepcidin expression prevents iron overload in a mouse model of hemochromatosis. Nat Genet 2003; 34:97.
  117. Nemeth E, Roetto A, Garozzo G, et al. Hepcidin is decreased in TFR2 hemochromatosis. Blood 2005; 105:1803.
  118. Camaschella C. Understanding iron homeostasis through genetic analysis of hemochromatosis and related disorders. Blood 2005; 106:3710.
  119. Muckenthaler MU, Rivella S, Hentze MW, Galy B. A Red Carpet for Iron Metabolism. Cell 2017; 168:344.
  120. Latour C, Kautz L, Besson-Fournier C, et al. Testosterone perturbs systemic iron balance through activation of epidermal growth factor receptor signaling in the liver and repression of hepcidin. Hepatology 2014; 59:683.
  121. Lesbordes-Brion JC, Viatte L, Bennoun M, et al. Targeted disruption of the hepcidin 1 gene results in severe hemochromatosis. Blood 2006; 108:1402.
  122. Zumerle S, Mathieu JR, Delga S, et al. Targeted disruption of hepcidin in the liver recapitulates the hemochromatotic phenotype. Blood 2014; 123:3646.
  123. Roetto A, Papanikolaou G, Politou M, et al. Mutant antimicrobial peptide hepcidin is associated with severe juvenile hemochromatosis. Nat Genet 2003; 33:21.
  124. Nicolas G, Bennoun M, Porteu A, et al. Severe iron deficiency anemia in transgenic mice expressing liver hepcidin. Proc Natl Acad Sci U S A 2002; 99:4596.
  125. Roy CN, Mak HH, Akpan I, et al. Hepcidin antimicrobial peptide transgenic mice exhibit features of the anemia of inflammation. Blood 2007; 109:4038.
  126. Gardenghi S, Ramos P, Marongiu MF, et al. Hepcidin as a therapeutic tool to limit iron overload and improve anemia in β-thalassemic mice. J Clin Invest 2010; 120:4466.
  127. Billesbølle CB, Azumaya CM, Kretsch RC, et al. Structure of hepcidin-bound ferroportin reveals iron homeostatic mechanisms. Nature 2020; 586:807.
  128. Malerba M, Louis S, Cuvellier S, et al. Epidermal hepcidin is required for neutrophil response to bacterial infection. J Clin Invest 2020; 130:329.
  129. Bessman NJ, Mathieu JRR, Renassia C, et al. Dendritic cell-derived hepcidin sequesters iron from the microbiota to promote mucosal healing. Science 2020; 368:186.
  130. Schmidt PJ, Fleming MD. Modulation of hepcidin as therapy for primary and secondary iron overload disorders: preclinical models and approaches. Hematol Oncol Clin North Am 2014; 28:387.
  131. Andriopoulos B Jr, Corradini E, Xia Y, et al. BMP6 is a key endogenous regulator of hepcidin expression and iron metabolism. Nat Genet 2009; 41:482.
  132. Meynard D, Kautz L, Darnaud V, et al. Lack of the bone morphogenetic protein BMP6 induces massive iron overload. Nat Genet 2009; 41:478.
  133. Canali S, Zumbrennen-Bullough KB, Core AB, et al. Endothelial cells produce bone morphogenetic protein 6 required for iron homeostasis in mice. Blood 2017; 129:405.
  134. Koch PS, Olsavszky V, Ulbrich F, et al. Angiocrine Bmp2 signaling in murine liver controls normal iron homeostasis. Blood 2017; 129:415.
  135. Xiao X, Dev S, Canali S, et al. Endothelial Bone Morphogenetic Protein 2 (Bmp2) Knockout Exacerbates Hemochromatosis in Homeostatic Iron Regulator (Hfe) Knockout Mice but not Bmp6 Knockout Mice. Hepatology 2020; 72:642.
  136. Heeney MM, Finberg KE. Iron-refractory iron deficiency anemia (IRIDA). Hematol Oncol Clin North Am 2014; 28:637.
  137. Enns CA, Jue S, Zhang AS. The ectodomain of matriptase-2 plays an important nonproteolytic role in suppressing hepcidin expression in mice. Blood 2020; 136:989.
  138. Finberg KE, Whittlesey RL, Fleming MD, Andrews NC. Down-regulation of Bmp/Smad signaling by Tmprss6 is required for maintenance of systemic iron homeostasis. Blood 2010; 115:3817.
  139. Chambers JC, Zhang W, Li Y, et al. Genome-wide association study identifies variants in TMPRSS6 associated with hemoglobin levels. Nat Genet 2009; 41:1170.
  140. Ganesh SK, Zakai NA, van Rooij FJ, et al. Multiple loci influence erythrocyte phenotypes in the CHARGE Consortium. Nat Genet 2009; 41:1191.
  141. Soranzo N, Spector TD, Mangino M, et al. A genome-wide meta-analysis identifies 22 loci associated with eight hematological parameters in the HaemGen consortium. Nat Genet 2009; 41:1182.
  142. Tanaka T, Roy CN, Yao W, et al. A genome-wide association analysis of serum iron concentrations. Blood 2010; 115:94.
  143. Nai A, Pagani A, Silvestri L, et al. TMPRSS6 rs855791 modulates hepcidin transcription in vitro and serum hepcidin levels in normal individuals. Blood 2011; 118:4459.
  144. Arezes J, Foy N, McHugh K, et al. Erythroferrone inhibits the induction of hepcidin by BMP6. Blood 2018; 132:1473.
  145. Wang CY, Xu Y, Traeger L, et al. Erythroferrone lowers hepcidin by sequestering BMP2/6 heterodimer from binding to the BMP type I receptor ALK3. Blood 2020; 135:453.
  146. Arezes J, Foy N, McHugh K, et al. Antibodies against the erythroferrone N-terminal domain prevent hepcidin suppression and ameliorate murine thalassemia. Blood 2020; 135:547.
  147. Ganz T, Jung G, Naeim A, et al. Immunoassay for human serum erythroferrone. Blood 2017; 130:1243.
  148. Coffey R, Jung G, Olivera JD, et al. Erythroid overproduction of erythroferrone causes iron overload and developmental abnormalities in mice. Blood 2022; 139:439.
  149. Shayeghi M, Latunde-Dada GO, Oakhill JS, et al. Identification of an intestinal heme transporter. Cell 2005; 122:789.
  150. Qiu A, Jansen M, Sakaris A, et al. Identification of an intestinal folate transporter and the molecular basis for hereditary folate malabsorption. Cell 2006; 127:917.
  151. Stoffel NU, Zeder C, Brittenham GM, et al. Iron absorption from supplements is greater with alternate day than with consecutive day dosing in iron-deficient anemic women. Haematologica 2020; 105:1232.
  152. Moretti D, Goede JS, Zeder C, et al. Oral iron supplements increase hepcidin and decrease iron absorption from daily or twice-daily doses in iron-depleted young women. Blood 2015; 126:1981.
  153. Disler PB, Lynch SR, Charlton RW, et al. The effect of tea on iron absorption. Gut 1975; 16:193.
  154. Hallberg L, Rossander L, Skånberg AB. Phytates and the inhibitory effect of bran on iron absorption in man. Am J Clin Nutr 1987; 45:988.
  155. Cook JD, Monsen ER. Food iron absorption in human subjects. III. Comparison of the effect of animal proteins on nonheme iron absorption. Am J Clin Nutr 1976; 29:859.
  156. Storcksdieck genannt Bonsmann S, Hurrell RF. Iron-binding properties, amino acid composition, and structure of muscle tissue peptides from in vitro digestion of different meat sources. J Food Sci 2007; 72:S019.
  157. Armah CN, Sharp P, Mellon FA, et al. L-alpha-glycerophosphocholine contributes to meat's enhancement of nonheme iron absorption. J Nutr 2008; 138:873.
  158. Hutchinson C, Geissler CA, Powell JJ, Bomford A. Proton pump inhibitors suppress absorption of dietary non-haem iron in hereditary haemochromatosis. Gut 2007; 56:1291.
  159. Das NK, Schwartz AJ, Barthel G, et al. Microbial Metabolite Signaling Is Required for Systemic Iron Homeostasis. Cell Metab 2020; 31:115.
  160. Mastrogiannaki M, Matak P, Keith B, et al. HIF-2alpha, but not HIF-1alpha, promotes iron absorption in mice. J Clin Invest 2009; 119:1159.
  161. Shah YM, Matsubara T, Ito S, et al. Intestinal hypoxia-inducible transcription factors are essential for iron absorption following iron deficiency. Cell Metab 2009; 9:152.
  162. Das NK, Jain C, Sankar A, et al. Modulation of the HIF2α-NCOA4 axis in enterocytes attenuates iron loading in a mouse model of hemochromatosis. Blood 2022; 139:2547.
  163. Prentice AM, Doherty CP, Abrams SA, et al. Hepcidin is the major predictor of erythrocyte iron incorporation in anemic African children. Blood 2012; 119:1922.
  164. Finch CA, Huebers H. Perspectives in iron metabolism. N Engl J Med 1982; 306:1520.
  165. Cook JD. Clinical evaluation of iron deficiency. Semin Hematol 1982; 19:6.
  166. Alyanakian MA, Taes Y, Bensaïd M, et al. Monoclonal immunoglobulin with antitransferrin activity: A rare cause of hypersideremia with increased transferrin saturation. Blood 2007; 109:359.
  167. van Raaij SEG, Srai SKS, Swinkels DW, van Swelm RPL. Iron uptake by ZIP8 and ZIP14 in human proximal tubular epithelial cells. Biometals 2019; 32:211.
  168. Scindia PhD Y, Leeds Md J, Swaminathan Md S. Iron Homeostasis in Healthy Kidney and its Role in Acute Kidney Injury. Semin Nephrol 2019; 39:76.
  169. Korolnek T, Hamza I. Macrophages and iron trafficking at the birth and death of red cells. Blood 2015; 125:2893.
  170. Miyajima H. Aceruloplasminemia. In: GeneReviews, Pagon RA, Adam MP, Ardinger HH, et al. (Eds), University of Washington, Seattle, Seattle 1993-2017.
  171. Tanno T, Bhanu NV, Oneal PA, et al. High levels of GDF15 in thalassemia suppress expression of the iron regulatory protein hepcidin. Nat Med 2007; 13:1096.
  172. Weizer-Stern O, Adamsky K, Margalit O, et al. Hepcidin, a key regulator of iron metabolism, is transcriptionally activated by p53. Br J Haematol 2007; 138:253.
  173. Pinto JP, Ribeiro S, Pontes H, et al. Erythropoietin mediates hepcidin expression in hepatocytes through EPOR signaling and regulation of C/EBPalpha. Blood 2008; 111:5727.
  174. Huang H, Constante M, Layoun A, Santos MM. Contribution of STAT3 and SMAD4 pathways to the regulation of hepcidin by opposing stimuli. Blood 2009; 113:3593.
  175. Robach P, Recalcati S, Girelli D, et al. Alterations of systemic and muscle iron metabolism in human subjects treated with low-dose recombinant erythropoietin. Blood 2009; 113:6707.
  176. Piperno A, Galimberti S, Mariani R, et al. Modulation of hepcidin production during hypoxia-induced erythropoiesis in humans in vivo: data from the HIGHCARE project. Blood 2011; 117:2953.
  177. Fraenkel PG, Gibert Y, Holzheimer JL, et al. Transferrin-a modulates hepcidin expression in zebrafish embryos. Blood 2009; 113:2843.
  178. Finberg KE. Regulation of systemic iron homeostasis. Curr Opin Hematol 2013; 20:208.
  179. Ganz T. Anemia of Inflammation. N Engl J Med 2019; 381:1148.
  180. Weiss G, Ganz T, Goodnough LT. Anemia of inflammation. Blood 2019; 133:40.
  181. Peyssonnaux C, Zinkernagel AS, Schuepbach RA, et al. Regulation of iron homeostasis by the hypoxia-inducible transcription factors (HIFs). J Clin Invest 2007; 117:1926.
  182. Finch C. Regulators of iron balance in humans. Blood 1994; 84:1697.
  183. Casanovas G, Vujić Spasic M, Casu C, et al. The murine growth differentiation factor 15 is not essential for systemic iron homeostasis in phlebotomized mice. Haematologica 2013; 98:444.
  184. Kautz L, Jung G, Valore EV, et al. Identification of erythroferrone as an erythroid regulator of iron metabolism. Nat Genet 2014; 46:678.
  185. Nai A, Rubio A, Campanella A, et al. Limiting hepatic Bmp-Smad signaling by matriptase-2 is required for erythropoietin-mediated hepcidin suppression in mice. Blood 2016; 127:2327.
  186. Sonnweber T, Nachbaur D, Schroll A, et al. Hypoxia induced downregulation of hepcidin is mediated by platelet derived growth factor BB. Gut 2014; 63:1951.
  187. Pasricha SR, Lim PJ, Duarte TL, et al. Hepcidin is regulated by promoter-associated histone acetylation and HDAC3. Nat Commun 2017; 8:403.
  188. Camaschella C. Iron and hepcidin: a story of recycling and balance. Hematology Am Soc Hematol Educ Program 2013; 2013:1.
  189. Kautz L, Nemeth E. Molecular liaisons between erythropoiesis and iron metabolism. Blood 2014; 124:479.
  190. Steinbicker AU, Bartnikas TB, Lohmeyer LK, et al. Perturbation of hepcidin expression by BMP type I receptor deletion induces iron overload in mice. Blood 2011; 118:4224.
  191. Wang RH, Li C, Xu X, et al. A role of SMAD4 in iron metabolism through the positive regulation of hepcidin expression. Cell Metab 2005; 2:399.
  192. Corradini E, Schmidt PJ, Meynard D, et al. BMP6 treatment compensates for the molecular defect and ameliorates hemochromatosis in Hfe knockout mice. Gastroenterology 2010; 139:1721.
  193. Truksa J, Gelbart T, Peng H, et al. Suppression of the hepcidin-encoding gene Hamp permits iron overload in mice lacking both hemojuvelin and matriptase-2/TMPRSS6. Br J Haematol 2009; 147:571.
  194. Theurl I, Schroll A, Sonnweber T, et al. Pharmacologic inhibition of hepcidin expression reverses anemia of chronic inflammation in rats. Blood 2011; 118:4977.
  195. Porras CA, Rouault TA. Iron Homeostasis in the CNS: An Overview of the Pathological Consequences of Iron Metabolism Disruption. Int J Mol Sci 2022; 23.
  196. Goldsteins G, Hakosalo V, Jaronen M, et al. CNS Redox Homeostasis and Dysfunction in Neurodegenerative Diseases. Antioxidants (Basel) 2022; 11.
  197. Costa I, Barbosa DJ, Benfeito S, et al. Molecular mechanisms of ferroptosis and their involvement in brain diseases. Pharmacol Ther 2023; 244:108373.
  198. Sun Y, Xia X, Basnet D, et al. Mechanisms of Ferroptosis and Emerging Links to the Pathology of Neurodegenerative Diseases. Front Aging Neurosci 2022; 14:904152.
  199. Delatycki MB, Bidichandani SI. Friedreich ataxia- pathogenesis and implications for therapies. Neurobiol Dis 2019; 132:104606.
  200. Cheng R, Dhorajia VV, Kim J, Kim Y. Mitochondrial iron metabolism and neurodegenerative diseases. Neurotoxicology 2022; 88:88.
  201. Lane DJR, Metselaar B, Greenough M, et al. Ferroptosis and NRF2: an emerging battlefield in the neurodegeneration of Alzheimer's disease. Essays Biochem 2021; 65:925.
  202. Dichtl S, Haschka D, Nairz M, et al. Dopamine promotes cellular iron accumulation and oxidative stress responses in macrophages. Biochem Pharmacol 2018; 148:193.
  203. Trenkwalder C, Allen R, Högl B, et al. Comorbidities, treatment, and pathophysiology in restless legs syndrome. Lancet Neurol 2018; 17:994.
  204. Manconi M, Garcia-Borreguero D, Schormair B, et al. Restless legs syndrome. Nat Rev Dis Primers 2021; 7:80.
  205. Haschka D, Volani C, Stefani A, et al. Association of mitochondrial iron deficiency and dysfunction with idiopathic restless legs syndrome. Mov Disord 2019; 34:114.
  206. Ma L, Gholam Azad M, Dharmasivam M, et al. Parkinson's disease: Alterations in iron and redox biology as a key to unlock therapeutic strategies. Redox Biol 2021; 41:101896.
  207. Devos D, Labreuche J, Rascol O, et al. Trial of Deferiprone in Parkinson's Disease. N Engl J Med 2022; 387:2045.
Topic 7105 Version 72.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟