SciELO - Scientific Electronic Library Online

 
vol.97 issue4Clinical evaluation of drug-induced hepatitisLiver metastases from uveal melanoma author indexsubject indexarticles search
Home Pagealphabetic serial listing  

My SciELO

Services on Demand

Journal

Article

Indicators

Related links

Share


Revista Española de Enfermedades Digestivas

Print version ISSN 1130-0108

Rev. esp. enferm. dig. vol.97 n.4 Madrid Apr. 2005

 

POINT OF VIEW


Non-HFE hemochromatosis

J. A. Solís Herruzo and P. Solís Muñoz

Service of Digestive Diseases. Hospital Universitario 12 de Octubre. Madrid, Spain


Solís Herruzo JA, Solís Muñoz P. Non-HFE hemochromatosis. Rev Esp Enferm Dig 2005; 97: 266-286.


Recibido: 01-12-04.
Aceptado: 06-02-05.

Correspondencia: J.A. Solís-Herruzo. Servicio de Medicina de Aparato Digestivo. Hospital Universitario 12 de Octubre. Avda. de Córdoba, s/n. 28041 Madrid.
e-mail: jsolis.hdoc@salus.madrid.org

 

INTRODUCTION

Hereditary hemochromatosis (HH), a designation first used by v. Recklinghausen in 1888 (1), is an inherited, autosomal recessive disease characterized by excessive iron deposition in the liver, pancreas, heart, and other organs as a result of excessive intestinal iron absorption. In organs where iron excess accumulates, lesions develop that are responsible for clinical manifestations (liver cirrhosis, joint disease, heart failure, arrhythmias, endocrine disease, skin pigmentation, etc.). The search for a genetic cause led first to acknowledge an association with the short arm of chromosome 6, the area coding for the HLA-A3 molecule (2), and then to recognize the hemochromatosis gene-HFE (3). Two mutations were initially identified in the HFE gene-845G→A (C282Y), and the substitution 187G→C (H63D). The search for the C282Y mutation in the HFE gene of Northern European patients with HH has revealed that more than 80% of such individuals are homozygous for this mutation (4). In Northern European countries some C282Y non-homozygous patients are C282Y/H63D double heterozygotes. In these same countries the frequency of the C282Y mutation amongst the general population is 1 in 100 inhabitants in Ireland, and 1 in 400 inhabitants in the United States (5,6). In Southern European countries the frequency of this mutation is lower. In Greece, for instance, it is to be found not even in 1 in 100,000 population (7). This mutation is presumed to have developed in a Celt individual during the Bronze Age, later to spread in countries where Celtic penetration was greatest (8).

Protein HFE belongs in the family of HLA class 1 histocompatibility molecules. Similar to these molecules, it is 343-aminoacid glycoprotein located at the plasma membrane in some cells. It is made up of three extracellular loops (α1, α2, α3), a transmembrane portion, and an intracytoplasmic portion. It interacts with transferrin receptors at a gap between loops α1 and α2 (9). A thiol bridge between two cysteine molecules gives rise to loop α3. This loop is crucial for its non-covalent binding of β2-microglobulin (β2MG), and its expression in the surface of cells (10). It is currently accepted that this protein plays a relevant role in iron metabolism. However, its precise mechanism of action is unknown. Experiments performed in cultures of cells transfected with an HFE-expression plasmid revealed that iron passage into cells, and thus cell iron and ferritin content, decreased, while the number of transferrin receptors 1 (TfR1) increased. Hence, HFE seemed to behave as a negative regulator of TfR1 function (11-13) and a blocker of iron passage into cells. HFE was suggested to compete with transferrin (Tf) in binding TfR1 (14). However, these experiments have been criticized - overexpression of HFE alone is ineffective, as the presence of β2MG is required for its expression on the surface of cells. When an overexpression of both proteins - HFE and β2MG - occurs, cell iron and ferritin content dramatically increases (15). That is, HFE behaves as a facilitator of iron passage into cells. In fact, HFE mutations involving the region through which the binding of TfR1 occurs block internalization of the Tf/TfR1/HFE complex (16).

Two mechanisms have been suggested through which HFE may regulate intestinal iron absorption. One is related to the role ascribed to duodenal crypt cells as sensors of total body iron stores. HFE is found at the base membrane of these cells together with β2MG and TfRs. Upon arrival, iron-saturated Tf incorporates itself into the HFE/β2MG/TfR complex, which is then wholly internalized within an endosome (17). Tf releases iron within duodenal crypt cells. These cells maturate into enterocytes the way iron-rich cells should, that is, by inhibiting the synthesis of all those proteins favoring iron passage (Dcytb, DMT, ferroportin, hephaestin). Under such conditions intestinal iron absorption decreases (Fig. 1). In contrast, when little iron is transported by Tf the amount of iron Tf may deliver to duodenal crypt cells is small. These cells then maturate into enterocytes as iron-poor cells. That is, the synthesis of proteins favoring iron passage increases (Fig. 2) (reviewed in 18). More recently, it has been suggested that the synthesis of hepcidin, a hormone of liver origin that slows intestinal iron absorption, may be stimulated by HFE, as well as by infection, iron, TfR2, and hemojuvelin (reviewed in 19). TfR2 and HFE favor the passage of iron into liver cells, which would be responsible for hepcidin induction. Thus, normal HFE would slow intestinal iron absorption after inducing the synthesis of hepcidin.

The C282Y mutation results in the substitution of tyrosine for cysteine 282. This brings about a severe structural distortion in the HFE molecule. Specifically, it blocks the formation of the α3 loop upon the disappearance of its constituent thiol bond. This molecular malformation prevents HFE - following its synthesis in the endoplasmic reticule - from binding β2MG and therefore reaching the base membrane of cells. The absence of this protein may increase intestinal iron absorption through either of the above-mentioned mechanisms: a) its absence at the base membrane of duodenal crypt cells blocks the passage of iron into these cells. Iron deficiency in these cells determines their maturation through increasing iron-binding proteins and favoring intestinal iron absorption; and b) similarly, a lack of HFE determines a decrease of hepcidin formation in the liver, and hence an increase in intestinal iron absorption. In fact, patients with HH have been found to have greatly reduced hepcidin concentrations in their urine (20).

Mutation H63D is very common in Europe, particularly in Spain. It has been estimated that 22% of the European population are heterozygous for this mutation, 2% are homozygous for this mutation, and 2% are compound heterozygous in combination with C282Y (21). Even with homozygous status this mutation has little impact on iron homeostasis. Only patients with C282Y/H63D may exhibit morbid iron overload. Overall these are moderate, low-penetrance overloads (22).

While many other mutations in the HFE gene have been described of late, their clinical significance is uncertain. S65C is the most common of these mutations, and may result in some siderosis when in association with C282Y, particularly when alcohol abuse or other factors favoring iron overload are also present (23). Other mutations include G93R, I105T (24), Q127H (25), V272L (26), Q283P (27), E168X, W169X (28), V68dT, P160dC (29), IV53, and IG-T (30).

Despite our better understanding of HFE nowadays, many patients remain with unexplained iron overload. This is particularly common in Southern Italy, where up to 40% of patients with HH lack these mutations (4,27,31). In Spain the frequency of HFE-associated HH is similar to that of Nordic countries (32). In countries with non-European-origin populations (Asia, Africa, etc.), mutation C282Y is exceptional in frequency. It is for this reason that other genetic factors -in addition to those related to HFE- are thought to exist, which would influence the development of HH in countries where populations are not of Celtic descent. From studies in HFE knock-out mice it was also concluded that other genetic factors leading to iron overload must exist (33).

Once the secondary causes of iron overload (thalassemia, hemolytic anemia, chronic viral hepatitis, alcoholism, oral or parenteral iron administration, transfusion, porphyria cutanea tarda, portosystemic shunts, etc.) had been ruled out, the genetic study of patients with iron overload and absence of the C282Y mutation led to the identification of mutations in the genes coding for a number of proteins involved in the regulation of body iron stores. These new HH forms have been designated HH types II through IV, with type I being reserved for HFE-associated HH.

HEREDITARY TYPE II HEMOCHROMATOSIS OR HEREDITARY JUVENILE HEMOCHROMATOSIS

In 1932 Bezançon et al. (34) described in France the case of a 20-year-old male with liver cirrhosis, infantilism, and multiple endocrine insufficiencies who died from heart failure. More recently Goossens, in 1975 (35), and Lamon et al., in 1979 (36), described similar cases.

The presence of a rare, serious form of hereditary hemochromatosis involving children or adults of both genders younger than 30 years of age is currently admitted. It develops in Caucasian individuals of European descent (35,36), and is clinically characterized by patients having hypogonadotropic hypogonadism, miocardial disease, hepatomegaly, liver cirrhosis, and melanic skin pigmentation (36,37). These patients may also present with hypothyroidism with reduced response to TRH (35), adrenal insufficiency (38,39), and decreased prolactin and growth hormone (40). This form of hemochromatosis has been designated juvenile hemochromatosis (JH) or hereditary type II hemochromatosis. It results from iron overload secondary to excessive intestinal iron absorption (1-4 mg/day) (41). The disorder is inherited in an autosomic recessive fashion (36,37), and bears no relationship to the HFE gene, transferrin receptor 2, or ferroportin. It is currently acknowledged that differing genetic defects hide behind a single phenotype. In most cases JH is linked to chromosome 1q, which has been designated as type IIa hemochromatosis. In a minority of patients the disease is linked to chromosome 19, this being type IIb hemochromatosis.

The first mutations identified included those related to chromosome 19, where the HAMP gene coding for hepcidin (hepatic bactericidal protein) is located. It is a small, 25-aminoacid peptide with antimicrobial properties that is synthesized by liver cells in response to inflammation, iron overload, and interleukin 6, and is decreased during hypoxia or when the need for erythropoiesis increases (42-45). The gene of hepcidin, which has three exons and two introns, codes for a polypeptide with 83 aminoacids that is a precursor to mature hepcidin. The latter results from the separation of 25 aminoacids at the carboxylic end (43). Its genetic expression is regulated by transcription factors C/EBPα and HNF4 (Hepatocyte Nuclear Factor 4) (46), but not by iron regulatory proteins (IRPs). Hepcidin messenger RNA (mRNA) lacks the iron responsive element (IRE) (47). Its function is to block iron passage through bowel cells (43,48), and iron release from macrophage stores (49). The mechanism by which these effects occur is unknown. Two distinct mechanisms have been suggested. Nicolas et al. (50) suggested that hepcidin interacts with the transferrin-transferrin receptor-HFE protein complex in duodenal crypt cells, thus favoring iron entry through these cells' base wall. As a consequence these iron-rich cells maturate into enterocytes while inhibiting the expression of proteins involved in intestinal iron absorption. In contrast, Frazer and Anderson (19) suggested a direct effect on proteins involved in the passage of iron through enterocytes as an alternative mechanism. To this effect hepcidin has been seen to bind, internalize and inactivate ferroportin 1 at duodenal mature enterocytes (51). Intestinal iron absorption is thus blocked (Fig. 3). Whatever the mechanism of action of hepcidin, its absence favors intestinal iron absorption and the release of iron stored in the reticuloendothelial system (RES). This is seen in all situations where this hepatic hormone is low (iron-deficient diet, bleeding, hypoxia, types I, II, III hemochromatosis, etc). On the contrary, increased hepcidin (inflammation, infection, exogenic iron overload, liver adenomatosis, etc.) (52) results in decreased intestinal iron absorption and iron retention in RES cells. During inflammation and infection hepatic hepcidin synthesis increases (52), which translates into decreased intestinal iron absorption (53,54), iron retention within macrophages (55), and anemia (45,53,56).

A number of mutations in the HAMP gene have been found in some patients with JH (57,58). A change G→A in the sequence +14 at the 5'-untranslated end (5'-UTR) has been reported in a Portuguese family, which creates a new AUG sequence that inhibits the translation of normal hepcidin mRNA, and probably results in the formation of a new, abnormal, unstable and degradable peptide (59). Other mutations reported include R56X, which creates a "stop codon", the deletion of guanine 93, 175G→C (R59G), which precludes prohepcidin activation into hepcidin by convertases (60), particularly by furin, and 212G→A (G71D), which alters this peptide's structure and function (61).

In most patients with JH the disorder is linked to chromosome 1q (57), but the gene involved has remained unknown until very recently. In 2004, Papanikolaou et al. (62) published the results of a thorough study of chromosome 1q, where they unveiled a locus of previously unknown function, LOC148738, which was associated with JH. The gene involved was initially designated HFE2, and more recently HJV. In this gene, which was made up of four exons separated by three introns, they found numerous mutations, and one of them, G320V, was present in all patients of Greek, Canadian, and French descent with JH (62). This gene codes for a 426-aminoacid protein that has been called hemojuvelin. Various regions have been identified in its molecule, including a short transmembrane region, a signal peptide, a v-Willebrand-like region, another RGD region through which it relates to extracellular matrix proteins, and a receptor-binding region (62). The mechanism of action of hemojuvelin is unknown, but seems to be closely linked to that of hepcidin. It is known not to be a hepcidin receptor (62), as it is not expressed in organs where hepcidin acts (intestine, spleen) (62). When mutations exist in the HJV gene, urine hepcidin decreases (62). In JH urine hepcidin is deeply reduced despite the fact that body iron is strongly elevated. Hemojuvelin is therefore thought to be a hepcidin-modulating protein, so that the former's decreased levels or inactivity results in the latter's reduced presence. Such decreases would be responsible for the increased intestinal iron absorption and iron overload found in patients with JH (62).

Since first described by Papanikolaou et al. (62), many other authors have confirmed the presence of mutation 959G→T (G320V) in patients with type IIa HH, in addition to their finding a few more. Huang et al. (63) described mutations 962G→A, 963C→A (C321X), 18G→C (Q6H), and 842T→C (I281T). The former two determine early transcription termination, and the latter involves the signal peptide region. Lee et al. added 238T→C (C80R), 302T→C (L101P), 665T→A (I222N) (64) and C321W (65). Lanzara et al. also confirmed mutation G320V, and then identified 17 new mutations. Most of them were located in exons 3 and 4, particularly within the molecular region corresponding to the von Willebrand-like domain (66), and many were determinant of transcription termination. These mutations included a deletion of 13 base-pairs (CGGGGCCCCGCCC), which may be expected to result in a nil phenotype. They found two mutations in another patient - 220delG, which creates a transcription end signal at 113, and 806-807insA, which leads to molecule truncation at position 331 and the formation of a 310-aminoacid molecule. Mutation 1153C→T (R385X) originates a protein devoid of 42 aminoacids at its carboxylic end, precisely those corresponding to the transmembrane portion. Mutation 295G→A (G99R) affects the RGD region, as does mutation G99V, already reported by Papanikolaou et al. (62). Mutations 253T→C (S85P) and 302T→C (L101P) also occur around this region.

Patients must be diagnosed and treated during the early stages of disease. As in HFE-linked HH, treatment primarily relies on an intensive program of periodic bleeding (67). Even in the presence of heart disease, heart failure, and arrhythmia patients may recover with this therapy (36,68,69). Combining bloodletting and deferroxamine may be useful in the presence of severe, life-threatening heart disease (68). Cases have been reported where heart transplantation was life-saving (40,68,70). As expected, liver iron contents clearly diminishes with bleedings, and fibrosis has also been shown to decrease or disappear with this therapy (36,39). There is no evidence that these patients may have a higher risk for liver cancer, but it is only logical to think so. In fact, foci of iron-depleted hepatocytes have also been reported in these patients (70).

HEREDITARY TYPE III, TRANSFERRIN RECEPTOR 2-LINKED HEMOCHROMATOSIS

In a number of patients with iron overload in whom the presence of HFE gene mutations or its being secondary to other conditions was excluded genetic research demonstrated some homozygous mutation in the gene coding for transferrin receptor 2 (TfR2). A number of the families where these mutations were unveiled originated in Southern Italy (71-74), but they have also been detected Japan (75), Portugal (76), and Northern France (77).

Patient characteristics are varied, but may be similar to those of patients with classic, type I hemochromatosis. However, they commonly experience symptoms at younger ages, even before turning 30 (74,76,77), and cardiac manifestations (74), hypogonadotropic hypogonadism (74,76), joint pain, skin hyperpigmentation, and liver cirrhosis are common. Manifestations may resemble those of type II or juvenile hemochromatosis.

TfR2 is a transmembrane glycoprotein exhibiting a large extracellular part. Sixty percent of aminoacids in this molecular region are similar to those in the extracellular domain of transferrin receptor 1 (TfR1). This is an area through which it contacts iron-carrying transferrin (Tf) (78), albeit with a lower affinity as compared to TfR1 (79,80). A potential binding of the HFE protein is currently debated (81,82). It is preferentially expressed by liver cells (82), but has also been identified in duodenal crypt cells (83). The binding of iron-carrying Tf to TfRs represents the primary entry of iron into cells. This bond is modified by HFE. Diferric Tf, TfR, HFE, and β2MG all make up a complex that is internalized by endocytosis. Tf-bound Fe3+ is released in the endosome's acid pH, and is then transported into the cytoplasm by means of a reductase (Fe3+→ Fe2+) and of divalent metal transporter 1 (DMT1) (reviewed in 84). Both Tf and TfR return to the cell's surface for reuse.

In 1999 Kawabata et al. (78) cloned the TfR2 gene and showed that two transcripts resulted from it - TfR2-α and TfR2-β. The former is highly homologous to TfR1; its product is expressed on the surface of some cells, and plays a role in the regulation of iron metabolism. TfR2-β gives rise to a smaller protein that stays within cells. Iron plays no part in the regulation of TfR2 expression, since its mRNA lacks an IRE (82).

Genetic studies performed thus far allowed to identify the following mutations: 84-88insC (E60X) (72), R105X (77), 515T→A (M172K) (72), Y250X (71,85), Q317X (74), 1780-1791del (AVAQ 594-597) (75,73), and 2069A→C (Q690P) (76,86). Some of these mutations (E60X, R105X, Q317X) result in a transcription termination sequence, which blocks protein expression.

From what we know so far about this receptor's function, how its depletion or loss of activity may lead to increased intestinal iron absorption and iron overload is difficult to understand. However, its causal role in this condition has been demonstrated in tfr2-deleted mice, which also develop iron overload (87).

Two hypothetical mechanisms may explain the fact that TfR2 or HFE defects, both of them involved in iron entry into cells, may lead to iron overload through increased intestinal iron absorption. One of them is based on the role duodenal crypt cells play in the regulation of body iron (83). The other one relies on the role of HFE, hemojuvelin, and TfR2 in the regulation of hepcidin synthesis or function (74). One within the body, the passage of iron into cells is facilitated by TfR1s, which remain present in this disease (87).

HEREDITARY TYPE IV HEMOCHROMATOSIS. FERROPORTIN DISEASE

It has been known for years that people from the Solomon Islands commonly have abnormal iron overload (88). During the study of 81 living members of an extensive family of Melanesian descent 31 people were found to show evidence of iron overload. The fact that their disorder was inherited in an autosomal dominant fashion, and that HFE involvement could be ruled out stood out among these subjects' characteristics (88).

In 1999 Pietrangelo et al. (89) studied 53 members of an Italian family, some of which had iron overload in the absence of HFE gene mutations. Other authors have described similar families in various countries (The Netherlands, Canada, Italy).

These patients share common characteristics, which differ from those of patients with HFE-linked hemochromatosis. The disease is transmitted in an autosomal dominant pattern. Considerably high blood ferritin levels is a common feature in these patients, this increase being not paralleled by transferrin saturation. The latter is even normal or slightly high in a number of cases (89,90-92). Hemoglobin may be diminished in young women. This disproportion between serum ferritin rates and transferrin saturation is particularly noticeable in early disease. Liver biopsy confirms a big amount of iron within the liver, both inside hepatocytes and reticuloendothelial system cells, specifically Kupffer cells and macrophages located in portal spaces (90,92,93). During early disease iron preferentially settles in Kupffer cells, but siderosis increases in hepatocytes as the disorder progresses. The predominantly periportal location of hepatocyte siderosis that is common in classic hemochromatosis is not seen in such cases. Iron rather distributes itself evenly throughout the lobule. Anyway, this siderosis is well tolerated, and liver fibrosis is mild or nonexistant (90,91,93,94). Such iron overload may be treated using periodic bleedings, and tolerance may be normal. However, a number of patients do not tolerate bleedings and develop anemia even in the presence of high serum ferritin levels (90,92).

HFE gene mutations and iron overload secondary to other diseases were all excluded in patients with this conditions.

The study of the ferroportin 1 (FP-1) gene, also designated SLC11A3, IREG1, and MTP1, has revealed a number of mutations. Njajou et al. (95) found mutation 430A→C in exon 5 (N144H) in a Dutch family. Arden et al. (88), in members of a family of Melanesian descent, identified a similar mutation - 431A→C (N144T). In the extensive Italian family studied by Pietrangelo et al. (89), Montosi et al. (96) unveiled mutation A77D in exon 3. In turn, Wallace et al. (93), Devalia et al. (90), Cazzola et al. (97), and Reotto et al. (91) found a deletion of three base pairs (485_487delTTG) in a region of exon 5 containing three repeat TTGs in the populations of various countries. This mutation results in the loss of 1 of the 3 valines in positions 160 to 162 (V162del) at FP1. This valine triplet is maintained by most species (98), and is thus supposed to play a relevant role in iron binding or transportation. Other mutations reported include 190T→C (Y64N) (99), 774A→G (D177G), 850G→T (E182H), and 1272G→T (G323V) (100). Jouanolle et al. (92) found a mutation in exon 8 that results in a G490D change between FP1 helices 8 and 9. This change must disrupt the molecule's packing and structure. While Njajou et al. (95) suggested that these mutations would increase the functional capabilities of FP1, most other authors agree that they result in loss of function (88,96).

The gene coding for FP-1, located in chromosome 2q32, responds to inflammation, hypoxia, and iron demands by the bone marrow (101-103). Its mRNA has an IRE in region 5'-UTR (104). This element binds an IRP (Iron-Regulatory Protein). The expression of this mRNA in the presence of iron is contrary to expectations for an mRNA having an IRE near the 5'-UTR end. In all mRNAs having an IRE in this area, the presence of iron determines increased expression (105). Such is the case of ferritin (18). In contrast, FP1 mRNA decreases its expression in the presence of iron, only to increase it under iron deficiency conditions (106). The cause of this abnormal behavior in FP-1 mRNA is unknown. FP-1 is located in cell membranes, piercing them in up to 9 different sites (90). While the above-mentioned mutations occur throughout the FP-1 molecule, most of them develop between the first and fourth transmembrane domains, in extracellular helices 1 and 3. FP1 has been presumed to establish functional relations with apotransferrin, ceruloplasmin or hephaestin via these helices (107). Their function is not accurately understood, but is thought to be essential in allowing iron release from RES cells. These cells play a primary role in the reuse of iron from old RBC destruction. The loss of its function determines iron retention in these cells, and thus potential iron scarcity regarding erythropoiesis. Secondary to this deficiency intestinal iron absorption would increase (108). Why this protein's malfunction impacts iron release from RES cells while not preventing iron passage through duodenal enterocytes remains unclear (109). It may possibly result from the fact that iron flow through RES is much more intensive than iron flow through enterocytes (110). Thus, a failure in FP1 would have a greater impact on RES versus the bowel.

While these patients have been usually treated with periodic bleedings, some authors have questioned their need. Although the extent of iron overload may be very significant, iron excess is generally well tolerated, and relevant medical injuries are usually few. Conversely, some patients do not tolerate bleedings (90,92) and develop anemia despite still elevated levels of serum ferritin. If bleedings are eventually used, caution and frequent hemoglobin and hematocrit monitoring are recommended.

OTHER HEREDITARY HEMOCHROMATOSES

Congenital atransferrinemia

In 1961 Heilmeyer et al. (111) described the case of a young woman with serious hypochromic anemia associated with generalized iron overload. Eight additional patients have been subsequently described in Slovakia (112), Japan (113), Mexico (114,115), France (116), Samoa Islands (117), and the USA (118). All these patients shared similar clinical features. They had severe hypochromic, sideropenic, refractory anemia since childhood in association with severe siderosis in the liver and other organs. Some of them were prone to infection (111), and two died from pneumonia (111,115). Blood iron is usually very low, but serum ferritin rates are highly elevated (2000-8000 µg/L). Blood transferrin (Tf) is very low or undetectable (118). In patients having undergone liver biopsy extensive hepatic siderosis was seen to compromise both hepatocytes and Kupffer cells (118). Varying degrees of liver fibrosis were found in some patients (117). Other organs may also be damaged by siderosis (myocardium, pancreas, thyroid, kidneys); in contrast, iron is absent from the bone marrow.

A similar condition was found in mice with atransferrinemia. A Tf gene mutation has been identified in these mice (119,120). Genetic studies available on this disease are few in human beings. The condition is transmitted with an autosomal recessive pattern. Beutler et al. (118) found two Tf gene mutations in a female patient - one in exon 5 and one in exon 12 (double heterozygote). The former was a 562_571 deletion followed by a 572_580 duplication, which represented a transcription interruption point. The second mutation was a 1429G→C (A477P) substitution, which probably determined the synthesis of an unstable Tf. A mutation 1180G→A (E394K) (121) was identified in other patient, and a homozygous mutation 229G→A in exon 3 (D77N) was identified in the Slovakian patient (122).

Tf functions to carry iron in the plasma and then deliver it to the erythron and other tissues. Tf absence determines the development of severe anemia from iron deficiency (113). Intestinal iron absorption and deposition in tissues increases as a result of anemia (118).

Treatment for this condition includes periodic i.v. infusion of apoTf (113) or standard fresh plasma (118). Bleedings prior to plasma infusion have been used to control iron overload (118), and deferrioxamine has also been administered (112).

Hereditary aceruloplasminemia

In 1987, Miyajima et al. (123) reported on a Japanese 52-year-old patient who has diabetes, retinal degeneration, extrapyramidal symptoms, and total lack of serum ceruloplasmin. Shortly afterwards similar cases were reported in Ireland and Japan (124,125), where blood ceruloplasmin (CP) was undetectable. The ceruloplasmin gene, located in chromosome 3q23-q24 (126), has 20 exons and codes for 1046-aminoacid protein (127). Various mutations associated with CP loss of function or plasma CP absence have been identified in this gene (128,129). Many other cases have been reported with this disorder ever since, most of them in Japan, but also in Caucasian populations (124,130-132). The disease has an autosomal recessive pattern of inheritance.

Genetic studies have uncovered numerous mutations, homozygous on occasion, double heterozygous in other instances. Mutations commonly result in transcription termination points, and thus the synthesized CP is abnormal, inactive, and degradation-bound (128,129,131-133). Harris et al. (128) found in their patient a 5-base pair insert that gave rise to a truncated protein. So was also the result of change G→A in the sequence corresponding to amino acid 991 as found by Yoshida et al. (129). In the case reported by Bosio et al. (131) two mutations were encountered. The first was a 436C→G (Q146E) change, and the other was an adenine insertion at position 2917. The latter induced transcription termination, and hence the production of a truncated protein in amino acid 983. The missing area just includes the sites where copper binds apoceruloplasmin, which then acquires ferroxidase activity. Okamoto et al. (133) also described an adenine insertion in exon 3, the area corresponding to amino acid 184, which translates into early transcription termination. In the patient reported by Loréal et al. (132), these authors found a 2-base pair deletion in one allele's exon 11, which brought about a transcription termination signal (TGA) in codon 632. In the other allele, they found change 694T→A (TAT→TAA), which also results in transcription termination.

When symptomatic, these patients are usually in the fourth of fifth decade of life, and show neurologic changes including dementia, dysarthria, and dystonia (123-125). Retinal degeneration and insulin-dependent diabetes mellitus are common findings (125). All these manifestations result from iron deposition in the central nervous system, primarily in the base nuclei, retina, and pancreas ß-cells. Liver biopsy commonly demonstrates the presence of intense siderosis (hepatic iron > 1.500 mg/g) compromising both hepatocytes and RES cells (134). Despite such iron overload and elevated serum transferrin levels, fibrosis or hepatocellular necrosis is uncommon for most patients. In contrast, hyposideremia, and normocytic, normochromic anemia are common, since iron is not bound by transferrin and cannot reach the bone marrow (123,134). Intestinal iron absorption increases secondary to anemia, as hephaestin -a molecule with a structure similar to ceruloplasmin and sharing the latter's iron oxidizing role- is found in enterocytes (135).

CP is essential for Fe2+ oxidation into Fe3+ (136,137), which allows iron to leave cells and then be transported by Tf to organs in need of it (138). In the absence of CP Tf cannot bind iron, and iron cannot leave RES cells (136,139,140), thus remaining confined in them. In addition, plasma unbound Fe2+ deposits itself in tissues (liver, pancreas, etc.), as occurs in atransferrinemia and classic hemochromatosis when Tf saturation reaches 100%. The expression of TfR1 and DMT1 diminishes in iron-laden hepatocytes, since their mRNAs have an IRE at their 3'-end whose expression is inhibited by iron through IRP (141). Neurologic lesions result from iron preferential deposition in astrocytes. CP cannot go across the blood-brain barrier, but is usually synthesized by said cells. In aceruloplasminemia astrocytes synthesize no CP, and iron is retained within these cells (142).

Aceruloplasminemia may be mistaken for Wilson's disease, but differs from the latter in that serum ferritin levels are very high -resulting from insulin-dependent diabetes- and magnetic resonance imaging shows excessive iron in base nuclei of the brain. It differs from hemochromatosis in that Tf saturation is low, as are serum iron rates. This differentiation is relevant to avoid bleedings. Some patients have received subcutaneous deferroxamin (2 g/day; 5 days/week) (132). This managed to decrease serum ferritin and liver iron deposition, as well as to stabilize diabetes; however, it had to be discontinued on occasion because of aggravated anemia. Iron deposits in the nervous system remained unaltered.

Iron overload associated with H-ferritin mutation (type V HH)

Only one Japanese family with iron overload resulting from a mutation of ferritin's H subunit has been reported so far (143). Subunit L mutations result in the hereditary hyperferritinemia-cataract syndrome, where no iron overload is present (144-146).

In 2001, Kato et al. (143) described a Japanese family with 4 out of 8 members suffering from hyperferritinemia. Moreover, hypersideremia increased transferrin saturation, and iron deposition in tissues were all demonstrated in some of them. Liver biopsy showed siderosis to be distributed around hepatocytes in lobular areas 1 and 2. Other causes of iron overload were ruled out, including HFE and TfR2 mutations, as well as aceruloplasminemia. The study of ferritin subunits L and H mRNA revealed the former to be normal, but the latter had a heterozygous point mutation at position 49, corresponding to the 5'-IRE loop, with an U replacing A (A49U). The study of genomic DNA showed a mutation 49A→T. This change was only found in family members with hyperferritinemia, and the mutation is thus seemingly transmitted in an autosomal dominant fashion. Such mutation in this area confers IRPs a higher affinity for IRE, and hence a suppression of H-ferritin mRNA translation. Conversely, IRP binds L-ferritin mRNA no as firmly as H-ferritin mRNA, which translates into an increased expression of the former subunit (143).

H-ferritin has ferroxidase activity, which is necessary for Fe3+ to be able to incorporate L-ferritin capsules (147). Mutation A49U conditions a decrease in the binding of iron to L-ferritin (143), and this metal's deposition in the cytoplasm of cells. H-ferritin-knockout mice die during embryo stages from excessive body iron accumulation (148). In the hyperferritinemia-cataract syndrome there is no iron overload despite severely increased L-ferritin levels (144). This likely due to the fact that H-ferritin is normal and retains its iron-oxidizing activity.

Iron overload in Subsaharian Africa (Bantu siderosis)

The fact that iron overload is very common among Subsaharian Africans has been known for over 70 years now (149,150). In some rural areas the frequency of this condition exceeds 10% of the total population (151). Iron overload has also been seen to be common in urban areas and among Afro-Americans (152,153). Historically this iron overload was attributed to the consumption of iron-rich food, specifically beer fermented in non-galvanized iron drums (154). However, it has been shown in recent years that besides this diet-related factor -currently non-existant in African cities- a genetic non-HFE-related factor must be involved in the pathogenesis of this iron overload (151,154,155). Whereas heterozygotes for this factor need an iron-rich diet for iron overload to develop, homozygotes may have this condition even in the absence of such foods. Studies performed in the Afro-American population reached this same conclusion (153). To this day no gene suspected of being involved in the pathogenesis of this disease has been identified.

Iron overload initially compromises RES cells, both in the liver and the marrow or spleen (155,156). Thus, during early disease liver biopsy shows that iron deposition preferentially involves Kupffer cells, with the typical iron gradient present in other hemochromatosis forms remaining unrecognized in this one (151). In later stages of disease iron also deposits itself inside hepatocytes, fibrosis is encountered in varying degrees, and cirrhosis or even hepatocellular carcinoma may develop (157). Furthermore, there is excess iron in the heart, lungs, spleen, and other organs (153,157). As in other iron overloads preferentially compromising the RES, transferrin saturation may be normal or slightly high. The characteristics of this iron overload resemble those of FP1-induced disease. A mutation in the FP1 gene has been recently identified in the Subsaharian and Afro-American populations, which may well explain this disease (158).

To conclude, new mutations in the genes coding for various proteins involved in iron metabolism and accounting for some non-HFE hereditary hemochromatosis have been identified in recent years. Besides HFE, these proteins include hepcidin, hemojuvelin, transferrin receptor 2, transferrin, ceruloplasmin, and ferritin H subunit. In upcoming years we shall no doubt witness the finding of new changes in other proteins to account for all hereditary hemochromatosis types. Also in upcoming years will insight be gained into the mechanisms of action of all these proteins.

REFERENCES

1. von Reckinghausen FD. Über Hämochromatose. Tagesblatt Versammlung deutsche Naturforscher Ärzte Heidelberg 1888; 62: 324-5.         [ Links ]

2. Simon M, Le Mignon L, Fauchet R, Yaouanq J, David V, Edan G, et al. A study of 609 HLA haplotypes marking for the hemochromatosis gene.Am J Hum Genet 1987; 41: 89-105.        [ Links ]

3. Feder JN, Gnirke A, Thomas W, Tsuchihashi Z, Ruddy DA, Basava A, et al. A novel MHC class I-like gene is mutated in patients with hereditary haemochromatosis. Nat Genet 1996; 13: 399-408.        [ Links ]

4. Merryweather-Clarke AT, Pointon JP, Jouanolle AM, Rochette J, Robson KJH. Geography of HFE C282Y and H63D mutations. Genet Test 2000; 4: 183-98.         [ Links ]

5. Murphy S, Curran MD, McDougall N, Callender ME, O'Brien CJ, Middleton D. High incidence of the Cys 282 Tyr mutation in the HFE gene in the Irish population - implications for haemochromatosis. Tissue Antigens 1998; 52: 484-8.         [ Links ]

6. Steinberg KK, Cogswell ME, Chang JC, Caudill SP, McQuillan GM, Bowman BA, et al. Prevalence of C282Y and H63D mutations in the hemochromatosis (HFE) gene in the United Status. JAMA 2001; 285: 2216-22.        [ Links ]

7. Papanikolaou G, Politou M, Terpos E, Fourlemadis S, Sakellaropoulos N, Loukopoulos D. Hereditary hemochromatosis: HFE mutation analysis in Greeks reveals genetic heterogeneity. Blood Cells Mol Dis 2000; 26: 163-8.        [ Links ]

8. Milman N, Pedersen P. Evidence that the Cys282Tyr mutation of the HFE gene originated from a population in Southern Scandinavia and spread with the Vikings. Clin Genet 2003; 64: 36-47.        [ Links ]

9. Bennet MJ, Lebron JA, Bjorkman PJ. Crystal structure of the haemochromatosis protein HFE complexed with transferrin receptor. Nature 2000; 403: 46-53.        [ Links ]

10. Ehrlich R, Lemonnier FA. HFE: a novel nonclassic class I molecule that is involved in iron metabolism. Immunity 2000; 13: 585-8.        [ Links ]

11. Riedel HD, Muckenthaler MU, Gehrke SG, Mohr I, Brennan K, Herrmann T, et al. HFE downregulates iron uptake from transferrin and induces iron-regulatory protein activity in stably transfected cells. Blood 1999; 94: 3915-21.        [ Links ]

12. Corsi B, Levi S, Cozzi A, Corti A, Altimare D, Albertini A, et al. Overexpression of the hereditary hemochromatosis protein, HFE, in HeLa cells induces an iron-deficient phenotype. FEBS Lett 1999; 460: 149-52.        [ Links ]

13. Roy CN, Penny DM, Feder JN, Enns CA. Hemochromatosis protein, HFE, specifically regulates transferrin-mediated iron uptake in HeLa cells. J Biol Chem 1999; 274: 9022-8.        [ Links ]

14. Lebron JA, West AJ, Bjorkman PJ. The hemochromatosis protein HFE competes with transferrin for binding to the transferrin receptor. J Mol Biol 1999; 294: 239-45.        [ Links ]

15. Waheed A, Grubb JH, Zhou XY, Tomatsu S, Fleming RE, Costaldi ME, et al. Regulation of transferrin-mediated iron uptake by HFE, the protein defective in hereditary hemochromatosis. Proc Natl Acad Sci. USA 2002; 99: 3117-22.        [ Links ]

16. Ramalingam TS, West AP Jr, Lebron JA, Nangia JS, Hogan TH, Enns CA, et al. Binding to the transferrin receptors is required for endocitosis of HFE and regulation of iron homeostasis. Nature Cell Biol 2001; 2: 953-7.        [ Links ]

17. Lebron JA, Bennet MJ, Vaughan DE, Chirino AJ, Show PM, Mintier GA, et al. Crystal structure of the hemochromatosis protein HFE and characterization of its interaction with transferrin receptor. Cell 1998; 93: 111-23.        [ Links ]

18. Philpott CC. Molecular aspects of iron absorption: insight into the role of HFE in hemochromatosis. Hepatology 2002; 35: 993-1001.        [ Links ]

19. Frazer DM, Anderson GJ. The orchestration of body iron intake: how and where do enterocytes receive their cues? Blood Cells Mol Dis 2003; 30: 288-97.        [ Links ]

20. Bridle KR, Frazer DM, Wilkins SJ, Dixon JL, Purdie DM, Crawford DH, et al. Disrupted hepcidin regulation in HFE-associated haemochromatosis and the liver as a regulator of body iron homeostasis. Lancet 2003; 361: 669-73.         [ Links ]

21. Hanson EH, Imperatore G, Burke W. HFE gene and hereditary hemochromatosis: a HuGE review. Am J Epidemiol 2001; 154: 193-200.        [ Links ]

22. Risch N. Haemochromatosis, HFE, and genetic complexity. Nat Genet 1997; 17: 375-6.        [ Links ]

23. Wallace DF, Walter AP, Pietrangel A, Clare M, Bomford AB, Dixon JL, et al. Frequency of the S65C mutation of HFE and iron overload in 309 subjects heterozygous for C282Y. J Hepatol 2002; 36: 474-9.        [ Links ]

24. Barton EH, Sawada-Hirai R, Rothenberg BE, Acton RT. Two novel missense mutations of the HFE gene (I105T and G93R) and identification of the S65C mutation in Alabama hemochromatosis probands. Blood Cell Mol Dis 1999; 25: 147-55.        [ Links ]

25. de Villiers JN, Hillermann R, Loubser L, Kotze MJ. Spectrum of mutations in the HFE gene implicated in haemochromatosis and porphyria. Hum Mol Gen 1999; 8: 1517-22.        [ Links ]

26. Worwood M, Jackson HA, Feeney GP, Edwards C, Bowen DJ. A single tube heteroduplex PCR for the common HFE genotype. Blood 1999; 94 Supl. 405a.        [ Links ]

27. Le Gac G, Dupradeau F, Mons F, Jacolot S, Scotet V, Esnault G et al. Phenotypic expression of the C282Y/Q283P compound heterozygosity in HFE and molecular modelling of the Q283P mutation effect. Bood Cells Mol Dis 2003; 30: 231-7.        [ Links ]

28. Piperno A, Arosio C, Fossati L, Vigano M, Trombini P, Vergani A, Mancia G. Two novel nonsense mutations of HFE gene in five unrelated Italian patients with hemochromatosis. Gastroenterology 2000; 119: 441-5.        [ Links ]

29. Pointon JJ, Wallace D, Merryweather-Clarke AT, Robson KJ. Uncommon mutations and polymorphisms in the hemochromatosis gene. Gen Testing 2000; 4: 151-61.        [ Links ]

30. Wallace DF, Dooley JS, Walter AP. A novel mutation of HFE explains the classical phenotype of genetic haemochromatosis in a C282Y heterozygote. Gastroenterology 1999; 116: 1409-12.        [ Links ]

31. De Marco F, Liguori R, Giardina MG, D'Armiento M Angelucci E, Lucariello A, et al. High prevalence of non-HFE gene associated haemochromatosis in patients from southern Italy. Clin Chem Lab Med 2004; 42: 17-24.        [ Links ]

32. Pardo A, Quintero E, Barrios Y, Bruguera M, Rodrigo L, Vila C, et al. Expresión genotípica y fenotípica de la hemocromatosis hereditaria en España. Gastroenterol Hepatolol 2004; 27: 437-443.        [ Links ]

33. Fleming R, Holden C, Tomatsu S, Waheed A, Brunt E, Britton R, et al. Mouse strain differences determine severity of iron accumulation in Hfe knockout model of hereditary hemochromatosis. Proc Natl Acad Sci USA 2001; 98: 2707-11.         [ Links ]

34. Bezançon F, De Gennes L, Delarue J, Oumensky D. Cirrhosis pigmentaire avec infantilisme et insuffisance cardiaque et aplasies endocriniennes multiples. Bull Mém Soc Méd Hop Paris 1932; 48: 967-74.        [ Links ]

35. Goossens JP. Idiopathic haemochromatosis: Juvenile and familiar type -Endocrine aspects. Neth J Med 1975; 18: 161-9.        [ Links ]

36. Lamon JM, Marynick SP, Roseblatt R, Donnelly S. Idiopathic hemochromatosis in a young female. A case study and review of the syndrome in young people. Gastroenterology 1979; 76: 178-83.         [ Links ]

37. Kaltwasser JP. Juvenile hemochromatosis. En: Barton JC, Edwards CQ, eds. Hemochromatosis: genetics, pathophysiology, diagnosis, and treatment. Cambridge: Cambridge University Press, 2000. p. 318-28.        [ Links ]

38. Varkonyi J, Kaltwasser JP, Seidl C, Kollai G, Andrikovics H, Tordai A. A case of non-HFE juvenile haemochromatosis presenting with adrenocortical insufficiency. Br J Haematol 2000; 109: 252-3.        [ Links ]

39. Ross CE, Muir WA, Ng ABP, Graham RC Jr, Kellermeyer RW. Hemochromatosis, pathophysiologic and genetic considerations. Am J Pathol 1975; 63: 179-91.        [ Links ]

40. Jensen PD, Baggett J, Jensen FT, Baadrup U, Christensen T, Ellegard J. Heart transplantation in a case of juvenile hereditary hemochromatosis followed up by MRI and endomyocardial biopsies. Eur J Haematol 1993; 51: 199-205.        [ Links ]

41. Cazzola M, Cerani P, Rovati A, Iannone A, Claudiani G, Bergamashi G. Juvenile hemochromatosis is clinically and genetically distinct from the classical HLA-related disorder. Blood 1998; 92: 2979-81.        [ Links ]

42. Krause A, Neitz S, Magert HJ, Schulz A, Forssmann WG, Schulz-Knappe P, et al. LEAP-1, a novel highly disulfide-bounded human peptide, exhibits antimicrobial activity. FEBS Lett 2000; 480: 147-50.        [ Links ]

43. Pigeon C, Ilyin G, Courselaud B, Leroyer P, Turlin B, Brissot P, 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: 7811-9.        [ Links ]

44. Nicolas G, Chauvet C, Viatte L, Danan JL, Bigard X, Devaux I, et al. The gene encoding the iron regulatory peptide hepcidin is regulated by anemia, hypoxia, and inflammation. J Clin Invest 2002; 110: 1037-44.        [ Links ]

45. Nemeth E, Valore EV, Territo M, Schiller G, Lichtenstein A, Ganz T. Hepcidin, a putative mediator of anemia of inflammation, is a type II acute-phase protein. Blood 2003; 101: 2461-3.        [ Links ]

46. Courselaud B, Pigeon C, Inoue Y, Gonzalez FJ, Leroyer P, Gilot D, et al. C/EBPa regulates hepatic transcription of hepcidina, an antimicrobial peptide and regulator of iron metabolism. Cross-talk between C/EBP pathway and iron metabolism. J Biol Chem 2002; 277: 41163-70.        [ Links ]

47. Leong W, Lönnerdal B. Hepcidin, the recently identified peptide that appears to regulate iron absorption. J Nutr 2004; 134: 1-4.        [ Links ]

48. Nicolas G, Bennoun M, Porteu A, Mativet S, Beaumont C, Grandchamp B, et al. Severe iron deficiency anemia in transgenic mice expressing liver hepcidin. Proc Natl Acad Sci USA 2002; 99: 4596-601.        [ Links ]

49. Ganz T. Hepcidin, a key regulator of iron metabolism and mediator of anemia of inflammation. Blood 2003; 102: 783-8.         [ Links ]

50. Nicolas G, Bennoun M, Devaux I, Beaumont C, Grandchamp B, Kahn A, et al. Lack of hepcidin gene expression and severe tissue overload in upstream stimulatory factor 2 (USF2) knockout mice. Proc Natl Acad Sci USA 2001; 98: 8780-5.        [ Links ]

51. Nemeth E, Tuttle MS, Powelson J, Vaughn MB, Donovan A, Ward DM, et al. Hepcidin regulates cellular iron efflux by binding to ferroportin and inducing its internalization. Science 2004; 306: 2090-3.         [ Links ]

52. Weinstein DA, Roy CN, Fleming MD, Loda MF, Wolfsdorf JI, Andrews NC. Inappropriate expression of hepcidin is associated with iron refractory anemia: implications for anaemia of chronic disease. Blood 2002; 100: 3776-81.        [ Links ]

53. Roy CN, Weinstein DA, Andrews NC. 2002 E. Mead Johnson Award for Research in Pediatrics Lecture: the molecular biology of the anemia of chronic disease: a hypothesis. Pediatr Res 2003; 53: 507-12.        [ Links ]

54. Weiss G. Pathogenesis and treatment of anaemia of chronic disease. Blood Rev 2002; 16: 87-96.        [ Links ]

55. Weinstein DA, Roy CN, Fleming MD, Loda MF, Wolfsdorf JI, Andrews NC. Inappropriate expression of hepcidin is associated with iron refractory anemia: implications for the anemia of chronic disease. Blood 2002; 100: 3776-81.        [ Links ]

56. Means RT Jr. The anaemia of infection. Baillieres Best Pract Res Clin Haematol 2002; 13: 151-62.        [ Links ]

57. Roetto A, Papanikolaou G, Politou M, et al. Mutant antimicrobial peptide hepcidin is associated with severe juvenile hemochromatosis. Nat Genet 2003; 33: 21-2.         [ Links ]

58. Roetto A, Daraio F, Porporato P, Caruso R, Cox TM, Cazzola M, et al. Screening hepcidin for mutations in juvenile hemochromatosis: identification of a new mutation (C70R). Blood 2004; 103: 2407-9.        [ Links ]

59. Matthes T, Aguilar-Martinez P, Pizzi-Bosman L, Darbellay R, Rubbia-Brandt L, Giosta E, et al. Severe hemochromatosis in a Portuguese family associated with a new mutation in the 5'-URT of the HAMP gene. Blood 2004; 104: 2181-3.        [ Links ]

60. Van de Loo JWHP, Creemers JWM, Bright NA, Young BD, Roebroek AJM, Van de Ven WJM. Biosynthesis, distinct post-translational modifications, and functional characterization of lymphoma proprotein convertase. J Biol Chem 1997; 272: 27116-23.        [ Links ]

61. Jacolot S, La Gac G, Scotet V, Quere I, Mura C, Ferec C. HAMP as a modifier gene that increases the phenotypic expression of the HFE pC282Y homozygous genotype. Blood 2004; 103: 2835-40.        [ Links ]

62. Papanikolaou G, Samuels ME, Ludwig EH, MacDonald MLE, Franchini PL, Dubé MP, et al. Mutations in HFE2 cause iron overload in chromosome 1q-linked juvenile hemochromatosis. Nat Genet 2004; 36: 77-82.        [ Links ]

63. Huang FW, Rubio-Aliaga I, Kushner JP, Andrews NC, Fleming MD. Identification of a novel mutation (C321X) in HJV. Blood 2004; 104: 2176-7.        [ Links ]

64. Lee PL, Beutler E, Rao SV, Barton JC. Genetic abnormalities and juvenile hemochromatosis: mutations of the HJV gene encoding hemojuvelin. Blood 2004; 103: 4669-71.        [ Links ]

65. Lee PL, Barton JC, Brandhagen D, Beutler E. Hemojuvelin (HJV) mutations in persons of European, African-American and Asian ancestry with adults onset haemochromatosis. Br J Haematol 2004; 127: 224-9        [ Links ]

66. Lanzara C, Roetto A, Daraio F, Rivard S, Ficarella R, Simard H, et al. Spectrum of hemojuvelin gene mutations in 1q-linked juvenile hemochromatosis. Blood 2004; 103: 4317-21.        [ Links ]

67. Barton JC, McDonnell SE, Adams PC, et al. Management of hemochromatosis. Ann Intern Med 1998; 129: 932-9.        [ Links ]

68. Nelly AL, Rhodes DA, Roland JM, Schofield P, Cox TM. Hereditary juvenille haemochromatosis: A genetically heterogeneous life-threatening iron-storage disease. Q J Med 1998; 91: 607-18.        [ Links ]

69. Cazzola M, Ascari E, Barosi G, Claudiani G, Dacha M, Kaltwasser JP, et al. Juvenile idiopathic hemochromatosis: a life-threatening disorder presenting as hypogonadotropic hypogonadism. Human Genet 1983; 65: 149-54.        [ Links ]

70. Case Records of the Massachussets General Hospital. Weekly clinicopathological exercises. Case 31-1994. A 25-year-old man with the recent onset of diabetes mellitus and congestive heart failure. N Engl J Med 1994; 331: 460-6.        [ Links ]

71. Camaschella C, Roetto A, Cali A, De Gobbi M, Garozzo G, Carella M, et al. The gene TFR2 is mutated in a new type of haemochromatosis mapping to 7q22. Nat Genet 2000; 25: 14-5.        [ Links ]

72. Roetto A, Totaro A, Piperno A, Piga A, Longo F, Garozzo G, et al. New mutations inactivating transferring receptor 2 in hemochromatosis type 3. Blood 2001; 97: 2555-60.        [ Links ]

73. Girelli D, Bozzini C, Roetto A, Alberti F, Daraio F, Colombari R, et al. Clinical and pathologic findings in hemochromatosis type 3 due to a novel mutation in transferrin receptor 2 gene. Gastroenterology 2002; 122: 1295-302.        [ Links ]

74. Pietrangelo A, Caleffi A, Henrion J, Ferrara F, Corradini E, Kulaksiz H, et al. Juvenile hemochromatosis associated with pathogenic mutations of adult hemochromatosis genes. Gastroenterology 2005; 128: 470-9.        [ Links ]

75. Hattori A, Wakusawa S, Hayashi H, Harashima A, Sanae F, Kawanaka M, et al. AVAQ 594-597 deletion of the TfR2 gene in a Japanese family with hemochromatosis. Hepatol Res 2003; 26: 154-6.        [ Links ]

76. Mattman A, Huntsman D, Lockitch G, Langlois S, Buskard N, Ralston D, et al. Transferrin receptor 2 (TfR2) and HFE mutational análisis in non-C282Y iron overload: identification of a novel TfR2 mutation. Blood 2002; 100: 1075-7.        [ Links ]

77. Le Gac G, Mons F, Jacolot S, Scotet V, Férec C, Frébourg T. Early onset hereditary hemochromatosis resulting from a novel TFR2 gene nonsense mutation (R105X) in two siblings of north French descent. Br J Haematol 2004; 125: 674-8.        [ Links ]

78. Kawabata H, Yang R, Hirama T, Vuong PT, Kawano S, Gombart AF, et al. Molecular cloning of transferrin 2. A new member of the transferrin receptor-like family. J Biol Chem 1999; 274: 20826-32.        [ Links ]

79. Kawabata H, Germain RS, Vuong PT, Nakamaki T, Said JW, Koeffler HP. Transferrin receptor 2-a supports cell growth both in iron-chelated cultured cells and in vivo. J Biol Chem 2000; 275: 16618-25.        [ Links ]

80. Kawabata H, Germain RS, Ikezoe T, Tong X, Green EM, Gombart AF, et al. Regulation of expression of murine transferrin receptor 2. Blood 2001; 98: 1949-54.        [ Links ]

81. West AP, Bennet MJ, Sellers VM, Andrews NC, Enns CA, Bjorkman PJ, et al. Comparison of the interactions of transferrin receptor and transferrin receptor 2 with transferrin and the hereditary hemocromatosis protein HFE. J Biol Chem 2000; 275: 38135-8.        [ Links ]

82. Kawabata H, Nakamaki T, Ikonomi P, Smith RD, Germain RS, Koeffler HP. Expression of transferring 2 in normal and neoplastic hematopoietic cells. Blood 2001; 98: 2714-9.        [ Links ]

83. Griffiths WJH, Cox TM. Co-localization of the mammalian hemochromatosis gene product (HFE) and a newly identified transferrin receptor (TfR2) in intestinal tissue and cells. J Histochem Cytochem 2003; 51: 613-23.        [ Links ]

84. Richardson DR, Ponka P. The molecular mechanisms of the metabolism and transport of iron in normal and neoplastic cells. Biochim Biophys Acta 1997; 1331: 1-40.        [ Links ]

85. Piperno A, Roetto A, Mariani R, Pelucchi S, Corengia C, Daraio F, et al. Heterozygosity for transferrin receptor-2 Y250X mutation induces early iron overload. Haematologica 2004; 89: 359-60.        [ Links ]

86. Biasiotto G, Belloli S, Ruggeri G, Zanella I, Gerardi G, Corrado M, et al. Identification of new mutations of the HFE, hepcidin, and transferrin receptor 2 genes by denaturing HPLC analysis of individuals with biochemical indications of iron overload. Clin Chem 2003; 49: 1981-8.        [ Links ]

87. Fleming RE, Ahmann JR, Migas MC, Waheed A, Koeffler HP, Kawabata H, et al. Targeted mutagenesis of the murine transferrin receptor 2 gene produces hemocromatosis. Proc Natl Acad Sci USA 2002; 99: 10653-8.        [ Links ]

88. Eason RJ, Adams PC, Aston CE, Searle J. Familiar iron overload with possible autosomal dominant inheritance. Aust NZ J Med 1990; 20: 226-30.        [ Links ]

89. Pietrangelo A, Montosi G, Totaro A, Garuti C, Conte D, et al. Hereditary hemochromatosis in adults without pathogenic mutations in the hemochromatosis gene. N Engl J Med 1999; 341: 725-32.        [ Links ]

90. Devalia V, Carter K, Walker AP, Perkins SJ, Worwood M, May A, et al. Autosomal dominat reticuloendothelial iron overload associated with a 3-base pair deletion in the ferroportin 1 gene (SLC11A3). Blood 2002; 100: 695-7.        [ Links ]

91. Roetto A, Merryweather-Clarke AT, Daraio F, Livesey K, Pointon JJ, Barbabietola G, et al. A valine deletion of ferroportin 1: a common mutation in hemochromatosis type 4? Blood 2002; 100: 733-4.        [ Links ]

92. Jouanolle AM, Douabin-Gicquel V, Halimi C, Loréal O, Fergelot P, Delacour T, et al. Novel mutation in ferroportin 1 gene is associated with autosomal dominant iron overload. J Hepatol 2003; 39: 286-9.        [ Links ]

93. Wallace DF, Pedersen P, Dixon JL, Stephenson P, Searle JW, Powell LW, et al. Novel mutation in ferroportin 1 is associated with autosomal dominant hemochromatosis. Blood 2002; 100: 692-4.        [ Links ]

94. Arden KE, Wallace DF, Dixon JL, Summerville L, Searle JW, Anderson GJ, et al. A novel mutation in ferroportin 1 is associated with haemochromatosis in a Solomon Islands patient. Gut 2003; 52: 1215-7.        [ Links ]

95. Njajou OT, Vaessen N, Joosse M, Gerghnis B, van Dongen JW, Breuning NH, et al. A mutation in SLC11A3 is associated with autosomal dominant hemochromatosis. Nat Genet 2001; 28: 213-4.        [ Links ]

96. Montosi G, Donovan A, Totaro A, Garuti C, Pignatti E, Cassanelli S, et al. Autosomal-dominant hemochromatosis is associated with a mutation in the ferroportin (SLC11A3) gene. J Clin Invest 2001; 108: 619-23.        [ Links ]

97. Cazzola M, Cremonesi L, Papaioannou M, Soriani N, Kioumi A, Charalambidou A, et al. Genetic hyperferritinaemia and reticuloendothelial iron overload associated with a three base pair deletion in the coding region of the ferroportin gene (SLC11A3). Br J Haematol 2002; 119: 539-46.        [ Links ]

98. Donovan A, Brownlie A, Zhou Y, Shepard J, Pratt SJ, Moynihan J, et al. Positional cloning of zebrafish ferroportin 1 identifies a conserved vertebrate iron exporter. Nature 2000; 403: 776-81.        [ Links ]

99. Rivard SR, Lanzara C, Grimard D, Carella M, Simard H, Ficarella R, et al. Autosomal dominant reticuloendothelial iron overload (HFE type 4) due to a new missense mutation in the FERROPORTIN 1 gene (SLC11A3) in a large French-Canadian family. Haematologica 2003; 88: 824-5.        [ Links ]

100. Hetet G, Devaux I, Soufir N, Grandchamp B, Beaumont C. Molecular analysis of patients with hyperferritinemia and normal serum iron values reveal both L ferritin IRE and 3 new ferroportin (slc11A3) mutations. Blood 2003; 102: 1904-10.        [ Links ]

101. Abboud S, Haile DJ. A novel mammalian iron-regulated protein involved in intracellular iron metabolism. J Biol Chem 2000; 275: 19906-12.        [ Links ]

102. Yang F, Wang X, Haile DJ, Piantadosi CA, Ghio AJ. Iron increases expression of iron-export protein MTP1 in lung cells. Am J Physiol Lung Cell Mol Physiol 2002; 283: 932-9.        [ Links ]

103. Martini LA, Tchack L, Wood RJ. Iron treatment downregulates DMT1 and IREG1 mRNA expression in Caco-2 cells. J Nutr 2002; 132: 693-6.        [ Links ]

104. Abboud S, Haile DJ. A novel mammalian iron-regulated protein involved in intracellular iron metabolism. J Biol Chem 2000; 275: 19906-12.        [ Links ]

105. La Vaute T, Smith S, Cooperman S, Iwai K, Land W, Meyron-Holtz E, et al. Targeted deletion of the gene encoding iron regulatory protein-2 causes misregulation of iron metabolism and neurodegenerative disease in mice. Nat Genet 2001; 27: 209-14.        [ Links ]

106. Enns CA. Pumping iron: the strange partnership of the hemochromatosis protein, a class I MHC homolog, with the transferrin receptor. Traffic 2001; 2: 167-74.        [ Links ]

107. Vulpe CD, Kuo YM, Murphy TL, Cowley L, Askwith C, Libina N, et al. Hephaestin, a ceruloplasmin homologue implicated in intestinal iron transport, is defective in the sla mouse. Nat Genet 1999; 21: 195-9.        [ Links ]

108. Pietrangelo A. Non-HFE hemochromatosis. Hepatology 2004; 39: 21-9.        [ Links ]

109. Cazzola M. Genetic disorders of iron overload and the novel "ferroportin disease". Haematologica 2003; 88: 721-4.         [ Links ]

110. Brittenham GM. The red cell cycle. En: Brock JH, Halliday JW, Pippard MJ, Powell LW, eds. Iron metabolism in health and disease. London, United Kingdom; WB Saunders, 1994. p. 31-62.        [ Links ]

111. Heilmeyer L, Keller W, Vivel O, Betker K, Wöhler F, Keiderling W. Die kongenitale Atransferrinämie. Schweiz Med Wochenschr 1961; 91: 1203.        [ Links ]

112. Hromec A, Payer J Jr, Killinger Z, Rybar I, Rovensky J. Kongenitale Atransferrinämia. Dtsch Med Wochenschr 1994; 119: 663-6.        [ Links ]

113. Goya N, Miyazaki S, Kodate S, Ushio B. A family of congenital atransferrinemia. Blood 1972; 40: 239-45.        [ Links ]

114. Loperana L, Dorantes S, Medrano E. Atransferrinemia hereditaria. Bol Med Hosp. Infant Mex 1974; 31: 519.        [ Links ]

115. Dorantes-Mesa S, Márquez JL, Valencia-Mayoral P. Sobrecarga de hierro en atranferrinemia hereditaria. Bol Med Hosp. Infant Mex 1986; 43: 99-101.        [ Links ]

116. Walbaum R. Déficit congénital en transferrine. Lille Med 1971; 16: 1122-4.        [ Links ]

117. Hamill RL, Woods JC, Cook BA. Congenital atransferrinemia: a case report and review of the literature. Am J Clin Pathol 1991; 96: 215-8.        [ Links ]

118. Beutler E, Gelbart T, Lee P, Trevino R, Fernandez MA, Fairbanks VF. Molecular characterization of a case of atransferrinemia. Blood 2000; 96: 4071-4.        [ Links ]

119. Trenor CC, Campagna DR, Sellers VM, Andrews NC, Fleming MD. The molecular defect in hypotransferinemic mice. Blood 2000; 96: 1113-8.         [ Links ]

120. Huggenvik JI, Craven CM, Idzerda RL, Bernstein S, Kaplan J, McKnight GS. A splicing defect in the mouse transferrin gene leads to congenital atransferrinemia. Blood 1989; 74: 482-6.        [ Links ]

121. Asasa-Senju M, Maeda T, Sakata T, Hayashi A, Suzuki T. Molecular analysis of the transferrin gene in a patient with hereditary hypotransferrinemia. J Hum Genet 2002; 47: 355-9.        [ Links ]

122. Knisely AS, Gelbart T, Beutler E. Molecular characterization of a third case of human atransferrinemia. Blood 2004; 104: 2607.        [ Links ]

123. Miyajima H, Nishimura Y, Mizoguchi K, Sakanoto M, Shimizu T, Honda N, et al. Familiar apoceruloplasmin deficiency associated with blepharospasm and retinal degeneration. Neurology 1987; 37: 761-7.         [ Links ]

124. Logan JL, Harveyson KB, Wisdom GB, Hughes AE, Archibold GP. Hereditary caeruloplasmin deficiency, dementia and diabetes mellitus. Q J Med 1994; 87: 663-70.         [ Links ]

125. Morita H, Ikeda S, Yamamoto K, Morita S, Yoshida K, Nomoto M, et al. Hereditary ceruloplasmin deficiency with hemosiderosis: a clinicopathological study of a Japanese family. Ann Neurol 1995; 37: 646-56.        [ Links ]

126. Royle NJ, Irwin DM, Koschinsky ML, MacGillivray RT, Hamerton JL. Human genes encoding prothrombin and ceruloplasmin map to 11p11-q12 and 3q21-24, respectively. Somat Cell Mol Genet 1998; 13: 285-92.        [ Links ]

127. Koschinsky ML, Funk WD, Van Oost BA, MacGillivray RT. Complete cDNA sequence of human preceruloplasmin. Proc Natl Acad Sci USA 1986; 83: 5086-90.        [ Links ]

128. Harris ZL, Takahashi Y, Miyajima H, Serizawa M, MacGillivray RT. Aceruloplasminemia: molecular characterization of this disorder of iron metabolism. Proc Natl Acad Sci USA. 1995; 92: 2539-43.        [ Links ]

129. Yoshida K, Furihata K, Takeda S, Nakamura A, Yamamoto K, Morita H, et al. A mutation in the ceruloplasmin gene is associated with systemic hemosiderosis in humans. Nat Genet 1995; 9: 267-72.        [ Links ]

130. Hellman NE, Schaefer M, Gehrke S, Stegen P, Hoffman WJ, Gitlin JD, et al. Hepatic iron overload in aceruloplasminemia. Gut 2000; 47: 858-60.         [ Links ]

131. Bosio S, de Gobbi M, Roetto A, Zecchina G, Leonardo E, Rizzetto M, et al. Anemia and iron overload due to compound heterozygosity for novel ceruloplasmin mutations. Blood 2002; 100: 2246-8.        [ Links ]

132. Loréal O, Turlin B, Pigeon C, Moisan A, Ropert M, Morice P, et al. Aceruloplasminemia: new clinical, pathophysiological and therapeutic insights. J Hepatol 2002; 36: 851-6.         [ Links ]

133. Okamoto N, Wada S, Oga T, Kawabata Y, Baba Y, Habu D, et al. Hereditary ceruloplasmin deficiency with hemosiderosis. Human Genet 1996; 97: 755-8.        [ Links ]

134. Morita H, Ikeda S, Yamamoto K, Morita S, Yoshida K, Nomoto S, et al. Hereditary ceruloplasmin deficiency with hemosiderosis: a clinicopathological study of a Japanese family. Ann Neurol 1995; 37: 646-56.        [ Links ]

135. Vulpe CD, Kuo YM, Murphy TL, Cowley L, Askwith C, Libina N, et al. Hephaestin, a ceruloplasmin homologue implicated in intestinal iron transport, is defective in the sla mouse. Nat Genet 1999; 21: 195-9.        [ Links ]

136. Kaplan J, O'Halloran TV. Iron metabolism in eukaryotes: Mars and Venus at it again. Science 1996; 271: 1510-2.        [ Links ]

137. Osaki S, Johnson DEF. The possible significance of the ferrous oxidase activity of ceruloplasmin in normal human serum. J Biol Chem 1996; 241: 2746-51.        [ Links ]

138. Yang F, Friedrichs WE, Cupplest RL, Bonifacio MJ, Sanford JA, Horton WA, et al. Human ceruloplasmin. Tissue specific expression of transcripts produced by alternative splicing. J Biol Chem 1990; 285: 10780-5.        [ Links ]

139. Ragan HA, Nacht S, Lee GR, Bishop CR, Cartwright GE. Effect of ceruloplasmin on plasma iron in copper-deficient swine. Am J Physiol 1969; 217: 1320-3.        [ Links ]

140. Roeser HP, Lee GR, Nacht S, Cartwright GE. The role of ceruloplasmin in iron metabolism. J Clin Invest 1970; 49: 2408-17.        [ Links ]

141. Yamamoto K, Yoshida K, Miyagoe Y, Ishikawa A, Hanaoka K, Nomoto S, et al. Quantitative evaluation of expression of iron-metabolism genes in ceruloplasmin-deficient mice. Biochim Biophys Acta 2002; 1588: 195-202.        [ Links ]

142. Klomp LWJ, Farhangrazi ZS, Dugan LL, Gitlin JD. Ceruloplamin gene expression in the murine central nervous system. J Clin Invest 1996; 98: 207-15.        [ Links ]

143. Kato J, Fujikawa K, Kanda M, Fukuda N, Sasaki K, Takayama T, et al. A mutation, in the iron-responsive element of H ferritin mRNA, causing autosomal dominant iron overload. Am J Hum Genet 2001; 69: 191-7.        [ Links ]

144. Beaumont C, Leneuve P, Devaux I, Scoazec JY, Berthier M, Loiseau MN, et al. Mutation of the iron responsive element of the L ferritin mRNA in a family with dominant hyperferritinemia and cataract. Nat Genet 1995; 11: 444-6.         [ Links ]

145. Levi S, Girelli D, Perrone F, Pasti M, Beaumont C, Corrocher R, et al. Analysis of ferritins in lymphoblastoid cell lines and in the lens of subjects with hereditary hyperferritinemia-cataract syndrome. Blood; 1998; 91: 4180-7.         [ Links ]

146. Ladero JM, Balas A, García-Sánchez F, Vicario JL, Díaz-Rubio M. Hereditary hyperferritinemia-cataract syndrome. Study of a new family in Spain. Rev Esp Enferm Dig 2004; 96: 507-11.        [ Links ]

147. Harrison PM, Arosio P. The ferritins: molecular properties, iron storage function and cellular regulation. Biochim Biophys Acta 1996; 1275: 161-203.        [ Links ]

148. Ferreira C, Bucchini D, Martin ME, Levi S, Arosio P, Grandchamp B, et al. Early embryonic lethality of H ferritin gene deletion in mice. J Biol Chem 2000; 275: 3021-3.         [ Links ]

149. Gordeuk VR. Hereditary and nutritional iron overload. Balliere's Clin Haemat 1992; 5: 169-86.         [ Links ]

150. Strachan AS. Haemosiderosis and haemochromatosis in South African natives with a comment on the etiology of haemochromatosis. MD Thesis, University of Glasgow. Scotland, 1929        [ Links ]

151. Gordeuk VR, Mukiibi J, Hasstedt SJ, Samowitz W, Edwards CQ, West G, et al. Iron overload in Africa. Interaction between a gene and dietary iron content. N Engl J Med 1992; 326: 95-110.        [ Links ]

152. Gordeuk VR, McLaren CE, Looker A, Hasselblad V, Brittenham GM. Distribution of transferrin saturation in the African-American population. Blood 1998; 91: 2175-9.         [ Links ]

153. Gangaidzo IT, Moyo VM, Saungweme T, Áhumalo H, Charakupa RM, Gomo ZAR, et al. Iron overload in urban Africans in the 1990s. Gut 1999; 45: 278-83.        [ Links ]

154. Bothwell TH, Seftel H, Jacobs P. Iron overload in Bantu subjects. Studies on the availability of iron in Bantu beer. Am J Clin Nutr 1964; 14: 47-51.        [ Links ]

155. Moyo VM, Mandishona E, Hasstedt SJ, Gangaidzo IT, Gomo ZA, Khumalo H, et al. Evidence of genetic transmission in African iron overload. Blood 1998; 91: 1076-82.        [ Links ]

156. Brink B, Disler P, Lynch S, Jacobs P, Charlton R, Bothwell T, et al. Patterns of iron storage in dietary iron in idiopathic hemochromatosis. J Lab Clin Med 1976; 88: 727-31.         [ Links ]

157. Bothwell TH, Abrahams C, Bradlow BA. Idiopathic and Bantu haemochromatosis: comparative histological study. Arch Pathol 1965; 79: 163-8.        [ Links ]

158. Gordeuk VR, Caleffi A, Corradine E, Ferrera F, Jones RA, Castro O et al. Iron overload in Africans and African-Americans and a common mutation in the sc140a1 (ferroportin 1) gene. Blood Cell Mol Dis 2003; 31: 299-304.        [ Links ]

Creative Commons License All the contents of this journal, except where otherwise noted, is licensed under a Creative Commons Attribution License