You are on page 1of 12

review

Modern diagnosis and management of the porphyrias

Shigeru Sassa
Laboratory of Biochemical Hematology, The Rockefeller University, New York, USA

are not associated with neurological symptoms. In contrast,


Summary
acute hepatic porphyrias are characterised by neurological
Recent advances in the molecular understanding of the symptoms. Some of them may have additional photosensi-
porphyrias now offer specific diagnosis and precise definition tivity (Fig 1).
of the types of genetic mutations involved in the disease.
Molecular diagnostic testing is powerful and very useful in
The haem biosynthetic pathway
kindred evaluation and genetic counselling when a disease-
responsible mutation has been identified in the family. It is The enzymatic steps and intermediates in the haem biosyn-
also the only way to properly screen asymptomatic gene thetic pathway are illustrated in Fig 2. In eukaryote cells, the
carriers, facilitating correct treatment and appropriate genetic first enzymatic step and the last three steps occur in
counselling of family members at risk. However, it should be mitochondria; the other four steps take place in the cytosol.
noted that DNA-based testing is for the diagnosis of the gene The two major cell types that are active in haem synthesis are
carrier status, but not for the diagnosis of clinical syndrome hepatocytes and bone marrow erythroblasts, and inherited
or severity of the disease, e.g. an acute attack. For the enzymatic defects in the porphyrias are chiefly manifested in
diagnosis of clinically expressed porphyrias, a logical stepwise these cells.
approach including the analysis of porphyrins and their The first intermediate of the haem biosynthetic pathway is
precursors should not be underestimated, as it is still very d-aminolaevulinic acid (ALA), a 5-carbon aminoketone, which
useful, and is often the best from the cost-effective point of is formed in mitochondria by the condensation of glycine and
view. succinyl CoA by d-aminolaevulinate synthase (ALAS). Two
molecules of ALA are then condensed in the cytosol to form a
Keywords: porphyria, porphyrin, haem, d-aminolaevulinate monopyrrole, porphobilinogen (PBG), by ALA dehydratase
synthase, molecular diagnosis. (ALAD). Four molecules of PBG are combined by PBG
deaminase (PBGD), to form the first cyclic tetrapyrrole,
The porphyrias are uncommon, complex, and fascinating uroporphyrinogen I, which is then converted to uroporphyri-
metabolic conditions, caused by deficiencies in the activities nogen III by uroporphyrinogen synthase (UROS). Uro-
of the enzymes of the haem biosynthetic pathway. While porphyrinogen III is decarboxylated by uroporphyrinogen
most of them are inherited, some may also occur as acquired decarboxylase (UROD) to form coproporphyrinogen III.
diseases. In addition, not all gene carriers of inherited Coproporphyrinogen III enters into the mitochondria, where
porphyrias develop clinical disease and there is a significant it is oxidatively decarboxylated by coproporphyrinogen oxid-
interplay between the primary gene defect and the secondary ase (CPO) to form protoporphyrinogen IX. Protoporphyrino-
acquired or environmental factors. The enzyme deficiencies gen IX is then oxidised to protoporphyrin IX by
can be either partial or nearly complete depending on the protoporphyrinogen oxidase (PPO). Finally, ferrous iron is
types of genetic mutations. Depending on deficient enzymatic inserted into protoporphyrin IX by ferrochelatase to form
steps, various porphyrins and their precursors are accumu- haem. Protoporphyrin IX is the immediate precursor of the
lated in tissues and are excreted in urine and/or stool. various haems and also of the chlorophylls. Information on
Porphyrias can be classified either as (i) erythropoietic enzyme proteins, and genes for haem pathway enzymes is
porphyria, (ii) acute hepatic porphyria, or (iii) chronic hepatic summarised in Table I.
porphyria. Both erythropoietic porphyria and chronic hepatic There is significant tissue-specific regulation for enzymes
porphyria accompany cutaneous photosensitivity, but they in the haem biosynthetic pathway (Sassa, 2006a,b). For
example, there are two separate genes for ALAS, i.e. the
housekeeping and the erythroid-specific ALAS genes, which
are termed ALAS1 (or ALAS-N) and ALAS2 (or ALAS-E)
Correspondence: Prof. S. Sassa, MD, PhD, Laboratory of Biochemical
respectively. In addition, there are the housekeeping and the
Hematology, The Rockefeller University, New York, NY 10021, USA.
erythroid-specific mRNAs for ALAD, the gene for PBGD
E-mail: sassa@mail.rockefeller.edu and ssassa@jcom.home.ne.jp

ª 2006 The Author


Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292 doi:10.1111/j.1365-2141.2006.06289.x
Review

Enzyme Disease Type Symptoms Products


Glycine + Suc.CoA
ALAS2 XLSA Erythroid Microcytic anaemia Sideroblasts
δ-Aminolaevulinic acid
ALAD ADP Hepatic Neurovisceral Urinary ALA

Porphobilinogen
HMBS AIP Hepatic Neurovisceral Urinary ALA, PBG

Hydroxymethylbilane
Photosensitivity Urinary & RBC
UROS CEP Erythropoietic
(Non- Haemolytic anaemia U’ gen I, C ’gen I
enzymatic) (UROS)

U’gen I U’gen III PCT Hepatic/ Photosensitivity 7-C porphyrin; faecal


UROD isocoproporphyrin
HEP Erythropoietic Haemolytic anaemia

C’gen I C’gen III


Neurovisceral & Urinary ALA, PBG,
CPOX HCP Hepatic
Photosensitivity coproporphyrin

P’gen IX Neurovisceral Urinary ALA, PBG;


PPOX VP Hepatic
Photosensitivity faecal protoporphyrin
Proto IX
Fe2+ RBC protoporphyrin
FECH EPP Erythropoietic Photosensitivity
faecal protoporphyrin
Haem

Fig 1. Classification of porphyrias Enzymatic defects, associated diseases, major symptoms and principal accumulation products are shown. ALAS2
defect is responsible for X-linked sideroblastic anaemia (XLSA) but is not associated with any porphyria, since the enzymatic defect blocks production
of ALA, the obligatory precursor for porphyrin formation. ALA dehydratase porphyria (ADP) and acute intermittent porphyria (AIP) are accom-
panied by acute hepatic porphyria but not by photocutaneous porphyria, because their enzymatic defects do not result in an increase in porphyrin
synthesis. Enzymatic defects beyond uroporphyrinogen synthase (UROS) are all associated with photocutaneous porphyrias, because they produce
excessive amounts of various porphyrins. Hereditary coproporphyria (HCP) and variegate porphyria (VP) are additionally associated with acute
hepatic porphyria. Suc.CoA, succinyl coenzyme A; P’gen, rotoporphyrinogen; Proto, protoporphyrin; U’gen, uroporphyrinogen; C’gen, cop-
roporphyrinogen;. Adapted from Sassa S & Shibahara S. Disorders of Heme Production and Catabolism. In Handin RI, Lux SE, and Stossel TP, eds,
Blood: Principles and Practice of Hematology, 2nd ed, Philadelphia, Lippincott Williams & Wilkins, 2003, with permission).

(HMBS), and UROS, and the housekeeping and the • Porphyrins and their precursors are excreted in urine or
erythroid-specific enzymes for ALAS as well as for PBGD stool depending on their solubility. The water solubility of
(Table I). Haem-mediated regulation of ALAS is also tissue- porphyrins is directly attributable to the number of
specific; namely, ALAS1 expression in the liver is carboxyl groups in each molecule (Falk, 1964). Accord-
repressed by haem, while ALAS2 in the erythroid bone ingly, the water-soluble uroporphyrin is excreted into
marrow is not. urine, while the water-insoluble protoporphyrin is
excreted into bile and stool. Coproporphyrin is excreted
into both urine and stool because of its intermediate
General considerations
solubility. Porphyrin precursors are essentially all excreted
Pathogenesis into urine.
• During an acute attack, ALAS1, the hepatic isoform of the
• All of the haem pathway intermediates are potentially toxic.
first enzyme in the haem biosynthetic pathway is induced.
Their overproduction causes the characteristic neurovisceral
ALAS1 formation in normal hepatocytes is repressed by
and/or photosensitizing symptoms.
feedback inhibition by the final product, haem. Metabolic
• Porphyrins produce free radicals when exposed to ultravi-
inhibition along the pathway in the liver leads to reduced
olet light ( 400 nm). As a result, skin damage ensues in
production of haem, resulting in derepression of ALAS1.
the light-exposed areas, resulting in cutaneous porphyrias.
This leads to increased production of haem precursors in an
• In contrast to porphyrins, their precursors, e.g. ALA and
effort to overcome the metabolic block, and contributes to
PBG, are associated with neurological symptoms of acute
the accumulation of intermediates prior to the deficient
hepatic porphyrias.

ª 2006 The Author


282 Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292
Review

Fig 2. Haem biosynthetic pathway Enzymes and intermediates of the haem biosynthetic pathway are shown. Step : ALAS. Step : ALAD. Step :
PBGD. Step : UROS. Step : UROD. Step : CPO. Step : PPO. Step : Ferrochelatase. The carbon atom that is derived from the a carbon of
glycine is shown as a bold red dot. The structure that is denoted by the brackets after step is the presumed intermediate whose pyrrole ring D
becomes rearranged to yield uroporphyrinogen III. At step , 1 mole of oxygen is consumed per 1 mole of water produced. CoA, coenzyme A.
(Adapted from Clinical Hematology, Philadelphia, Mosby Elsevier. Sassa S. Porphyrias. In Young NS, Gerson SL, and High KA, eds, Copyright 2006,
with permission from Elsevier).

enzymatic step. This abnormality continues until sufficient • The logical stepwise approach is most useful when there are
haem synthesis is restored. clinical symptoms of the porphyrias (Sassa, 2004).
• There is significant interaction between the primary gene • Molecular diagnostic testing is powerful and very useful in
defect in the haem biosynthetic pathway and environmental kindred evaluation and genetic counselling when a disease-
factors for ALAS1 induction. Namely, patients with acute responsible mutation has been identified in the family. It is
hepatic porphyrias may not become symptomatic unless also the only proper way to screen asymptomatic gene
these subjects are exposed to certain drugs, liver damage, carriers.
hormonal changes during the menstrual cycle, stress, or
starvation, which result in the induction of ALAS1.
Treatment
• Recognition and avoidance of precipitating events is the
Diagnosis
first key part of treatment.
• Demonstration of porphyrin precursors, such as ALA and/or • Acute attacks of hepatic porphyrias should be treated
PBG, is essential for the diagnosis of acute hepatic porphyrias. similarly by (i) avoiding the precipitating factors; (ii) pro-
• Porphyrin analysis is necessary for the diagnosis of viding sufficient amounts of calories as carbohydrates (glu-
porphyrias with cutaneous photosensitivity. cose infusion); and (iii) intravenous infusion of haematin.

ª 2006 The Author


Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292 283
Review

Table I. Enzymes and genes for haem biosynthesis.

Genome

Enzyme Gene symbol Chromosomal location cDNA (bp) Protein (aa) Size (kb) Organization*

d-Aminolaevulinate synthase:
Housekeeping ALAS1 3p21.1 2199 640 17 11 exons
Erythroid-specific ALAS2 Xp11.21 1937 587 22 11 exons
d-Aminolaevulinate dehydratase: ALAD 9q34 13 exons
Housekeeping 1149 330 15Æ9 Exons 1A + 2-12
Erythroid-specific 1154 330 Exon 1B + 2-12
Porphobilinogen deaminase: HMBS 11q23.3 11 15 exons
Housekeeping 1086 361 Exons 1 + 3-15
Erythroid-specific 1035 344 Exons 2-15
Uroporphyrinogen Ill synthase: UROS 10q25.2–q26.3 34 10 exons
Housekeeping 1296 265 Exon 1 + 2B-10
Erythroid-specific 1216 265 Exon 2A + 2B-10
Uroporphyrinogen decarboxylase UROD 1p34 1104 367 3 10 exons
Coproporphyrinogen oxidase CPOX 3q12 1062 354 14 7 exons
Protoporphyrinogen oxidase PPOX 1q23 1431 477 5Æ5 13 exons
Ferrochelatase FECH 18q21.3 1269 423 45 11 exons

*Number of exons and those encoding housekeeping and erythroid-specific forms.

• Haemolytic anaemia in erythropoietic porphyrias may be uroporphyrin I and coproporphyrin I. This is the only
treated by blood transfusion. porphyria that produces type I isomers in excess.
• Cutaneous photosensitivity of erythropoietic protoporphy- Uroporphyrinogen synthase catalyses the cyclization of the
ria may be treated by oral b-carotene, while that of linear tetrapyrrole, hydroxymethylbilane, to yield a tetrapyr-
porphyria cutanea tarda (PCT) by phlebotomy, or oral role, uroporphyrinogen III, the physiological isomer, which
chloroquine. ultimately leads to the formation of haem. This step involves
inversion of the pyrrole D ring of hydroxymethylbilane and
cyclization to uroporphyrinogen III (Fig 2) (Battersby et al,
Erythropoietic porphyrias 1980). In the absence of UROS, as in CEP, hydroxymethylbi-
lane is converted non-enzymatically to the non-physiological
The characteristic features of each porphyric disorder are
porphyrin isomer, uroporphyrin I. Uroporphyrinogen I is then
described below. Porphyrins in red cells can cause photosen-
enzymatically converted to coproporphyrinogen I via the
sitive cell lysis, resulting in haemolytic anaemia. The two
activity of UROD, but it cannot be metabolized further.
homozygous erythropoietic porphyrias, congenital erythropoi-
Mild to severe haemolysis in CEP is characterised by
etic porphyria (CEP) and hepatoerythropoietic porphyria
anisocytosis, poikilocytosis, polychromasia, basophilic stip-
(HEP), are associated with haemolytic anaemia of varying
pling, reticulocytosis, increased nucleated red cells, absence of
degrees. In contrast, erythropoietic protoporphyria (EPP), a
haptoglobin, increased unconjugated bilirubin, increased
heterozygous disease, rarely has accompanying haemolytic
faecal urobilinogen and increased plasma iron turnover.
anaemia. The effect of life-long anaemia in CEP or HEP may
Secondary splenomegaly may contribute to the anaemia, and
lead to compensatory expansion of erythroid marrow, which
may also result in leucopenia and thrombocytopenia.
may result in pathological fractures, vertebral compression or
Anaemia can be so severe that some patients are transfu-
collapse, and shortness of stature. The haemolysis is also
sion-dependent. Splenectomy may reduce the need for
associated with varying degrees of splenomegaly and the
transfusions, although signs of ineffective erythropoiesis and
production of pigment-laden gallstones.
gallstones may continue.
Severe cutaneous photosensitivity usually begins in early
Congenital erythropoietic porphyria infancy and is manifested by increased friability and blistering
of the epidermis on the hands and face and other sun-exposed
Congenital erythropoietic porphyria is an erythropoietic
areas. Pink or red-brown staining of nappies due to markedly
porphyria inherited in an autosomal recessive fashion. It is
increased urinary porphyrins may be the first clue to the
one of the most severely affected photosensitive disorders. The
disease. Bullae and vesicles contain serous fluid and are prone
primary abnormality is an almost complete absence of UROS
to rupture and infection. The skin may be thickened, with
activity, previously termed uroporphyrinogen III cosynthase,
areas of hypo- and hyperpigmentation. Hypertrichosis of the
which results in massive accumulation and excretion of

ª 2006 The Author


284 Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292
Review

face and extremities is often prominent. Sunlight, other common porphyria after acute intermittent porphyria (AIP)
sources of ultraviolet light, and minor skin trauma increase and PCT, and the most common erythropoietic porphyria.
the severity of the cutaneous manifestations. Recurrent vesicles Molecular analysis of ferrochelatase mutations has revealed
and secondary infection can lead to cutaneous scarring and missense mutations, splicing abnormalities, intragenic dele-
deformities, as well as loss of finger nails and digits and severe tions, and possible nonsense mutations associated with func-
damage to the eyelids, nose and ears. Corneal scarring can lead tional deficiency of ferrochelatase (Ostasiewicz et al, 1995;
to blindness. Porphyrins deposited in the teeth produce a Todd, 1998). Among them, exon skipping was the most
reddish brown colour in natural light, termed erythrodontia. predominant. It has been reported that (i) a C fi T transition
Erythrodontia shows intense red fluorescence of porphyrins on at position $-$23 in intron 1 (NT) reduces ferrochelatase
exposure to long wavelength ultraviolet light. activity; but (ii) the NT allele alone is not sufficient to bring
A variety of mutations that cause CEP have been identified about EPP, even in its homozygous state; and (iii) both a
in the UROS gene, including missense and nonsense muta- mutant allele (M) and the Nr allele in trans are necessary for
tions, large and small deletions and insertions, splicing defects the clinical expression of EPP (Gouya et al, 1999). Thus,
and intronic branch point mutations (Fontanellas et al, 1996; ferrochelatase activity in patients can be defined as M-M, or
Desnick et al, 1998). UROS knockout in mice is embryonic M-NT, silent gene carriers as M-NC, and normal controls as
lethal, while some UROS mutants knocked into these animals NC-NC, NC-NT, or NT-NT (increasingly reduced ferrochelatase
not only support the survival of the animals, but also develop activity). This model accounts for the fact (i) why normal
cutaneous photosensitivity similar to those observed in CEP controls have a wide range of ferrochelatase activity; (ii) why
(Bishop et al, 2006). These findings suggest that this animal silent gene carriers have higher ferrochelatase activity than
model is useful for the study of the effects of UROS patients; and (iii) why patients with EPP have less than 50% of
mutations. ferrochelatase activity. This model explains the majority of EPP
Urinary porphyrin excretion is markedly increased (up to cases, the rest being true autosomal recessive disease.
50–100 mg/d, normal range: <0Æ2 mg, or 0Æ3 lmol/d), and Cutaneous photosensitivity begins in childhood, affects sun-
consists mostly of uroporphyrin, hepta-carboxyl porphyrin, exposed areas, such as face and dorsum of hands, and is
and coproporphyrin, with lesser increases in hexa- and penta- generally worse in spring and summer. Common symptoms
carboxyl porphyrins (Fritsch et al, 1997). Urinary ALA and include itching, painful erythema and swelling, which can
PBG excretion are not increased. Faecal porphyrins are develop within minutes of sun exposure (Poh-Fitzpatrick,
markedly increased due to increased coproporphyrin I. 1984). Diffuse oedema of the skin in sun-exposed areas may
Circulating erythrocytes contain large amounts of uroporphy- resemble angioneurotic oedema. On occasion, burning and
rin I and lesser, but still excessive, amounts of coproporphyrin itching can occur without obvious skin damage. Petechiae and
I. Red cell protoporphyrin may also be increased, which purpuric lesions may occur. Skin lichenification, leathery
reflects increased erythropoiesis. HEP is clinically very similar pseudovesicles and nail changes can be pronounced.
to CEP but can be distinguished by the predominance of Bone marrow reticulocytes are the primary source of the
protoporphyrin in red cells and isocoproporphyrin in stool. excess protoporphyrin that accumulates in tissues and is
There are three main components in the treatment of CEP, excreted in faeces in EPP. Bone marrow fluorescence is almost
(i) protection from sunlight and UV exposure; (ii) meticulous entirely in reticulocytes rather than nucleated erythroid cells.
skin care; and (iii) red cell transfusions and other haemato- Protoporphyrin is markedly increased in erythrocytes and
logical supportive care. Other therapeutic interventions plasma, and excreted in stool. Unlike other porphyrias, there is
include, (i) treatment with hydroxyurea to reduce bone no increase in porphyrin excretion in urine. In contrast to iron
marrow porphyrin synthesis (Guarini et al, 1994); (ii) splen- deficiency anaemia and lead poisoning, in which increased
ectomy to reduce transfusion requirements in patients with erythrocyte protoporphyrin is chelated with zinc, erythrocytic
hypersplenism (Piomelli et al, 1986); and (iii) oral charcoal porphyrin in EPP is exclusively free protoporphyrin (Lamola
treatment to facilitate faecal excretion of porphyrins (Tishler & et al, 1975). There is no increase in plasma porphyrin
Winston, 1990). Allogeneic bone marrow transplantation has concentrations in the two former conditions (Poh-Fitzpatrick
also proved curative for patients with CEP (Kauffman et al, & Lamola, 1976).
1991; Thomas et al, 1996; Zix-Kieffer et al, 1996; Tezcan et al, Oral administration of b-carotene is useful for treating EPP
1998). Successful marrow transplantation results in marked (Mathews-Roth et al, 1977). b-Carotene doses of 120–180 mg
reduction of the photosensitivity and porphyrin levels. daily in adults are usually required to maintain serum carotene
levels in the recommended range of 11Æ16–14Æ88 lmol/l, but
doses up to 300 mg daily may sometimes be needed. Suntan-
Erythropoietic protoporphyria
ning, resulting from better tolerance of sunlight, may lead to
Erythropoietic protoporphyria is characterised by a partial further protection. Cholestyramine and other porphyrin
deficiency of ferrochelatase activity. Cutaneous photosensitiv- absorbents, such as activated charcoal, may be helpful. Other
ity characteristically begins in childhood, but there is no therapeutic options include red blood cell transfusions,
neurological involvement. This is probably the third most exchange transfusion and intravenous haematin to suppress

ª 2006 The Author


Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292 285
Review

erythroid and hepatic protoporphyrin production, as well as asymptomatic, they may be at a higher risk than normal
liver transplantation. While liver transplantation can be individuals for developing ADP when exposed to environ-
temporary beneficial, the new liver is susceptible to proto- mental chemicals or conditions that further influence deficient
porphyrin-induced damage. ALAD activity (Akagi et al, 2000).
In addition to the inherited enzymatic deficiency, ALAD
activity can be inhibited by a number of substances including
Acute hepatic porphyrias
divalent heavy metal ions. By virtue of its zinc displacing
The most common neurovisceral complaints in acute hepatic activity from the enzyme, lead potently inhibits ALAD activity
porphyrias are abdominal pain (occurring in 85–95% cases), (Granick et al, 1978). Over 99% of lead in blood is found in
vomiting (43–88%), constipation (48–84%), muscle weakness erythrocytes, of which over 80% is bound to ALAD (Bergdahl
(42–60%), mental symptoms (40–58%), limb, head, neck, et al, 1997), and patients with lead poisoning are associated
chest pain (50–52%), hypertension (36–54%), tachycardia with marked inhibition of ALAD activity in red cells. They also
(28–80%), convulsion (10–20%), sensory loss (9–38%), fever develop neurological symptoms, some of which resemble those
(9–37%), respiratory paralysis (5–12%) and diarrhoea (5– seen in ADP. The most potent inhibitor of ALAD is
12%) (Waldenstrom, 1957; Goldberg, 1959; Stein & Tschudy, succinylacetone (Sassa & Kappas, 1983), a structural analogue
1970). While the exact mechanism underlying these com- of ALA, which is found in the urine and blood of patients with
plaints is not yet well understood, various hypotheses have hereditary tyrosinemia type 1 (Lindblad et al, 1977). As a
been put forward. For example, (i) excess amounts of PBG or result, approximately 40% of children with hereditary tyros-
ALA may cause neurotoxicity (Meyer et al, 1998); (ii) inaemia develop the signs and symptoms of ADP, suggesting
increased ALA concentrations in the brain may inhibit that succinylacetone results in the significant inhibition of
gamma-aminobutyric acid release (Mueller & Snyder, 1977; haem synthesis in the liver.
Brennan & Cantrill, 1979); (iii) haem deficiency may result in The ADP is characterised by the markedly increased
degenerative changes in the central nervous system (Whetsell production of ALA, such that ALA production is 10–20 times
et al, 1984); (iv) decreased haem synthesis in the liver results in that of PBG. AIP, on the other hand, is characterised by
decreased activity of hepatic tryptophan pyrrolase (TP), a increased production of both ALA and PBG, often to a
haem-dependent enzyme, possibly resulting in increased levels comparable extent. The fact that AIP and ADP are clinically
of brain tryptophan and increased turnover of 5-hydroxytryp- indistinguishable suggests that ALA or its metabolites, rather
tamine, a neurotransmitter (Litman & Correia, 1985); (v) as all than PBG, may be the neurotoxic agent. This is further
of the porphyrias that are associated with neurovisceral supported by the finding, that patients with hereditary
complaints show increased urinary excretion of ALA (ADP), tyrosinaemia frequently develop ADP along with its neurolog-
or of ALA and PBG [AIP, hereditary coproporphyria (HCP) ical abnormalities, most likely via the marked inhibition of
and variegate porphyria (VP)], and ALA increases lipid ALAD by succinylacetone.
peroxidation, ALA-mediated lipid peroxidation may under- Treatment of ADP, as with other acute hepatic porphyrias, is
score the acute crisis of porphyria (Bechara, 1996); and (vi) TP two-fold: (i) acute attacks are treated by withdrawal of the
deficiency results in decreased plasma melatonin levels (Puy offending agent (if any), infusion of glucose, and, if the patient
et al, 1996), which may result in the loss of protection against is still symptomatic, intravenous infusion of haem prepara-
ALA-mediated lipid peroxidation. tions or administration of inhibitors of haem oxygenase, (ii)
future attacks are prevented by avoiding agents known to
stimulate ALAS activity and by avoiding those agents known to
ALA dehydratase porphyria (ADP)
inhibit ALAD activity.
ALA dehydratase porphyria (ADP) is the rarest form of the
inherited porphyrias. Only six cases have been reported that
Acute intermittent porphyria
were molecularly confirmed to be due to ALAD mutations:
three German males (Ishida et al, 1992; Doss et al, 2004), one Acute intermittent porphyria is due to inherited (autosomal
Swedish baby boy (Thunell et al, 1987; Plewinska et al, 1991), dominant) mutations of the PBGD gene (HMBS) leading to
one elderly Belgian man (Hassoun et al, 1989; Akagi et al, deficient activity of the enzyme. This is the most important
2000), and an American male (Akagi et al, 2006). It is an acute hepatic porphyria, both in its incidence and clinical
autosomal recessive disorder resulting from a homozygous severity. The enzyme activity is 50% of normal in those who
deficiency of ALAD in five reported patients, while it was due inherit the genetic trait. There is no difference in erythrocyte
to a clonal expansion of a mutant ALAD allele in a late-onset PBGD activity between patients and latent gene carriers. The
disease in the Belgian patient. While clinical ADP is rare, the prevalence of AIP in the United States is thought to be 5–10
frequency of genetic carriers for ALAD deficiency, i.e. hetero- per 100 000. It is more common in northern European
zygotes, in the normal population is as high as 2%, according countries, such as Sweden (60–100 per 100 000), Britain and
to a study performed in Sweden (Thunell et al, 1987). Ireland. More than 200 mutations of the PBGD gene have been
Although subjects heterozygous for ALAD deficiency are described to date in AIP.

ª 2006 The Author


286 Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292
Review

Acute intermittent porphyria is more common in women Nutritional factors. Reduced energy intake, usually institu-
than in men, and very rare in children. Symptoms may appear ted in an effort to lose weight, commonly contributes to
at or any time after puberty. The major clinical manifestations attacks of AIP. Thus, even brief periods of starvation during
of AIP, including abdominal pain and other neurovisceral and weight reduction, postoperative periods, or intercurrent illness
circulatory disturbances, are due to effects on the nervous should be avoided. Starvation induces haem oxygenase-1 (HO-
system. Neurological and visceral symptoms are almost always 1) activity, which can lead to depletion of regulatory hepatic
intermittent and usually occur in acute attacks that develop haem pools and contribute to ALAS induction. Transcription
over a few hours or days. The disease can be disabling but is of ALAS1 can also be upregulated by the peroxisome
only occasionally fatal. Abdominal pain has been reported in proliferators-activated receptor c-coactivator 1a (PGC-1a).
85–95% of cases, followed by tachycardia (80%). Abdominal Under conditions of low glucose, PGC-1a production increa-
pain is usually severe, steady and poorly localised, but may be ses, leading to increased levels of ALAS1, creating conditions
cramping. A variety of mental symptoms, pain in limbs, head, for attacks of acute porphyrias (Handschin et al, 2005).
neck, or chest, muscle weakness and sensory loss can occur. Glucose and other forms of carbohydrate are effective in
Weakness most commonly begins in the proximal muscles and treating acute attacks of porphyria.
more often in the arms than in the legs. Smoking. Chemicals in tobacco smoke, such as polycyclic
The course of the neurological manifestations is highly aromatic hydrocarbons, are known inducers of hepatic cyto-
variable. Sudden death may occur, presumably due to cardiac chrome P450 enzymes and haem synthesis. An association
arrhythmia. Acute attacks of porphyria may resolve quite between cigarette smoking and repeated attacks of porphyria
rapidly. The disease may be complicated by electrolyte was found in a survey of 144 patients with AIP in Britain (Lip
abnormalities. Hyponatraemia is common during acute et al, 1991). As a result, smoking cessation may have particular
attacks, and may help to suggest the diagnosis. This is health benefits in patients with acute porphyrias.
sometimes due to the syndrome of inappropriate antidiuretic Infections, surgery and stress. Attacks of porphyria may
hormone secretion. The urine is often dark red in colour, due develop during intercurrent infections and other illnesses, after
to the presence of porphobilin, the oxidised product of PBG. major surgery, and psychological stress. Most of these condi-
Mice with PBGD deficiency induced by gene targeting tions are known to increase hepatic HO-1 activity (Rodgers &
display some of the symptoms observed in patients with AIP, Stevenson, 1990).
such as impaired motor function, ataxia, increased levels of All individuals with AIP, whether active or latent, have a
ALA in plasma and brain, and decreased haem saturation of 50 % deficiency of PBGD activity in the liver. Urinary ALA
liver tryptophan pyrrolase (Lindberg et al, 1996). This model and PBG are always markedly increased in symptomatic
should be useful to further define the mechanisms of patients with AIP (Aarsand et al, 2006) and even in some
neurological damage in acute porphyrias. asymptomatic individuals with the inherited enzyme deficiency
(Floderus et al, 2006). PBG in urine is oxidised to porphobilin
Precipitating factors. An inherited deficiency of PBGD is not in upon standing, which gives a dark-brown colour to urine, and
itself sufficient to cause clinical expression of AIP, as the great often referred to as ‘port-wine reddish urine’. Urinary ALA
majority ( 90%) of individuals who inherit a deficiency of and PBG levels may decrease to normal if the disease becomes
PBGD never develop porphyric symptoms. There is inactive for a prolonged period. Urinary porphyrins are also
considerable evidence that endocrine factors and steroid markedly increased. Circulating concentrations of ALA and
hormones play an important role in precipitating the acute PBG in plasma (or serum) are substantially increased (Flode-
attack of AIP (Sassa, 1996). Thus, acute hepatic porphyrias are rus et al, 2006).
rarely seen before puberty, but are quite common at puberty or Diagnosis can be established by the demonstration of
after, and particularly often in the premenstrual phase in reduced PBGD activity in erythrocytes (about 50% of normal)
women. in type I and type III AIP patients. Type II AIP patients show
Drugs. Drugs are among the most important factors that normal PBGD activity in erythrocytes, but have reduced PBGD
precipitate acute attacks of acute porphyria, with barbiturates activity in non-erythroid cells, such as fibroblasts or lympho-
and sulphonamide antibiotics being most common (Anderson cytes (Sassa, 1996). Patients with the clinically expressed disease
et al, 2001). A significant number of commonly used drugs are excrete increased amounts of ALA and PBG in the urine, and
widely agreed to be either safe or unsafe for use in patients with often even during clinical remission. During an acute attack of
AIP. Many of these can be found on a dedicated web site AIP, there are further massive increases in excretion of these
(http://www.porphyria-europe.org). However, knowledge precursors (ALA 25–100 mg/d; PBG 50–200 mg/d). Some
about the safety of many drugs and other over-the-counter latent gene carriers who are characterised by reduced PBGD
preparations in acute porphyrias is yet incomplete. Most drugs activity similar to patients also show increased ALA and PBG
that exacerbate AIP have the capacity to induce ALAS activity excretion. The Watson–Schwartz test has been widely used as a
in the liver. As mentioned above, this process is closely screening test for urinary PBG (Watson & Schwartz, 1941). It is
associated with the induction of cytochrome P450 enzymes, a useful for evaluating the efficacy of treatment on acute crisis.
process that increases the demand for hepatic haem synthesis. While the test may occasionally be associated with false-positive

ª 2006 The Author


Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292 287
Review

findings, it is quick and quite sufficient for the rapid diagnosis (CPO) activity. HCP is markedly less frequent than AIP and
of an acute attack of AIP. The column method of Mauzerall and VP. HCP has identical clinical symptoms as AIP, except for its
Granick is quantitative and more specific, and can determine cutaneous photosensitivity, which is entirely absent in AIP.
the amounts of both ALA and PBG (Anderson et al, 2001). For It is generally thought that HCP may be a milder disease
this reason, the column method of Mauzerall and Granick is than AIP (Brodie et al, 1977), but there have also been a few
now more widely used than the Watson–Schwartz test to avoid fatalities from respiratory paralysis in HCP. Molecular
possible biological false-positive findings. All participating analysis of several families with HCP, including the homo-
laboratories in the UK External Quality Assurance now use zygous dominant form and the harderoporphyria variant, has
the column method, instead of the Watson–Schwartz test. revealed a variety of mutations in the CPO gene (Martasek,
Elevated levels of urinary ALA may additionally be seen in ADP, 1998). In a series of 53 cases of HCP in Germany, the
HCP and VP, while elevated levels of both ALA and PBG may incidence of symptoms was 89%, 33%, 28%, 25% and 14 %,
be seen in HCP and VP. for abdominal pain, neurological, psychiatric, cardiovascular
Detection of PBGD mutations in AIP provides 95% and skin symptoms respectively (Kuhnel et al, 2000). The
sensitivity and around 100% specificity (Kauppinen, 2004), disease is latent before puberty, and symptoms are more
and has quickly been incorporated into good clinical practice. common in adult women than men. This porphyria is
It can definitively confirm or exclude the diagnosis of AIP. exacerbated by many of the same factors that precipitate
While it is powerful and specific, and is very useful in pedigree attacks in AIP, including barbiturates and other drugs, and
screening and in genetic counselling, its application at the endogenous or exogenous steroid hormones. Symptoms can
population level is not recommended, unless the frequency of occur in association with the menstrual cycle. The risk of
gene carriers is locally very high (Kauppinen, 2004). DNA hepatocellular carcinoma may be increased in this porphyria,
testing of AIP also determines the gene carrier status, rather as in AIP and VP. A few homozygous HCP cases, with
than the severity of the disease. The same consideration applies cutaneous lesions beginning in early childhood, have also
to the diagnosis of most other acute hepatic porphyrias. been reported.
The treatment of acute attack is essentially the same for AIP, The most prominent biochemical feature of HCP is a
ADP, HCP and VP. Intravenous administration of carbohy- marked increase in urinary and faecal coproporphyrin,
drate (as dextrose) should be given to provide a minimum of predominantly isomer type III, typically 10–200 times com-
300 g of carbohydrate per day. Glucose and insulin together pared with controls (Martasek, 1998). Urinary ALA, PBG and
might also be more effective than glucose alone, as insulin uroporphyrin are also increased during acute attacks. In
blunts the PGC-1a effect (Handschin et al, 2005). Intravenous harderoporphyria, a variant form of HCP, faecal excretion of
haematin administration is the treatment of choice, which harderoporphyrin (tricarboxylate porphyrin) and copropor-
curtails urinary excretion of ALA and PBG, acute attacks, and phyrin is increased. Treatment of acute attacks of HCP is
perhaps the severity of neuropathy. Reconstitution of haema- identical to the treatment of AIP. The identification and
tin for intravenous infusion should be made according to the avoidance of precipitating factors is also essential for HCP.
protocol described recently (Anderson et al, 2006). Normosang
[Orphan Europe (UK) Ltd, Henley-on-Thames, UK], is a
Variegate porphyria
stable haematin solution complexed with arginine (Mustajoki
et al, 1989). It is known to cause lesser vascular complications, Variegate porphyria is an autosomal dominant acute hepatic
such as phlebitis than the conventional haematin solution. porphyria, due to a deficiency of PPO activity, the penultimate
Liver transplantation was attempted in a 19-year-old woman enzyme in the haem biosynthetic pathway. The disease is
with severe AIP. It successfully reduced urinary porphyrin termed variegate because it can present itself either with
precursor excretion to normal, and improved her quality of life neurological manifestations, cutaneous photosensitivity, or
(Soonawalla et al, 2004). Thus liver transplantation might be both.
considered for selected patients with the severest forms of AIP, In most countries this porphyria is less commonly recog-
though it has not been successful in ADP. Nasal or subcuta- nised than AIP. However, VP is quite common in South Africa,
neous administration of long-acting agonist of luteinizing where it was first reported in 1945. The high prevalence among
hormone-releasing hormone inhibits ovulation and greatly South African whites (approximately 3 of 1000) is due to
reduces the incidence of perimenstrual attacks of AIP in such genetic drift or the ‘founder effect’, and most VP cases in South
women (Anderson et al, 1990a). Pain, which is invariably Africa can be traced to a man or his wife who emigrated from
present and severe, can be treated with frequent regular doses Holland and were married in 1688 (Dean, 1982). As many as
of narcotic analgesics. 20 000 South Africans may carry this gene, which is now found
to be R59W mutation of the PPO gene (Meissner et al, 1996).
The acute attack is identical to that seen in AIP, and may
Hereditary coproporphyria
include abdominal pain, tachycardia, vomiting, constipation,
Hereditary coproporphyria is an autosomal dominant hepatic hypertension, neuropathy, back pain, confusion, bulbar para-
porphyria due to a deficiency of coproporphyrinogen oxidase lysis, psychiatric symptoms, fever, urinary frequency and

ª 2006 The Author


288 Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292
Review

dysuria. Hyponatraemia with evidence of sodium depletion or A variety of mutations of UROD have been identified in
inappropriate antidiuretic hormone secretion can occur during PCT (McManus et al, 1996; Mendez et al, 1998; Christiansen
acute attacks. Attacks of VP are generally milder than in AIP, et al, 2000). All but two were each found in only one family.
and recurrent attacks are less common. Cutaneous photosen- UROD mutations have also been found in HEP, the homo-
sitivity is more common than in HCP. Skin manifestations zygous form of UROD deficiency, but they were mostly unique
generally occur, and are usually of longer duration. They are and usually not found in PCT (Meguro et al, 1994).
very similar to those seen in PCT and HCP, and include Chronic blistering lesions develop on sun-exposed areas of
increased fragility, vesicles, bullae, erosions, milia, hyperpig- skin. These are more common in the summer than winter. The
mentation and hypertrichosis of sun-exposed areas. Photosen- fluid-filled vesicles rupture easily and the denuded areas
sitivity may be less commonly associated with VP in northern become crusted and heal slowly. Previous areas of blisters may
countries, where sunlight is less intense (Kirsch et al, 1998). appear atrophic, or brownish. Facial hypertrichosis and
Faecal protoporphyrin and coproporphyrin and urinary hyperpigmentation are also common. All types of PCT
coproporphyrin are markedly increased when VP is clinically respond readily to repeated phlebotomy (Ippen, 1977). PCT
active. Urinary ALA, PBG and uroporphyrin are also increased can also be treated with of chloroquine or hydroxychloroquine
during acute attacks but may be normal or only slightly (administered in very low dosage regimen). This is the
increased during remission. Plasma porphyrin is commonly preferred therapy at some centres, especially for patients
increased in VP, and its analysis is more sensitive than that of without marked iron overload (Valls et al, 1994).
faecal porphyrins (Hift et al, 2004). Increased plasma porphy- Porphyrins accumulate in large amounts in liver, and are
rin in VP is readily detected as a fluorescence emission increased in plasma. Uroporphyrin and heptacarboxylporph-
maximum at 626–628 nm. This is a specific finding in VP yrin are predominantly increased in urine.
(Enriquez et al, 1993), as fluorescence emission peaks in EPP Multiple factors can contribute to inactivation or inhibition
and PCT are found at 636 nm and 618–622 nm respectively of hepatic UROD in this disease, probably by an iron-
(Enriquez et al, 1993). Biliary porphyrins are also increased. dependent oxidative mechanism, including alcohol (Grossman
The risk of gallstones may be increased in VP, which consist et al, 1979), hepatitis C infection (Fargion et al, 1992),
mainly of protoporphyrin. The X porphyrin fraction (ether- oestrogen (Grossman et al, 1979), human immunodeficiency
acetic acid-insoluble porphyrins extractable from faeces by virus (HIV) (Egger et al, 2002), and factors that increase
urea-Triton) is increased in VP. Rare homozygous VP patients hepatic iron content, such as mutations of the HFE gene (HFE)
have markedly increased erythrocyte Zn-protoporphyrin associated with haemochromatosis (Roberts et al, 1997a,b).
(Whatley et al, 1999). Liver biopsy frequently shows haemosiderosis, and serum iron
and ferritin concentrations are increased. Patients often
develop cirrhosis, and may occasionally develop hepatoma
Chronic hepatic porphyria
(Kordac, 1972).
Patients with chronic hepatic porphyria, such as PCT, are HFE mutation is found in PCT more commonly than in
associated with significant chronic cutaneous photosensitivity, control populations, indicating that a tendency to iron
but they do not accompany neurological symptoms. Petechiae overload can predispose to PCT (Bonkovsky et al, 1998).
and purpuric lesions may occur. Skin lichenification, leathery There is no increase in ALA or PBG in this disorder. Faecal
pseudovesicles and nail changes can be pronounced. Chronic isocoproporphyrin is the biochemical stigmata of the UROD
blistering lesions may develop; the fluid-filled vesicles rupture deficiency, and its detection establishes the diagnosis of PCT
easily and the denuded areas become crusted and heal slowly; (or HEP) (Elder, 1998). Namely, isocoproporphyrin III is
secondary infection is common. Previous areas of blisters may formed from dehydroisocoproporphyrinogen III that is
appear atrophic, or brownish. Facial hypertrichosis, scarring normally a minor product but can accumulate from penta-
and hyperpigmentation may result. carboxylate porphyrinogen III if UROD activity is reduced
(Elder & Chapman, 1970).
The identification and avoidance of precipitating factors is
Porphyria cutanea tarda
the first line of treatment, particularly for patients with type I
Porphyria cutanea tarda is due to a profound deficiency of PCT. Abstinence from alcohol should be recommended in all
uroporphyrinogen decarboxylase (UROD) activity in liver. types of PCT. Phlebotomy is usually effective in reducing
The disease has been classified into three subtypes. Type I urinary porphyrin concentration, serum ferritin concentration,
PCT has decreased hepatic UROD activity, but normal serum transferrin saturation, and in induction of clinical
erythrocyte UROD activity, and is found in sporadic fashion remission (Ippen, 1977). Phlebotomy should be undertaken
without family history. Type II PCT has decreased UROD weekly or biweekly until serum transferrin saturation and
activity both in red cells and in liver, and occurs in multiples serum ferritin levels are normalised. It usually requires blood
in a family. Type III PCT is similar to type II with respect withdrawal ranging from 2Æ5 l to 7 l. Serum transferrin
to familial occurrence, but erythrocyte UROD activity is saturation and ferritin levels are correlated with urinary
normal. porphyrin excretion, and the clinical syndrome, and should

ª 2006 The Author


Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292 289
Review

be used to assess the exhaustion of hepatic iron stores during in a patient on long-term hemodialysis [published erratum appears
phlebotomy treatment, as well as to detect an early replenish- in N Engl J Med 1990 May 31;322(22):1616]. New England Journal of
ment after remission (Rocchi et al, 1985). If phlebotomy is Medicine, 322, 315–317.
ineffective, or contraindicated because of the presence of other Anderson, K.E., Sassa, S., Bishop, D.F. & Desnick, R.J. (2001) Disor-
ders of heme biosynthesis: X-linked sideroblastic anemia and the
diseases, such as anaemia, low-dose chloroquine therapy may
porphyrias. In: The Metabolic & Molecular Bases of Inherited Disease
be considered (Valls et al, 1994). Chloroquine is thought to
(ed. by C.R. Scriver, A.L. Beaudet, W.S. Sly & D. Valle), p. 2991.
chelate porphyrins to facilitate their water-solubility, thus to McGraw-Hill Medical Publishing Division, New York.
enhance their excretion into urine. Treatment of PCT associ- Anderson, K.E., Bonkovsky, H.L., Bloomer, J.R. & Shedlofsky, S.I.
ated with end-stage renal disease can be difficult, because (2006) Reconstituion of hematin for intravenous infusion. Annals of
phlebotomy is contraindicated and chloroquine is not effect- Internal Medicine, 144, 537–538.
ive, presumably because kidney failure precludes excretion of Battersby, A.R., Fookes, C.J., Matcham, G.W. & McDonald, E. (1980)
excess porphyrins mobilized from liver. Recombinant erythro- Biosynthesis of the pigments of life: formation of the macrocycle.
poietin administration is now the treatment of choice for these Nature, 285, 17–21.
patients, as it can correct anaemia, mobilize iron and support Bechara, E.J. (1996) Oxidative stress in acute intermittent porphyria
phlebotomy in this condition (Anderson et al, 1990b; Yaqoob and lead poisoning may be triggered by 5-aminolevulinic acid.
Brazilian Journal of Medical & Biological Research, 29, 841–851.
et al, 1992; Peces et al, 1994).
Bergdahl, I.A., Grubb, A., Schutz, A., Desnick, R.J., Wetmur, J.G.,
Sassa, S. & Skerfving, S. (1997) Lead binding to delta-aminolevulinic
Hepatoerythropoietic porphyria acid dehydratase (ALAD) in human erythrocytes. Pharmacology &
Toxicology, 81, 153–158.
Hepatoerythropoietic porphyria is a rare form of porphyria, Bishop, D.F., Johansson, A., Phelps, R., Shady, A.A., Ramirez, M.C.,
which is due to a homozygous or compound heterozygous Yasuda, M., Caro, A. & Desnick, R.J. (2006) Uroporphyrinogen III
deficiency of UROD. Only some 20 cases of HEP were reported synthase knock-in mice have the human congenital erythropoietic
(Meguro et al, 1994). HEP resembles CEP clinically and porphyria phenotype including the characteristic light-induced
usually presents in infancy or childhood with red urine, cutaneous lesions. American Journal of Human Genetics, 78, 645–
blistering skin lesions, hypertrichosis and scarring. Scleroder- 658.
moid skin changes may be prominent. Erythrocyte porphyrins Bonkovsky, H.L., Poh-Fitzpatrick, M., Pimstone, N., Obando, J., Di
Bisceglie, A., Tattrie, C., Tortorelli, K., LeClair, P., Mercurio, M.G.
are increased, but they are predominantly type III isomers.
& Lambrecht, R.W. (1998) Porphyria cutanea tarda, hepatitis C,
Some patients may be associated with haemolytic anaemia and
and HFE gene mutations in North America. Hepatology, 27, 1661–
splenomegaly. 1669.
The biochemical findings in HEP, however, resemble those Brennan, M.J. & Cantrill, R.C. (1979) Delta-aminolaevulinic acid is a
that are observed in PCT, and include predominant accumu- potent agonist for GABA autoreceptors. Nature, 280, 514–515.
lation and excretion of uroporphyrin, heptacarboxylate Brodie, M.J., Thompson, G.G., Moore, M.R., Beattie, A.D. & Goldberg,
porphyrin and isocoproporphyrins. Erythrocyte Zn-proto- A. (1977) Hereditary coproporphyria. Demonstration of the ab-
porphyrin concentration is increased in HEP. Treatment of normalities in haem biosynthesis in peripheral blood. Quarterly
HEP is identical to that for CEP. Unlike PCT, phlebotomy is Journal of Medicine, 46, 229–241.
not effective in this disease. Christiansen, L., Bygum, A., Jensen, A., Brandrup, F., Thomsen, K.,
Horder, M. & Petersen, N.E. (2000) Uroporphyrinogen decarbox-
ylase gene mutations in Danish patients with porphyria cutanea
References tarda. Scandinavian Journal of Clinical & Laboratory Investigation,
60, 611–615.
Aarsand, A.K., Petersen, P.H. & Sandberg, S. (2006) Estimation and
Dean, G. (1982) Porphyria variegata. Acta Dermato-Venereolog-
application of biological variation of urinary d-aminolevulinic acd
ica.Supplementum (Stockh), 100, 81–85.
and porphobilinogen in healthy individuals and in patients with
Desnick, R.J., Glass, I.A., Xu, W., Solis, C. & Astrin, K.H. (1998)
acute intermittent porphyria. Clinical Chemistry, 52, 650–656.
Molecular genetics of congenital erythropoietic porphyria. Seminars
Akagi, R., Nishitani, C., Harigae, H., Horie, Y., Garbaczewski, L.,
in Liver Disease, 18, 77–84.
Hassoun, A., Mercelis, R., Verstraeten, L. & Sassa, S. (2000) Mole-
Doss, M.O., Stauch, T., Gross, V., Renz, M., Akagi, R., Doss-Frank, M.,
cular analysis of d-aminolevulinate dehydratase deficiency in a
Seelig, H.P. & Sassa, S. (2004) The third case of doss porphyria (d-
patient with an unusual late-onset porphyria. Blood, 96, 3618–3623.
aminolevulinic acid dehydratase deficiency) in Germany. Journal of
Akagi, R., Kato, N., Inoue, R., Anderson, K.E., Jaffe, E.K. & Sassa, S.
Inherited Metabolic Disease, 27, 529–536.
(2006) d-Aminolevulinate dehydratase (ALAD) porphyria: the first
Egger, N.G., Goeger, D.E., Payne, D.A., Miskovsky, E.P., Weinman,
case in North America with two novel ALAD mutations. Molecular
S.A. & Anderson, K.E. (2002) Porphyria cutanea tarda: multiplicity
Genetics & Metabolism, 87, 329–336.
of risk factors including HFE mutations, hepatitis C, and inherited
Anderson, K.E., Spitz, I.M., Bardin, C.W. & Kappas, A. (1990a) A
uroporphyrinogen decarboxylase deficiency. Digestive Diseases &
gonadotropin releasing hormone analogue prevents cyclical attacks
Sciences, 47, 419–426.
of porphyria. Archives of Internal Medicine, 150, 1469–1474.
Elder, G.H. (1998) Porphyria cutanea tarda. Seminars in Liver Disease,
Anderson, K.E., Goeger, D.E., Carson, R.W., Lee, S.M. & Stead, R.B.
18, 67–75.
(1990b) Erythropoietin for the treatment of porphyria cutanea tarda

ª 2006 The Author


290 Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292
Review

Elder, G.H. & Chapman, J.R. (1970) Isolation of a hydroxylated por- Ishida, N., Fujita, H., Fukuda, Y., Noguchi, T., Doss, M., Kappas, A. &
phyrin from urine and faeces of patients with cutaneous hepatic Sassa, S. (1992) Cloning and expression of the defective genes from a
porphyria. Biochimica et Biophysica Acta, 208, 535–537. patient with d-aminolevulinate dehydratase porphyria. Journal of
Enriquez de Salamanca, R., Sepulveda, P., Moran, M.J., Santos, J.L., Clinical Investigation, 89, 1431–1437.
Fontanellas, A. & Hernandez, A. (1993) Clinical utility of fluoro- Kauffman, L., Evans, D.I., Stevens, R.F. & Weinkove, C. (1991) Bone-
metric scanning of plasma porphyrins for the diagnosis and typing marrow transplantation for congenital erythropoietic porphyria.
of porphyrias. Clinical & Experimental Dermatology, 18, 128–130. Lancet, 337, 1510–1511.
Falk, J.E. (1964) Porphyrins and Metalloporphyrins, vol. 2. Elsevier, New Kauppinen, R. (2004) Molecular diagnostics of acute intermittent
York. porphyria. Expert Review of Molecular Diagnostics, 4, 243–249.
Fargion, S., Piperno, A., Cappellini, M.D., Sampietro, M., Fracanzani, Kirsch, R.E., Meissner, P.N. & Hift, R.J. (1998) Variegate porphyria.
A.L., Romano, R., Caldarelli, R., Marcelli, R., Vecchi, L. & Fiorelli, Seminars in Liver Disease, 18, 33–41.
G. (1992) Hepatitis C virus and porphyria cutanea tarda: evidence of Kordac, V. (1972) Frequency of occurrence of hepatocellular carci-
a strong association. Hepatology, 16, 1322–1326. noma in patients with porphyria cutanea tarda in long-term follow-
Floderus, Y., Sardh, E., Moeller, C., Andersson, C., Rejkjaer, L., up. Neoplasma, 19, 135–139.
Andersson, D.E.H. & Harper, P. (2006) Variations in porphobili- Kuhnel, A., Gross, U. & Doss, M.O. (2000) Hereditary coproporphyria
nogen and 5-aminolevulinic acid concentrations in plasma and ur- in Germany: clinical-biochemical studies in 53 patients. Clinical
ine from asymptomatic carriers of the acute intermittent porphyria Biochemistry, 33, 465–473.
gene with increased porphyrin precursor excretion. Clinical Chem- Lamola, A.A., Piomelli, S., Poh-Fitzpatrick, M.G., Yamane, T. &
istry, 52, 701–707. Harber, L.C. (1975) Erythropoietic protoporphyria and lead in-
Fontanellas, A., Bensidhoum, M., Enriquez de Salamanca, R., Moruno, toxication: the molecular basis for difference in cutaneous photo-
T.A., de Verneuil, H. & Ged, C. (1996) A systematic analysis of the sensitivity. II. Different binding of erythrocyte protoporphyrin to
mutations of the uroporphyrinogen III synthase gene in congenital hemoglobin. Journal of Clinical Investigation, 56, 1528–1535.
erythropoietic porphyria. European Journal of Human Genetics, 4, Lindberg, R.L., Porcher, C., Grandchamp, B., Ledermann, B., Burki, A.,
274–282. Brandner, S., Aguzzi, A. & Meyer, U.A. (1996) Porphobilinogen
Fritsch, C., Bolsen, K., Ruzicka, T. & Goerz, G. (1997) Congenital deaminase deficiency in mice causes a neuropathy resembling that of
erythropoietic porphyria. Journal of the American Academy of Der- human hepatic porphyria. Nature Genetics, 12, 195–199.
matology, 36, 594–610. Lindblad, B., Lindstedt, S. & Steen, G. (1977) On the genetic defects in
Goldberg, A. (1959) Acute intermittent porphyria: a study of 50 cases. hereditary tyrosinemia. Proceedings of the National Academy of Sci-
Quarterly Journal of Medicine, 28, 183–209. ences, USA, 74, 4641–4645.
Gouya, L., Puy, H., Lamoril, J., Da Silva, V., Grandchamp, B., Lip, G.Y., McColl, K.E., Goldberg, A. & Moore, M.R. (1991) Smoking
Nordmann, Y. & Deybach, J.C. (1999) Inheritance in erythropoietic and recurrent attacks of acute intermittent porphyria. British Med-
protoporphyria: a common wild-type ferrochelatase allelic variant ical Journal, 302, 507–507.
with low expression accounts for clinical manifestation. Blood, 93, Litman, D.A. & Correia, M.A. (1985) Elevated brain tryptophan and
2105–2110. enhanced 5-hydroxytryptamine turnover in acute hepatic heme
Granick, J.L., Sassa, S. & Kappas, A. (1978) Some biochemical and deficiency: clinical implications. Journal of Pharmacology and
clinical aspects of lead intoxication. In: Advances in Clinical Chem- Experimental Therpeutics, 232, 337–345.
istry (ed. by O. Bodansky and A.L. Latner), p. 287. Academic Press, Martasek, P. (1998) Hereditary coproporphyria. Seminars in Liver
New York. Disease, 18, 25–32.
Grossman, M.E., Bickers, D.R., Poh-Fitzpatrick, M.B., DeLeo, V.A. & Mathews-Roth, M.M., Pathak, M.A., Fitzpatrick, T.B., Harber, L.H. &
Harber, L.C. (1979) Porphyria cutanea tarda: clinical features and Kass, E.H. (1977) Beta carotene therapy for erythropoietic proto-
laboratory findings in forty patients. American Journal of Medicine, porphyria and other photosensitivity diseases. Archives of Derma-
67, 277–286. tology, 113, 1229–1232.
Guarini, L., Piomelli, S. & Poh-Fitzpatrick, M.B. (1994) Hydroxyurea McManus, J.F., Begley, C.G., Sassa, S. & Ratnaike, S. (1996) Five new
in congenital erythropoietic porphyria. New England Journal of mutations in the uroporphyrinogen decarboxylase gene identified in
Medicine, 330, 1091–1092. families with cutaneous porphyria. Blood, 88, 3589–3600.
Handschin, C., Lin, J., Rhee, J., Peyer, A.K., Chin, S., Wu, P.H., Meyer, Meguro, K., Fujita, H., Ishida, N., Akagi, R., Kurihara, T., Galbraith,
U.A. & Spiegelman, B.M. (2005) Nutritional regulation of hepatic R.A., Kappas, A., Zabriskie, J.B., Toback, A.C., Harber, L.C. & Sassa,
heme biosynthesis and porphyria through PGC-1alpha. Cell, 122, S. (1994) Molecular defects of uroporphyrinogen decarboxylase in a
505–515. patient with mild hepatoerythropoietic porphyria. Journal of
Hassoun, A., Verstraeten, L., Mercelis, R. & Martin, J.-J. (1989) Bio- Investigative Dermatology, 102, 681–685.
chemical diagnosis of an hereditary aminolaevulinate dehydratase Meissner, P.N., Dailey, T.A., Hift, R.J., Ziman, M., Corrigall, A.V.,
deficiency in a 63-year-old man. Journal of Clinical Chemistry and Roberts, A.G., Meissner, D.M., Kirsch, R.E. & Dailey, H.A. (1996) A
Clinical Biochemistry, 27, 781–786. R59W mutation in human protoporphyrinogen oxidase results in
Hift, R.J., Meissner, D. & Meissner, P.N. (2004) A systematic study of decreased enzyme activity and is prevalent in South Africans with
the clinical and biochemical expression of variegate porphyria in a variegate porphyria. Nature, 13, 95–97.
large South African family. British Journal of Dermatology, 151, Mendez, M., Sorkin, L., Rossetti, M.V., Astrin, K.H., del C. Batlle,
465–471. A.M., Parera, V.E., Aizencang, G. & Desnick, R.J. (1998) Familial
Ippen, H. (1977) Treatment of porphyria cutanea tarda by phlebot- porphyria cutanea tarda: characterization of seven novel
omy. Seminars in Hematology, 14, 253–259. uroporphyrinogen decarboxylase mutations and frequency of

ª 2006 The Author


Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292 291
Review

common hemochromatosis alleles. American Journal of Human Seligsohn, K. Kaushansky & J.T. Prchal), pp. 803–822. McGraw-Hill,
Genetics, 63, 1363–1375. Inc., New York.
Meyer, U.A., Schuurmans, M.M. & Lindberg, R.L. (1998) Acute por- Sassa, S. (2006b) Porphyrias. In: Clinical Hematology (ed. by N.S.
phyrias: pathogenesis of neurological manifestations. Seminars in Young, S.L. Gerson & K.A. High), pp. 706–720. Mosby Elsevier,
Liver Disease, 18, 43–52. Philadelphia, PA.
Mueller, W.E. & Snyder, S.H. (1977) d-Aminolevulinic acid: influences Sassa, S. & Kappas, A. (1983) Hereditary tyrosinemia and the heme
on synergistic GABA receptor binding may explain CNS symptoms biosynthetic pathway. Profound inhibition of d-aminolevulinic acid
of porphyria. Annals of Neurology, 2, 340–342. dehydratase activity by succinylacetone. Journal of Clinical Investi-
Mustajoki, P., Tenhunen, R., Pierach, C. & Volin, L. (1989) Heme in gation, 71, 625–634.
the treatment of porphyrias and hematological disorders. Seminars Soonawalla, Z.F., Orug, T., Badminton, M.N., Elder, G.H., Rhodes,
in Hematology, 26, 1–9. J.M., Bramhall, S.R. & Elias, E. (2004) Liver transplantation as a cure
Ostasiewicz, L.T., Huang, J.L., Wang, X., Piomelli, S. & Poh-Fitzpa- for acute intermittent porphyria. Lancet, 363, 705–706.
trick, M.B. (1995) Human protoporphyria: genetic heterogeneity at Stein, J.A. & Tschudy, D.P. (1970) Acute intermittent porphyria: a
the ferrochelatase locus. Photodermatology, Photoimmunology & clinical and biochemical study of 46 patients. Medicine, 49, 1–16.
Photomedicine, 11, 18–21. Tezcan, I., Xu, W., Gurgey, A., Tuncer, M., Cetin, M., Oner, C., Yetgin,
Peces, R., Enriquez de Salamanca, R., Fontanellas, A., Sanchez, A., de la S., Ersoy, F., Aizencang, G., Astrin, K.H. & Desnick, R.J. (1998)
Torre, M., Caparros, G., Ferreras, I. & Nieto, J. (1994) Successful Congenital erythropoietic porphyria successfully treated by allo-
treatment of haemodialysis-related porphyria cutanea tarda with geneic bone marrow transplantation. Blood, 92, 4053–4058.
erythropoietin. Nephrology, Dialysis, Transplantation, 9, 433–435. Thomas, C., Ged, C., Nordmann, Y., de Verneuil, H., Pellier, I., Fis-
Piomelli, S., Poh-Fitzpatrick, M.B., Seaman, C., Skolnick, L.M. & cher, A. & Blanche, S. (1996) Correction of congenital erythropoietic
Berdon, W.E. (1986) Complete suppression of the symptoms of porphyria by bone marrow transplantation. Journal of Pediatrics,
congenital erythropoietic porphyria by long-term treatment with 129, 453–456.
high-level transfusions. New England Journal of Medicine, 314, 1029– Thunell, S., Holmberg, L. & Lundgren, J. (1987) Aminolevulinate
1031. dehydratase porphyria in infancy. A clinical and biochemical
Plewinska, M., Thunell, S., Holmberg, L., Wetmur, J.G. & Desnick, R.J. study. Journal of Clinical Chemistry and Clinical Biochemistry, 25,
(1991) d-Aminolevulinate dehydratase deficient porphyria: identi- 5–14.
fication of the molecular lesions in a severely affected homozygote. Tishler, P.V. & Winston, S.H. (1990) Rapid improvement in the che-
American Journal of Human Genetics, 49, 167–174. mical pathology of congenital erythropoietic porphyria with treat-
Poh-Fitzpatrick, M.B. (1984) Diagnosis and treatment of erythropoi- ment with superactivated charcoal. Methods & Findings in
etic protoporphyria. Comprehensive Therapy, 10, 12–17. Experimental & Clinical Pharmacology, 12, 645–648.
Poh-Fitzpatrick, M.B. & Lamola, A.A. (1976) Direct spectro- Todd, D.J. (1998) Molecular genetics of erythropoietic protoporphyria.
fluorometry of diluted erythrocytes and plasma: a rapid diagnostic Photodermatology, Photoimmunology & Photomedicine, 14, 70–73.
method in primary and secondary porphyrinemias. Journal of Valls, V., Ena, J. & Enriquez de Salamanca, R. (1994) Low-dose oral
Laboratory & Clinical Medicine, 87, 362–370. chloroquine in patients with porphyria cutanea tarda and low-
Puy, H., Deybach, J.C., Bogdan, A., Callebert, J., Baumgartner, M., moderate iron overload. Journal of Dermatological Science, 7, 169–
Voisin, P., Nordmann, Y. & Touitou, Y. (1996) Increased delta 175.
aminolevulinic acid and decreased pineal melatonin production. A Waldenstrom, J. (1957) The porphyrias as inborn errors of metabo-
common event in acute porphyria studies in the rat. Journal of lism. American Journal of Medicine, 22, 758–773.
Clinical Investigation, 97, 104–110. Watson, C.J. & Schwartz, S. (1941) A simple test for urinary por-
Roberts, A.G., Whatley, S.D., Morgan, R.R., Worwood, M. & Elder, phobilinogen. Proceedings of Society for Experimental Biology and
G.H. (1997a) Increased frequency of the haemochromatosis Cy- Medicine, 47, 393–394.
s282Tyr mutation in sporadic porphyria cutanea tarda. Lancet, 349, Whatley, S.D., Puy, H., Morgan, R.R., Robreau, A.M., Roberts, A.G.,
321–323. Nordmann, Y., Elder, G.H. & Deybach, J.C. (1999) Variegate por-
Roberts, A.G., Whatley, S.D., Nicklin, S., Worwood, M., Pointon, J.J., phyria in Western Europe: identification of PPOX gene mutations in
Stone, C. & Elder, G.H. (1997b) The frequency of hemochromatosis- 104 families, extent of allelic heterogeneity, and absence of corre-
associated alleles is increased in British patients with sporadic por- lation between phenotype and type of mutation. American Journal of
phyria cutanea tarda. Hepatology, 25, 159–161. Human Genetics, 65, 984–994.
Rocchi, E., Cassanelli, M., Gibertini, P., Pietrangelo, A., Casalgrandi, G. Whetsell, W.O.J., Sassa, S. & Kappas, A. (1984) Porphyrin-heme bio-
& Ventura, E. (1985) Standard or branched-chain amino acid in- synthesis in organotypic cultures of mouse dorsal root ganglia. Ef-
fusions as short-term nutritional support in liver cirrhosis? Journal fects of heme and lead on porphyrin synthesis and peripheral
of Parenteral & Enteral Nutrition, 9, 447–451. myelin. Journal of Clinical Investigation, 74, 600–607.
Rodgers, P.A. & Stevenson, D.K. (1990) Developmental biology of Yaqoob, M., Smyth, J., Ahmad, R., McClelland, P., Fahal, I., Kumar,
heme oxygenase. Clinics in Perinatology, 17, 275–291. K.A., Yu, R. & Verbov, J. (1992) Haemodialysis-related porphyria
Sassa, S. (1996) Diagnosis and therapy of acute intermittent porphyria. cutanea tarda and treatment by recombinant human erythropoietin.
Blood Reviews, 10, 53–58. Nephron, 60, 428–431.
Sassa, S. (2004) Understanding the porphyrias. In: UpToDate (ed. by Zix-Kieffer, I., Langer, B., Eyer, D., Acar, G., Racadot, E., Schlaeder, G.,
B.D. Rose), pp. 1–6. UpToDate, Wellesley, MA. Oberlin, F. & Lutz, P. (1996) Successful cord blood stem cell
Sassa, S. (2006a) The hematologic aspects of porphyria. In: Williams transplantation for congenital erythropoietic porphyria (Gunther’s
Hematology (ed. by M.A. Lichtman, E. Beutler, T.J. Kipps, U. disease). Bone Marrow Transplantation, 18, 217–220.

ª 2006 The Author


292 Journal Compilation ª 2006 Blackwell Publishing Ltd, British Journal of Haematology, 135, 281–292

You might also like