You are on page 1of 90

Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

PART 13: PORPHYRINS

Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism


Jayanta Roy Chowdhury, Allan W. Wolkoff, Namita Roy Chowdhury, Irwin M. Arias

Abstract

1. Bilirubin is an orange pigment derived from the degradation of the heme moiety of hemoproteins,
particularly the hemoglobin of mature circulating erythrocytes.
2. Bilirubin is a potentially toxic waste product that is normally rendered harmless by binding to serum
albumin, conjugation in the liver, and efficient excretion into bile by the liver. Bilirubin is an
antioxidant, and a protective role of bilirubin against oxidant damage has been suggested. On the
other hand, patients with profound unconjugated hyperbilirubinemia are at risk for bilirubin
encephalopathy (kernicterus). Accumulation of bilirubin in plasma and tissues results in jaundice,
which has attracted the attention of patients and clinicians since antiquity.
3. Following formation in the reticuloendothelial system, bilirubin is released into the circulation, where it
avidly binds to serum albumin and is rapidly cleared by the liver. Extraction of bilirubin from the
circulation is a specific hepatic function involving facilitated diffusion. Within the hepatocyte, bilirubin
binds to cytosolic proteins, primarily to glutathione-S-transferases. The water-insoluble bilirubin
molecule is transformed into polar bilirubin monoglucuronide and diglucuronide by the action of
bilirubin-UDP-glucuronosyltransferase (bilirubin-UGT) and is excreted into the bile canaliculus against
a concentration gradient by energy-consuming mechanisms.
4. Inherited disorders of bilirubin metabolism result in hyperbilirubinemia. These include disorders
resulting in predominantly unconjugated hyperbilirubinemia (Crigler-Najjar syndrome types I and II,
and Gilbert syndrome) and those resulting in predominantly conjugated hyperbilirubinemia
(Dubin-Johnson syndrome, Rotor syndrome, and benign recurrent intrahepatic cholestasis).
5. Bilirubin-UGT (protein/gene for UDP-glucuronosyltransferase 1, family member 1A equivalent to
bilirubin-UGT (UGT1A1)) and several additional isoforms of the UGT1A subfamily that mediate the
glucuronidation of other aglycone substrates are expressed from the locus UGT1A, located on
chromosome 2q37. Genetic lesions in any of the five exons that encode UGT1A1 may lead to
complete absence (Crigler-Najjar syndrome type I; MIM 218800) or incomplete deficiency
(Crigler-Najjar syndrome type II) of bilirubin glucuronidation. In contrast, Gilbert syndrome (MIM
143500) is associated with an abnormality of the TATAA box within the promoter region upstream to
exon 1 of UGT1A1 that results in reduced expression of structurally normal UGT1A1. Dubin-Johnson
syndrome (MIM 237500) is caused by a genetic abnormality of bile canalicular multispecific organic
anion transporter, which is involved in the excretion of many non–bile salt organic anions by an
adenosine triphosphate (ATP)-requiring active process. Dubin-Johnson syndrome is associated with
a characteristic accumulation of pigments in the liver and an abnormality of porphyrin metabolism in
which over 80 percent of urinary coproporphyrin is coproporphyrin I, as compared with less than 35
percent in normal individuals. Rotor syndrome (MIM 237450) is primarily a disorder of hepatic storage
and differs from Dubin-Johnson syndrome by the lack of hepatic pigmentation, urinary coproporphyrin
excretion pattern, and hepatic sulfobromophthalein (BSP) metabolism. Dubin-Johnson syndrome is
caused by genetic lesions of the gene MRP2 (also termed cMOAT). Other genetic abnormalities can
cause hyperbilirubinemia, secondary to structural abnormalities of the biliary system or derangement
of specific excretory functions of the bile canaliculus. These include progressive familial intrahepatic

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -1- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

cholestasis, type I (MIM 211600) and benign recurrent intrahepatic cholestasis (MIM 243300), both of
which are associated with mutations of the protein/gene for familial intrahepatic cholestasis-1 (FIC1)
gene on chromosome 18q21. Progressive familial intrahepatic cholestasis type II (MIM 601847) is
characterized by abnormality of bile salt excretion and is associated with mutations of the gene
SPGP, which is located on chromosome 2q24. A third type of progressive familial intrahepatic
cholestasis (MIM 602347) involves mutations of the MDR3 gene, the products of which are needed
for phospolipid excretion in bile. Several heritable developmental disorders of the biliary system have
been described. Of these, Alagille syndrome (MIM 118450) has been found to be caused by lesions of
JAG1, a gene located on chromosome 20p12.
6. Crigler-Najjar syndrome types I and II have an autosomal-recessive pattern of inheritance. Patients
with Gilbert syndrome are homozygous for the specific promoter abnormality, but all subjects homozygous
for this genotype do not exhibit the clinical picture of Gilbert syndrome. Studies of urinary coproporphyrin
excretion reveal autosomal-recessive patterns of inheritance for Dubin-Johnson and Rotor syndromes.
7. Several animal models of inherited disorders of bilirubin metabolism are important in understanding
the pathophysiology of their human counterparts. These models include the Gunn rat (Crigler-Najjar
syndrome type I); the Bolivian population of squirrel monkeys (Gilbert syndrome); and mutant albino rats
with organic anion excretion defect (TR −\−), golden lion tamarin monkeys, and mutant Corriedale
sheep (Dubin-Johnson syndrome).

Bilirubin is an orange pigment derived from the degradation of heme proteins, particularly the hemoglobin
of mature circulating erythrocytes and hepatic hemoproteins. Bilirubin is a potentially toxic waste product
that is generally harmless because of binding to serum albumin. However, patients with profound
unconjugated hyperbilirubinemia are at risk for bilirubin encephalopathy (kernicterus).

Studies of bilirubin chemistry, synthesis, transport, metabolism, distribution, and excretion have attracted
the attention of generations of chemists, biologists, and clinical investigators. Because bilirubin is an
organic anion of limited aqueous solubility, it has proved to be a model for the study of the transport,
metabolism, and excretion of other biologically important organic anions. Defects in bilirubin formation or
disposal are usually manifested by hyperbilirubinemia and jaundice. A number of inherited disorders
affecting these pathways have been described in both humans and animals. Study of these disorders has
provided important information regarding normal and abnormal metabolic pathways.

Crigler-Najjar syndrome type I, a potentially lethal disorder, requires liver transplantation as a definitive
treatment, and transplantation of isolated hepatocytes also has been used. Development of strategies for
noninvasive therapy for this condition remains a therapeutic challenge and continues to stimulate
research.

BILIRUBIN

Formation of Bilirubin
Bilirubin is exclusively derived from heme. In humans, 250 to 400 mg of bilirubin is formed daily by the
breakdown of hemoglobin, other hemoproteins, and free heme. 1 Approximately 80 percent is derived from
the hemoglobin of senescent erythrocytes. 2 After injection of radiolabeled porphyrin precursors (glycine or
δ-aminolevulinic acid) in humans or rats, radioactivity is incorporated into bile pigments in two peaks 3
(Fig. 125-1). The first peak (early-labeled peak of bilirubin, ELB) appears within 3 days and contains an
initial component and a slow later phase. The initial component comprises two thirds of the ELB in
humans and is largely derived from hepatic hemoproteins such as cytochromes, catalase, peroxidase,
and tryptophan pyrrolase. 4 The labeled bilirubin that appears in bile within 15 min after administration of
the precursor may be derived from a rapidly turning over pool of free heme in the cytosol of hepatocytes

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -2- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

that is degraded without incorporation into hemoproteins. 5 Myoglobin has a relatively long half-life and is
an unlikely source. Induction of hepatic cytochrome P450 enhances the ELB. 6 The slower phase of the
ELB is derived from both erythroid and nonerythroid sources and is enhanced in conditions associated
with ineffective erythropoiesis, such as congenital dyserythropoietic anemias, megaloblastic anemias,
iron-deficiency anemia, and lead poisoning. 4 The ELB is also increased in erythropoietic porphyria 3 but
not in porphyria cutanea tarda 7 or acute intermittent porphyria. 8 The erythroid phase is increased in
accelerated erythropoiesis, probably because of intramedullary destruction of normoblasts, destruction of
reticulocytes in the peripheral circulation, and injury to reticulocytes during maturation. 9 δ-Aminolevulinic
acid is preferentially incorporated into hepatic hemoproteins. 10 A late-labeled peak appears at
approximately 50 days in rats and by 110 days in humans and is derived from the hemoglobin of
senescent erythrocytes.

Labeling of plasma bilirubin in UDP-glucuronosyltransferase–deficient rats (Gunn strain) after the injection
of [14C]glycine. The early (0 to 3 days) peak has an initial “sharp” and a slower component. [Reprinted
with permission from Robinson SH: In Stohlman F Jr (ed): Hemopoietic Cellular Proliferation. New York:
Grune & Stratton, 1970, p 180.]

In the liver, heme derived from exogenously administered hemoglobin is quantitatively converted to
bilirubin. 11 A portion of heme associated with hepatic hemoproteins may not be converted to bilirubin. 12
This suggests that exogenous heme and hepatocellular heme may be processed differently by the liver.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -3- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Mechanism of Opening of the Heme Ring.


Ferroprotoporphyrin IX is the heme prosthetic group (Fig. 125-2) in vertebrate hemoproteins. The
porphyrin ring is selectively cleaved at the α-methene bridge. The first step is catalyzed by microsomal
heme oxygenase and requires an electrophilic attack at Fe(II) by a reducing agent, such as NADPH and
oxygen. This reaction results in formation of α-oxyheme (Fig. 125-2). 13 Heme oxygenase activity is
highest in the spleen, which is involved in sequestration of senescent erythrocytes, and is enhanced in
hemolytic states. Within the liver, hepatocytes and Kupffer cells have heme oxygenase activity. In Kupffer
cells, the enzyme activity is comparable with that in the spleen. 14 Heme stimulates the synthesis of heme
oxygenase, 15 which enhances heme degradation, 16 suggesting that heme oxygenase is rate limiting in
heme oxidation and bilirubin formation. 16 However, evidence also has been presented supporting a
contrasting view that heme oxygenase is present in excess as compared with its substrate. 17 Heme
oxygenase has been purified from microsomal fractions of pig 18 and bovine 19 spleen and rat liver. 20 The
binding of purified heme oxygenase to heme requires heme iron and the propionic acid substituents in the
C6 and C7 positions. Protoporphyrins that contain other metals, such as tin, bind with even greater
affinity 19 but are not degraded by heme oxygenase. Thus, noniron hemes may serve as dead-end
inhibitors of heme oxygenase. 19

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -4- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Mechanism of heme ring opening and subsequent reduction of biliverdin to bilirubin.

The second step in opening the heme ring involves oxidation by molecular oxygen and probably occurs
nonenzymatically. 21 Carbons at the angular positions of the porphyrin ring neighboring the α-methene
bridge are oxidized, and CO is eliminated. During this step, two oxygen atoms, derived from two different
oxygen molecules, are added. 22 These oxygen atoms appear as the lactam oxygens of biliverdin and
bilirubin. Release of iron occurs after addition of electrons, suggesting that conversion of ferric to ferrous
iron is required. 23 The resulting green pigment is biliverdin.

Conversion of Biliverdin to Bilirubin.


Biliverdin is readily excreted by the liver and is the major bile pigment in many amphibian, avian, and fish
species. However, in most mammals, biliverdin is converted to bilirubin, which requires several
energy-consuming metabolic steps for biliary excretion. The physiologic benefits of conversion of biliverdin
to bilirubin are not clear. Bilirubin is less polar than is biliverdin and crosses the placental membranes
more readily than does biliverdin. 24 However, some placentate animals, such as nutria and rabbits,
excrete biliverdin as the main bile pigment, 25 whereas many nonplacentate vertebrates, such as some
species of fish, excrete predominantly bilirubin in bile. 26 Bilirubin is an antioxidant and is thought to play
an important antioxidant defense for the body. 27–29 The antioxidant activity of bilirubin may be particularly
important during the neonatal period, when concentrations of other antioxidants are low in body fluids.

Conversion of biliverdin to bilirubin is catalyzed by a cytosolic enzyme, biliverdin reductase, which requires
NADH or NADPH for activity. 30, 31 Guinea pig liver biliverdin reductase is a 70-kDa protein. 32 Three
interconverting molecular forms of biliverdin reductase have been described in rat liver. 33

Because of the specificity of heme oxygenase for the α-carbon bridge, the most abundant bile pigment is
bilirubin IXα, and only minute amounts of non-α isomers (Fig. 125-3) have been detected in human and
animal bile. 34

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -5- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Nonenzymatic cleavage of heme in vitro results in the formation of four isomeric forms of biliverdin owing
to the nonequivalence of the four methene bridge positions (α, β, γ, and δ). P, CH2CH2COOH.

Quantification of Bilirubin Production.


Bilirubin production reflects turnover of biologically important hemoproteins. At a steady-state condition of
blood hemoglobin levels, the rate of bilirubin production approximates the rate of heme synthesis. For
these reasons, the quantification of bilirubin production is important in clinical and physiologic
investigations. 35 Normally, bilirubin is almost quantitatively excreted in bile; therefore, bilirubin production
can be quantified in biliary excretion in animals, but this is not practical in humans. Bilirubin is converted to
urobilinogen by bacteria in the gastrointestinal tract, and fecal urobilinogen excretion approximates daily
bilirubin production, 36 although conversion to urobilinogen may not be quantitative.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -6- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

In humans, bilirubin production is conveniently quantified from the turnover of radioisotopically labeled
bilirubin. Radiolabeled bilirubin bound to albumin is injected intravenously, blood samples are collected at
frequent intervals, and plasma bilirubin concentration and radioactivity are measured. 37 Plasma bilirubin
clearance (the fraction of plasma from which bilirubin is irreversibly extracted) is proportional to the
reciprocal of the area under the radiobilirubin disappearance curve. 38 Bilirubin removal is quantified as
the product of plasma bilirubin concentration and clearance. When plasma bilirubin concentrations remain
constant, removal of bilirubin equals the amount of newly synthesized bilirubin entering the plasma pool.
This method does not take into account a small portion of bilirubin that is produced in the liver and
excreted directly into bile without appearing in the circulation and, therefore, slightly underestimates
bilirubin production.

Bilirubin formation also can be quantified from carbon monoxide production. The subject is placed in a
closed rebreathing system to prevent CO excretion. CO production is calculated from the CO
concentration in the breathing chamber or from an increment in blood carboxyhemoglobin saturation. 39
This method assumes that body CO stores rapidly equilibrate, blood carboxyhemoglobin reflects total
body CO, and metabolism of CO is insignificant compared with its rate of production. However, under
certain circumstances, such as anoxia, assumption of a steady equilibrium of body stores of CO with
blood carboxyhemoglobin may not be correct. 40 CO production exceeds plasma bilirubin turnover by 12
to 18 percent. This discrepancy is partly due to a small portion of bilirubin produced in the liver and
excreted into bile without appearing in serum. A portion of CO in expired air may be produced from
nonheme sources, such as halogenated methane 41 and polyphenolic compounds, including
catecholamines. 42 A small fraction of the CO may be formed by intestinal bacteria. 43

Pharmacologic Inhibition of Bilirubin Production.


Administration of nonmetabolized dead-end inhibitors of heme oxygenase, such as tin-protoporphyrin or
tin-mesoporphyrin, results in marked inhibition of the enzyme activity in various organs. 19 A single dose of
tin-mesoporphyrin administered in neonates, shortly after birth, resulted in an average of 76 percent
reduction of serum bilirubin levels and abolished the need for phototherapy. 44, 45

Chemistry of Bilirubin
The systemic name given to bilirubin IXα is
1,8-dioxo-1,3,6,7-tetramethyl-2,8-divinylbiladiene-a,c-dipropionic acid. 4, 5 The gross chemical structure
(Fig. 125-3) assigned to bilirubin by Fischer and Plieninger 46 has been confirmed by x-ray diffraction
analysis (Fig. 125-4). 47 The bonds between pyrrolenone rings A and B (C4 to C5) and C and D (C15 to
C16) are in the Z or trans-configuration. The oxygen attached to the outer pyrrolenone ring is in a lactam
rather than lactim configuration. Titration of bilirubin in aqueous solutions suggests a pK value of 7.0 to
8.0. 48 Because bilirubin tends to form insoluble aggregates below pH 8.0, determination of pK by titration
of aqueous solutions of bilirubin may be misleading. 49 Studies using 13 C NMR spectra and potentiometric
and spectrophotometric titrations in aqueous solutions indicate that bilirubin has four acidic groups. The
pK value of the two carboxyl groups is 4.4 and that of the two lactam groups is 13.0. 49

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -7- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

X-ray crystallographic structure of bilirubin showing a ridge-tile configuration caused by internal hydrogen
bonding of the propionic acid carboxyls to the amino groups and the lactam oxygen of the pyrrolenone
rings of the opposite half of the molecule. The bonds between pyrrolenone rings A, B, C, and D are in the
Z (trans) configuration.

Physical Conformation and Solubility of Bilirubin IXα.


Crystallized bilirubin IXα-ZZ with two protonated carboxyl groups is virtually insoluble in water. At acidic,
neutral, or mildly alkaline pH, bilirubin can be extracted from aqueous solutions into water-immiscible
solvents such as chloroform, ethyl acetate, or methylethyl ketone. After intravenous injection, bilirubin
stains phospholipid membranes of brain and those covering adipose tissues. However, bilirubin is not truly
lipophilic. Determination of solubility in progressively nonpolar solvents indicates that bilirubin is readily
soluble in polar solvents, provided the intramolecular hydrogen bonds can be interrupted. 50 Bilirubin and
polar ligands, such as sulfonamides, share a binding site within a polar domain of albumin. 49, 51 Thus,
despite its insolubility in water at physiologic pH, bilirubin should be considered a relatively polar
substance, and its mechanism of toxicity may differ from that of truly lipid-soluble toxins, such as DDT. 51
Although both bilirubin and biliverdin have two propionic acid side chains, bilirubin IXα is less polar than is

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -8- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

biliverdin IXα at physiologic pH (Fig. 125-2). An explanation for this was suggested initially by Fog and
Jellum 52 and Kuenzle et al., 53 who proposed that bilirubin IXα may be internally stabilized by hydrogen
bonding between the carboxyl and the two external pyrrolenone rings (Fig. 125-5). X-ray diffraction
studies of crystalline bilirubin confirm hydrogen bonding between each propionic acid side chain and the
pyrrolic and lactam sites in the opposite half of the molecule. 47 The molecule takes the form of a ridge tile
in which the two dipyrrolic halves of the molecule lie in two different planes with an interplanar angle of 98
to 100 degrees (Fig. 125-4). The integrity of the hydrogen bonded structure requires the interpyrrolic
bridges at the 5 and 15 position of bilirubin to be in trans- or Z configuration. In nonpolar solvents, the
structure of unconjugated bilirubin oscillates between that shown in Fig. 125-5 and its mirror image. 54 A
similar conformation has been proposed for bilirubin dianions in aqueous solutions. 55 The hydrogen
bonded structure of bilirubin may explain many of its physicochemical properties. The two carboxylic
groups, all four NH groups, and the two lactam oxygens are engaged by hydrogen bonding, making the
molecule insoluble in water. Addition of methanol, ethanol, or 6 M urea disrupts the hydrogen bonds,
thereby making bilirubin water soluble and more labile. 56 The central methene bridge becomes accessible
to diazo reagents after disruption of hydrogen bonds, making bilirubin readily reactive to these reagents.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. -9- Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Ionic species of bilirubin. A. Internally hydrogen bonded form. B. Bilirubin acid with hydrogen bonds
disrupted. C. Bilirubin dianion.

Ultraviolet/Visible Absorption Spectra.


The position of the main absorption band (λ max ) of bilirubin depends on the bile pigment and the solvent.
Unconjugated bilirubin IXα has a λ max of 450 to 474 nm in most organic solvents (Fig. 125-2) and an
extinction coefficient (E max ) of 48.0 to 63.4/mM. In alkaline aqueous solutions, there is a 10- to 30-nm shift

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 10 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

of the λ max toward shorter wavelengths (hypsochromic shift) and a weaker absorption band at 280 to
300 nm.

Helical Conformation of Bilirubin and Biliverdin.


Circular dichroism spectroscopy shows that bilirubin preferentially adopts a plus-helicity when bound to
human serum albumin, whereas biliverdin prefers a minus helicity. 57

Fluorescence.
Pure bilirubin does not fluoresce. When it is dissolved in detergent, albumin solution, or alkaline methanol,
an intense fluorescence is observed at 510 to 530 nm. 58 Determination of fluorescence of bilirubin can be
used for rapid quantification of blood bilirubin concentrations and the unsaturated bilirubin binding capacity
of albumin.

Geometric Isomerization and Cyclization.


As mentioned earlier (Fig. 125-3), the 5 and 15 bridges of bilirubin are in a trans- or Z configuration.
Exposure of circulating bilirubin to light changes the configuration of one or both of the interpyrrolic
bridges at the 5 and 15 positions to an E or cis-configuration. The resulting ZE, EZ, or EE isomers lack
one or more internal hydrogen bonds, are more polar than is bilirubin IXα-ZZ, and can be excreted in bile
without conjugation. 59 The vinyl substituent in the endovinyl half of bilirubin IXα-EZ is subsequently
cyclized with the methyl substituent on the internal pyrrole ring, forming the structural isomer
E-cyclobilirubin. 60 Although cyclization of bilirubin occurs at a slower rate than formation of configurational
isomers, because of the relative stability of cyclobilirubin, this form may be quantitatively more important in
phototherapy of neonatal jaundice. 60

Photooxidation and Degradation.


Whether in aqueous solution or bound to proteins or lipids, bilirubin undergoes gradual bleaching in light
and oxygen. 61 Bleaching results in formation of colorless fragments, chiefly maleimides and
propentdyopent adducts, owing to a self-sensitized reaction involving singlet oxygen. A small amount of
biliverdin is also formed by mechanisms that are not established. 61

Dipyrrolic Scrambling.
When bilirubin IXα is irradiated in deoxygenated aqueous solution, free radical disproportionation results
in formation of bilirubin IIIα and bilirubin XIIIα (Fig. 125-6), which are nonphysiologic symmetrical isomers
of bilirubin. 61 The reaction is faster in the presence of oxygen and is catalyzed by acid 61 and inhibited by
ascorbic acid.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 11 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Isomerization of unconjugated and conjugated bilirubin. Upper panel. Geometric isomerization: The bond
between pyrrolenone rings A and B or C and D can change into an E (cis) configuration, as shown here
on the left half of the bilirubin molecule, resulting in the EZ, ZE, or EE isomers. E configuration of the bond
between the pyrrolenone rings interferes with hydrogen bonding and renders the molecule relatively polar.
Middle panel. Nonenzyma...

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 12 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Toxicity of Bilirubin
A toxic effect of bilirubin on the brain of neonates has been known for at least five centuries. 62 Yellow
discoloration of basal ganglia in babies with intense jaundice was described in 1847, and the term
kernicterus was coined to describe these changes in 1903. 63, 64

Biochemical Mechanisms of Bilirubin Toxicity.


Bilirubin inhibits DNA synthesis in a mouse neuroblastoma cell line. 65 Bilirubin also may uncouple
oxidative phosphorylation and inhibit adenosine triphosphatase (ATPase) activity of brain mitochondria. 66
Studies in vitro reveal that bilirubin inhibits hydrolytic enzymes, 67 dehydrogenases, 68 and enzymes
involved in electron transport. 69 In mutant rats (Gunn strain) with congenital nonhemolytic
hyperbilirubinemia, bilirubin inhibited RNA synthesis, protein synthesis, and carbohydrate metabolism in
brain and inhibited protein synthesis in liver. 70, 71 Bilirubin also decreased respiration of isolated brain
mitochondria, uncoupled oxidative phosphorylation, inhibited ATPase activity, and induced swelling in the
brain. 72–74 Bilirubin has been shown to inhibit protein kinase-mediated phosphorylation of neural
proteins. 75, 76 In a cell-free system, bilirubin irreversibly inhibited Ca 2+ -activated, phospholipid-dependent
protein kinase (protein kinase C) activity and cyclic adenosine monophosphate (cAMP)-dependent protein
kinase activity. 77 All toxic effects of bilirubin are reduced or reversed by albumin in vivo and in vitro.
Increased plasma concentrations of bilirubin increase the risk of bilirubin encephalopathy in newborn
babies. A serum concentration of 20 mg/dl is often quoted as the highest limit of safety, although
kernicterus can occur at lower serum bilirubin concentrations. 78 Individual differences in the susceptibility
to bilirubin encephalopathy is underscored by the finding that Gunn rats bred against different normal
background strains have similar serum bilirubin levels but markedly different incidences of kernicterus and
mortality. 79 Serum albumin concentrations, pH, and substances that compete for albumin binding are
important in the pathogenesis of bilirubin encephalopathy. 80

Clinical Features of Bilirubin Encephalopathy.


Kernicterus occurs in infants with severe unconjugated hyperbilirubinemia and in young adults with high
serum unconjugated bilirubin levels due to severe inherited deficiency of bilirubin-UGT activity.
Kernicterus usually presents between the third and sixth days of life, with poor feeding and feeble suck
reflex, high-pitched cry, hypertonia or hypotonia, reflex opisthotonus in response to a startling stimulus,
convulsion, incomplete Moro reflex, and instability of thermal regulation (hypothermia or hyperthermia). 78
This may progress to lethargy, atonia, and death. Long-term sequelae include delay in motor
development, chorioathetosis, asymmetric spasticity, sensorineural hearing loss, paralysis of upward
gaze, dental dysplasia, cognitive dysfunction, and mental retardation. The cochlear nucleus is particularly
sensitive to bilirubin-induced damage. The affected areas in the cochlear nucleus and superior olive are
innervated by large axosomatic end bulbs or caliceal endings. Cells receiving synaptic input from end
bulbs or calices appear to be early targets of bilirubin toxicity in the auditory system. 81 Brainstem auditory
evoked responses correlate well with plasma free bilirubin levels in humans 82 and with morphologic
changes in the cochlear nucleus in Gunn rat pups. 83 There is prolongation of interval between the
auditory stimulus and appearance of wave I (latency), indicating delayed auditory nerve conduction. The
intervals between wave I and wave III (reflecting the superior olive) and wave I and wave V (reflecting
inferior colliculus) are also prolonged (interpeak latency). This test may be made more sensitive by
recording binaural difference waves obtained by subtracting the sum of two monaural brainstem auditory
evoked potentials from a binaural brainstem auditory evoked potential. 84 Moderately high serum
unconjugated bilirubin levels may result in a higher incidence of impaired neurologic or intellectual
performance in later life. 85, 86 In some children with Crigler-Najjar syndrome type I there may be late
clinical presentation of bilirubin encephalopathy with cerebellar symptoms as the presenting feature. 87

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 13 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Magnetic resonance imaging of the brain provides a sensitive means of evaluating kernicterus. 86–90
Abnormally high intensity signals are observed over the globus pallidus, particularly over the
posteromedial border. In many cases, increased intensity is also seen over the thalamus, internal capsule,
and hippocampi. These magnetic resonance imaging abnormalities are also found in several other types
of metabolic encephalopathy and must be interpreted in the context of the clinical presentation. Bilirubin
staining of the hippocampus, basal ganglia, and nuclei of the cerebellum and brainstem is observed in
infants who die from acute kernicterus. 91 In infants who die within 72 h after the onset of kernicterus,
there may be no cellular damage of the brain seen by light microscopy. Early histologic changes occur
after this period and include cytoplasmic degeneration, loss of Nissl substance, and fine vacuolation and
swelling of nuclear chromatin. 92 Evidence of cell death may be present. In children who die in the chronic
stage of the disorder, bilirubin staining is not found in the brain, 91 but focal necrosis of neurons and glia
are found. Gliosis of the affected areas occurs in later cases. 92

Role of the Blood–Brain Barrier.


Tight junctions between capillary endothelial cells and foot processes of astroglial cells restrict the
exchange of water-soluble substances and proteins between blood and brain. 93 In addition, specific
transport processes for ions, water, and nutrients from plasma to brain may provide a functional
blood–brain barrier. Conventionally, the immaturity of the blood–brain barrier in neonates has been
thought to contribute to kernicterus. However, a more rapid passage of labeled markers 94 or lipophilic
substances 95 into the immature brain has been difficult to confirm, and there is little evidence to support
the concept of immaturity of the blood–brain barrier in the neonate. Moreover, opening of the blood–brain
barrier is expected to permit both bilirubin and albumin to enter the brain. Because albumin binding
neutralizes the toxic effects of bilirubin, this should not result in increased bilirubin toxicity. Current
evidence indicates that the non–albumin-bound (free) fraction of bilirubin enters the brain, and such entry
is independent of the intactness of the blood–brain barrier. A special proclivity of the neonatal brain cells
to bind bilirubin may facilitate its retention in the brain of newborns with severe unconjugated
hyperbilirubinemia. Hyperosmolarity-associated shrinkage of capillary endothelial cells results in
temporary and reversible opening of the tight junctions. When the blood–brain barrier is opened in
newborn rats by infusion of hypertonic urea 96 or arabinose, 93 intravenously administered
albumin–bilirubin complex rapidly enters the brain. Following reversal of opening of the blood–brain
barrier, bilirubin is rapidly cleared from the brain in parallel with clearance from the serum, suggesting that
bilirubin is cleared by diffusion or transport back into the general circulation. 97 However, damaged and
edematous brain may bind bilirubin, 98 may be unable to clear it rapidly, and may be more vulnerable to
bilirubin toxicity for that reason.

Binding of Bilirubin to Albumin


In the circulation, bilirubin is tightly but reversibly bound to albumin. Albumin binding plays a critical role in
the disposition of bilirubin by the body. It keeps bilirubin in solution and transports the pigment from its
sites of production, primarily the spleen and the bone marrow, to its site of excretion, the liver. The tight
binding of unconjugated bilirubin to albumin prevents its excretion by the kidney, except during
albuminuria. Conjugated bilirubin is bound much less tightly to albumin, and the relatively larger unbound
fraction undergoes glomerular filtration and is excreted in the urine. During disease states that are
associated with prolonged increase in plasma-conjugated bilirubin levels, a fraction of the pigment
becomes irreversibly bound to albumin. 99 This fraction, termed δ-bilirubin, is not excreted in the bile or
urine and disappears slowly, reflecting the long half-life of albumin.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 14 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Albumin-binding protects against all known toxic effects of bilirubin, and a small unbound fraction of
bilirubin is thought to be responsible for its toxicity. Coadministration of equimolar amounts of albumin
protects against otherwise lethal effects of unconjugated bilirubin following intravenous injection in
puppies. 100

In normal plasma, bilirubin is bound to a primary binding site on albumin, almost exclusively as the
dianion. Normal molar concentration of albumin (500–700 µM) exceeds that of bilirubin (upper limit
17 µM). However, during neonatal jaundice, in patients with Crigler-Najjar syndrome, and occasionally in
acquired liver diseases, the molar ratio of unconjugated bilirubin to albumin can exceed 1. The molar
excess of bilirubin can be accentuated by reduction of serum albumin levels due to inflammatory states,
chronic malnutrition, or liver diseases. In these circumstances, bilirubin also can bind weakly to a second,
third, and a fourth site. However, such binding is much weaker, so the unbound fraction of bilirubin
increases sharply. Use of sulfonamides in newborn babies enhances bilirubin encephalopathy, 101 as a
result of dissociation of bilirubin by sulfonamide from its binding to albumin. 102 Infusion of albumin
increases the plasma bilirubin concentration because of transfer of bilirubin from tissues to plasma. 103
Because of the clinical importance of estimation of the unbound fraction of unconjugated bilirubin, the
binding of bilirubin to albumin has been evaluated by separating bound from free bilirubin by ultrafiltration,
ultracentrifugation, gel chromatography, affinity chromatography on albumin agarose polymers, dialysis,
and electrophoresis. Unbound bilirubin is rapidly destroyed by treatment with H 2 O 2 and horseradish
peroxidase, as compared with bound bilirubin. Binding of bilirubin to albumin induces bilirubin
fluorescence, circular dichroism, quenching of protein fluorescence, and a shift in the absorbance spectra.
In most studies, the primary binding constant at physiologic pH and temperature is slightly below 10 8 M −1 .
The binding constant for the secondary site is believed to be lower by one order of magnitude. 49, 104
Enzymatic hydrolysis and analysis of albumin covalently bound to bilirubin indicate that bilirubin binds to
lysine 240 in human albumin and to lysine 238 in bovine serum albumin. 105 Binding of other ligands to
albumin plays a major role in determining bilirubin binding capacity. The other ligand may bind at the
same site as does bilirubin, resulting in competitive displacement, or it may bind noncompetitively at a
different site. Noncompetitive binding may not affect bilirubin binding or may produce conformational
changes that enhance (cooperative binding) or decrease (anticooperative) bilirubin binding. Sulfonamides,
antiinflammatory drugs, and contrast media used for cholangiography displace bilirubin competitively from
albumin and increase the risk of kernicterus in jaundiced newborn babies. 106 Some benzodiazepine drugs
and long-chain fatty acids in low concentration bind to human albumin without affecting bilirubin
binding. 107, 108 Albumin binding of medium-chain fatty acids, such as laureate and myristate, increase the
binding constant for bilirubin. 109 Short-chain fatty acids bind to albumin anticooperatively with bilirubin. 110
When large amounts of fatty acid bind to albumin, major conformational changes occur that generally
decrease the binding of other ligands, including bilirubin. Acidosis increases the risk of brain damage in
neonatal jaundice 111, 112 but does not influence bilirubin binding to the primary site of albumin. The
increased risk of kernicterus may result from enhanced transport of bilirubin from plasma to selected
areas of the central nervous system. 49 Because of the influence of many metabolites and drugs on
albumin binding of bilirubin and on its transfer from plasma to the central nervous system, measurement
of plasma bilirubin concentration does not accurately estimate the risk of brain damage from unconjugated
bilirubin. It is generally believed, although it has not been verified, that unbound bilirubin is transferred
from plasma to the central nervous system. 102 Efforts have been made to quantify unbound bilirubin in
serum by gel chromatography, 113 peroxidase treatment, 114 electrophoresis on cellulose acetate, 115 and
fluorimetry of serum with or without detergent treatment. 116

Free bilirubin concentration is determined from the equilibrium equation

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 15 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

where [F] is the free bilirubin concentration, [B] is albumin-bound bilirubin concentration, [RA] is the
concentration of reserve bilirubin binding sites on albumin, and K is the association constant for bilirubin.
Equilibrium between free and bound bilirubin is assumed, and binding of bilirubin to tissues and secondary
binding sites on albumin are ignored. The numerical values for binding constants, as determined from
experiments with pure albumin and bilirubin, are assumed to be valid in serum. These assumptions may
not be valid with icteric serum, so it is not possible to calculate reliably the concentration of unbound
bilirubin. The alternative approach is to determine the amount of unoccupied bilirubin binding sites on
albumin. Titration of serum with bilirubin or a dye that binds to albumin has been used to estimate
unoccupied bilirubin binding sites.

Binding to secondary binding sites begins before primary sites are saturated, and some dyes bind at sites
other than the bilirubin site. Binding of bilirubin to erythrocytes depends on the albumin:bilirubin ratio in
serum and indirectly reflects reserve bilirubin binding sites on albumin. 117 Competitive binding by a
14 C-labeled ligand (monoacetyl-4,4′-diaminodiphenyl sulfone) 118 or a spin-labeled ligand [1-N-(2,2,6,6,

tetramethyl-1-oxyl-4-piperidinyl)5-N-(1-aspartate)-2,4,-dinitrobenzene] 119 has been used to determine


reserve binding capacity. A fluorimetric method has been described for determination of bound albumin
and reserve bilirubin binding capacity. 116 Despite inaccuracies, several empirical tests for determination
of reserve bilirubin binding capacity of serum albumin correlate clinically with brain damage 120 and may
be useful in assessing the risk of bilirubin toxicity.

Uptake of Bilirubin by the Liver


Although tightly bound to albumin, bilirubin is rapidly removed from the circulation by the liver (Fig. 125-7).
Kinetic studies of bilirubin uptake in isolated perfused livers of dogs 121 and rats 122 and in intact rats in
vivo 123 reveal that the process is saturable. Following an intravenous loading dose of bilirubin, the plasma
disappearance of a subsequent tracer dose of [ 3 H]-bilirubin is enhanced, 123 suggesting a
carrier-mediated facilitated diffusion. Countertransport of bilirubin (i.e., efflux of radiolabeled ligand from
liver after subsequent infusion with unlabeled ligand) has been claimed, 123 but the data also could be
interpreted to represent efflux of ligand from intracellular binding sites. Mutual competition for hepatic
uptake in vivo has been described with respect to bilirubin and other organic anions, such as indocyanine
green (ICG), 123, 124 sulfobromophthalein (BSP), 123, 124 and conjugated bilirubin. 124 Bile acids do not
compete with these compounds for hepatic uptake. 123, 125

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 16 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Summary of hepatic metabolism of bilirubin (B). Bilirubin is strongly bound to albumin in the circulation.
This complex dissociates, and bilirubin enters hepatocytes by a specific uptake mechanism (1). A fraction
of the bilirubin is also derived from catabolism of hepatocellular heme proteins. Within the hepatocyte,
bilirubin binds to a group of cytosolic proteins, termed ligandins; this inhibits the efflux of bilirubin from the
cell. UDP-gl...

Bilirubin dissociates from albumin before entering the hepatocyte, and albumin does not accompany the
pigment into the hepatocyte. Five minutes after intravenous injection of a mixture of [ 3 H]-bilirubin and
131 I-labeled albumin into rats, approximately 60 percent of injected bilirubin is internalized by the liver,

whereas only 10 percent of the injected albumin is present in the liver, probably in the vascular space. 122
In isolated perfused rat and dog livers, simultaneous injection of [ 125 I]-albumin and [ 3 H]-bilirubin discloses
rapid bilirubin uptake, with no removal of albumin from the perfusate. 121, 126 It is not clear whether free or
albumin-bound bilirubin interacts with the hepatocyte. Early studies suggested that the unbound fraction of
bilirubin is taken up by the hepatocyte. 127 This view was challenged by investigators who found that
uptake of albumin-bound ligands by the perfused rat liver correlated poorly with the expected
concentration of unbound ligand. 128, 129 For example, increasing the albumin concentration tenfold from
0.5 g/dl to 5 g/dl reduced the concentration of free taurocholate by a factor of five, but reduced uptake by
only 50 percent. 129 Similar results were found for uptake of fatty acids, bilirubin, and BSP. 128, 130 In
addition, uptake of a 1:1 complex of one of these ligands with albumin was saturable and competitively
inhibited by albumin. 128 Based on these observations, it was suggested that albumin mediated hepatic
uptake of these ligands, 129 and the presence of a receptor for albumin on the liver cell surface was
postulated. 128 However, an alternative hypothesis to explain albumin receptor-like kinetics suggests that
the unbound ligand interacts with the liver cell plasma membrane, but the rate of dissociation of the ligand
from albumin may limit uptake. 131, 132 This new model of dissociation-limited uptake of bilirubin is
compatible with the existence of an albumin receptor and adequately describes BSP uptake in perfused
elasmobranch liver. 133

To evaluate the role of albumin binding on solute distribution within the zones of the hepatic acinus,
Gumucio et al. studied BSP transport in isolated rat liver perfused with 0.01 to 1.0 mM BSP with or without
albumin. 134 When steady-state conditions of BSP excretion were established, the liver was frozen and
relative BSP concentrations in various zones of the liver acinus were estimated. Without albumin, there

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 17 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

was 95 percent extraction of BSP in a single pass; a decreasing concentration gradient from zone 1
(periportal) to zone 3 (pericentral) was observed. Inclusion of 4.5 percent or 1 percent albumin in the
perfusate resulted in single-pass extraction of only 8 to 22 percent of BSP, and the zonal gradient of BSP
content was abolished. The results demonstrate that albumin binding produces more homogeneous
distribution of organic anions within the liver acinus. When the liver was perfused retrogradely (through the
hepatic vein) in the absence of albumin, BSP was taken up predominantly by hepatocytes of zone 3 and a
decreasing gradient from zone 3 to zone 1 was produced. BSP-glutathione conjugates appeared in bile
during antegrade and retrograde perfusion, indicating that hepatocytes of both zones have the ability to
conjugate and excrete BSP.

To elucidate the mechanism and driving forces responsible for hepatic organic anion uptake, a number of
studies have been performed on isolated rat hepatocytes 135–138 or liver sinusoidal vesicles. 139 These
experiments, conducted in the absence of albumin, suggested temperature-dependent,
sodium-independent uptake. Studies of [ 35 S]-BSP uptake in short-term cultured rat hepatocytes,
performed in the presence of a molar excess of bovine albumin, revealed linear uptake of BSP over at
least 15 min with little formation of its glutathione (GSH) conjugate over this time. 140 The initial uptake of
[ 35 S]-BSP was depressed by isosmotic substitution of NaCl by sucrose. A specific cation requirement for
BSP uptake is unlikely because uptake was unaffected by substitution for NaCl by KCl or LiCl. However,
substitution of Cl − by HCO 3 − or gluconate − markedly inhibited BSP uptake. Similar observations were
made for the uptake of [ 3 H]-bilirubin glucuronides. 141 Studies in rat liver perfused with NaCl − or Na
gluconate-substituted mediums revealed similar inhibition of bilirubin influx. 140 The mechanism by which
hepatocyte organic anion transport is stimulated by inorganic anions does not appear to be related to
transport of the inorganic anion in or out of the cell. 142 Studies using 36 Cl revealed that in short-term
cultured rat hepatocytes, BSP uptake requires external Cl − and is not stimulated by unidirectional Cl −
gradients. 142 Thus, there was no evidence for linkage of chloride transport with organic anion transport.
Studies of binding of [ 35 S]-BSP to hepatocytes at 4°C revealed an approximately tenfold higher affinity in
the presence of Cl − as compared with its absence. 142 Whether this is the complete explanation for
Cl − -dependent kinetics of BSP transport remains to be determined. In the transfer of organic anions from
the space of Disse to the hepatocyte, the liver cell plasma membrane is the first barrier to entry into the
cell and presumably is the site of interaction with organic anions that results in carrier-mediated uptake
kinetics. Although kinetic evidence has been presented that bilirubin has rapid transit across lipid
bilayers, 132, 143 the cellular specificity and driving forces of this event suggest a hepatocyte transporter for
this ligand.

Several studies of organic anion interaction with liver plasma membrane preparations have been
performed in an attempt to describe the nature of the putative carrier. A number of early studies
demonstrated saturable binding of BSP to liver plasma membrane. 143–145 Several putative plasma
membrane organic binding proteins were isolated. One isolated by Berk, Stremmel, and colleagues was
termed a “BSP/bilirubin binding protein.” 143, 146 These investigators reported that polyclonal antibodies to
this protein inhibited uptake of BSP and bilirubin by isolated rat hepatocytes. 147 However, interpretation of
these studies is difficult because a relatively large amount of antibody was needed for even a partial
inhibition of BSP uptake. A second protein termed bilitranslocase was isolated by Tiribelli, Sottocasa, and
colleagues. 148 Bilitranslocase is a 170-kDa protein composed of 37-kDa and 35-kDa subunits and has
been reported to reconstitute BSP transport in liposomes 148 and in erythrocyte membrane vesicles. 149 In
studies by Wolkoff and Chung, a photoaffinity probe was devised in which [ 35 S]-BSP was covalently
bound to liver cell plasma membrane after exposure to ultraviolet light. 144 Subsequent sodium dodecyl
sulfate polyacrylamide gel electrophoresis and fluorography revealed radioactivity predominantly
associated with a single 55-kDa protein. 150 This organic anion-binding protein was purified and
characterized immunologically. 144 It appears to differ from the BSP/bilirubin binding protein and
bilitranslocase. Using an antibody against this organic acid- binding protein to screen a rat liver λgt11

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 18 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

expression library, a 1550-bp complementary DNA (cDNA) was cloned. 151 On further characterization, it
was found that this cDNA encoded the β subunit of mitochondrial F 1 -ATPase.

Based on the characteristics of [ 35 S]-BSP extraction from albumin by hepatocytes, 140 an assay was
devised that enabled detection of transport in Xenopus laevis oocytes that had been injected with rat liver
poly(A)+ RNA. 152 These oocytes were able to extract BSP from albumin in a chloride-dependent fashion.
Subsequently, a single complementary RNA (cRNA) was isolated that when injected into oocytes resulted
in marked enrichment of BSP transport activity. The corresponding cDNA encodes a rat liver protein that
has been named organic anion-transporting polypeptide (oatp). The oatp cDNA contains an open reading
frame of 2010 nucleotides and suggests that oatp is a hydrophobic protein. The best computer-generated
model of oatp predicts that the protein has 12 transmembrane domains and three potential
N-glycosylation sites. Northern blot analysis revealed that oatp is expressed in the liver and kidney.
Recent data indicate that it is also present in the choroid plexus, a tissue that also expresses several
otherwise liver-specific proteins. At low stringency, oatp cDNA also hybridizes with messenger RNA
(mRNA) extracted from other organs, including lung, skeletal muscle, and proximal colon. Studies using
HeLa cells stably transfected with oatp indicate that oatp-mediated taurocholate transport is Na +
independent, saturable, and associated with HCO 3 − exchange. 153 However, the role of oatp or one of the
related proteins in bilirubin transport has not been directly established, and is being investigated. Oatp
appears to be the first member of a family of sodium-independent plasma membrane transport proteins.
Other members include transporters for prostaglandins, 154 digoxin, and folic acid.

Intrahepatocellular Storage of Bilirubin


Fifteen minutes following intravenous injection of [ 3 H]-bilirubin into rats, over 90 percent disappeared from
plasma and 25 to 30 percent of the injected dose remained in liver. 155 Radioactivity does not appear in
bile until 3 to 4 min after injection and subsequently appears at a rate of approximately 3 percent of the
injected dose per minute. 156 Thus, from the time bilirubin is cleared from plasma and subsequently
excreted into bile, it is stored within hepatocytes. At all times after intravenous injection, a large proportion
of [ 3 H]-bilirubin is associated with the cytosolic fraction of liver homogenates. 155, 156 Because the water
solubility of unconjugated bilirubin is very low at physiologic pH, it is kept in solution by binding to cytosolic
proteins. Gel filtration of cytosol containing [ 3 H]-bilirubin or [ 35 S]-BSP reveals that radioactivity is
associated with two protein peaks that were originally termed Y and Z (Fig. 125-8). 157 Tracer quantities of
anions bind almost exclusively to the Y protein; with larger amounts, binding to the Z protein becomes
apparent. 157 This suggests that, under physiologic conditions, Y protein is the principal cytoplasmic
protein to which organic anions bind. Purified Y protein was shown to bind many compounds, including
drugs, hormones, and organic anions. 157, 158 Similar proteins with the ability to bind a cortisol
metabolite 159 and an azo-dye carcinogen 160 were identified by various laboratories. These proteins were
found to be immunologically cross-reactive and were termed ligandin. Ligandin accounts for 5 percent of
liver cytosol proteins. 161 Subsequently, ligandin was found to be a family of proteins, identical to the α
class of glutathione-S-transferases in the rat liver. 162, 163 There are corresponding proteins in human
hepatocytes as well. These proteins bind bilirubin and a wide variety of organic anions as nonsubstrate
ligands. 163 The high affinity of these proteins for organic anions suggests that they may play a role in
transport by the liver. The role of ligandin in transport of bilirubin was studied in isolated perfused liver
from rats in which ligandin levels were increased following thyroidectomy or treatment with
phenobarbital. 164 There was no correlation between hepatic ligandin concentration and the influx rate of
bilirubin. However, the efflux rate of bilirubin from liver back to plasma varied inversely with hepatic
ligandin concentration. Thus, intracellular protein binding of bilirubin appears to play no role in its
extraction from serum albumin and subsequent influx into the hepatocyte. However, binding to ligandin
increases the net uptake of bilirubin by reducing efflux from the hepatocyte. With respect to organic anion
transport, ligandin may function within the hepatocyte much as albumin does in the circulation, binding

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 19 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

bilirubin and preventing efflux from the hepatocyte back into the circulation and nonspecific diffusion of
bilirubin into compartments of the hepatocyte in which it may do harm. This hypothesis is supported by the
finding that bilirubin inhibits mitochondrial respiration in vitro, an effect that is prevented by ligandin. 165

Binding of bilirubin to cytosolic proteins. Sephadex-G75 gel chromatography of 110,000 × g rat liver
supernatant to which [14C]bilirubin has been added reveals association of radioactivity with two protein
peaks, Y and Z. Y protein was determined to be quantitatively more important in organic anion binding
and was named ligandin. Subsequently, ligandin was found to consist of several proteins belonging to the
glutathione...

Conjugation of Bilirubin

Bilirubin Conjugates.
Efficient excretion of bilirubin across the bile canaliculus requires its conversion to polar conjugates by
esterification of the propionic acid carboxyl groups. Esterification of one or both propionic acid side chains
forms monoconjugates or diconjugates, respectively. Glucuronic acid is by far the major conjugating group
in normal mammalian bile pigments, 166 although smaller amounts of glucosyl and xylosyl conjugates are
also found. 167 Bilirubin glucuronides are present as monoconjugates and diconjugates (Fig. 125-9). 168
Bilirubin IXα is asymmetrical; therefore, bilirubin IXα monoglucuronide exists as two isomers, depending
on where the glucuronosyl group is attached. 168–171 Bilirubin diglucuronide is the major pigment in normal
human bile. 169–171

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 20 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Bilirubin glucuronides. Both propionic acid side chains are glucuronidated in bilirubin diglucuronide.
Bilirubin monoglucuronide can exist as two molecular species, depending on whether the C12 or C8
propionic acid is conjugated.

Enzyme-Catalyzed Glucuronidation of Bilirubin.


Conjugation of bilirubin with glucuronic acid is catalyzed by a specific form of uridine
diphosphoglucuronate glucuronosyltransferase, abbreviated as UDP-glucuronosyltransferase or UGT.
Protein isolation 172–177 and molecular cloning studies 178–184 revealed that UGTs are a family of
enzymes 185 present in the endoplasmic reticulum and nuclear envelope of hepatocytes and many other
cells. 186 All UGT isoforms accept uridine diphosphoglucuronic acid as the donor substrate and use a
broad range of aglycone acceptor substrates; they catalyze the transfer of the glucuronic acid moiety from
uridinediphosphoglucuronic acid (UDP-glucuronic acid), forming ether, ester, thiol, and N-glucuronides. 187
Aglycone substrates of UGTs are diverse, including steroid hormones, thyroid hormones, bile salts,
neurotransmitters, bilirubin, and various exogenous substrates, such as drugs, environmental toxins, and
laboratory xenobiotics. UGTs are integral membrane proteins that require specific membrane lipids for
their function. 188, 189 Delipidation results in loss of enzyme activity, which is restored upon addition of
appropriate phospholipids. 188, 189 UGT activity in microsomal vesicles is partially latent. Membrane
perturbation by detergent treatment, sonication, or brief incubation with phospholipase A removes the
latency, and full activity of the enzyme is expressed. 188–191 UGT activity in native microsomes is also

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 21 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

enhanced by low concentrations of UDP-N-acetylglucosamine, which may be a physiologic activator of the


transferase. Two mechanisms have been proposed to explain this activation. In the compartmental model,
microsomal lipid membranes are thought to pose a partial barrier between UDP-glucuronic acid and the
catalytic site of UGT, which is thought to be located within the endoplasmic reticular lumen. The amino
acid sequences of UGTs have been interpreted to indicate a single transmembrane domain of the protein.
Transport of UDP-glucuronic acid into the endoplasmic reticulum cisternae by specific mechanisms has
been proposed. 192 This model also postulates the presence of a transporter in the microsomal
membranes, which facilitates access of UDP-glucuronic acid to the catalytic site of the transferase.
UDP-N-acetylglucosamine is envisioned as an activator of the permease. Membrane perturbation is
thought to enhance enzyme activity in vitro by increasing the permeability of the lipid membranes to
UDP-glucuronic acid. 193 The alternative hypothesis, termed the allosteric model, postulates that the
enzyme activity is constrained by the membrane. Activating agents release the enzyme from constraint
and thereby enhance enzyme activity. 194

Structure and Expression of UGT Isoforms.


UGT isoforms are expressed from multiple loci. The C-terminal domain of UGT isoforms have a high
degree of homology and is responsible for binding of the common donor substrate, UDP-glucuronic
acid, 195 whereas the N-terminal domains are more heterogeneous and account for the aglycone substrate
specificity of individual isoforms. 196 Based on the degree of homology of the mRNA sequences, UGT
genes have been classified among multiple families, and each family into several subfamilies. 185 Only
one UGT isoform is physiologically relevant in bilirubin glucuronidation. 196 The locus that expresses
bilirubin-UGT is termed UGT1A (Fig. 125-10). 197 This locus expresses several UGT isoforms and
contains four consecutive exons (exons 2–5) at the 3′ end that are used in all mRNAs expressed from this
locus. Upstream from these four common region exons are a series of unique exons, each preceded by a
separate promoter. Depending on which promoter is used, transcripts of various sizes are generated. The
unique exon, present at the 5′ end of the transcript, is spliced to exon 2, and the intervening sequence is
spliced out. Thus, all isoforms expressed from the UGT1A locus have identical C-terminal domains
(encoded by exons 2–5) but variable N-terminal domains (encoded by a single unique exon). Within the
UGT1A locus, genes encoding individual isoforms are named according to the unique exon used in the
expressed mRNA. Thus, the gene for the bilirubin-UGT mRNA, which consists of the first unique region
exon of the UGT1A locus (plus exons 2–5), is termed UGT1A1 according to current terminology, and the
enzyme is termed UGT1A1. Another isoform that catalyzes phenolic substrates is encoded by the unique
region exon 6 (plus exons 2–5); therefore, this gene is termed protein/gene for
UDP-glucuronosyltransferase 1, family member A6 (UGT1A6), and the expressed isoform is termed
UGT1A6.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 22 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Schematic representation of the human UGT1A locus, located at 2q37. This locus contains multiple genes
that express bilirubin-UGT and several other UGT isoforms. Exons 2, 3, 4, and 5, located at the 3′ end of
UGT1A, encode the identical C-terminal domains of all UGT isoforms expressed from this locus.
Upstream to these common region exons are a series of unique exons (exons 1A1 through 1A12), each of
which encodes the variable N-termi...

Differential Expression of UGT Isoforms.


The presence of a separate promoter upstream from each unique region exon (Fig. 125-10) permits
differential regulation of individual UGT isoforms during ontogenic development 198 and enzyme
induction. 199, 200 Enzyme activity toward 4-nitrophenol and other simple phenolic substrates (catalyzed by
UGT1A6) develops in late fetal life, whereas UGT activity toward bilirubin (UGT1A1) develops after
birth. 198 Treatment of rats with 3-methylcholanthrene induces UGT1A6. 199, 200 In contrast, UGT activity
toward bilirubin is specifically induced by clofibrate. 199 Treatment of rats with triiodothyronine results in a
threefold increase in transferase activity toward 4-nitrophenol, whereas activity toward bilirubin is
decreased by 80 percent. 198

Quantification of Bilirubin and Its Conjugates


Bile pigments are quantified in body fluids as native or derivatized tetrapyrroles, or after conversion to
azoderivatives. Total bilirubin also can be quantified indirectly by quantification of the intensity of yellow
discoloration of the skin.

Methods Involving Conversion of Bilirubin to azo Derivatives.


Conversion to azodipyrroles by reaction with diazo reagents is used commonly for determination of serum
bilirubin levels for clinical purposes. Reaction of bilirubin with diazo reagents begins with electrophilic
attack by a diazonium ion at the 9 and 11 positions of bilirubin 201 and converts 1 mole of the tetrapyrrole
to 2 moles of diazotized azodipyrrole and 1 mole of formaldehyde (Fig. 125-11). The azoderivatives are

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 23 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

more stable than is bilirubin, and its conjugates and can be quantified colorimetrically. Unconjugated
bilirubin is converted to two unconjugated dipyrroles. Bilirubin diconjugates form two conjugated
azodipyrroles, and bilirubin monoconjugates form one conjugated and one unconjugated azodipyrrole. In
1916, van den Bergh and Muller 202 showed that, on the basis of diazo reaction, serum bile pigments can
be classified into a direct and an indirect reacting species. The direct reaction occurs within minutes, and
the indirect reaction occurs rapidly only in the presence of accelerator substances such as methanol or
caffeine. Subsequently, the direct and indirect reacting components were identified as conjugated and
unconjugated bilirubin, respectively. 203 The basis of the direct and indirect van den Bergh reactions can
be understood from the crystal structure of bilirubin. Because of internal hydrogen bonding, the central
methene bridge of bilirubin is not readily accessible to diazo reagents. Addition of accelerators results in
disruption of the hydrogen bonds, and completes the reaction. Bilirubin glucuronides lack some or all of
the hydrogen bonds and, therefore, react immediately with diazo reagents, without requiring the addition
of accelerators. For confirmation, the conjugated and unconjugated azodipyrroles can be extracted and
analyzed by thin-layer chromatography 204 or high-pressure liquid chromatography (HPLC) 205 (Fig.
125-12).

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 24 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Reaction of bilirubin tetrapyrrole with the diazonium salt of ethylanthranilate results in the formation of
equimolar amounts of two azodipyrroles. The central methenyl bridge carbon is converted to
formaldehyde. GA, glucuronic acid.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 25 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Separation of ethylanthranilate azodipyrroles by HPLC. Wistar rat bile was diazotized with
ethylanthranilate diazo reagent, azodipyrroles were extracted, 208 organic solvents were eliminated in
reduced pressure, and the pigments were dissolved in methanol and separated by reverse-phase HPLC
(µ-Bondapak C-18 column, Waters Associates) using a concave gradient (dashed line) of methanol (80 to
100 percent) in sodiu...

Because 10 to 15 percent of unconjugated bilirubin may give direct diazo reaction, the direct-reacting
fraction slightly overestimates the levels of conjugated bilirubin. In addition, the irreversibly albumin-bound
fraction of serum bilirubin, which is formed in the serum of patients with prolonged conjugated
hyperbilirubinemia, exhibits direct diazo reaction. 99 Because irreversibly protein-bound bilirubin is cleared
slowly, it persists in serum for a relatively long period after correction of biliary obstruction. Finding of
direct-reacting bilirubin during this period may give a false impression of continued biliary obstruction. In
patients with renal failure, indican accumulates in serum and may interfere with the diazo reaction of
bilirubin. 206 Finally, the diazo method cannot be applied to separately quantify bilirubin monoconjugates
and diconjugates when they are present in a complex mixture. The latter requires analysis of intact
bilirubin tetrapyrroles.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 26 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Separation and Quantification of Intact Bilirubin Tetrapyrroles.


Chromatographic separation of unmodified bile pigments was attempted by Cole et al. as early as in
1954. 207 Subsequently, Heirwegh and associates developed highly resolving thin-layer chromatographic
systems for separation of bilirubin and its conjugates. 208 Higher resolution and quantitative recovery of
bile pigments is possible by using HPLC. Methyl esters formed by alkaline methanolysis of bilirubin
monoconjugates and diconjugates have been separated and quantified by HPLC. 209 Because the
conjugating sugars are replaced by methyl groups, the pigments cannot be separated on the basis of their
conjugating moieties. Therefore, methods for separation and quantitation of underivatized bilirubin
tetrapyrroles by HPLC have been developed 169–171, 210 (Fig. 125-13). Reverse-phase HPLC of
incompletely deproteinated serum has been used to quantify simultaneously irreversibly protein-bound
and other bilirubin fractions. 99 These studies indicate that the irreversibly protein-bound serum bilirubin
fraction is present in conditions associated with conjugated hyperbilirubinemia. After successful surgical
correction of biliary obstruction, the reversibly protein-bound fraction of serum bilirubin is rapidly excreted
in bile; this results in an increase in the proportion of the irreversibly protein-bound fraction of serum
bilirubin. If biliary obstruction persists, both reversibly protein-bound and irreversibly protein-bound
fractions are retained, and no increase in the proportion of the latter is observed.

HPLC of intact underivatized bilirubin tetrapyrroles in human bile from a normal individual (A) and from
patients with Crigler-Najjar syndrome type I (B), Crigler-Najjar syndrome type II (Arias syndrome) (C),
Gilbert syndrome (D), and Dubin-Johnson syndrome (E). Bile pigments were separated by reverse-phase

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 27 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

chromatography. Absorbance at 436 nm (ordinate) and retention time (abscissa) are shown. Peaks are as
follows: 1, bilirubin diglu...

Slide Tests.
Two slide tests have been introduced for determination of conjugated, unconjugated, and irreversibly
protein-bound bilirubin. One slide (Ektachem TIBL) is used for measurement of total bilirubin by a diazo
technique. 211 The other slide has a special coating that allows only the free and reversibly protein-bound
bilirubins to come in contact with the diazo reagent; conjugated and unconjugated bilirubin are separately
quantified by reflectometric measurements at two wavelengths. 212 The difference between total bilirubin
and the sum of conjugated and unconjugated bilirubin gives the value for irreversibly protein-bound
bilirubin. These results have been verified by HPLC and indicate that the results obtained by the
Ektachem slide tests are consistent and reliable.

Transcutaneous Bilirubinometry.
The current emphasis on early discharge of neonates makes it imperative to assess the risk of severe
neonatal hyperbilirubinemia by evaluating the rate of increase of serum bilirubin levels during the first 24
to 48 hours of life. This requires repeated measurement of serum bilirubin levels, which is painful and
expensive. Measurement of the yellow color of the skin by analysis of reflected light provides a
noninvasive and relatively inexpensive method for estimating serum bilirubin levels without drawing
samples. 213, 214 The analyzers use on-board computers programmed to measure the yellow color without
interference by underlying skin pigmentation or degree of erythema. In 900 term and premature infants of
various races, bilirubin levels estimated by transcutaneous bilirubinometry correlated well with serum
billirubin concentrations measured by a standard diazo method. 215

Fluorimetric Analysis.
Fluorescence characteristics of bilirubin have been used in the development of a method for determination
of total bilirubin, albumin-bound bilirubin, and reserve bilirubin-binding capacity from as little as 0.1 ml of
whole blood. Bilirubin bound to the high-affinity site of albumin has a fluorescence peak at 520 nm when
excited at 430 nm. Unbound bilirubin and bilirubin bound to other proteins have negligible fluorescence.
Addition of a saturating amount of bilirubin to blood results in maximum fluorescence, allowing
determination of total bilirubin-binding capacity. Addition of a detergent, dodecylmethylamine oxide, to
whole blood results in hemolysis and quantitative incorporation of bilirubin into the detergent micelles.
Fluorescence of detergent-bound bilirubin is used for quantitation of total bilirubin. These parameters can
be readily determined using a digital hematofluorometer. Because fluorescence also depends on
hemoglobin concentration, hemoglobin values are independently determined by the hematofluorometer
and taken into account in calculation of displayed values for total bilirubin and albumin-bound bilirubin and
reserve bilirubin binding capacity. 216

Biliary Secretion of Bilirubin


In mammals, conjugation is essential for the secretion of bilirubin acoss the bile canaliculus. Gunn rats
and patients with Crigler-Najjar syndrome type 1 lack bilirubin UGT activity and manifest lifelong
unconjugated hyperbilirubinemia. Their bile contains only a small amount of bilirubin. Normally, the
conjugating capacity and biliary excretory capacity of bilirubin are closely matched. Thus, drugs and
chemicals that reduce the maximal secretory capacity for bilirubin cause the retention of conjugated
bilirubin. On the other hand, when bilirubin-UGT activity is partially deficient, conjugation becomes rate
limiting in bilirubin secretion. 217 In rats there is approximately a 15-fold functional reserve in the ability to
secrete conjugated bilirubin into bile.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 28 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Prior to 1991, the major driving force for the transport of bilirubin conjugates from the hepatocyte into the
bile was thought to be the electrochemical gradient of −35 mV. 218 Relative intracellular negativity is
generated by the sodium pump, which tranports three molecules of sodium out of the cell while
transporting two molecules of potassium into the cell. The sodium pump is functional mainly in the
basolateral domain of the hepatocyte plasma membrane, and is present but inactive in the canalicular
domain. The observed potential difference is too small to account for the large bilirubin glucuronide
concentration gradients between the hepatocyte and bile, which may be as high as 150-fold. The energy
for this uphill transport is provided by an ATP-dependent system in the canalicular membranes that is
specific for non–bile acid organic anions, including bilirubin and other glucuronides and glutathione
conjugates. 218–223 In contrast to the bidirectional transport at the sinusoidal aspect of the hepatocyte,
canalicular transport of organic anions is unidirectional from the cytoplasmic aspect of the hepatocyte into
the bile. In addition to the energy-consuming process, the canalicular transport may be assisted by the
membrane potential, but the contribution of membrane potential in organic anion transport has not been
quantified. The membrane potential–driven canalicular transport of non–bile acid organic anions is distinct
from that driven by ATP hydrolysis. 223 Mutant animals that lack ATP-dependent canalicular transport of
non–bile acid organic anions retain normal activity with respect to potential-driven canalicular transport of
non–bile acid organic anions, including bilirubin glucuronides. 224, 225 The benign phenotype of the defect
in human and animal mutants may reflect persistence of the potential-driven transport system in the bile
canaliculus.

The ATP-dependent canalicular non–bile acid organic anion transporter (also called MOAT, multiorganic
anion transporter) is functionally distinct from other canalicular ATP-dependent transporters for organic
cations, 226–228 bile acids, 229, 230 and phospholipids. 231 Attempts to purify the ATP-dependent transporter
for bilirubin glucuronide were unsuccessful. Its identification by cloning followed the serendipitous
observation that a drug-resistant cancer cell line expressed an ATP-dependent multidrug resistance–like
protein (MRP) that transports various organic cations. 232 The MRP family of ATP-dependent membrane
transporters differs from the previously described multidrug resistance (MDR) family of ATP-dependent
membrane proteins in amino acid sequence, molecular weight, hydropathy plots, and substrates. Both
have multiple transmembrane domains and two nucleotide binding sites that are separated by distinct
linker domains. Whereas MRP1 is present in the plasma membrane of many cell types, a family member,
MRP2 (initially termed multispecific organic anion transporter, MOAT), is restricted to the bile canalicular
membrane, where it is responsible for ATP-dependent transport of a wide variety of non-bile acid organic
anions, which are primarily glucuronides and glutathione conjugates. 233

Patients with the Dubin-Johnson syndrome, 234, 235 mutant Corriedale sheep, 224 transport-deficient mutant
rat strain (TR) and Eisai hyperbilirubinemic rat (EHBR) rats 225 and mutant golden lion tamarin
monkeys 236 manifest conjugated hyperbilirubinemia that is transmitted as an autosomal-recessive trait.
These mutants share defective capacity to transport bilirubin glucuronide, other organic anionic
metabolites (including leukotriene C4 and metanephrine glucuronide), and other organic anions, such as
BSP, ICG, iopanoic acid, and phylloerythrin, from hepatocytes into the bile. Affected patients and mutant
animals have a normal transport maximum for infused bile acids. 235, 237 These observations
demonstrated that there are at least two mechanisms for organic anion secretion by the liver, one for bile
acids and another for other organic anions. In canalicular membrane vesicles, these processes were
functionally distinct and required ATP hydrolysis. 230 The major, and possibly only, canalicular bile acid
transporter has been demonstrated to be sister-p-glycoprotein (SPGP), a member of the MDR family of
transmembrane transporters. 238

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 29 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

The predominant member (approximately 90 percent) of MDR family gene products in the canalicular
membrane is MDR3 (the human homologue of murine mdr2). MDR3 couples ATP hydrolysis to the
selective transfer of phosphatidylcholine from the inner to the outer leaflet of the bile canalicular
membrane. 239 Phospholipids are required to form mixed micelles with bile acids, which protects small bile
ducts from the detergent action of bile acids.

Maximal bilirubin secretory capacity (T max ) depends on bile flow. Flow is increased by infusion of bile
acids 240 or by phenobarbital treatment, which enhances bile flow rate by a non–bile acid dependent
mechanism. 241 The T max of bilirubin is enhanced in both cases. Several other bile acid–independent
choleretics increase bile flow but not the T max for organic anions. 242 The maximal ATP-dependent biliary
secretion of bile acids and non–bile acid organic anions greatly exceeds the apparent capacity and
amount of SPGP and MRP2 present in the canalicular membrane under basal conditions. 243 Bile acids
increase the transfer of MRP2, SPGP, and MDR3 from the Golgi to the apical domain, thereby more than
doubling the amount of each transporter in the canalicular membrane. 244 Recruitment of canalicular
ATP-dependent transporters involves microtubular-dependent vesicular trafficking, which requires
association with and activity of phosphoinositide 3-kinase and its lipid products.

A small amount of unconjugated bilirubin (up to 3% of total bile pigments) is found in normal bile. This may
be explained by hydrolysis of bilirubin glucuronide by biliary β-glucuronidase, canalicular ATP-dependent
transport of bilirubin, or self-aggregation and incorporation of bilirubin in mixed micelles.

Although bilirubin glucuronides, bile acids, and phospholipids are transported across the canalicular
membrane by different ATP-dependent transporters, heritable defects in bile acid or phospholipid
transport produce bile secretory failure (i.e., cholestasis), which is manifested by conjugated
hyperbilirubinemia. Molecular defects in heritable cholestatic disorders remained unknown until the
mechanisms responsible for biliary secretion of bilirubin glucuronide, bile acids, and phospholipids were
identified. The discovery of bile canalicular ATP-dependent transporters for bile acids (SPGP), non–bile
acid organic anions (MRP2), and phospholipids (MDR3) enabled the molecular characterization of several
heritable hyperbilirubinemias, cholestatic or not.

Fate of Bilirubin in the Gastrointestinal Tract


Bilirubin reaches the intestinal tract mainly conjugated and is not substantially absorbed. 245 In some
circumstances, there may be enhanced excretion of unconjugated bilirubin into the intestine. Absorption of
unconjugated bilirubin from the intestine may contribute to neonatal hyperbilirubinemia. 246 Milk inhibits
intestinal absorption of unconjugated bilirubin. However, human milk inhibits this reabsorption less than
does infant milk formula. 247 These observations suggest that intestinal absorption of bilirubin may
contribute to jaundice associated with breast-feeding. Absorption of bilirubin from the gallbladder occurs in
animals. 248 Bilirubin is degraded by intestinal bacteria into a series of urobilinogen and related
products. 249 The specific products may relate to strains of bacteria present in the intestine. 250
Urobilinogens are present in the deconjugated state. It is not known whether deconjugation precedes or
follows bilirubin degradation, but bacterial β-glucuronidase plays a role in the deconjugation. 246, 251 Most
of the urobilinogen reabsorbed from the intestine is reexcreted in the bile. A small fraction is excreted by
the kidney. Enhanced tubular absorption and instability of the pigment in acid urine makes urobilinogen
excretion in urine an unreliable indicator of the status of bilirubin metabolism. Absence of urobilinogen in
stool and urine indicates complete obstruction of the bile duct. In liver disease and states of increased
bilirubin production, urinary urobilinogen excretion is increased. Urobilinogen is colorless. Oxidation leads
to formation of urobilin, which contributes to the color of normal urine and stool.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 30 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Alternative Pathways of Bilirubin Elimination


After injection of labeled unconjugated bilirubin, only 3 percent of radioactivity is normally excreted by the
kidney in humans. Even in the presence of marked hyperbilirubinemia, bile remains the main route of
bilirubin excretion. In patients with Crigler-Najjar syndrome and in Gunn rats, a small amount of
unconjugated bilirubin is secreted in bile. Additional unconjugated bilirubin may reach the intestinal lumen
by passage across the intestinal wall or by desquamation of intestinal epithelial cells. 252 Ambient light or
phototherapy forms geometric isomers of bilirubin (EE, EZ, or ZE forms), which are excreted in
unconjugated forms and converted to bilirubin IXα-ZZ in the bile. 253 Considerable amounts of bilirubin are
degraded to polar diazo-negative compounds, which are excreted in both bile and urine. 252

The presence of hydroxylated products of bilirubin in Gunn rat bile suggests a role of enzyme-catalyzed
oxidation in the disposition of bilirubin. 254 Induction of a specific isoform of microsomal P450s by
administration of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in Gunn rats results in a sevenfold increase
in the fractional turnover of bilirubin and reduction of the bilirubin pool. 255 A mitochondrial bilirubin oxidase
found in rat liver, 256 intestine, 257 and kidney consumes 1 to 1.5 moles of oxygen per mole of bilirubin and
forms propentdyopents. The enzyme does not require NADP, NAD, or ATP and is inhibited by potassium
cyanide (KCN), thiol reagents, NADH, and albumin. 256 Enzyme-mediated oxidation of bilirubin also has
been reported to occur in mitochondrial fractions of brain, lung, heart, and skeletal muscle.

Disposition of Bilirubin by the Kidney.


In intrahepatic or extrahepatic cholestasis, the plasma-conjugated bilirubin concentration increases. After
injection of radiolabeled bilirubin in animals with experimentally ligated bile ducts 258 and in children with
biliary atresia, 259 50 to 90 percent of injected radioactivity is excreted in urine. In total biliary obstruction,
urinary excretion becomes the major pathway of bilirubin excretion. 260 Renal excretion of conjugated
bilirubin depends on glomerular filtration of a small non–protein-bound fraction of conjugated
bilirubin. 260, 261 There is evidence for tubular reabsorption but none for tubular secretion of bilirubin. 261

Antioxidative Property of Bilirubin


Until recently, bilirubin has been considered a waste product with no known physiologic utility. However,
both unconjugated 262 and conjugated 263 bilirubin have been shown to be inhibitors of lipid peroxidation.
This property of bilirubin depends on its ability to donate hydrogen ions. 264 Bilirubin bound to hepatic
cytosolic proteins has a greater free-radical scavenging effect than does the bilirubin–albumin complex. It
has been postulated that bilirubin is converted to its peroxidative product, biliverdin, thereby sparing
hepatic endoplasmic reticulum membranes of lipid peroxidation. 265 Regeneration of bilirubin by the action
of cytosolic biliverdin reductase may enhance the protective effect of bilirubin. Albumin binding increases
the cytoprotective activity of bilirubin, 265 suggesting a possible beneficial role of bilirubin on tissues other
than the liver.

DISORDERS OF BILIRUBIN METABOLISM RESULTING IN UNCONJUGATED


HYPERBILIRUBINEMIA

General Considerations
The hepatic transport of bilirubin involves four distinct but probably interrelated stages: (1) uptake from the
circulation; (2) intracellular binding or storage; (3) conjugation, largely with glucuronic acid; and (4) biliary
excretion. Abnormalities in any of these processes may result in hyperbilirubinemia. Complex clinical
disorders, such as hepatitis or cirrhosis, may affect multiple processes. In several inherited disorders, the
transfer of bilirubin from blood to bile is disrupted at a specific step. Study of these disorders has permitted

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 31 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

better understanding of bilirubin metabolism in health and disease. Each disorder is characterized by
varied degrees of hyperbilirubinemia of the unconjugated or conjugated type.

Neonatal Jaundice
By adult standards, every newborn baby has hyperbilirubinemia, and about half of all neonates become
clinically jaundiced during the first 5 days of life. Serum bilirubin is predominantly unconjugated.
Exaggeration of this physiologic jaundice can result in marked hyperbilirubinemia, with an attendant risk of
kernicterus (see earlier section on Toxicity of Bilirubin). In 4000 consecutive infants, 16 percent had
maximal serum bilirubin concentrations of 10 mg/dl or above, and in 5 percent, bilirubin concentrations
exceeded 15 mg/dl. 266 In the normal, full-term human neonate, the serum bilirubin concentration
increases rapidly from 1 to 2 to 5 to 6 mg/dl in approximately 72 h and subsequently decreases until
normal levels are attained in 7 to 10 days. 267 Physiologic jaundice of the newborn appears to result from
a combination of increased bilirubin production and delayed maturation in the capability of the liver to
dispose of bilirubin. Severe neonatal unconjugated hyperbilirubinemia results from exaggeration in one or
more of the regularly occurring developmental restrictions that are characteristic of the newborn period or
from superimposition of additional mechanisms. Although the incidence of cerebral toxicity from neonatal
jaundice had decreased markedly after the introduction of immunoglobulin therapy for maternal–fetal Rh
blood group incompatibility, the incidence of neonatal kernicterus may be on the increase again. This may
be partly related to early discharge after delivery, which is being practiced in a majority of hospitals in the
United States and elsewhere. This issue and the current concepts of neonatal jaundice have been
reviewed elsewhere. 78 The physiologic basis of neonatal hyperbilirubinemia and mechanisms of its
exaggeration to potentially harmful states are discussed briefly below.

Increased bilirubin production in the newborn period is evidenced by increased endogenous carbon
monoxide production, 268 increased early-labeled peak from erythroid and nonerythroid sources, and
decreased erythrocyte half-life. 269 Meconium contains unconjugated bilirubin derived primarily from
hydrolysis of conjugated bilirubin by intestinal β-glucuronidase. 270 As compared with adults, newborns
lack intestinal bacteria that degrade bilirubin to urobilinogen and have a greater surface-to-volume ratio of
the bowel. As a result, intestinal absorption of unconjugated bilirubin in neonates may be increased. 270
Hemolytic diseases of the fetus increase bilirubin production and may lead to severe neonatal
unconjugated hyperbilirubinemia. Rh incompatibility between mother and fetus was formerly a common
cause of severe neonatal unconjugated hyperbilirubinemia and kernicterus (erythroblastosis fetalis). This
disease can be prevented by treatment of the mother with anti-Rh immunoglobulins. 271 Major blood group
(ABO) incompatibility remains a common cause of exaggerated neonatal hyperbilirubinemia that often
requires treatment. 272

Hepatic Bilirubin Uptake During the Neonatal Period.


Cumulative hepatic bilirubin uptake capacity is reduced during the first 25 h of life in the rhesus monkey.
Relative hepatic uptake deficiency extends beyond the second day of life and correlates with maturation of
hepatic ligandin, 273 which influences the net hepatic uptake of bilirubin (see earlier section on Uptake of
Bilirubin by the Liver). Delayed closure of the ductus venosus may permit portal blood, which is enriched
in unconjugated bilirubin from the intestine, to bypass the liver. Reduced caloric intake, which reduces
hepatic bilirubin clearance in adults (see later section on Gilbert Syndrome), may have a similar effect in
neonates.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 32 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Postnatal Development of Bilirubin-Glucuronidating Activity.


In many mammals, including humans, UGT activity toward bilirubin is deficient in fetal liver and rapidly
develops to adult levels during the first few days of life. 274 Deficiency of UGT activity may be prolonged
and exaggerated in some inherited disorders due to inhibitory factor(s) in maternal milk or serum (see
later section on Transient Familial Neonatal Hyperbilirubinemia).

Recently, a variant TATAA element within the promoter region of UGT1A1 has been found to be
associated with Gilbert syndrome. 275 This variant promoter reduces the expression of bilirubin-UGT
(UGT1A1). The Gilbert genotype has been found to accentuate neonatal hyperbilirubinemia 276 and
prolong the duration of neonatal jaundice. 277

Inhibition of Bilirubin Glucuronidation by Maternal Milk.


Plasma bilirubin concentrations tend to be higher in breast-fed infants as compared with formula-fed
babies 278 and occasionally rise to maximum concentrations of 15 to 24 mg/dl within 10 to 19 days of life.
This transient nonhemolytic, unconjugated hyperbilirubinemia is promptly ameliorated by discontinuation
of breast-feeding; otherwise, the hyperbilirubinemia may take up to a month to disappear. Kernicterus is
rare, but isolated cases have been reported. 279 Neonatal unconjugated hyperbilirubinemia related to
breast-feeding is associated with an inhibitor of UGT activity in maternal milk but not maternal serum. 280
A progestational steroid, 3α,20β-pregnanediol, was isolated from the milk of mothers of infants who had
the syndrome. The steroid inhibited bilirubin glucuronidation by rat and rabbit liver, 280 but activity with
human liver was not inhibited. 281 Experimental feeding of the steroid to healthy infants yielded
contradictory results. 282, 283 The free fatty acid concentration in maternal milk correlates positively with its
inhibitory effect on human hepatic bilirubin-UGT activity, and free fatty acids inhibit UGT activity in vitro in
proportion to the number of double bonds in unsaturated fatty acids and in inverse proportion to the chain
lengths of saturated fatty acids (C10 to C18). 284 Odievre and associates have postulated that a lipolytic
enzyme, which is present in some maternal milk samples, may be responsible for the increased
concentration of free fatty acids in the milk. 284 The inhibitory effect of maternal milk on UGT increases on
storage and is destroyed by heating at 56°C.

Inhibition of Bilirubin Glucuronidation by Factors Derived from Maternal Plasma.


Transient familial neonatal hyperbilirubinemia (Lucey-Driscoll syndrome) was described by Lucey and
associates in 1961. 285, 286 This syndrome is characterized by jaundice occurring within the first 4 days of
life. In 24 infants, 286 peak serum bilirubin concentrations of 8.9 to 65 mg/dl were reached within 7 days.
An unidentified inhibitor of UGT was found in the serum of mothers of these infants. One infant died at
36 h. This condition is clinically distinguished from maternal milk jaundice by earlier onset of severe
hyperbilirubinemia, a more prolonged course, and occasional kernicterus.

Immaturity of Canalicular Excretion.


During the late newborn period, hepatic bilirubin uptake, conjugation, and canalicular excretion attain adult
levels, even though the bilirubin load remains increased. In this period of life, canalicular excretion may be
rate limiting in the hepatic disposition of bilirubin. Consequently, when the bilirubin load is further
increased, conjugated bilirubin accumulates in serum. 287

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 33 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Management of Neonatal Unconjugated Hyperbilirubinemia.


In the great majority of cases, neonatal hyperbilirubinemia is innocuous, and indeed may provide an
important antioxidant defense for the newborn. However, it is critical to be vigilantly prepared for the
occasional case in which neonatal jaundice can reach levels that may cause cerebral toxicity. 78 Although
a plasma bilirubin concentration of 20 mg/dl is usually considered dangerous, cerebral toxicity can occur
at lower concentrations of bilirubin (see earlier section on Toxicity of Bilirubin). The goal of treatment is to
decrease serum bilirubin concentrations to an acceptable level until the capacity of the liver to dispose of
bilirubin matures. Phototherapy is the most common treatment modality used. In severe cases, exchange
transfusion is used to reduce serum bilirubin levels rapidly (see later section on Treatment of
Crigler-Najjar Syndrome Type I). Although phototherapy is useful and safe, concern persists about its
potential side effects. Ingestion of agar to bind unconjugated bilirubin in the intestine also decreases
serum bilirubin concentrations, 270 but the efficacy of this treatment is not certain. 269 Inhibition of heme
oxygenase activity by the administration of tin-mesoporphyrin at birth has been shown to prevent the
development of significant levels of neonatal jaundice, thereby abolishing the need for phototherapy or
exchange transfusion. 44, 45 However, the routine use of this compound in all newborns is not generally
recommended.

Hyperbilirubinemia Due to Bilirubin Overproduction


Hyperbilirubinemia in the presence of normal liver function often occurs in disorders associated with
increased bilirubin production. The serum bilirubin is unconjugated and rarely exceeds 3 to 4 mg/dl.
Higher levels usually indicate hepatobiliary dysfunction in addition to bilirubin overproduction. 4 The most
common cause of increased bilirubin production is hemolysis such as occurs in sickle cell anemia,
hereditary spherocytosis, and toxic or idiosyncratic drug reactions in susceptible individuals. These
disorders are associated with premature destruction of erythrocytes; red cell morphology and life span are
often abnormal. In the absence of any hepatobiliary disorder, a small amount of conjugated bilirubin,
mainly bilirubin monoglucuronide, may accumulate in the serum, in addition to unconjugated bilirubin. In
these cases, the proportion of conjugated bilirubin does not usually exceed the normal limits
(approximately 4 percent of total bilirubin). The conjugated bilirubin probably appears in the serum by
diffusion out of the hepatocyte. In contrast to previous interpretation, 288 this phenomenon does not
necessarily indicate that the rate of bilirubin production has exceeded the excretory transport maximum for
biliary excretion of conjugated bilirubin. Ineffective erythropoiesis occurs in thalassemia and other
hematologic disorders and is often associated with hyperbilirubinemia. 289 Congenital dyserythropoietic
anemias are a group of rare hereditary anemias characterized by ineffective erythropoiesis, intramedullary
normoblastic hyperplasia, secondary hemochromatosis, and unconjugated hyperbilirubinemia. 290–293

Crigler-Najjar Syndrome Type I

Clinical Findings.
Crigler-Najjar syndrome type I is a rare disorder in which hepatic bilirubin-UGT activity is absent or barely
detectable (Table 125-1). The syndrome was described by Crigler and Najjar in 1952 in six infants in three
families. 294 All infants manifested severe nonhemolytic icterus within the first few days of life. Jaundice
was characterized by increased plasma concentration of indirect-reacting bilirubin and was lifelong. Five
of the six infants died of kernicterus by the age of 15 months. Although icteric, the single surviving infant
was free of neurologic disease until 15 years of age, when kernicterus suddenly developed, and he died 6
months later. 295 A female cousin also had Crigler-Najjar syndrome; she developed neurologic symptoms
at 18 years of age and died at the age of 24. 296 This family had increased consanguinity and several
other recessively inherited traits, such as Morquio syndrome, homocystinuria, metachromatic
leukodystrophy, and bird-headed dwarfism. 297 However, association with these disorders was not
observed in the subsequently reported cases of Crigler-Najjar syndrome type I. The syndrome occurs in

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 34 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

all races, is transmitted as an autosomal-recessive trait (Fig. 125-14), 295–298 and is associated with
known consanguinity in some, but not all, cases. Until the introduction of phototherapy, almost all patients
died with kernicterus during the first 18 months of life. 299 Several individuals have survived only to
succumb to kernicterus later in life. 291, 294, 296, 297 With the advent of phototherapy and intermittent
plasmapheresis, survival until puberty, without significant brain damage, is not unusual. However, the risk
of bilirubin encephalopathy persists, and kernicterus is common around the time of adolescence, when
phototherapy becomes less effective. Orthotopic liver transplantation or auxiliary transplantation of a
single liver lobe has resulted in long-term survival in several cases. 300 As discussed below, isolated
hepatocyte transplantation has been used in a single case with partial amelioration of hyperbilirubinemia.

Table 125-1: Principal Differential Characteristics of Inherited Unconjugated Hyperbilirubinemia

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 35 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Crigler-Najjar
Crigler-Najjar
Syndrome Type Gilbert Syndrome
Syndrome Type II
I

Histology of liver Normal Normal Normal

Serum bilirubin concentration 20–50 mg/dl <20 mg/dl Usually <3 mg/dl

Routine liver function test


Normal Normal Normal
results

Usually normal; may be


45-min plasma BSP retention Normal Normal elevated in some
patients

Usually pale;
Increased
contains small Increased proportion of
proportion of
Bile amounts of bilirubin
bilirubin
unconjugated monoglucuronide
monoglucuronide
bilirubin

Hepatic bilirubin
UDP-glucuronosyltransferase Absent Markedly reduced Reduced
activity

Effect of phenobarbital on
None Reduction Reduction
serum bilirubin

Autosomal Autosomal
Mode of inheritance Autosomal recessive
recessive recessive

Common (~9% of
caucasians are
homozygous for a
Prevalence Rare Rare
variant TATAA box;
4–5% have
hyperbilirubinemia)

Kernicterus,
Usually benign,
unless
Prognosis kernicterus occurs Benign
vigorously
rarely
treated

Homozygous Bolivian squirrel


Animal model —
Gunn rat monkey

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 36 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Inheritance of Crigler-Najjar syndrome type I. This family, originally described by Crigler and Najjar in
1962, is unique in that two cousins, J.D.H. and M.E.H., escaped kernicterus in infancy only to die at ages
16 and 24, respectively. (Reprinted with permission from Blaschke et al. 363 )

Laboratory Tests.
Laboratory test results in Crigler-Najjar syndrome type 1 are normal except for the serum bilirubin level,
which is usually 20 to 25 mg/dl, but may be as high as 50 mg/dl. 294–297, 299 Virtually all the serum bilirubin
is unconjugated, and no serum conjugated bilirubin has been found. There is no bilirubinuria, but the urine
may be yellow due to a chloroform-soluble pigment of unknown structure. 300 The level of icterus in a
given patient varies—it is lower in summer and on exposure to sun and higher during intercurrent
illness. 301 Stool color is normal, but fecal urobilinogen excretion is reduced. 294–301 Bilirubin production,
hematocrit, bone marrow morphology, and red cell survival are normal. 250, 302 Results of routine liver
function tests are normal, including studies of plasma disappearance of BSP and ICG. 294, 303 Because the
canalicular excretion mechanism is normal in these patients, radiologic visualization of the biliary tree by
cholecystographic agents is normal. Jaundice and occasional neurologic impairment are the only
abnormal physical findings. Liver biopsy reveals normal histology. In several patients, pigment plugs were
observed in bile canaliculi and bile ducts (Fig. 125-15). 294, 301, 303 Pigment stones have been found in
several cases. The pigment plugs and stones probably result from biliary excretion of unconjugated
bilirubin as an effect of long-term phototherapy. Electron microscopy of the liver reveals no specific
pathologic change. 304

High-power view (hematoxylin and eosin; magnification ×650) of a liver biopsy obtained from a patient
with Crigler-Najjar syndrome type I during pregnancy. A portal area is shown with portal vein (PV) and a
bile ductule (B) containing amorphous material, which appeared to be bilirubin. (Reprinted with permission
from Wolkoff et al. 299 )

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 37 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Molecular Defect in Crigler-Najjar Syndrome Type 1.


Of the numerous isoforms of UGT, UGT1A1 (bilirubin-UGT1) is the only isoform that contributes
significantly to bilirubin metabolism in humans (GenBank AJ005162). 196 Genetic lesions of exons
constituting the UGT1A1 gene that cause Crigler-Najjar syndrome type 1 were first described in
1992. 196, 305–307 Subsequently, several laboratories have reported the genetic lesions in over 60 patients
with Crigler-Najjar syndrome type 1 by polymerase chain reaction–based DNA sequence analysis. These
genetic lesions and their relationship to Crigler-Najjar syndrome have been reviewed. 308 The genetic
lesions can be deletions, insertions, missense mutations, or premature stop codons, and can be located in
any of the five exons comprising the UGT1A1 mRNA (Fig. 125-16). 308 In one case the mutation is located
in the hydrophobic signal peptide that is cleaved off from the enzyme during synthesis. 309 In two other
cases, the mutation was located within the intronic regions of the gene at a splice donor and a splice
acceptor region, respectively. 310 These mutations result in the utilization of a cryptic splice site within an
exon, which leads to the splicing out of a part of the exon. In most families, with or without known
consanguinity, both alleles carry identical mutations, although different genetic lesions on the two alleles
are not rare. 308 A 13-nucleotide deletion in exon 2 of UGT1A1 has been observed in three unrelated
families of different ethnic origins. Crigler-Najjar syndrome type 1 is relatively common among the Amish
and Mennonite communities of Lancaster County, Pennsylvania. 311 All Crigler-Najjar syndrome type 1
patients within these related communities carry identical mutations in exon 1, indicating a founder effect.
Similarly, four apparently unrelated families, originating from Punjab, India, have been fround to carry an
identical mutation (unpublished observation, J. Roy Chowdhury and N. Roy Chowdhury). In cases where
the genetic lesion is located in exon 1, only bilirubin-UGT (UGT1A1) activity is expected to be abnormal,
whereas when the mutation affects one of the common region exons (exons 2 to 5), all UGT isoforms
expressed from the UGT1A locus are expected to be abnormal.

Genetic lesions causing Crigler-Najjar syndrome type I, Crigler-Najjar syndrome type II, and Gilbert
syndrome. Crigler-Najjar syndrome type I is produced by mutations, deletions, or insertions within the five
exons that constitute the UGT1A1 mRNA. These genetic lesions may cause premature stop codons or
substitution of a single amino acid. In two cases, there were mutations in the splice donor sequences on

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 38 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

intron 1 and splice acceptor region...

Diagnosis.
The combination of very high levels of serum unconjugated bilirubin and the absence of any other
abnormality of routine liver function tests is diagnostic of Crigler-Najjar syndrome type 1. Hemolysis alone
does not increase serum bilirubin levels beyond 6 to 8 mg/dl. Differential diagnosis includes Crigler-Najjar
syndrome type 2, with or without coexisting hemolysis. Although serum bilirubin levels are relatively lower
in Crigler-Najjar syndrome type 2, ranges of bilirubin concentration in the two disorders overlap. In most
cases, serum bilirubin concentrations are reduced by more than 25 percent after phenobarbital
administration (60 to 120 mg for 14 days) in Crigler-Najjar syndrome type 2, but not in type 1. 301 The two
types of Crigler-Najjar syndrome can be differentiated conveniently by chromatographic analysis of bile
collected from the duodenum through a perorally placed duodenal catheter or an upper gastrointestinal
endoscope. In bile from patients with Crigler-Najjar syndrome type 1, bilirubin glucuronides are absent or
are present in traces only (in concentrations less than that of unconjugated bilirubin). In contrast, in
Crigler-Najjar syndrome type 2 significant amounts of conjugated bilirubin are excreted in bile, although
the proportion of bilirubin diglucuronide is reduced (see below in Crigler-Najjar syndrome type 2). Liver
biopsy is not necessary for diagnosis, unless a coexisting liver disease is suspected. If a biopsy is
performed, UGT activity toward bilirubin can be determined, and is expected to be undetectable. The
diagnosis can be made also on the basis of genetic analysis of DNA extracted from blood, buccal
scrapings, or any tissue. The five exons of the UGT1A1 gene, and the flanking intronic sequences are
amplified by polymerase chain reaction and the nucleotide sequences are determined. 305 If the genetic
lesion matches one of the previously identified lesions associated with Crigler-Najjar syndrome type 1, the
diagnosis is established. If a new mutation is found that predicts the truncation of a major portion of the
enzyme or causes a frame-shift, the diagnosis is also established. When a novel mutation predicts the
substitution of a single amino acid residue, the mutation can be generated in an expression plasmid by
site-directed mutagenesis and the effect of the mutation can be determined after transfection of the
plasmid into African Green monkey kidney cell lines (COS cells). 196 Genetic analysis can be utilized to
detect heterozygous carriers of one Crigler-Najjar syndrome type 1 allele. Genetic analysis of chorionic
villus samples has been used successfully for prenatal identification of the Crigler-Najjar syndrome type 1
genotype in four fetuses in three families (J. Roy Chowdhury and N. Roy Chowdhury, unpublished
observation).

Animal Model for Crigler-Najjar Syndrome Type 1: The Gunn Rat.


The description by Gunn in 1938 of mutant Wistar rats with nonhemolytic unconjugated
hyperbilirubinemia 312 and the wisdom of the late Professor William E. Castle, Emeritus Professor of
Genetics at Berkeley, who maintained the mutants for over 15 years, have resulted in major advances in
understanding bilirubin metabolism, transport, and encephalopathy. 313, 314 Jaundice in these animals is
inherited as an autosomal recessive trait. 312 Heterozygotes are anicteric. Depending on the background
strain against which a Gunn rat colony is maintained, homozygous Gunn rats have bilirubin levels that
range from 3 to 20 mg/dl. Serum bilirubin is all unconjugated, 313 there is no bilirubinuria, and bile is light
yellow in color due to the excretion of small amounts of unconjugated bilirubin. 313 Liver histology is
normal. 315

Neurologic Lesions.
Homozygous Gunn rats are prototypes of Crigler-Najjar syndrome type I and frequently develop
kernicterus. 299, 315–317 The Gunn rat is the only experimental model in which endogenously produced
bilirubin results in neuropathologic lesions and neurologic deficits. Cytoplasmic neuronal changes develop
in these rats on the third day of life, and by 2 weeks, degeneration of Purkinje cells and other neurons

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 39 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

occurs. The degenerative changes begin by enlargement of mitochondria and formation of membranous
cytoplasmic bodies. By 8 days of age, many mitochondria contain glycogen. 316, 317 When a clinically
healthy Gunn rat is killed and rapidly perfused with saline or formalin, the brain does not show yellow
staining. Administration of sulfadimethoxine, a drug that competes with bilirubin for binding to albumin, to
14-day-old animals results in neurologic deterioration and yellow staining in the brain. 318

Urinary Concentration Defect.


Gunn rats cannot concentrate urine and do not tolerate water deprivation. 319 The renal medullary bilirubin
concentration is high and interferes with sodium and water transport. 319 Occasionally, renal papillary
necrosis occurs. 320 Treatment of rats with agents or methods designed to lower serum bilirubin, such as
cholestyramine, agar, or phototherapy, may ameliorate the renal lesion. 319 Similar concentrating
problems have not been described in patients with Crigler-Najjar syndrome type I, although bilirubin is
deposited in the kidney. 321

Abnormalities of UGT Activity and Their Relationship to Genetic Lesions.


The rat UGT1A locus closely resembles its human counterpart in exon organization. 322 The genetic lesion
in Gunn rats consists of the deletion of a single guanosine residue in the common region exon, exon 4,
which creates a frameshift and deletes a large segment of the C-terminal domain of all UGT isoforms that
are expressed from this locus. 322–324 Consequently, not only bilirubin glucuronidation is absent in this
strain, other UGT isoforms expressed from the UGT1A locus are also truncated and nonfunctional. 325 As
in patients with Crigler-Najjar syndrome type 1, Gunn rats lack bilirubin conjugates in bile. 252, 313, 315 Gunn
rat livers lack UGT activity toward digitoxigenin monodigitoxoside. 326 The transferase activity toward
4-nitrophenol is present at a lower level, 327 suggesting that this activity is partly provided by UGT isoforms
expressed from other loci. Several functionally normal forms of UGT have been isolated from the Gunn rat
liver, 328 indicating that UGT isoforms expressed from genes that do not belong to the UGT1A locus are
normal. UGT activities for aniline, 329 steroid substrates, 330 and thyroid hormone 331 are normal in Gunn
rat liver. Biliary excretion of substances that do not require glucuronidation, such as BSP, 313 phenol
red, 332 and exogenously administered conjugated bilirubin, 125 are normal in Gunn rats.

Treatment of Crigler-Najjar Syndrome Type 1.


Unconjugated hyperbilirubinemia in Crigler-Najjar syndrome type 1 is usually associated with bilirubin
encephalopathy (kernicterus). Conventional treatment is designed to reduce serum bilirubin levels. Unlike
in patients with Crigler-Najjar syndrome type II and Gilbert syndrome, the serum bilirubin levels are not
significantly reduced by the administration of UGT enzyme inducing agents, such as phenobarbital. 301, 302

Phototherapy has received widespread acceptance. It is the major treatment for icteric newborns whose
serum bilirubin concentrations place them at risk for kernicterus. 253, 299, 333 Experience with phototherapy
in older children and adults is limited to patients with Crigler-Najjar syndrome type 1 and occasional cases
of Crigler-Najjar syndrome type 2. An array of 140-W fluorescent lamps with devices for shielding the eyes
has been used effectively. However, about the time of puberty, phototherapy becomes relatively less
effective because of thickening of the skin, increased skin pigmentation, and decreased surface area in
relation to body mass. 299 Phototherapy converts a fraction of bilirubin IXα-ZZ into geometric and
structural isomers, that are excreted in bile (see earlier section on Chemistry of Bilirubin). A portion of the
unconjugated bilirubin excreted in bile may be reabsorbed in the small intestine. Oral administration of
agar, cholestyramine, or calcium salts 333 enhances the effect of phototherapy, probably by inhibiting the
reabsorption of unconjugated bilirubin.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 40 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Plasmapheresis is the most efficient method for rapidly reducing serum bilirubin concentration during
crisis (Fig. 125-17). 296, 299, 303 This procedure takes advantage of the fact that bilirubin is tightly bound to
serum albumin and removal of albumin results in the removal of equimolar amounts of bilirubin.

Summary of the hospital course of a 19-year-old patient with Crigler-Najjar syndrome type I who was
admitted with acute bilirubin encephalopathy. Before hospitalization, the patient’s serum bilirubin ranged
between 35 and 45 mg/dl. After an initial course of plasmapheresis and maintenance phototherapy, serum
bilirubin was maintained between 10 and 15 mg/dl. (Reprinted with permission from Wolkoff et ...

Liver Transplantation.
Orthotopic or auxiliary liver transplantation rapidly normalizes serum bilirubin levels. Currently, liver
transplantation is considered the only definitive treatment for Crigler-Najjar syndrome type 1. 334 Although
these procedures are not without risk in these individuals, some investigtors have suggested prophylactic
liver transplantation in patients with Crigler-Najjar syndrome type 1 to avoid the risk of kernicterus, which
once established, may not be fully reversible. 334

Hepatocyte Transplantation.
Because the liver is structurally normal in Crigler-Najjar syndrome type 1 and in Gunn rats, alternatives to
the irreversible, expensive, and risky procedure of liver transplantation are being sought. Transplantation
of congeneic normal isolated hepatocytes into Gunn rats by infusion into the portal vein, 335 intrasplenic
injection, 336, 337 or intraperitoneal injection after attachment to microcarrier beads 338 has been used
successfully to provide partial correction of bilirubin-UGT deficiency in Gunn rats. It has been shown that
after intrasplenic injection, a great majority of the hepatocytes rapidly translocate to the liver, where, in the
absence of immune rejection, they exhibit long-term persistence. 336 The liver is the preferred site for
long-term survival and function of isolated hepatocytes. Following transplantation by intraportal infusion or

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 41 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

intrasplenic injection, hepatocytes migrate out of the sinusoidal space and integrate into the liver chords
within days. The transplanted cells survive and function for prolonged periods and respond to normal
proliferative stimuli. 337

Based on the extensive experience in Gunn rats, an 11-year-old girl with Crigler-Najjar syndrome was
transplanted with 7.5 billion isolated allogeneic primary human hepatocytes by infusion through a
percutaneously placed portal venous catheter. 339, 340 Tacrolimus was used for prevention of allograft
rejection. The procedure did not cause portal hypertension, and resulted in excretion of bilirubin
glucuronides in bile and led to approximately 50 percent reduction of serum bilirubin concentration over
the course of several months. The hypobilirubinemic effect persists to date 16 months after the
transplantation. Although the experience is limited to one case only, transplantation of isolated
hepatocytes appears to be a safe and relatively inexpensive alternative to liver transplantation in patients
with Crigler-Najjar syndrome type 1.

Degradation of Bilirubin by Bilirubin Oxidase.


Bilirubin oxidase from Myrothecium verrucaria catalyzes the oxidation of bilirubin with oxygen to a
colorless derivative. Perfusion of filters packed with bilirubin oxidase immobilized on agarose with rat or
human blood containing bilirubin resulted in the degradation of 90 percent of the pigment in a single
pass. 341 When blood from a Gunn rat was passed through the column and returned to the venous
circulation, the serum bilirubin level was reduced to 50 percent in 30 min. However, the use of such
columns may be associated with removal of formed elements of blood. As an alternative, systemic
administration of bilirubin oxidase has been considered. To circumvent the short half-life of bilirubin
oxidase in circulation (2.5 min), the enzyme has been covalently linked to polyethyleneglycol. 342 The
linked enzyme has a plasma half-life of 190 min in rats. A single intravenous injection of
polyethyleneglycol-conjugated bilirubin oxidase in Gunn rats resulted in substantial reduction of serum
bilirubin level for 3 h.

Induction of P450c.
As mentioned above, the induction of P450c with TCDD results in a decrease of serum bilirubin levels in
Gunn rats, presumably due to oxidation of bilirubin in the liver. This observation has stimulated the search
for more innocuous drugs for induction of this enzyme. Several naturally occurring indoles extracted from
cruciferous vegetables, such as cabbage, cauliflower, and brussels sprouts induce P4501A1 and
P4501A2 in rat liver and intestine. 343 Administration of indole-3-carbinol, an inducer of P4501A2, results
in a short-term reduction of serum bilirubin levels in children with Crigler-Najjar syndrome type 1. 343

Gene Therapy.
Because the metabolic defect in Gunn rats and in patients with Crigler-Najjar syndrome type I is caused
by molecular lesions of a single gene, introduction of a normal bilirubin-UGT would be an elegant potential
therapeutic method. Although there has been no clinical trial of gene therapy for this disease as yet,
significant progress has been made by experiments on Gunn rats. Introduction of a normal bilirubin-UGT
gene can be performed by ex vivo methods, introduction of the gene into the liver by in situ perfusion, or
systemic administration of vectors that are capable of carrying the gene to the liver. These approaches
are briefly mentioned below.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 42 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

In ex vivo gene therapy, liver cells harvested from a mutant subject by partial hepatectomy are established
in primary culture and transduced with a therapeutic gene using a method that permits stable gene
expression. 344 These cells are then transplanted into the same mutant subject. This method has been
used in low-density lipoprotein (LDL) receptor-deficient rabbits (Watanabe heritable hyperlipidemic strain),
with long-term reduction of plasma LDL cholesterol concentration. Because the cells are autologous,
immunosuppression is not needed. However, the number of hepatocytes that can be harvested and
transduced is limited, and because surgical resection is required, this procedure cannot be readily
repeated. Therefore, for the treatment of Gunn rats, hepatocytes have been conditionally immortalized
before transduction with the bilirubin-UGT gene and transplantation. 345 In addition to improving gene
transduction, this strategy could potentially assure a long-term supply of phenotypically corrected
autologous hepatocytes. Additional research is required to assure that the transplanted cells should not
become transformed into malignant cells.

Methods are also being developed to directly introduce a normal bilirubin-UGT gene into the liver of
mutant Gunn rats, using viral or nonviral vectors. Recombinant murine leukemia viruses have been used
to transfer the gene into the liver by perfusing the liver with the recombinant vector after transiently
occluding inflow and outflow vessels. 346 Because gene transfer by the murine leukemia viruses requires
cell division, these vectors are not very efficient in transferring genes to the intact liver.

Adenoviral vectors have the advantage of spontaneously localizing to the liver after systemic
administration and the ability to transfer genes into nondividing cells with high efficiency. However, these
viruses are episomal and would require repeated administration for long-term gene therapy. Unfortunately,
the vectors evoke both cellular and humoral immunity in the host, precluding repeated injection. Several
approaches are being developed to tolerize the host to antigens contained in the vector. These include
administration in neonatal animals, 347 intrathymic inoculation, 348 or oral tolerization to adenoviral
antigens. 349, 350 Specific tolerization to the vector proteins permits long-term gene therapy in Gunn rats
using adenoviral vectors. Whether these methods would be effective in nonhuman primates and humans
remains to be examined. Newer viral vectors, such as recombinant SV40 and recombinant lentiviruses
that can transfer genes into nondividing cells are being developed for trial in Gunn rats.

Receptor-mediated gene delivery to the liver using carrier proteins that deliver systemically administered
DNA into the liver also have been used in the treatment of Gunn rats. 351 Transgenes introduced in this
manner are expressed transiently. The expression can be prolonged to several months by inducing cell
proliferation by partial hepatectomy 352 or by pharmacologic disruption of microtubules. 353

Site-directed gene conversion is a recently described strategy. RNA-DNA chimeric molecules have been
used in an effort to repair genetic mutations. 354 These chimeric molecules are designed to align with
specific sequences within the genome and to create a single mismatch. This triggers the host mismatch
repair system to correct the mutation. One study has reported that it may be possible to insert the missing
guanosine residue into the Gunn rat bilirubin-UGT (UGT1A1) gene by this method. 355

Crigler-Najjar Syndrome Type II (Arias Syndrome)

Clinical Findings.
Crigler-Najjar syndrome type II, otherwise known as Arias syndrome, is phenotypically similar to
Crigler-Najjar syndrome type I, except that the serum bilirubin concentration is usually below 20 mg/dl, the
prognosis is much less severe, serum bilirubin levels are usually reduced after administration of
bilirubin-UGT inducing agents, such as phenobarbital, and the bile contains significant amounts of bilirubin

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 43 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

glucuronides (Table 125-1). This disorder was first described by Arias in 1962 in a study of chronic
unconjugated hyperbilirubinemia in eight patients between 14 and 52 years of age. 356 Although half the
patients were icteric before the age of 1 year, one patient was 30 years old before jaundice was noted. In
these patients, serum bilirubin concentration ranged from 8 to 18 mg/dl. Each had reduced hepatic
glucuronosyltransferase activities using bilirubin, o-aminophenol, or 4-methylumbelliferone as glucuronide
acceptor. Survival of 51 Cr-labeled red cells was normal. 301 All patients were clinically normal, apart from
icterus, except for a 43-year-old woman with a neurologic syndrome resembling kernicterus. The patient
died at the age of 44. Autopsy revealed a histologically normal liver. The brain was small and lacked
bilirubin staining but demonstrated the typical histology of kernicterus.

Several other cases of neurologic abnormality in Crigler-Najjar syndrome type II have been described
subsequently. Three brothers had Crigler-Najjar syndrome type II for over 50 years. 357 Two were
neurologically normal. The third had a slight bilateral intention tremor and nonspecific abnormalities on
electroencephalogram. These nonspecific neurologic changes had not been noted previously. Another
patient was a 15-year-old boy who was icteric from the second day of life. 358 Total serum bilirubin was
24 mg/dl at 10 months and averaged approximately 15 mg/dl thereafter. Development was normal,
although psychological testing revealed a perceptual deficit and slightly subnormal intelligence. At age 13,
following surgery for acute appendicitis, the serum bilirubin increased to 40 mg/dl, and diplopia,
generalized seizures, confusion, and an abnormal electroencephalogram developed. He was treated for
hyperbilirubinemia and, after recovering from surgery, resumed a bilirubin level of 15 mg/dl. His neurologic
status returned to baseline and he has remained well.

Laboratory Tests.
As in Crigler-Najjar syndrome type I, results of laboratory examination are normal except for elevated
serum bilirubin, which is usually less than 20 mg/dl but may be as high as 40 mg/dl during fasting 357 or
intercurrent illness. 358 Serum bilirubin is unconjugated, and there is no bilirubinuria. The bile is pigmented,
although less than 50 percent of estimated daily bilirubin production is excreted into bile. 301, 358 Although
over 90 percent of conjugated bilirubin in normal bile is bilirubin diglucuronide, the major pigment in this
syndrome is bilirubin monoglucuronide. 358, 359 The liver has markedly reduced bilirubin-UGT activity. 358

Effect of Phenobarbital.
The reduced levels of hepatic bilirubin-UGT activity in Crigler-Najjar syndrome type II suggested that an
inducer of microsomal enzymes could ameliorate the hyperbilirubinemia. 356 Subsequent studies revealed
that serum bilirubin concentrations are reduced significantly (greater than 25 percent) following treatment
with phenobarbital (Fig. 125-18). 301 Similar results were obtained with other liver microsomal enzyme
inducers. 360–364 The response to phenobarbital treatment differentiates Crigler-Najjar syndrome type 1, in
which there is no response, from Crigler-Najjar syndrome type II (Fig. 125-19). 301 Although phenobarbital
has been used commonly for inducing hepatic bilirubin-UGT activity, clofibrate is equally effective and is
associated with fewer side effects. 364 In some patients, the differentiation from Crigler-Najjar syndrome
type 1 may be difficult on the basis of serum bilirubin levels and phenobarbital response. 363 In these
cases, the differentiation can be made on the basis of chromatographic analysis of pigments excreted in
bile.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 44 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Effect of phenobarbital administration on serum bilirubin concentration, menthol tolerance test, and fecal
urobilinogen excretion in a patient with Crigler-Najjar syndrome type II. (Reprinted with permission from
Arias et al. 299 )

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 45 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Differentiation of types I and II Crigler-Najjar syndrome on the basis of response to phenobarbital. All
patients had chronic unconjugated hyperbilirubinemia and were treated for at least several weeks with
phenobarbital. (Reprinted with permission from Arias et al. 301 )

Inheritance.
Crigler-Najjar syndrome type II runs in families. 301, 356 There is no sex predilection. Although the pattern
of inheritance was not certain for many years, genetic analysis clearly establishes an autosomal-recessive
pattern of inheritance. In some heterozygous carriers, the coexistence with a variant promoter associated
with Gilbert syndrome may lead to intermediate levels of jaundice (see later section on Gilbert Syndrome),
which caused confusion about the mode of inheritance in the past.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 46 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Molecular Mechanism.
As in Crigler-Najjar syndrome type I, Crigler-Najjar syndrome type II is caused by mutations of one of the
five exons that encode bilirubin-UGT (UGT1A1) (Fig. 125-16). 308, 365 However, in Crigler-Najjar syndrome
type II, the genetic lesion for at least one allele always consists of point mutations that result in
substitution of a single amino acid. Such substitutions result in marked reduction, but not a total loss of
bilirubin-UGT activity. 365 Phenobarbital works presumably by induction of the residual bilirubin-UGT
activity.

Gilbert Syndrome

Clinical Findings.
The syndrome, described by Gilbert in 1901, also has been called constitutional hepatic dysfunction and
familial nonhemolytic jaundice. 366 It is characterized by mild, chronic, unconjugated hyperbilirubinemia
(Table 125-1). Familial occurrence is common, 367 although many patients present as isolated cases.

Typically, Gilbert syndrome is diagnosed in young adults, who present with mild, predominantly
unconjugated hyperbilirubinemia. Serum bilirubin levels are usually less than 3 mg/dl and fluctuate with
time. Bilirubin concentrations can increase during intercurrent illness, but can be normal at other times.
Aside from icterus, physical examination is normal. Some patients complain of vague constitutional
symptoms, including fatigue and abdominal discomfort 368 ; these symptoms are probably unrelated to
bilirubin metabolism and may be manifestations of anxiety. Newly presenting patients are rarely
symptomatic. Results of routine laboratory tests are normal except for elevated serum bilirubin
concentrations. There is no elevation of serum alkaline phosphatase or aminotransferase activities. Oral
cholecystography allows visualization of the gallbladder. Although percutaneous liver biopsy is not
routinely indicated in patients with Gilbert syndrome, liver histology is normal, except for a nonspecific
accumulation of lipofuscin pigment in the centrilobular zones. Electron microscopic studies have not
revealed consistent ultrastructural alterations. Hepatic bilirubin-UGT activity is reduced to approximately
30 percent of normal (Fig. 125-20). 369, 370

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 47 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Hepatic bilirubin-UGT activity in patients with hepatitis, cirrhosis, and Gilbert syndrome. The hatched area
indicates the normal range. (Reprinted with permission from Black and Billing. 369 )

Organic Anion Transport.


Several studies of the disappearance of plasma bilirubin after intravenous injection into patients with
Gilbert syndrome have demonstrated reduced clearance (Fig. 125-21). Multicompartmental analysis
suggests that reduced plasma clearance results from reduction in hepatic bilirubin uptake as well as
bilirubin conjugation. 371 Goresky and associates determined the initial plasma disappearance of
radiolabeled bilirubin and then determined an initial space of distribution by dividing the injected dose by
the plasma volume as determined after radiolabeled albumin injection. 372 The initial plasma
disappearance of bilirubin was as rapid in patients with Gilbert syndrome as in normal subjects,
suggesting that bilirubin uptake is normal in Gilbert syndrome. Although plasma disappearance of organic
anions other than bilirubin is usually normal in Gilbert syndrome (Fig. 125-22), two subsets were
described in which BSP 373 and ICG 374 plasma disappearance is abnormal (Fig. 125-23). Because the
excretion of neither of these compounds depends on bilirubin-UGT activity, these organic anion clearance
abnormalities may not be related to the reduced bilirubin-UGT activity, which is a constant feature of
Gilbert syndrome.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 48 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Plasma disappearance of a trace dose of [3H]bilirubin after intravenous administration to patients with
Gilbert syndrome and to normal volunteers. There is no overlap between the two groups for the first 16 h
after injection. (Reprinted with permission from Berk et al. 371 )

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 49 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Plasma concentration of BSP 45 min after intravenous administration of 5 mg/kg to normal individuals and
patients with Gilbert syndrome. In one subset of patients, BSP retention was elevated. (Reprinted with
permission from Berk et al. 373 )

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 50 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Plasma BSP disappearance curves in three patients with Gilbert syndrome. The shaded area indicates
the normal range. Curve 1 is indistinguishable from normal. Curves 2 and 3 are representative of the two
subtypes of abnormal BSP disappearance seen in some patients with otherwise typical Gilbert syndrome.
(Reprinted with permission from Berk et al. 373 )

The serum bilirubin levels fluctuate in patients with Gilbert syndrome. Factors such as intercurrent illness,
physical exertion, and stress have been implicated, and a relationship to the menstrual cycle has been
reported in two women. 375 A 48-h fast exaggerates the unconjugated hyperbilirubinemia of Gilbert
syndrome. 376 Serum bilirubin levels in normal individuals 377 and in individuals with other hepatobiliary
disorders also increase with fasting. 378 Thus, the fasting test appears to be of limited use in the
differential diagnosis of Gilbert syndrome. The mechanism of fasting-induced hyperbilirubinemia is
unclear. Studies in normal rats revealed no change in hepatic bilirubin-UGT activity during fasting, 378
although there was reduced activity of UDP-glucose dehydrogenase resulting in reduced hepatic content
of UDP-glucuronic acid. 379 Fasting also must affect hepatic disposition of bilirubin at a step other than
conjugation, because fasting exacerbates hyperbilirubinemia in homozygous Gunn rats. 380 It may be a
result of several factors, and a role for increased serum nonesterified fatty acid concentration has been
suggested. 381

Intravenous administration of nicotinic acid also has been proposed as a provocative test for the diagnosis
of Gilbert syndrome. 382 Its diagnostic value is controversial, and it does not clearly separate patients with
Gilbert syndrome from normal subjects or those with hepatobiliary disease. Unconjugated
hyperbilirubinemia following nicotinic acid administration does not occur after splenectomy, 382 suggesting
that nicotinic acid–induced unconjugated hyperbilirubinemia may result from increased erythrocyte fragility
and enhanced splenic heme oxygenase activity, leading to augmentation of splenic bilirubin formation. 383

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 51 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Conventionally, The diagnosis of Gilbert syndrome has been applied to individuals with mild unconjugated
hyperbilirubinemia without evidence of hemolysis or structural liver disease. However, the coexistence of
clinical or subclinical hemolysis may exacerbate the hyperbilirubinemia, thereby bringing the patient to the
attention of the physician. The relative content of bilirubin monoglucuronide and diglucuronide in bile is of
use in the diagnosis of Gilbert syndrome. Similar to findings in patients with Crigler-Najjar syndrome type
II and heterozygous Gunn rats, the proportion of bilirubin monoglucuronide is increased in the bile from
patients with Gilbert syndrome. Normally approximately 90 percent of bilirubin excreted in bile is in the
form of bilirubin diglucuronide, 7 percent is bilirubin monoglucuronide, and 4 percent is unconjugated
bilirubin. 359

The other significant pigment is a bilirubin monoglucoside-monoglucuronide diester. In Gilbert syndrome


the percentage of bilirubin monoglucuronide increases to 14 to 34 percent. 359

Molecular Mechanism of Gilbert Syndrome.


A variant TATAA element within the promoter region upstream to exon 1 of the gene encoding bilirubin
UGT (UGT1A1) has been found to be associated with Gilbert syndrome. 384 Sequence of the TATAA
element in normal subjects is A[TA] 6 TAA, whereas patients with Gilbert syndrome are homozygous for a
longer TATAA element, A[TA] 7 TAA (Fig. 125-16). Although this variant promoter was found in all Gilbert
syndrome patients studied in the United States and Europe, all subjects who are homozygous for the
variant TATAA box do not exhibit hyperbilirubinemia. Expression of the Gilbert genotype appears to
require a relatively high level of bilirubin production, in addition to the reduced expression of bilirubin-UGT.
For example, most patients diagnosed to have Gilbert syndrome are men, probably because the daily
production of bilirubin is greater in men than in women. The incidence of the Gilbert genotype is common
in the United States and Europe, approximately 9 percent of the general population being homozygous for
the variant promoter, and approximately half the population carrying at least one copy of the variant
promoter.

Because of the frequency of the Gilbert type promoter, some heterozygous carriers of a Crigler-Najjar
syndrome type I or II mutation are likely to carry the variant TATAA box. If the Gilbert type TATAA element
is present on the structurally normal allele of a heterozygous carrier of a Crigler-Najjar syndrome type I or
II mutation, the expression of the only normal allele will be reduced to approximately 30 percent of normal,
resulting in an intermediate level of jaundice. 385 This explains the frequent finding of intermediate levels of
hyperbilirubinemia in the family members of patients with Crigler-Najjar syndrome types I and II.

Inheritance.
Subjects who are heterozygous for the Gilbert genotype have higher average serum bilirubin
concentrations than subjects who are homozygous for the normal TATAA box. However, all patients in the
United States and Europe who were clinically diagnosed to have Gilbert syndrome were homozygous for
the variant promoter. On the basis of this, Gilbert syndrome may be considered to have an
autosomal-recessive mode of transmission.

Animal Model of Gilbert Syndrome: The Bolivian Squirrel Monkey.


The Bolivian population of squirrel monkeys (Saimiri siureus ) have a higher postcibal serum
unconjugated bilirubin concentration and a greater degree of fasting hyperbilirubinemia than does a
closely related Brazilian population. 386 Compared with the Brazilian population, Bolivian monkeys have
slower plasma clearance of intravenously administered bilirubin, a lower level of hepatic bilirubin-UGT

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 52 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

activity, and an increased bilirubin monoglucuronide to diglucuronide ratio in bile. 386 The two populations
of squirrel monkeys have a comparable erythrocyte life span and hepatic glutathione-S-transferase
activity. 386 In these respects, the Bolivian squirrel monkeys are a model of human Gilbert syndrome.
Fasting hyperbilirubinemia is rapidly reversed by oral or intravenous administration of carbohydrates, but
not by lipid administration. 387

DISORDERS OF BILIRUBIN METABOLISM RESULTING IN PREDOMINANTLY CONJUGATED


HYPERBILIRUBINEMIA

Dubin-Johnson Syndrome

Clinical Findings.
In 1954, Dubin and Johnson 236 and Sprinz and Nelson 388 described patients with chronic nonhemolytic
jaundice. The liver was grossly black, but the histology was normal except for an unidentified pigment in
hepatocytes. Subsequently, this disorder has been described in both sexes in virtually all nationalities and
races. 381–391 Dubin-Johnson syndrome is rare, except in Jews of Middle Eastern origin. Most reports
consist of individual cases or small groups. The largest series of 101 cases 391 was reported from Israel
on the basis of hospital records from 1955 to 1969. Of these 101 cases, 74 came from families that
immigrated from Iran, Iraq, and Afghanistan; nine were of Moroccan origin, and seven were
European-Ashkenazim. Among Jews of Persian origin, the incidence is 1:1300. 391 In this population,
Dubin-Johnson syndrome is associated with clotting factor VII deficiency. 392

The syndrome is clinically characterized by mild, predominantly conjugated hyperbilirubinemia (Table


125-2). Except for jaundice, physical examination is normal. Occasionally a patient may have
hepatosplenomegaly. Mild constitutional complaints such as vague abdominal pains and weakness occur,
but for the most part patients are asymptomatic. 389–391 Pruritus is absent in Dubin-Johnson syndrome,
and serum bile acid levels are normal. 390, 393 The degree of icterus is increased by intercurrent illness,
oral contraceptives, and pregnancy (Fig. 125-24). 390 Dubin-Johnson syndrome is rarely detected before
puberty, although cases have been reported in neonates. 394, 395 Often the disorder is not noted until a
woman becomes pregnant or receives oral contraceptives with conversion of mild chemical
hyperbilirubinemia into overt jaundice. 390

Table 125-2: Principal Differential Characteristics of Inherited Chronic Conjugated


HyperbilirubinemiasSeparate Window

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 53 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Dubin-Johnson Syndrome Rotor Syndrome

Appearance of liver Grossly black Normal

Dark pigment; predominantly in


Histology of liver Normal; no increase in pigmentation
centrilobular areas; otherwise normal

Usually 1.5–5 mg/dl, occasionally as Usually elevated, occasionally as


Serum bilirubin high as 20 mg/dl; predominantly direct high as 20 mg/dl; predominantly
reacting direct reacting

Routine liver
Normal except for bilirubin Normal except for bilirubin
function tests

45-min plasma BSP Normal or elevated; secondary rise at Elevated; no secondary increase at
retention 90 min 90 min

BSP infusion T max Virtually zero; S normal T max and S both reduced
studies

Oral Usually does not visualize the


Usually visualizes the gallbladder
cholecystogram gallbladder

Urinary Normal total; >80% as coproporphyrin Elevated total; elevated proportion


coproporphyrin I of coproporphyrin I but <80%

Mode of inheritance Autosomal recessive Autosomal recessive

Prevalence Uncommon (1:1300 in Persian Jews) Rare

Prognosis Benign Benign

Animal model Mutant TR – or EHBR rats None

Mutant Corriedale sheep

Mutant golden lion tamarind monkeys

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 54 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Exacerbation of conjugated hyperbilirubinemia by oral contraceptive administration in a patient with


Dubin-Johnson syndrome. Similar findings may be seen in pregnancy.

Laboratory Tests.
Routine laboratory examination 389, 391 reveals normal complete blood count, serum albumin, cholesterol,
transaminases, alkaline phosphatase, and prothrombin time. Serum bilirubin is usually between 2 and
5 mg/dl but can be as high as 20 to 25 mg/dl. Bilirubinuria is frequent, and 50 percent or more of total
serum bilirubin is conjugated. The serum bilirubin level fluctuates, and frequently individual determinations
may be normal. Both unconjugated and conjugated bilirubin accumulate in serum, and more than half of
the bilirubin gives a direct van den Bergh reaction. Prolonged retention of bilirubin glucuronides in plasma
results in the formation of irreversible adducts of bilirubin with plasma proteins, particularly albumin. The
bilirubin-albumin adduct, termed δ-bilirubin, is found in the serum of patients with Dubin-Johnson
syndrome, and various hepatobiliary disorders that are associated with prolonged conjugated
hyperbilirubinemia. δ-Bilirubin is not excreted in bile and urine and gives a direct van den Bergh reaction.

Excretion of Dyes Used for Imaging of the Biliary System.


Oral cholecystography, even using a double dose of contrast material, usually does not allow visualization
of the gallbladder, although visualization may occur 4 to 6 h after intravenous injection of iodipamide. 396

Hepatic Pigmentation.
On direct inspection, the liver is black. Light microscopy reveals normal histology except for accumulation
of a dense pigment, which on electron microscopy appears to be contained within lysosomes. 397, 398
Histochemical staining characteristics and physicochemical properties of extracted pigment resemble
those of melanin. 399, 400 In the mutant Corriedale sheep, an animal model of Dubin-Johnson syndrome,
the hepatic pigment resembles melanin histochemically. Studies performed in mutant Corriedale sheep
infused with [ 3 H]-epinephrine revealed reduced biliary excretion of radioactivity and demonstrated
incorporation of the isotope into the hepatic pigment, 401 which is consistent with the pigment being a
melanin-like derivative. However, electron spin resonance spectroscopy suggests that the Dubin-Johnson
pigment differs from authentic melanin, and could be composed of polymers of epinephrine
metabolites. 402 The TR rat, another animal model for the Dubin-Johnson syndrome, has an identical
phenotype except for the absence of pigmentation in the liver. 225 Biliary excretion of [ 3 H]-epinephrine is
also disturbed in this rat. 403 When these animals were fed a diet enriched in aromatic amino acids
(phenylalanine, tyrosine, and tryptophan), lysosomal pigmentation developed, which was absent in normal
rats. Impaired excretion of anionic metabolites of tyrosine, phenylalanine, and tryptophan in the TR − liver
may result in their retention, oxidation, polymerization, and subsequent lysosomal accumulation. 403 One
study of computerized tomography of the liver revealed that attenuation values were significantly higher in
patients with Dubin-Johnson syndrome as compared with normal controls, although there was
considerable overlap between the two groups. 404 The possible relationship of the liver cell pigment to this
finding is not known. The degree of hepatic pigmentation may be variable in individuals with the
Dubin-Johnson syndrome. Some variability in pigmentation may be due to occurrence of coincidental
disease such as acute viral hepatitis, in which the pigment is cleared from the liver only to reaccumulate
slowly after recovery. 405

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 55 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Organic Anion Transport Defect.


In Dubin-Johnson syndrome, initial plasma disappearance of
bilirubin, 406, 407 BSP, 391, 407, 408 dibromosulphthalein (DBSP), 408 ICG 407, 408 and 125 I-labeled rose
bengal 408 following intravenous administration are usually normal. Of diagnostic significance is that in
approximately 90 percent of patients, the plasma BSP concentration is higher 90 min after intravenous
administration than at 45 min (Fig. 125-25). 390, 391, 395 This is due to reflux of conjugated BSP from the
liver cell into the circulation. 409 This secondary increase is not seen following intravenous administration
of other organic anions such as DBSP, 125 I-labeled rose bengal, and ICG, which are not conjugated prior
to excretion by the hepatocytes. 407, 408 A similar secondary increase has been described following
intravenous administration of unconjugated bilirubin. 402, 407 Although the secondary rise of plasma BSP is
characteristic of Dubin-Johnson syndrome, it is not diagnostic and occurs in other hepatobiliary
disorders. 410

Typical BSP plasma disappearance curve in a patient with Dubin-Johnson syndrome. A secondary
increase occurs 45 min after the intravenous injection of the dye. (Reprinted with permission from Erlinger
et al. 408 )

Studies of BSP transport during constant intravenous infusion reveal that the T max is reduced to only 10
percent of normal, and the relative hepatic storage capacity is normal. 390, 391 This finding was also
demonstrated directly in a patient with Dubin-Johnson syndrome who had a biliary fistula. 237 In this
patient, dehydrocholate choleresis did not augment biliary BSP excretion. Similar studies of BSP transport
have been performed in phenotypically normal parents and children (i.e., carriers) of Dubin-Johnson
syndrome patients, and it was found to be normal. 390

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 56 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Inheritance and Urinary Coproporphyrin Excretion.


The familial nature of Dubin-Johnson syndrome was noted in its initial descriptions, but its mode of
inheritance was unclear. 389 Subsequently, it was observed that the ratio of coproporphyrin I to
coproporphyrin III excreted in the urine of patients with Dubin-Johnson syndrome is greater than that
found in other hepatobiliary disorders. 411 Of the two isomeric forms of coproporphyrin, isomer III is the
precursor of heme, whereas isomer I is a metabolic by-product without known function. 412 Coproporphyrin
isomers I and III are normally found in urine, where approximately 75 percent of total urinary
coproporphyrin is coproporphyrin III. In Dubin-Johnson syndrome, total urinary coproporphyrin excretion is
normal, but over 80 percent is coproporphyrin I. 411, 413 Urinary coproporphyrin excretion has been
determined in phenotypically normal relatives of patients with Dubin-Johnson syndrome (Fig.
125-26). 413–415 In obligate heterozygotes (i.e., unaffected parents and children of patients with
Dubin-Johnson syndrome), total urinary coproporphyrin excretion was reduced by 40 percent as
compared with normal control subjects. 413–415 This was due to a 50 percent reduction in coproporphyrin
III excretion. The proportion of coproporphyrin I in urine was intermediate between results in controls and
in patients with Dubin-Johnson syndrome. Analysis of data from studies revealed that with respect to
urinary coproporphyrin excretion, Dubin-Johnson syndrome is inherited as an autosomal-recessive
characteristic (Fig. 125-26 and Table 125-2). 414 A similar mode of inheritance was determined in a study
of BSP and bilirubin metabolism in 173 sibs of 44 patients with Dubin-Johnson syndrome. 416 No other
hepatobiliary disorder or porphyria has been described in which total urinary coproporphyrin excretion is
normal, with over 80 percent of the total as coproporphyrin I. In the presence of a consistent history and
physical examination, urinary coproporphyrin excretion appears to be diagnostic of this disorder.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 57 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Pedigree of a family in which consanguinity resulted in three children with Dubin-Johnson syndrome
(generation V). Solid symbols indicate individuals with Dubin-Johnson syndrome. Partially filled symbols
indicate phenotypically normal individuals with urinary coproporphyrin excretion in the heterozygous
range. Open symbols represent phenotypically normal individuals with normal urinary coproporphyrin
excretion. NT, individuals who were not t...

The overlap of results in carriers with those in controls 413–415 makes determination of urinary
coproporphyrin excretion less useful in deciding whether an individual carries the gene for the syndrome.
However, this disorder is benign, and genetic counseling is rarely required. Urinary coproporphyrin
excretion proved useful in diagnosing Dubin-Johnson syndrome in two neonates. 394, 395 Although
neonates normally have elevated urinary content of coproporphryin I as compared with adults, levels are
not as high as seen in Dubin-Johnson syndrome. 417

The pathogenesis of the abnormal urinary coproporphyrin excretion in this syndrome is unknown, as is its
relationship to conjugated hyperbilirubinemia. In addition to being present in urine, coproporphyrins are
also found in bile, where isomer I constitutes approximately 65 percent of the total. 412 Normally, total daily
biliary coproporphyrin excretion is approximately three times that of total daily urinary excretion. In most
hepatobiliary disorders, including cholestasis, coproporphyrin levels are increased in urine. 418 In these
disorders, total urinary coproporphyrin excretion is elevated and the proportion of isomer I in urine is

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 58 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

usually less than 65 percent. Dubin-Johnson syndrome is unique in that total urinary coproporphyrin is
normal, but the proportion of isomer I is over 80 percent. It seems unlikely that the abnormal pattern of
coproporphyrin isomers seen in Dubin-Johnson syndrome results simply from reduced biliary excretion,
and an alteration in hepatic porphyrin biosynthesis has been postulated (Fig. 125-27). 414, 417 Reduced
coproporphyrin III formation could result from decreased activity of hepatic uroporphyrin III
cosynthetase. 414 Enzyme activity as determined in blood cells and liver from four patients did not differ
from normal. 419 Following an intravenous load of δ-aminolevulinic acid, coproporphyrin III content of urine
and bile changed very little in patients with Dubin-Johnson syndrome, as compared with results in normal
control subjects. 420 Further study of porphyrin biosynthesis is required to elucidate the mechanism of
abnormal coproporphyrin excretion and the relationship of the porphyrin abnormality to the conjugated
hyperbilirubinemia that characterizes the syndrome.

Pathway of porphyrin biosynthesis. δ-Aminolevulinic acid (δ-ALA) condenses to form porphobilinogen


(PBG). In the presence of uroporphyrinogen synthetase, PBG forms the isomer I porphyrins, which are
excretory products without known function. On addition of uroporphyrinogen cosynthetase, PBG forms the
isomer III porphyrins, which are precursors of heme. (Reprinted with permission from Wolkoff AW, Cohen
LE, Arias IM: New Engl J M...

The differential diagnosis of Dubin-Johnson syndrome includes Rotor syndrome, another benign inherited
disorder characterized by the accumulation of both conjugated and unconjugated bilirubin in plasma (see
later section on Rotor Syndrome).

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 59 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Molecular Mechanism of Dubin-Johnson Syndrome.


As discussed before, the bile canalicular transport of bilirubin glucuronides occurs against a concentration
gradient via an energy-consuming mechanism that is shared with many other organic anions, except bile
acids. The functional characteristics of canalicular non–bile acid organic anion transporter, termed the
canalicular multispecific organic anion transporter (cMOAT), have been defined largely by studies in
mutant animal models with a transport defect for organic anions. 421 cMOAT is also termed MRP2
(GenBank NM_000392).

Animal Models.
The first animal model to be described was the mutant Corriedale sheep. The metabolic defect in this
strain closely resembles that found in Dubin-Johnson syndrome. Biliary excretion of a large number of
organic anions, including conjugated bilirubin, glutathione-conjugated BSP, iopanoic acid, and ICG is
decreased, whereas taurocholate transport is normal. 235, 422 Although the biliary excretion of the
glutathione conjugates of BSP is markedly reduced, the secretion of unconjugated BSP is unimpaired. 240
Serum bilirubin levels are mildly elevated, and 60 percent of the pigments in the serum is glucuronidated.
As in Dubin-Johnson syndrome, the liver is pigmented, and the histology is otherwise normal. 422 Total
urinary coproporhyrin excretion is normal with increased excretion of coproporphyrin isomer I and
decreased isomer III excretion.

The most important animal model for Dubin-Johnson syndrome is the TR rat, also known as GY
(Groningen yellow) rat. 421 These rats have been used extensively for elucidation of the mechanism of
canalicular excretion of conjugated bilirubin and other organic anions. As in Dubin-Johnson syndrome, the
biliary excretion of conjugated bilirubin and many other organic anions is impaired, and isomer I of
coproporphyrin constitutes the major fraction of porphyrins excreted in the urine. 421 Although the liver of
TR − rats maintained on standard laboratory chow does not contain black pigments, lysosomal pigment
accumulation occurs upon feeding a diet enriched in aromatic amino acids. 403 Breeding studies indicated
autosomal-recessive inheritance for this disorder and suggested that a single gene is responsible for the
defect. For organic anions, such as glutathione-conjugated leukotriene C 4 , the canalicular secretion
defect is nearly complete, whereas for bilirubin glucuronides, there is a residual transport activity (about
10 percent of normal). 422, 423 In contrast, secretion of the synthetic compound bilirubin ditaurate is nearly
normal. 422 These observations suggest the presence of additional transport mechanisms for organic
anions that may not be affected in TR − rats.

Energy requirements for cMOAT were investigated using dinitrophenylglutathione, the transport of which
is nearly absent in TR rats. 421 Dinitrophenylglutathione transport by isolated hepatocytes 421, 424 and rat
liver plasma membrane vesicles requires ATP. 224, 425, 426 Similarly, the transport of bilirubin
glucuronide, 223 BSP, 220 cysteinyl leukotrienes 221 and p-nitrophenylglucuronide 427 in liver canalicular
membrane vesicles is also stimulated by the addition of ATP.

Substrate Specificity of cMOAT.


The susbstrate specificity of cMOAT was studied by comparison of the biliary excretion of various
compounds in normal and TR rats. From these studies a picture emerged of a transporter that recognizes
a wide variety of compounds. 421 The common denominator of these substrates is that they are anionic
amphipaths. The majority of substrates has more than one negative charge in the molecule. Among the
recognized substrates are glucuronide conjugates, like that of bilirubin and triiodothyronine, but also of
xenobiotics like naphtol. Glutathione conjugates (including its own conjugate, GSSG) are high-affinity
subtrates. The endogenous glutathione conjugate leukotriene C 4 is actually the substrate with the highest

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 60 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

affinity known (K m 0.25 µM) for cMOAT. Sulfate conjugates are also transported but have much lower
affinity. There are also a number of anionic substrates that are not metabolized or conjugated in the
hepatocyte and that are excreted via cMOAT. Examples of this class are dibromosulphophtalein and the
cephalosporin ceftriaxone. 421

Studies in patients with Dubin-Johnson syndrome and in related animal models (Corriedale sheep and TR
rats) have shown that bile acids are secreted by a mechanism that is distinct from that of other organic
anions. Although both types of compounds undergo primary active (ATP-dependent) transport across the
canalicular membrane, 230 their transport processes are clearly mediated by different gene products. The
mutant animals as well as patients with Dubin-Johnson syndrome excrete bile salts normally, whereas the
biliary excretion of a wide variety of other organic anions is severely impaired. Taurocholate does not
compete with the transport of bilirubin glucuronides or BSP in canalicular membrane preparations, further
indicating that these compounds use separate transport systems. 219, 220 An exception to this is the bile
salts that are conjugated at the 3-OH position. In contrast to normal bile salts, these 3-OH conjugated bile
acids have a double-negative charge and behave as non–bile acid organic anions. 225, 428 There is an
ATP-driven and a membrane potential-dependent component of bilirubin glucuronide transport by
canalicular plasma membrane vesicles. 223 The ATP-dependent mechanism is absent in membrane
vesicles from the livers of TR − rats, but the membrane potential-dependent mechanism provides the
residual transport.

A mutant strain of golden lion tamarins (Leontopitheous rosalia rosalia ) with Dubin-Johnson–like
syndrome has been described. 236

Molecular Genetics of cMOAT.


cMOAT is a member of the family of ATP-binding cassette (ABC) transporters, which are, typically, single
polypeptides with two similar halves, each containing at least four and usually six transmembrane helices
and an ABC that mediates ATP hydrolysis. 429 The cDNA for cMOAT has been cloned. 430 cMOAT is a
1541–amino acid (200-kDa) integral membrane protein, localized in the apical membrane of the
hepatocyte. The protein is highly expressed in the liver and to a much lesser extent in kidney, duodenum,
and ileum. Among the other mammalian members of the ABC transporter family, the P-glycoproteins are
best characterized. The P-glycoprotein encoded by the human MDR1 gene mediates ATP-driven
extrusion of hydrophobic compounds from cells. 431 Overexpression of this protein in tumor cells makes
them multidrug resistant. 432 Another P-glycoprotein, encoded by the MDR2 gene, is concentrated in the
canalicular domain of hepatocyte plasma membranes and is required for the translocation of
phospholipids across the canalicular membrane into the bile. 239 The multidrug resistance–related protein,
MRP1, is distantly related to the P-glycoproteins and also confers resistance against cytotoxic drugs. 232
MRP1 mediates the canalicular transport of organic anions, such as dinitrophenylglutathione and
leukotriene C 4 , the transport of which is impaired in TR rats. 233, 236 However, the expression of MRP1 in
the liver is extremely low, and this protein is localized in the basolateral domain of epithelial cell plasma
membranes, 433 suggesting that it is not involved in the canalicular transport of organic anions. Direct
evidence for the role of the cMOAT in hepatocanalicular organic anion transport came from the discovery
of a mutation in this gene in the TR − rat. 430 The deletion of a single nucleotide leads to a frameshift and a
premature stop codon, generating a truncated protein. This mutation also markedly reduces the mRNA
level. cMOAT was immunologically undetectable in the plasma membrane preparations of TR − rats.

More recently, the human cMOAT cDNA has been isolated (GenBank U49248) on the basis of homology
with the rat cMOAT, 434 and the gene has been located on chromosome 10q23-q24. 435 Human cMOAT is
a 1545–amino acid protein with 78 percent amino acid homology to the rat protein. The first mutation that

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 61 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

was described in a Dubin-Johnson syndrome patient was a stop codon, predicted to express a truncated
protein. 434 In this patient and a subsequently reported case, 436 immunohistochemical staining of a liver
section was negative. Two additional patients with single-base deletions and one with a missense
mutation of the cMOAT gene have been reported. 437 Because of its homology with MRP1, the cMOAT
gene is also referred to as MRP2.

Inheritance.
The male-to-female ratio in clinically diagnosed cases of Dubin-Johnson syndrome is 1.5:1, 389 perhaps
because of a greater daily bilirubin production in postpubertal males than in females. With respect to
urinary coproporphyrin excretion pattern, Dubin-Johnson syndrome is inherited as an autosomal-recessive
characteristic. 412, 414 In the TR rat, which has a mutation in the same gene, the disease is clearly inherited
in an autosomal-recessive manner. This pattern is also confirmed in several patients with Dubin-Johnson
syndrome, in whom the genetic lesions have been identified. 435, 436

Clotting Factor VII Deficiency.


In a report from Israel, 60 percent of the patients with Dubin-Johnson syndrome were found to have
reduced prothrombin activity due to lower levels of clotting Factor VII. 437 Although Factor VII deficiency
was most common among the Jews originating from Iran, Iraq, and the neighboring areas, this
abnormality was also found in some patients belonging to other Jewish communities. However, in some
families the two disorders were found to segregate independently, indicating that the combined disorder
may be coincidental and does not represent a true association. The gene for Factor VII is located on
chromosome 13, 438 whereas the cMOAT gene (MRP2) is on chromosome 10, 435 excluding a primary
linkage between the two defects.

Rotor Syndrome

Clinical Findings.
In 1948, Rotor, Manahan, and Florentin described several individuals from two families in whom there was
chronic predominantly conjugated hyperbilirubinemia without any evidence of hemolysis. 439 Serum
alkaline phosphatase and cholesterol values were normal. Plasma disappearance of BSP was greatly
delayed. Liver histology was normal. Although previously Rotor and Dubin-Johnson syndromes were
thought to be variants of a single pathophysiologic disorder, 440 now these disorders are known to be
different entities (Table 125-2). 441 Rotor syndrome is benign. The liver is normal on histologic examination
and does not have excess pigmentation. 440 Although it has been described in several nationalities and
races, Rotor syndrome is rare.

Organic Anion Excretion.


Oral cholecystographic agents usually do not visualize the gallbladder in the Dubin-Johnson syndrome,
whereas roentgenologic visualization usually is possible in Rotor syndrome. 440 Unlike the findings in
Dubin-Johnson syndrome, patients with Rotor syndrome exhibit marked retention of BSP at 45 min after
injection, but biphasic plasma BSP peaks are not found (Fig. 125-25) and conjugated BSP does not
appear in plasma (Fig. 125-28). 442 There is also marked plasma retention of intravenously administered
unconjugated bilirubin and ICG. 443

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 62 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Plasma disappearance of BSP after intravenous injection of a 50 mg/kg dose into 11 patients with Rotor
syndrome, 11 phenotypically normal first-degree relatives defined as heterozygotes for the syndrome on
the basis of urinary coproporphyrin administration, and six normal controls. There was no secondary
increase of plasma BSP, and conjugated BSP was not found in plasma. (Reprinted with permission from
Wolpert et al. ...

With the use of a constant infusion technique, the transport maximum (T max ) for BSP and the relative
hepatic storage capacity have been determined in patients with Rotor and Dubin-Johnson syndromes
(Fig. 125-29). 442, 443 In Dubin-Johnson syndrome, the T max is virtually zero, whereas the hepatic storage
capacity is normal. In Rotor syndrome, the hepatic storage capacity was reduced by 75 to 90 percent and
T max was reduced by 50 percent. 442, 443 Determination of T max and relative hepatic storage capacity (S)
in phenotypically normal obligate heterozygotes revealed results intermediate between those in patients
with Rotor syndrome and controls. 442 The modest reduction in T max accompanied by a larger reduction in
hepatic storage is similar to observations in hepatic storage disease, a familial disorder manifested by
predominantly conjugated hyperbilirubinemia and normal liver histology. 444 Because there is little to
differentiate Rotor syndrome from hepatic storage disease, they may represent a single pathophysiologic
entity.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 63 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Hepatic relative storage capacity (S) and transport maximum (Tmax ) for BSP in six patients with Rotor
syndrome, five phenotypically normal heterozygotes, and six normal controls. (Reprinted with permission
from Wolpert et al. 442 )

Urinary Coproporphyrin Excretion.


Unlike results for Dubin-Johnson syndrome, total urinary coproporphyrin excretion is increased by 250 to
500 percent as compared with control subjects, and the proportion of coproporphyrin I in urine is
approximately 65 percent of total. 441, 445 In one report, however, two brothers with clinical Rotor syndrome
had over 80 percent of urinary coproporphyrins as isomer I. 446 These results are similar to those seen in
many other hepatobiliary disorders. 447 Phenotypically normal obligate heterozygotes have a
coproporphyrin excretory pattern that is intermediate between that of control subjects and patients with
Rotor syndrome. With respect to urinary coproporphyrin excretion, Rotor syndrome is inherited as an
autosomal-recessive characteristic and is distinct from Dubin-Johnson syndrome (Table 125-2, Figs.
125-30, 125-31). 445 The urinary coproporphyrin abnormality in Rotor syndrome, unlike that in
Dubin-Johnson syndrome, is most likely caused by a reduced biliary excretion of coproporphyrins, with
consequent increase in renal excretion. The nature of the organic anion transport defect in Rotor
syndrome is unknown.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 64 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Pedigree of a Philippine family originally described by Rotor in 1948. Solid symbols indicate individuals
with Rotor syndrome. Partial symbols indicate phenotypically normal individuals with urinary
coproporphyrin excretion in the heterozygote range. Open symbols represent phenotypically normal
individuals with normal urinary coproporphyrin excretion. NT, individuals who were not tested. (Reprinted
with permission from Wolkoff et al. ...

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 65 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Urinary coproporphyrin excretion in Dubin-Johnson and Rotor syndromes. The shaded bars represent the
percentage of total urinary coproporphyrin excreted as coproporphyrin I. The open bars represent total
urinary coproporphyrin excretion. Vertical bars represent 1 SEM. Total urinary coproporphyrin excretion is
normal in Dubin-Johnson syndrome (DJS), with a markedly elevated proportion of coproporphyrin I
(greater than 80 percent). Both variab...

Progressive Familial Intrahepatic Cholestasis


Progressive familial intrahepatic cholestasis (PFIC) is a heterogeneous group of autosomal-recessive
inherited diseases in infants and children. All are characterized by cholestasis associated with elevated
plasma levels of bile acids, defective bile acid secretion in bile, growth failure, and progressive liver
damage initially manifested by conjugated hyperbilirubinemia. On the basis of molecular genetic studies,
PFIC has been divided into three categories. The availability of DNA expressed sequence tags (ESTs)
generated by the Human Genome Project enabled identification of large and frequently inbred families
with individual disorders. The findings of the molecular genetic studies overlapped temporally with the
discovery of specific ATP-dependent transporters in the bile canaliculus.

Progressive familial intrahepatic cholestasis type I (Byler disease) was described in an old order Amish
pedigree of seven generations descended from Jacob Byler. 448 The disease is manifested by variable
conjugated hyperbilirubinemia, growth retardation, and progressive, fatal cholestasis. In some patients,
intestinal lipid malabsorption is present and not completely ameliorated by liver transplantation. Plasma
bile acids are elevated, particularly lithocholate. No unique cholestatic bile acids or metabolites have been
demonstrated in blood or urine.

The defective gene was mapped to chromosome 18q21 by screening of the genome for chromosomal
sequences shared by patients from the original Byler kindred. Probability calculations indicate that such
sharing is unlikely to occur by chance. A mutated gene, designated FIC1, that codes for a P-type ATPase
was identified (GenBank AF038007). 449 Northern hybridization reveals expression of the normal gene in
intestine, liver, and other tissues. Its cellular and subcellular sites have not yet been identified. How
defective FIC1 results in cholestasis and abnormal bile acid transport is not known. Normal P-ATPases
couple hydrolyis of ATP to the translocation of acidic phospholipids (i.e., phosphatidylserine,
phosphatidylethanolamine) from the outer to the inner layer of the plasma membrane of many different
cells. Alterations in hepatic or intestinal membrane lipids may affect the function of ATP-dependent
tranporters for bile acids, conjugated bilirubin, and other ligands.

Benign Recurrent Intrahepatic Cholestasis (BRIC).


This rare disease was described in 1959 450 and is characterized by recurrent episodes of cholestasis
followed by complete return to normalcy clinically, biochemically, and by liver histology. 451–453 The
disorder may begin in infancy or middle age but predominantly is manifested in adolescence or early
adulthood. 452, 454 Episodes last from several weeks to months, during which time patients manifest
malaise, anorexia, pruritus, weight loss, malabsorption, conjugated hyperbilirubinemia, and biochemical
evidence of cholestasis without severe hepatocellular injury. 453–455 Intervals between attacks may last
from a few weeks to several years. In a given patient, recurrent attacks resemble each other in symptoms,
signs, and duration. Liver histology reveals noninflammatory intrahepatic cholestasis without fibrosis
regardless of the number and severity of attacks. During remission, liver histology returns to normal
whether examined by light or electron microscopy. 456

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 66 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

The pathogenesis of this rare disorder is unknown. Recessive inheritance has been demonstrated by
family studies in isolated populations in which the disorder is frequently observed. There is no specific
treatment to prevent the occurrence of cholestatic episodes or to shorten their duration. Liver
transplantation has not been performed because of the episodic and nonprogressive course of the
disease.

Unexpectedly, mutated FIC1 was identified in BRIC by searching for chromosomal sequences shared by
only three distantly related patients. 449 Although BRIC and PFIC I have many clinical features in common,
the former is progressive whereas the latter is intermittent. How the different phenotypes result from the
same genetic defect is unknown.

Progressive Familial Intrahepatic Cholestasis Type II.


A second form of PFIC resembles Byler disease clinically but occurs in non-Byler families, mainly in the
Middle East and Europe. Homozygosity mapping and genome search in unrelated pedigrees associated
the disorder with chromosome 2q24, from which the gene for SPGP (sister of P-glycoprotein) was cloned
(GenBank AF091582). 457 Remarkably, at almost the same time, SPGP was shown to be a canalicular
ATP-dependent bile acid transporter. 238 Over 20 different point mutations in SPGP have been described
in patients with PFIC of this type. Liver transplantation ameliorates all manifestations of the disease in
concordance with the finding that SPGP is expressed only in the liver.

Progressive Familial Intrahepatic Cholestasis Type III.


A third PFIC group involves mutations in the loci encoding class III multidrug resistance (MDR)
P-glycoproteins (mdr2 in mice and MDR3 in humans), which mediate ATP-dependent translocation of
phosphatidylcholine from the inner to the outer leaflet of the bile canalicular plasma membrane. 231
Removal of mdr2 by homologous recombination in mice results in absence of phosphatidylcholine from
bile, progressive destruction of small bile ducts, cholestasis characterized by conjugated
hyperbilirubinemia, and eventual biliary cirrhosis. In mdr2 −/− mice, bile acids are secreted normally by
SPGP but are not incorporated into mixed micelles and progressively damage small bile ducts. 430

Two different point mutations in MDR3 resulting in a nonfunctioning protein were described in two children
with PFIC associated with elevated serum levels of γ-glutamyl transpeptidase activity. 458 Virtual absence
of MDR3 mRNA occurs in Navajo Indian children with PFIC and dysmyelinating peripheral and central
neuropathy. 459 The relationship between MDR3 and neuropathy is unknown. Elevated serum γ-glutamyl
transpeptidase activity is proposed to result from small bile duct damage and distinguishes this PFIC
group from PFIC I and BRIC.

Alagille Syndrome
Many heritable developmental disorders have been described; however, the Alagille syndrome is the first
to be described at the molecular level. The Alagille syndrome is transmitted as an autosomal-recessive
inherited characteristic that includes paucity or absence of small bile ducts resulting in progressive
intrahepatic cholestasis, and abnormalities of the eye, heart, and vertebrae. Identification of rare patients
with cytogenic deletions permitted mapping of the gene to chromosome 20p12, from which JAG1 was
identified. 460 JAG1 encodes an unidentified ligand that binds to the notch receptor, which is crucial for cell
fate development in Drosophila and mammals. Analysis of patients with Alagille syndrome who do not
have chromosomal deletions (see Chap. 65) revealed several point mutations in JAG1 (GenBank
U73936), each of which abolishes expression of the altered allele. 460

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 67 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

ACKNOWLEDGMENT
The authors thank Dr. Ajit Kadakol for assistance in preparation of this review. This work was supported in
part by National Institutes of Health Grants DK-46057, DK-39137, DK-41296, DK-23026, DK-35652, and
DK-34926.

REFERENCES

1. Berk PD, Howe RB, Bloomer JR, Berlin NI: Studies on bilirubin kinetics in normal adults. J Clin Invest.
48:2176, 1979.
2. London IM, West R, Shemin D, Rittenberg D: On the origin of bile pigment in normal man. J Biol
Chem. 184:351, 1950.
3. Schwartz S, Johnson JA, Stephenson BD, Anderson AS, Edmondson PR, Fusaro RM: Erythropoietic
defects in protoporphyria A study of factors involved in labelling of porphyrins and bile pigments from
ALA- 3 H and glycine- 14 C. J Lab Clin Med. 78:411, 1971. PMID 5092862
4. Berk PD, Jones EA, Howe RB, Berlin NI: Disorders of bilirubin metabolism, in Bondy PK, Rosenberg
LE (eds) Metabolic Control and Disease, 8th ed. Philadelphia. : WB Saunders, 1980, p 1009.
5. Grandchamp B, Bissel DM, Licko V, Schmidt R: Formation and disposition of newly synthesized
heme in adult rat hepatocytes in primary cultures. J Biol Chem. 256:11677, 1981. PMID 7298624
6. Levitt M, Schacter BA, Zipursky A, Israels LG: The nonerythropoietic component of early bilirubin. J
Clin Invest. 47:1281, 1968. PMID 5653209
7. Gray CH, Scott JJ: The effect of haemorrhage on the incorporation of [ 14 C]glycine into stercobilin.
Biochem J. 71:38, 1959.
8. Beattie AD, Goldberg A: Acute intermittent porphyria Natural history and prognosis, in Doss M (ed). :
Karger, 1976, p 245.
9. Come SE, Shohet SB, Robinson SH: Surface remodeling vs. whole-cell hemolysis of reticulocytes
produced with erythroid stimulation or iron deficiency anemia. Blood. 44:817, 1974. PMID 4429801
10. Robinson SH: Origins of the early-labeled peak, in Berk PD, Berlin NI (eds) Bile Pigments. :
Department of Health, Education, and Welfare, National Institutes of Health, 1977, p 175. PMID
4429801
11. Landaw SA: Quantitative recovery of 14 C-labeled carbon monoxide ( 14 CO) from viable heme-labeled
red blood cells in the rat. Blood. 40:257, 1972. PMID 5044593
12. Bissel DM, Guzelian PS: Degradation of endogenous hepatic heme by pathways not yielding carbon
monoxide Studies in normal rat liver and primary hepatocyte culture. J Clin Invest. 65:1135, 1980.
PMID 7364941
13. Tenhunen R, Marver HS, Schmid R: Microsomal heme oxygenase Characterization of the enzyme. J
Biol Chem. 244:6388, 1969. PMID 4390967
14. Bissel DM, Hammaker L, Schmidt R: Liver sinusoidal cells. Identification of a subpopulation for
erythrocyte catabolism. J Cell Biol. 54:107, 1972. PMID 5038868
15. Sassa S, Kappas A, Bernstein SE, Alvares AP: Heme biosynthesis and drug metabolism in mice with
hereditary hemolytic anemia. J Biol Chem. 254:729, 1979. PMID 762094
16. Ishizawa S, Yoshida T, Kikuchi G: Induction of heme oxygenase in rat liver. J Biol Chem. 258:4220,
1983. PMID 6403524
17. Posselt AM, Cowan BE, Kwong LK, Vreman HJ, Stevenson DK: Effect of tin protoporphyrin on the
excretion rate of carbon monoxide in newborn rats after hematoma formation. J Pediatr Gastroenterol
Nutr. 4:650, 1985. PMID 3839847
18. Yoshida T, Kikuchi G: Heme oxygenase purified to apparent homogeneity from pig spleen
microsomes. J Biochem (Tokyo. . 81:265, 1977. PMID 845137
19. Yosinga T, Sassa S, Kappas A: The occurrence of molecular interactions among NADPH-cytochrome
C reductase, heme oxygenase and biliverdin reductanse in heme degradation. J Biol Chem.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 68 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

257:7778, 1982. PMID 6806282


20. Maines MD, Ibrahim NG, Kappas A: Solubilization and partial purification of heme oxygenase from rat
liver. J Biol Chem. 252:5900, 1977. PMID 18477
21. Jackson AH, Kenner W: Recent developments in porphyrin chemistry, in Goodwin TW (ed)
Porphyrins and Related Compounds. London. : Academic, 1968, p 5.
22. Brown SB, King RFGJ: The mechanism of heme catabolism Bilirubin formation in living rats by
[ 18 O]oxygen labeling. Biochem J. 170:297, 1978. PMID 637844
23. Yoshida T, Kikuchi G: Features of the reaction of heme degradation catalyzed by the reconstituted
microsomal heme oxygenase system. J Biol Chem. 253:4230, 1978. PMID 96116
24. Cornelius CE: Comparative bile pigment metabolism in vertebrates, in Ostrow JD (ed) Bile Pigments
and Jaundice. New York. : Marcel Dekker, 1986, p 601.
25. McDonagh AF, Palma LA, Schmidt R: Reduction of biliverdin and placental transfer of bilirubin and
biliverdin in the pregnant guinea pig. Biochem J. 194:273, 1981. PMID 7305981
26. Roy Chowdhury J, Roy Chowdhury N, Arias IM: Bilirubin conjugates in the spiny dogfish, Squalus
acanthias, the small skate, Raja erinacea and the winter flounder Pseudopleuronectes americanas.
Comp Biochem Physiol [B]. 66:523, 1980.
27. Stocker R, Yamamoto Y, McDonagh AF: Bilirubin is an antioxidant of possible physiological
significance. Science. 235:1043, 1987. PMID 3029864
28. Halliwell B, Gutteridge JM: The antioxidants of human extracellular fluids. Arch Biochem Biophys.
280:1, 1990. PMID 2191627
29. Dannery PA, McDonagh AF, Spitz DR: Hyperbilirubinemia results in reduced oxidative injury in
neonatal Gunn rats exposed to hyperoxia. Free Radic Biol Med. 19:395, 1995 PMID 7590389
30. Colleran E, O’Carra P: Enzymology and comparative physiology of biliverdin reduction, in Berk PD,
Berlin NI (eds) Bile Pigments. : Department of Health, Education, and Welfare, National Institutes of
Health, 1977, p 69. PMID 7590389
31. Tenhunen R, Ross ME, Marver HS, Schmid R: Reduced nicotinamide-adenine dinucleotide
phosphate dependent biliverdin reductase Partial purification and characterization. Biochemistry.
9:298, 1970. PMID 4391687
32. Colleran E, O’Carra P: Enzymology and comparative physiology of biliverdin reduction, in Berk PD,
Berlin NI (eds) Bile Pigments. : Department of Health, Education, and Welfare, National Institutes of
Health, 1977, p 69. PMID 4391687
33. Frydman RBM, Tomaro ML, Awruch J, Frydman B: Interconversion of the molecular forms of
biliverdin reductase from rat liver. Biochem Biophys Acta. 759:257, 1983.
34. Heirwegh KPM, Blanckaert N, Compernolle F, Fevery J, Zaman Z: Detection and properties of the
non-α isomers of bilirubin-IX. Biochem Soc Trans. 5:316, 1977. PMID 892197
35. Bensinger TA, Maisels MJ, Mahmood L, McCurdy PR, Conrad MD: Effect of intravenous urea in
invert sugar on heme catabolism in sickle cell anemia. New Engl J Med. 285:995, 1971. PMID
5094820
36. Howe RB, Berlin NI, Berk PD: Estimation of bilirubin production in man, in Berk PD, Berlin NI (eds)
Bile Pigments. : Department of Health, Education, and Welfare, National Institutes of Health, 1977, p
105. PMID 5094820
37. Jones EA, Shrager R, Bloomer JR, Berk PD, Howe RB, Berlin NI: Quantitative studies of the delivery
of hepatic synthesized bilirubin to plasma utilizing δ-aminolevulinic acid-4- 14 C and bilirubin- 3 H in
man. J Clin Invest. 51:2450, 1972. PMID 4639027
38. Jones EA, Bloomer JR, Berk PD, Carson ER, Owens D, Berlin NI: Quantitation of hepatic bilirubin
synthesis in man, in Berk PD, Berlin NI (eds) Bile Pigments. : Department of Health, Education, and
Welfare, National Institutes of Health, 1977, p 189. PMID 4639027
39. Berk PD, Rodkey FL, Blaschke TF, Collison HA, Waggoner JG: Comparison of plasma bilirubin
turnover and carbon monoxide production in man. J Lab Clin Med. 83:29, 1974. PMID 4808654
40. Coburn RF, Gondrie P, Abboud F, Ploegmakers E: Myocardial myoglobin oxygen tension. Am J

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 69 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Physiol. 224:870, 1973. PMID 4698804


41. Stewart RD, Fisher TN, Hosko MJ, Peterson JE, Baretta ED, Dodd HC: Carboxyhemoglobin elevation
after exposure to dichloromethane. Science. 176:295, 1972. PMID 5019783
42. Engel RR: Alternative sources of carbon monoxide, in Berk PD, Berlin NI (eds) Chemistry and
Physiology. Washington, DC. : Department of Health, Education, and Welfare, National Institutes of
Health, 1977, p 148.
43. Westlake DWS, Roxburgh JM, Talbot G: Microbial production of carbon monoxide from flavinoids.
Nature. 189:510, 1961.
44. Valaes T, Petmezaki S, Henschke C: Control of jaundice in preterm newborns by an inhibitor of
bilirubin production. : 1, 1994. PMID 8265301
45. Kappas A, Drummond GS: Direct comparison of tin-mesoporphyrin, an inhibitor of bilirubin
production, and phototherapy in controlling hyperbilirubinemia in term and near-term newborns.
Pediatrics. 95:468, 1995. PMID 7700742
46. Fischer H, Plieninger H: Synthese des biliverdins (uteroverdins) und bilirubins der biliverdine XIII, und
III, sowie der Vinulneoxanthosaure. Hoppe Seyler Z Physiol Chem. 274:231, 1942.
47. Bonnet RJ, Davis E, Hursthouse MB: Structure of bilirubin. Nature. 262:326, 1976.
48. Krasner J, Yaffe SJ: The automatic titration of bilirubin. Biochem Med. 7:128, 1973. PMID 4684082
49. Broderson R: Binding of bilirubin to albumin. Crit Rev Clin Lab Sci. 11:305, 1980.
50. Brodersen R: Aqueous solubility, albumin binding and tissue distribution of bilirubin, in Ostrow JD
(ed). : Bile Pigments and Jaundice. New York; Marcel Dekker, 1986, p 157.
51. Brordersen R: Free bilirubin in the blood plasma of the newborn. Effects of albumin, fatty acids, pH,
displacing drugs and phototherapy. Appendix A Provisional survey of the bilirubin displacing effect of
150 drugs, in Stern L (ed). : Masson, 1978, p 331.
52. Fog J, Jellum E: Structure of bilirubin. Nature. 198:88, 1963.
53. Kuenzle CC, Weibel MH, Pelloni RR: A proposed novel structure for the metal chelates of bilirubin.
Biochem J. 130:1147, 1973.
54. Maritto P, Monti D: Free energy barrier of conformational inversion in bilirubin. J Chem Soc Chem
Commun. 4:122, 1976.
55. Knell AJ, Johnson B, Hutchinson DW: Intramolecular hydrogen bonds in bilirubin. Digestion. 6:288,
1972.
56. Kuenzle CC, Weibel MH, Pelloni RR: The reaction of bilirubin with diazomethane. Biochem J.
133:357, 1973. PMID 4723780
57. Trull FR, Ibars O, Lightner DA: Conformational inversion of bilirubin formed by reduction of the
biliverdin-human serum albumin complex Evidence from circular dichroism. Arch Biochem Biophys.
298:710, 1992. PMID 1416999
58. Krasner J, Yaffe SJ: Fluorescent properties of the bilirubin-albumin complex. Birth Defects. 12:168,
1976. PMID 953181
59. McDonagh AF, Palma LA, Lightner DA: Phototherapy for neonatal jaundice. Stereospecific and
regiospecific photoisomerization of bilirubin bound to human serum albumin and NMR characterization of
intromolecularly cyclized photoproducts. J Am Chem Soc. 104:6867, 1982.
60. Itho S, Onishi S: Kinetic study of the photochemical changes of (ZZ)-bilirubin IX bound to human
serum albumin. Demonstration of (EZ)-bilirubin IX as an intermediate in photochemical changes from
(ZZ)-bilirubin IX to (EZ)-cyclobilirubin IX. Biochem J. 226:251, 1985. PMID 3977868
61. McDonagh AF: Thermal and photochemical reactions of bilirubin IX. Ann NY Acad Sci. 244:553,
1975. PMID 1094886
62. Schenker S, Hoyumpa AM, McCandless DW: Bilirubin toxicity to the brain (kernicterus) and other
tissues, in Ostrow JD (ed) Bile Pigments and Jaundice. New York. : Marcel Dekker, 1986, p 395.
63. Hervieux J: De l’ictere des nouveau-nes. Paris These Med 1847
64. Schmorl G: Zur Kenntnis des ikterus neonatatorum, inbesondere der dabei auftreten den

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 70 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

gehirnveranderungen. Verh Dtsch Ges Pathol. 6:109, 1903.


65. Schiff D, Chan G, Poznasky MJ: Bilirubin toxicity in neural cell lines N115 and NBR10A. Pediatr Res.
19:908, 1985. PMID 4047759
66. Mustafa MG, Cowger ML, Kind TE: Effects of bilirubin on mitochondrial reactions. J Biol Chem.
244:6403, 1969. PMID 4982202
67. Strumia E: Effect of bilirubin on some hydrolases. Bull Soc Ital Biol Sper. 35:2160, 1959.
68. Flitman R, Worth NK: Inhibition of hepatic alcohol dehydrogenase by bilirubin. J Biol Chem. 251:669,
1966.
69. Cowger ML, Igo RP, Labbe RF: The mechanism of bilirubin toxicity studied with purified respiratory
enzyme and tissue culture systems. Biochemistry. 4:2763, 1965. PMID 5880684
70. Katoh R, Kashiwamata S, Niwa F: Studies on cellular toxicity of bilirubin. Effect on the carbohydrate
metabolism in the young rat brain. Brain Res. 83:81, 1975.
71. Greenfield S, Nandi Majumdar AP: Bilirubin encephalopathy. Effect on protein synthesis in the brain
of the Gunn rat. J Neurol Sci. 22:83, 1974. PMID 4857425
72. Nandi Majumdar AP: Bilirubin encephalopathy. Effect on RNA polymerase activity and chromatin
template activity in the brain of the Gunn rat. Neurobiology. 4:425, 1974. PMID 4453356
73. Diamond I, Schmid R: Oxidative phosphorylation in experimental bilirubin encephalopathy. Science.
155:1288, 1967. PMID 6018655
74. Zetterstrom R, Ernster L: Bilirubin, an uncoupler of oxidative phosphorylation in isolated mitochondria.
Nature. 178:1335, 1956.
75. Constantopoulos A, Matsaniotis N: Bilirubin inhibition of protein kinase Its prevention by cyclic AMP.
Cytobios. 17:17, 1976. PMID 196812
76. Morphis I, Constantopoulos A, Matsaniotis N, Papaphilis A: Bilirubin-induced modulation of cerebral
protein phosphorylation in neonate rabbits in vivo. Science. 218:156, 1982. PMID 7123226
77. Sano K, Nakamura H, Tamotsu M: Mode of inhibitory action of bilirubin on protein kinase C. Pediatr
Res. 19:587, 1985. PMID 2989761
78. Gourley GR: Bilirubin metabolism and kernicterus. Adv Pediatr. 44:173, 1997. PMID 9265971
79. Stobie PE, Hansen CT, Hailey JR, Levine RL: A difference in mortality between two strains of
jaundiced rats. Pediatrics. 87:88, 1991. PMID 1898623
80. Odell GB: Influence of binding on the toxicity of bilirubin. Ann NY Acad Sci. 226:225, 1973. PMID
4520394
81. Conlee JW, Shapiro SM: Morphological changes in the cochlear nucleus and nucleus of the trapezoid
body in Gunn rat pups. Hear Res. 57:23, 1991. PMID 1774208
82. Funato M, Tamai H, Shimada S: Viginitiphobia, unbound bilirubin and auditory brainstem responses.
Pediatrics. 93:50, 1994. PMID 8265323
83. Shapiro SW, Conlee JW: Brainstem auditory evoked potentials correlate with morphological changes
in Gunn rat pups. Hear Res. 57:16, 1991. PMID 1774207
84. Shapiro SM: Binaural effects in brainstem auditory evoked potentials of jaundiced Gunn rats. Hear
Res. 53:41, 1991. PMID 2066286
85. Naeye RL: Aminiotic fluid infections, neonatal hyperbilirubinemia and psychomotor impairment.
Pediatrics. 62:497, 1978. PMID 714581
86. Van De Bor M, Ens-Dokkum M, Schreuder AM, Veen S, Brand R, Verloove-Vanhorick SP:
Hyperbilirubinemia in low birth weight infants and outcome at 5 years of age. Pediatrics. 89:359,
1992. PMID 1371340
87. Labrune PH, Myara A, Francoual J, Trivin F, Odièvre M: Cerebellar symptoms as the presenting
manifestations of bilirubin encephalopathy in children with Crigler-Najjar type I disease. Pedicatrics.
89:768, 1992.
88. Penn AA, Enzmann DR, Hahn JS: Kernicterus in a full-term infant. Pediatrics. 93:1003, 1994. PMID
7794295

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 71 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

89. Matrich-Kriss V, Kollins SS, Ball WS: MR findings in kernicterus. AJNR. 16:817, 1995.
90. Yokochi K: Magnetic resonance imaging in children with kernicterus. Acta Pediatr. 84:937, 1995.
91. Zuelzer WW, Mudgett RT: Kernicterus. Etiologic study based on an analysis of 55 cases. Pediatrics.
6:452, 1950.
92. Vaughan VC, Allen FC, Diamond LK: Erythroblastosis fetalis. IV. Further observations on kernicterus.
Pediatrics. 6:706, 1950.
93. Rappaport SI: Blood–Brain Barrier in Physiology and Medicine. New York. : Raven, 1976.
94. Purpura DP, Carmichael MW: Characteristics of blood-brain barrier to gamma-aminobutyric acid in
neonatal cat. Science. 131:410, 1960.
95. Cornford EM, Braun LD, Oldendorp WH, Hill MA: Comparison of lipid-mediated blood–brain barrier
penetrability in neonates and adults. Am J Physiol. 243:C161, 1982. PMID 7114247
96. Laas R, Helmke K: Regional cerebral blood flow following unilateral blood–brain barrier alteration
induced by hyperosmolar perfusion in the albino rat, in Cervos-Navarro J, Fritsch E (eds) Cerebral
Circulation and Metabolism. New York. : Raven, 1981, p 317.
97. Levine RL, Fredericks WR, Rappaport SI: Clearance of bilirubin from rat brain after reversible osmotic
opening of the blood–brain barrier. Pediatr Res. 19:1040, 1985. PMID 4058976
98. Lee K-S, Gartner LM: Management of unconjugated hyperbilirubinemia in the newborn. Semin Liver
Dis. 3:52, 1983. PMID 6403991
99. Lauff JJ, Kasper ME, Ambros RT: Quantitative liquid chromatographic estimation of bilirubin species
in pathological serum. Clin Chem. 29:800, 1983. PMID 6839457
100. Bowen WR, Porter E, Waters WF: The protective action of albumin in bilirubin toxicity in newborn
puppies. Am J Dis Child. 98:568, 1959.
101. Silvermann WA, Andersen DH, Blanc WA, Crozier DN: A difference in mortality rate and incidence of
kernicterus among premature infants allotted to two prophylactic antibacterial regimens. Pediatrics.
18:614, 1956.
102. Odell GB: The dissociation of bilirubin from albumin and its clinical implications. J Pediatr. 55:268,
1959.
103. Bloomer JR, Berk PD, Vergalla J, Berlin NI: Influence of albumin on the extravascular distribution of
unconjugated bilirubin. Clin Sci Mol Med. 45:517, 1973. PMID 4751970
104. Jacobsen J: Binding of bilirubin to human serum albumin. Determination of the dissociation constants.
FEBS Lett. 5:112, 1969. PMID 11947253
105. Jacobsen C: Lysine residue 240 of human serum albumin is involved in high-affinity binding of
bilirubin. Biochem J. 171:453, 1978. PMID 656055
106. Harris RC, Lucey JF: Kernicterus in premature infants associated with low concentrations of bilirubin
in plasma. Pediatrics. 21:875, 1958.
107. Brodersen R, Sjodin T, Sjoholm I: Independent binding of benzodiazepines and bilirubin to human
serum albumin. J Biol Chem. 252:5067, 1977. PMID 873931
108. Woolley PW, Hunter M: Binding and circular dichroism data on bilirubin-albumin in the presence of
oleate and salicylate. Arch Biochem Biophys. 140:197, 1970. PMID 5466734
109. Brodersen R: Binding of bilirubin and other ligands of human serum albumin, in Peters T, Sjoholm I
(eds) Albumin, Structure, Biosynthesis, Function. FEBS 11th Meeting, Copenhagen, 1977. Oxford,
England. : Pergamon, 1978, p 61.
110. Rudman D, Bixler TJ, Del Rio AE: Effect of free fatty acid on binding of drugs by bovine serum
albumin, by human serum albumin and by rabbit serum. J Pharmacol Exp Ther. 176:261, 1971. PMID
4105545
111. Stern L, Denton RL: Kernicterus in small premature infants. Pediatrics. 35:483, 1965.
112. Diamond I, Schmid R: Experimental bilirubin encephalopathy. The mode of entry of bilirubin- 14 C into
the central nervous system. J Clin Invest. 45:678, 1966. PMID 5949114
113. Kapitulnik J, Valaes T, Kaufmann NA, Blondheim SH: Clinical evaluation of Sephadex gel filtration in

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 72 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

estimation of bilirubin binding in serum in neonatal jaundice. Arch Dis Child. 49:886, 1974. PMID
4474844
114. Brodersen R, Cashore W, Wennberg RP, Ahlfors CE, Rasmussen LF, Shusterman D: Kinetics of
bilirubin oxidation with peroxidase, as applied to studies of bilirubin-albumin binding. Scand J Clin Lab
Invest. 39:143, 1979. PMID 523962
115. Athanassiadis S, Chopra DR, Fisher M, McKenna J: An electrophoretic method for detection of
unbound bilirubin and reserve bilirubin binding capacity in serum of newborns. J Lab Clin Med.
83:968, 1974. PMID 4138574
116. Lamolla AA, Eisinger J, Blumberg WE, Palet SC, Flores J: Fluorometric study of the partition of
bilirubin among blood components Basis for rapid microassays of bilirubin and bilirubin binding capacity in
whole blood. Anal Biochem. 15:25, 1979.
117. Bratlid D: Reserve albumin binding capacity salicylate saturation index, and red cell binding of
bilirubin in neonatal jaundice. Arch Dis Child. 48:393, 1973. PMID 4703070
118. Brodersen R: Determination of the vacant amount of high affinity bilirubin binding site on serum
albumin. Acta Pharmacol Toxicol. 42:153, 1978.
119. Hsia JC, Kwan NH, Er SS, Wood DJ, Chance GW: Development of a spin assay for reserve bilirubin
loading capacity of human serum. Proc Natl Acad Sci U S A. 75:1542, 1978. PMID 206907
120. Porter EG, Waters WJ: A rapid micromethod for measuring the reserve albumin binding capacity in
serum for newborn infants with hyperbilirubinemia. J Lab Clin Med. 67:660, 1966.
121. Goresky CA: The hepatic uptake process Its implications for bilirubin transport, in Goresky CA, Fisher
MM (eds). : Plenum, 1975, p 159.
122. Bloomer JR, Zaccaria J: Effect of graded bilirubin loads on bilirubin transport by perfused rat liver. Am
J Physiol. 203:736, 1976.
123. Scharschmidt F, Waggoner JG, Berk PD: Hepatic organic anion uptake in the rat. J Clin Invest.
56:1280, 1975. PMID 1184749
124. Hunton DB, Bollman JL, Hoffman HN: The plasma removal of indocyanine green and
sulfobromophthalein Effect of dosage and blocking agents. J Clin Invest. 40:1648, 1961.
125. Shupeck M, Wolkoff AW, Scharschmidt BF, Waggoner JG, Berk PD: Studies of the kinetics of purified
conjugated bilirubin- 3 H in the rat. Am J Gastroenterol. 70:259, 1978. PMID 717377
126. Goresky CA: The hepatic uptake and excretion of sulfobromophthalein and bilirubin. Can Med Assoc
J. 92:851, 1965.
127. Barnhart JL, Clarenburg R: Factors determining clearance of bilirubin in perfused rat liver. Am J
Physiol. 225:497, 1973. PMID 4722415
128. Weisiger RA, Gollan JL, Ockner RK: Receptor for albumin on the liver cell surface may mediate
uptake of fatty acids and other albumin-bound substances. Science. 211:1048, 1981. PMID 6258226
129. Forker EL, Luxon BA: Albumin helps mediate removal of taurocholate by rat liver. J Clin Invest.
67:1517, 1981. PMID 7229036
130. Fleischer AB, Shurmantine WO, Luxon BA, Forker EL: Palmitate uptake by hepatocyte monolayers.
Effect of albumin binding. J Clin Invest. 77:964, 1986. PMID 3949983
131. Weisiger RA: Dissociation from albumin A potential rate-limiting step in the clearance of substances
by the liver. Proc Natl Acad Sci U S A. 82:1563, 1985. PMID 3856281
132. Zucker SD, Goessling W, Gollan JL: Kinetics of bilirubin transfer between serum albumin and
membrane vesicles. Insight into the mechanism of organic anion delivery to the hepatocyte plasma
membrane. J Biol Chem. 270:1074, 1995. PMID 7836362
133. Weisiger RA, Zacks CM, Smith ND: Effect of albumin binding on extraction of sulfobromophthalein by
perfused elasmobranch liver. : 492, 1984. PMID 6724517
134. Gumucio DL, Gumucio JJ, Wilson JAP: Albumin influences sulfobromophthalein transport by
hepatocytes of each acinar zone. Am J Physiol. 246:G86, 1984. PMID 6696072
135. Schwenk M, Burr R, Schwarz L, Pfaff E: Uptake of bromosulfophthalein by isolated liver cells. Eur J
Biochem. 64:189, 1976. PMID 6268

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 73 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

136. Van Bezooijen CF, Grell T, Knook DL: Bromosulfophthalein uptake by isolated liver parenchymal
cells. Biochem Biophys Res Commun. 69:354, 1976. PMID 1267791
137. Aperche T, Preaux AM, Berthelot P: Two systems are involved in the sulfobromophthalein uptake by
rat liver cells One is shared with bile salts. Biochem Pharmacol. 30:1333, 1982.
138. Schwarz LR, Gotz R, Klassen CD: Uptake of sulphobromophthalein-glutathione conjugates by
isolated hepatocytes. Am J Physiol. 239:C118, 1980. PMID 7435550
139. Potter BJ, Blades BF, Shepard MD, Thung SM, Berk PD: The kinetics of sulfobromophthalein uptake
by rat liver sinusoidal vesicles. Biochim Biophys Acta. 898:159, 1987. PMID 3828336
140. Wolkoff AW, Samuelson SC, Johansen KL, Nakata R, Withers D, Sosiak S: Influence of Cl − on
organic anion transport in short-term cultured rat hepatocytes and isolated perfused rat liver. J Clin
Invest. 79:1259, 1987. PMID 3031134
141. Adachi Y, Roy Chowdhury J, Roy Chowdhury N, Theilman L, Kinne R, Arias IM: Hepatic uptake of
bilirubin diglucuronide Analysis by using sinusoidal plasma membrane vesicles. J Biochem. 107:749,
1990. PMID 2398040
142. Min AD, Johansen KL, Campbell CG, Wolkoff AW: Role of chloride and intracellular pH on the activity
of the rat hepatocyte organic anion transporter. J Clin Invest. 87:1496, 1991. PMID 2022722
143. Reichen J, Berk PD: Isolation of an organic anion binding protein from rat liver plasma membrane
fractions by affinity chromatography. Biochem Biophys Res Commun. 91:484, 1979. PMID 518646
144. Wolkoff AW, Chung CT: Identification, purification and partial characterization of an organic anion
binding protein from rat liver cell plasma membrane. J Clin Invest. 65:1152, 1980. PMID 7364942
145. Tiribelli C, Lunazzi G, Luciani GL, Panfili E, Gazzin B, Liut G, Sandri G, Sottocasa G: Isolation of a
sulfobromophthalein-binding protein from hepatocyte plasma membrane. Biochim Biophys Acta.
532:105, 1978. PMID 620047
146. Stremmel W, Garber MA, Glazerov V, Thung SN, Kochwa S, Berk PD: Physiochemical and
immunohistological studies of a sulphobromophtalein and bilirubin-binding protein from rat liver plasma
membrane. J Clin Invest. 71:1797, 1983.
147. Stremmel W, Berk PD: Hepatocellular uptake of sulfobromophthalein and bilirubin is selectively
inhibited by an antibody to the liver plasma membrane sulfobromophthalein/bilirubin binding protein. J
Clin Invest. 78:822, 1986. PMID 3745441
148. Sottocasa GL, Baldini G, Sandri G, Lunazzi G, Tiribelli C: Reconstitution in vitro of
sulfobromophthalein transport by bilitranslocase. Biochim Biophys Acta. 685:123, 1982. PMID
6895854
149. Miccio M, Lunazzio GC, Gazzin B, Sottocasa GL: Reconstitution of sulfobromophthalein transport in
erythrocyte membranes induced by bilitranslocase. Biochim Biophys Acta. 1023:140, 1990. PMID
2317493
150. Wolkoff AW, Sosiak A, Greenblatt H, Van Renswoude J, Stockert RJ: Immunological studies of an
organic anion-binding protein isolated from rat liver cell plasma membrane. J Clin Invest. 76:454,
1985. PMID 3897285
151. Goeser T, Nakata R, Braly FL, Sosiak A, Campbell CG, Dermietzel R, Novikoff PM, Stockert RJ, Burk
RD, Wolkoff AW: The rat hepatocyte plasma membrane organic anion binding protein is immunologically
related to the mitochondrial F 1 -adenosine triphosphatase β-subunit. J Clin Invest. 86:220, 1990.
PMID 2142166
152. Jacquemin E, Hagenbuch B, Stieger B, Wolkoff AW, Meier PJ: Expression of the hepatocellular
chloride-dependent sulfobromophthalein uptake system in Xenopus laevis oocytes. J Clin Invest.
88:2146, 1991. PMID 1752967
153. Satlin LM, Amin V, Wolkoff AW: Organic anion transporting polypeptide mediates organic
anion/HCO 3 − exchange. J Biol Chem. 272:26340, 1997. PMID 9334206
154. Kanai N, Lu R, Striano JA, Bao Y, Wolkoff AW, Schuster VL: Identification and characterization of a
prostaglandin transporter. Science. 268:866, 1995. PMID 7754369

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 74 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

155. Bernstein LH, Ben-Ezzar JB, Gartner L, Arias IM: Hepatic intracellular distribution of tritium-labeled
unconjugated and conjugated bilirubin in normal and Gunn rats. J Clin Invest. 45:1194, 1966. PMID
5967697
156. Wolkoff AW, Ketley JN, Waggoner JG, Berk PD, Jakoby W: Hepatic accumulation and intracellular
binding of conjugated bilirubin. J Clin Invest. 61:142, 1978. PMID 618909
157. Levi AJ, Gatmaitan Z, Arias IM: Two hepatic cytoplasmic protein fractions, Y and Z, and their possible
role in the hepatic uptake of bilirubin, sulfobromophthalein, and other anions. J Clin Invest. 48:2156,
1969. PMID 4980931
158. Kamisaka K, Listowsky I, Gatmaitan Z, Arias IM: Interactions of bilirubin and other ligands with
ligandin. Biochemistry. 14:2175, 1975. PMID 1148165
159. Morey KS, Litwack G: Isolation and properties of cortisol metabolite binding proteins of rat liver
cytosol. Biochemistry. 8:4813, 1969. PMID 5365784
160. Ketterer B, Ross-Mansell P, Whitehead JK: The isolation of carcinogen-binding protein from livers of
rats given 4-dimethylaminoazobenzene. Biochem J. 103:316, 1967. PMID 4962081
161. Fleischner G, Robbins J, Arias IM: Immunological studies of Y protein A major cytoplasmic organic
anion binding protein in rat liver. J Clin Invest. 51:677, 1972. PMID 4622106
162. Hays JD, Pulford DJ: The glutathione S-transferase supergene family Regulation of GST and the
contribution of the isoenzymes to cancer chemoprotection and drug resistance. Crit Rev Biochem Mol
Biol. 30:445, 1995. PMID 8770536
163. Rowe JD, Nieves E and Listowsky I: Subunit diversity and tissue distribution of human glutathione
S-transferases Interpretations based on electrospray ionization-MS and peptide sequence-specific
antisera. Biochem J. 325:481, 1997. PMID 9230131
164. Wolkoff AW, Goresky CA, Sellin J, Gatmaitan Z, Arias IM: Role of ligandin in transfer of bilirubin from
plasma into liver. Am J Physiol. 236:E638, 1979. PMID 375751
165. Kamisaka K, Gatmaitan Z, Moore CL, Arias IM: Ligandin reverses bilirubin inhibition of liver
mitochondrial respiration in vitro. Pediatr Res. 9:903, 1975. PMID 1196709
166. Gordon ER, Goresky CA, Chan TH, Perlin AS: The isolation and characterization of bilirubin
diglucuronide, the major bilirubin conjugate in dog and human bile. Biochem J. 155:477, 1976. PMID
949313
167. Heirwegh KPM, Van Hees GP, Leroy P, Van Roy FP, Jansen FH: Heterogeneity of bile pigment
conjugates as revealed by chromatography of their ethyl anthranilate azopigments. Biochem J.
120:877, 1979.
168. Jansen FH, Billing BH: The identification of mono-conjugates of bilirubin in bile as amide derivatives.
Biochemistry. 1:917, 1971.
169. Onishi S, Itho S, Kawade N, Isobe K, Sugiyama S: An accurate and sensitive analysis by high
pressure liquid chromatography of conjugated and unconjugated bilirubin IXα and in various
biological fluids. Biochem J. 185:281, 1980. PMID 6769432
170. Spivak W, Carey MC: Reverse-phase h.p.l.c. separation, quantification and preparation of bilirubin
and its conjugates from native bile. Biochem J. 225:787, 1985. PMID 3919713
171. Roy Chowdhury J, Roy Chowdhury N: Quantitation of bilirubin and its conjugates by high pressure
liquid chromatography. Falk Hepatol. 11:1649, 1982.
172. Gorski JP, Kasper CB: Purification and properties of microsomal UDP glucuronyl transferase from rat
liver. J Biol Chem. 252:1336, 1977. PMID 402364
173. Burchell B: Purification of UDP glucuronyl transferase from untreated rat liver. FEBS Lett. 78:101,
1977. PMID 406146
174. Bock KW, Josling D, Lilenblum WM, Pfeil H: Purification of rat liver glucuronyl transferase-separation
of two enzyme forms inducible by 3-methyl-cholanthrene or phenobarbital. Eur J Biochem. 98:19,
1977.
175. Burchell B: Identification and purification of mutiple forms of UDP-glucuronosyltransferase. Rev
Biochem Toxicol. 3:1, 1981.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 75 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

176. Falany CN, Roy Chowdhury J, Roy Chowdhury N, Tephly TW: Steroid 3- and
17-OH-UDP-glucuronosyltransferase activities in rat and rabbit liver microsomes. Drug Metab Dispos.
11:426, 1983. PMID 6138227
177. Roy Chowdhury J, Roy Chowdhury N, Falany CN, Tephly TW, Arias IM: Isolation and
characterization of multiple forms of rat liver UDP-glucuronoate glucuronosyltransferase. Biochem J.
233:827, 1986. PMID 3085655
178. Harding D, Fournel-Gigleux S, Jackson MR, Burchell B: Cloning and substrate-specificity of a human
phenol-UDP-glucuronosyltransferase expressed in COS-7 cells. Proc Natl Acad Sci U S A. 85:8381,
1988. PMID 3141926
179. Ritter JK, Crawford JM, Owens IS: Cloning of two human liver bilirubin-UDP-glucuronosyltransferase
cDNAs with expression in COS-1 cells. J Biol Chem. 266:1043, 1991. PMID 1898728
180. Wooster R, Sutherland L, Ebner T, Clarke D, da Cruz E, Silva O, Burchell B: Cloning and stable
expression of a new member of a human liver phenol/bilirubin UDP-glucuronosyltransferase cDNA family.
Biochem J. 278:465, 1991. PMID 1910331
181. Iyanagi T, Haniu M, Sogawa K, Fujii-Kuriyama Y, Watanabe S, Shively JE, Anan KF: Cloning and
characterization of cDNA encoding 3-methylcholanthrene inducible rat messenger RNA for
UDP-glucronosyltransferase. J Biol Chem. 261:15607, 1986. PMID 3096993
182. Jackson MR, Burchell B: The full length coding sequence of rat liver androsterone
UDP-glucuronosyltransferase cDNA and comparison with other members of this gene family. Nucleic
Acids Res. 14:779, 1986. PMID 3003696
183. Jackson MR, McCarthy LR, Corser RB, Barr GC, Burchell B: Cloning of cDNAs coding for rat hepatic
microsomal UDP-glucuronosyltransferases. Gene. 34:147, 1985. PMID 2989105
184. Mackenzie PI: Rat liver UDP-glucuronosyltransferase cDNA sequence and expression of a form
glucuronidating 3-hydroxy androgens. J Biol Chem. 261:14112, 1986. PMID 2429951
185. Mackenzie PI, Owens IS, Burchell B, Bock KW, Bairoch A, Belanger A, Fournel-Gigleux S, Green M,
Jum DW, Iyanagi T, Lancet D, Louisot P, Magdalou J, Roy Chowdhury J, Ritter, JK, Schachter H, Tephly
TR, Tipton KF, Nebert DW: The UDP glycosyltransferase gene superfamily Recommended nomenclature
update based on evolutionary divergence. Pharmacogenetics. 7:255, 1997. PMID 9295054
186. Roy Chowdhury J, Novikoff PM, Roy Chowdhury N, Novikoff AB: Distribution of
uridinediphosphoglucuronate glucuronosyl transferase in rat tissues. Proc Natl Acad Sci U S A.
82:2990, 1985. PMID 3921970
187. Dutton GJ, Burchell B: Newer aspects of glucuronidation. Prog Drug Metab. 2:1, 1977.
188. Jansen PLM, Arias IM: Delipidation and reactivation of UDP glucuronosyl transferase from rat liver.
Biochim Biophys Acta. 391:28, 1975.
189. Zakim D, Eibl H: The influence of charge and the distribution of charge in the polar region of
phospholipids on the activity of UDP-glucuronosyltransferase. J Biol Chem. 267:13166, 1992. PMID
1618818
190. Henderson P: The activation in vitro of rat hepatic UDP glucuronyl transferase by ultrasound. Life Sci.
9:511, 1970.
191. Vassey DA, Zakim D: Regulations of microsomal enzymes by phospholipids. J Biol Chem. 246:4649,
1971. PMID 4327323
192. Hauser SC, Ziurys JC, Gollan JL: A membrane transporter mediates access of
uridine-5-diphosphoglucuronic acid from the cytosol into the endoplasmic reticulum of rat hepatocytes.
Implications for glucuronidation reactions. Biochim Biophys Acta. 967:149, 1988. PMID 3142526
193. Heirwegh KPM, Campbell M, Meuwissen JATP: Compartmentation of membrane bound enzymes.
Some basic concepts and consequences for kinetic studies, in Aitio A (ed) Conjugation Reactions in Drug
Biotransformation. Amsterdam. : Elsevier, 1978, p 191.
194. Hallinan T: Comparison of compartmented and of conformational phospholipid-constraint models of
the intramembranous arrangement of UDP-glucuronyltransferase, in Aitio A (ed) Conjugation Reactions in
Drug Biotransformation. Amsterdam. : Elsevier, 1978, p 257.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 76 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

195. Drake RR, Igari Y, Lester R, Elbein AD, Radominska A: Application of 5-azido-UDP-glucose and
5-azido-UDP-glucuronic acid photoaffinity probes for the determination of the active site orientation of
microsomal UDP-glucosyltransferases and UDP-glucuronosyltransferases. J Biol Chem. 267:11360,
1992. PMID 1534561
196. Bosma PJ, Seppen J, Goldhoorn B, Bakker C, Oude Elferink RPJ, Roy Chowdhury J, Roy
Chowdhury N, Jansen PLM: Bilirubin UDP-glucuronosyltransferase 1 is the only relevant bilirubin
glucuronidating isoform in man. J Biol Chem. 269:17960, 1994. PMID 8027054
197. Ritter JK, Chen F, Sheen YY, Tran HM, Kimura S, Yeatman MT, Owens IS: A novel complex locus,
UGT1, encodes human bilirubin, phenol and other UDP-glucuronosyltransferase isozymes with
identical carboxy termini. J Biol Chem. 267:3257, 1992. PMID 1339448
198. Wishart GF: Functional heterogeneity of UDP-glucuronosyl transferase as indicated by its differential
development and inducibility by glucocorticoids. Biochem J. 174:485, 1978. PMID 101211
199. Lillienblum W, Walli AK, Bock KW: Differential induction of rat liver microsomal
UDP-glucuronosyltransferase activities by various inducing agents. Biochem Pharmacol. 31:907,
1982. PMID 6805477
200. Roy Chowdhury J, Roy Chowdhury N, Moscioni AD, Tukey R, Tephley TR, Arias IM: Differential
regulation by triiodothyronine of substrate-specific uridinediphosphoglucuronate glucuronyl transferases in
rat liver. Biochim Biophys Acta. 761:58, 1983. PMID 6416299
201. Hutchinson DW, Johnson B, Knell AJ: The reaction between bilirubin and aromatic diazo compounds.
Biochem J. 127:907, 1972. PMID 5076239
202. van den Bergh AAH, Muller P: Ueber eine direkte und eine indirekte Diazoreaktion auf Bilirubin.
Biochem Z. 77:90, 1916.
203. Talafant E: Properties and composition of bile pigment giving direct diazo reaction. Nature. 178:312,
1956.
204. Heirwegh KPM, Fevery JB, Meuwissen JATP, De Groote J, Compernolle F, Desmet V, Van Roy FP:
Recent advances in the separation and analysis of diazo-positive bile pigments. Methods Biochem
Anal. 22:205, 1974. PMID 4437402
205. Trotman BW, Roy Chowdhury J, Wirt GD, Bernstein SE: Azodipyrole analysis of unconjugated and
conjugated bilirubin using diazotized ethylanthranilate in dimethylsulfoxide. Anal Biochem.
121:175,1982. PMID 7091678
206. Poon R, Hinberg IH: Indican interference with six commercial procedures for measuring total bilirubin.
Clin Chem. 31:92, 1985. PMID 3917384
207. Cole PG, Lathe GH, Billing BH: Separation of the bile pigments of serum, bile and urine. Biochem J.
57:514, 1954.
208. Heirwegh KPM, Fevery J, Michiels R, Van Hees GP, Compernolle F: Separation of thin layer
chromatography and structure elucidation of bilirubin conjugates isolated from dog bile. Biochem J.
145:185, 1975. PMID 1156357
209. Blanckaert N, Kabra PM, Farina FA, BE Stafford, Marton LJ, Schmidt R: Measurement of bilirubin and
its mono- and diconjugates in human serum by alkaline methanolysis and high performance liquid
chromatography. J Lab Clin Med. 96:198, 1980. PMID 7400659
210. Roy Chowdhury J, Roy Chowdhury N, Wu G, Shouval R, Arias IM: Bilirubin monoglucuronide and
diglucuronide formation by human liver in vitro assay by high pressure liquid chromatography.
Hepatology. 1:622, 1981. PMID 6796486
211. Dappen GM, Sundberg MW, Wu TW, Babb BE, Schaeffer JR: A diazo-based dry film for
determination of total bilirubin in serum. Clin Chem. 29:37, 1983. PMID 6848278
212. Kubasik NP, Mayer TK, Baskar AG, Sine HE, D’Souza JP: The measurement of fractionated bilirubin
by Ektachem film slides. Method validation and comparison of conjugated bilirubin measurements with
direct bilirubin in obstructive and hepatocellular jaundice. Am J Clin Pathol. 84:518, 1985. PMID
4036881
213. Schumacher RE, Thornbery JM, Gutcher GR: Transcutaneous bilirubinometry A comparison of old

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 77 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

and new methods. Pediatrics. 76:10, 1985. PMID 4011341


214. Maisels MJ, Kring E: Transcutaneous bilirubinometry decreases the need for serum billirubin
measurement and saves money. Pediatrics. 99:599, 1997. PMID 9093305
215. Tayba R, Gribetz D, Gribetz I, Holtzman IR: Non-invasive estimation of serum bilirubin. Pediatrics.
102:28, 1998.
216. Brown AK, Eisinger J, Blumberg WE, Flores J, Boyle G, Lamola AA: A rapid fluorometric method for
determining bilirubin levels and binding in the blood of neonates Comparison with other methods.
Pediatrics. 65:767, 1980. PMID 7367084
217. Robinson SH, Yannoni C, Nagasawa S: Bilirubin excretion in rats with normal and impaired bilirubin
conjugation. Effect of phenobarbital. J Clin Invest. 50:2606, 1971. PMID 5129312
218. Inoue M, Kinne R, Tran T, Biempica L, Arias IM: Rat liver canalicular membrane vesicles Isolation
and topological characterization. J Biol Chem. 258:5183, 1983. PMID 6833295
219. Kitamura T, Jansen PLM, Hardenbrook C, Kamimoto Y, Gatmaitan Z, Arias IM: Defective
ATP-dependent bile canalicular transport of organic anions in mutant (TR ) rats with conjugated
hyperbilirubinemia. Proc Natl Acad Sci U S A. 87:3557,1990. PMID 2333302
220. Nishida T, Hardenbrook C, Gatmaitan Z, Arias IM: ATP-dependent multi-specific organic anion
transport system in normal and TR rat liver canalicular membrane. Am J Physiol. 262:4, 1992.
221. Ishikaowa T, Muller M, Klunemann C, Schaub T, Keppler D: ATP-dependent primary active transport
of cysteinyl leukotrienes transport system for glutathione S-conjugates. J Biol Chem. 265:19279,
1990. PMID 2172249
222. Kobayashi K, Sogame Y, Hara H, Hayashi K: Mechanism of glutathione S-conjugate transport in
canalicular and basolateral rat liver plasma membranes. J Biol Chem. 265:7737, 1990. PMID
2335503
223. Nishida T, Gatmaitan Z, Roy Chowdhury J, Arias IM: Two distinct mechanisms for bilirubin
glucuronide transport by rat bile canalicular membrane vesicles. J Clin Invest. 90:2130, 1992. PMID
1430236
224. Cornelius CE: Organic anion transport in mutant sheep with congenital hyperbilirubinemia. Arch
Environ Health. 19:852, 1969. PMID 5389192
225. Jansen PLM, Peters WHM, Lamers WH: Hereditary chronic conjugated hyperbilirubinemia in mutant
rats caused by defective hepatic anion transport. Hepatology. 5:573 579, 1985. PMID 4018730
226. Kamimoto Y, Gatmaitan Z, Hsu J, Arias IM: The function of GP 170, the multidrug resistance gene
product in rat liver canalicular membrane vesicles. J Biol Chem. 264:11693, 1989. PMID 2568355
227. Arias IM: Multidrug resistance genes, P-glycoprotein and the liver. Hepatology. 12:159, 1990. PMID
1973680
228. Bushman E, Arceci RJ, Croop JM, Che M, Arias IM, Housman DE, Gross P: Mouse mdr2 encodes
P-glyprotein expressed in the bile canalicular membrane as determined by isoform-specific antibodies.
J Biol Chem. 267:18093, 1992. PMID 1381362
229. Meier PJ, Ruetz ST, Hugentobler G, Fricker G, Kurtz G: Identification and isolation of the putative
canalicular bile acid carrier from rat liver. Hepatology. 5:958, 1985.
230. Nishida T, Gatmaitan Z, Che M, Arias IM: Rat liver canalicular membrane vesicles contain an
ATP-dependent bile acid transport system. Proc Natl Acad Sci U S A. 88:6590, 1991. PMID 1862084
231. Nies AT, Gatmaitan Z, Arias IM: ATP-dependent phosphatidylcholine translocation in rat liver
canalicular plasma membrane vesicles. J Lipid Res. 37:1125, 1996. PMID 8725163
232. Cole SPC, Bhardwaj G, Gerlach JH, Mackie JE, Grant CE, Almquist KC, Stewart AJ, Kurz EU,
Duncan AMV, Deeley RG: Overexpression of a transporter gene in a multidrug-resistant human lung
cancer cell line. Science. 258:1650, 1992. PMID 1360704
233. Jedlitschky G, Leier I, Buchholz U, Center M, Keppler D: ATP-dependent transport of glutathione
S-conjugates by the multidrug resistance-associated protein. Cancer Res. 54:4833, 1994. PMID
7915193
234. Dubin IN, Johnson FB: Chronic idiopathic jaundice with unidentified pigment in liver cells A new

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 78 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

clinicopathologic entity with a report of 12 cases. Medicine (Baltimore). 33:155, 1954.


235. Alpert S, Mosher M, Shanske A, Arias IM: Multiplicity of hepatic excretory mechanism for organic
anions. J Gen Physiol. 53:238, 1969. PMID 5764745
236. Schulman FY, Montali RJ, Bush M, Citino SD, Tell LA, Ballou JD, Hutson TL, St-Pierre M, Dufour JF,
Gatmaitan Z, Arias IM, Johnson FP: Dubin-Johnson-like syndrome in golden lion tamarins
(Leontopithecus rosalia rosalia). Vet Pathol. 30:491, 1993. PMID 8116141
237. Gutstein S, Alpert S, Arias IM: Studies of hepatic excretory function. IV. Biliary excretion of
sulfobromophthalein in a patient with Dubin-Johnson syndrome and a biliary fistula. Isr J Med Sci.
4:46, 1968.
238. Gerloff T, Steiger B, Hagenbuch B, Madon J, Landmann L, Roth J, Hofmann AF, Meier PJ: The sister
of P-glycoprotein represents the canalicular bile salt export pump of mammalian liver. J Biol Chem.
273:10046, 1998. PMID 9545351
239. Smit JJM, Schinkel AH, Elferink RPJO, Groen AK, Wagenaar E, Vandeemter L, Mol CAAM, Ottenhoff
R, Vanderlugt NMT, Vanroon MA, Vandervalk MA, Offerhaus GJA, Berns AJM, Borst P: Homozygous
disruption of the murine mdr2 P glycoprotein gene leads to a complete absence of phospholipid from bile
and to liver disease. Cell. 75:451 462, 1993. PMID 8106172
240. Upson DW, Gronwall RR, Cornelius CE: Maximal hepatic excretion of bilirubin in sheep. Proc Soc
Exp Biol Med. 134:9, 1970. PMID 5423539
241. Klaassen CD, Plaa GL: Studies on the mechanism of phenobarbital-enhanced sulfobromophthalein
disappearance. J Pharmacol Exp Ther. 161:361, 1968. PMID 5650142
242. Barnhart J, Ritt S, Ware A, Coombes B: A comparison of the effects of taurocholate and theophylline
on BSP excretion in dogs, in Paumgartner G, Preisig R (eds) The Liver. : Karger, 1973, p 315. PMID
5650142
243. Gatmaitan ZC, Arias IM: ATP-dependent transport systems in the canalicular membrane of the
hepatocyte. Physiol Rev. 75:261,1995. PMID 7724663
244. Gatmaitan ZC, Nies AT, Arias IM: Regulation and translocation of ATP-dependent apical membrane
proteins in rat liver. Am J Physiol. 272:G1041, 1997. PMID 9176212
245. Lester R, Schmid R: Intestinal absorption of bile pigments. II. Bilirubin absorption in man. N Engl J
Med. 269:178, 1963.
246. Brodersen R, Herman LS: Intestinal reabsorption of unconjugated bilirubin A possible contributing
factor in neonatal jaundice. Lancet. 1:1242, 1963.
247. Alonso EM, Whitington PF, Whitington SH, Rivard WA, Given G: Enterohepatic circulation of
nonconjugated bilirubin in rats fed with human milk. J Pediatr. 118:425, 1991. PMID 1999786
248. Ostrow JD: Absorption of bile pigments by the gall bladder. J Clin Invest. 46:2035, 1967. PMID
6074006
249. Watson CJ: The urobilinoids Milestones in their history and some recent developments, in Berk PD,
Berlin NI (eds). : Department of Health, Education, and Welfare, National Institutes of Health, 1977, p
469. PMID 6074006
250. Moscowitz A, Weiner M, Lightner DA, Petryka ZJ, Davis H, Watson CJ: The in vitro conversion of bile
pigments to the urobilinoids by a rat clostridia species as compared with the human fecal flora. III. Natural
d-urobilin, synthetic I-urobilin, and synthetic I-urobilinogen. Biochem Med. 4:149, 1970.
251. Elder G, Gray CH, Nicholson DG: Bile pigment fate in gastrointestinal tract, in Schmidt R, Jaffe ER,
Miescher PA (eds) Physiology and Disorders of Hemoglobin Degradation. Seminars in Hematology.
New York. : Grune & Stratton, 1972, p 71.
252. Schmid R, Hammaker L: Metabolism and disposition of C 14 -bilirubin in congenital nonhemolytic
jaundice. J Clin Invest. 42:1720, 1963.
253. Lund HT, Jacobsen J: Influence of phototherapy on the biliary bilirubin excretion patterns in newborn
infants with hyperbilirubinemia. J Pediatr. 85:262, 1974. PMID 4842799
254. Berry CS, Zarembo JE, Ostrow JD: Evidence for conversion of bilirubin to dihydroxyl derivatives in
the Gunn rat. Biochem Biophys Res Commun. 49:1366, 1972. PMID 4645549

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 79 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

255. Kapitulnik J, Ostrow JD: Stimulation of bilirubin catabolism in jaundiced Gunn rats as an inducer of
microsomal mixed function mono oxygenases. Proc Natl Acad Sci U S A. 75:682, 1978. PMID
416438
256. Cardenas-Vazquez R, Yokosuka O, Billing BH: Enzymic oxidation of unconjugated bilirubin by rat
liver. Biochem J. 236:625, 1986. PMID 3790083
257. Yokosuka O, Billing BH: Catabolism of bilirubin by intestinal mucosa. Clin Sci. 58:13, 1980.
258. Cameron JL, Pulaski EJ, Abel T, Iber FL: Metabolism and excretion of bilirubin 14 C in experimental
obstructive jaundice. Ann Surg. 163:330, 1966. PMID 4955894
259. Cameron JL, Filler RM, Iber FL, Abel T, Randolph JG: Metabolism and excretion of 14 C labeled
bilirubin in children with biliary atresia. N Engl J Med. 274:231, 1966. PMID 5902217
260. Fulop M, Sandson J, Brazeau P: Dialyzabilty, protein binding, and renal excretion of plasma
conjugated bilirubin. J Clin Invest. 44:666, 1965.
261. Gollan JL, Dallinger KJC, Billing BH: Excretion of conjugated bilirubin in the isolated perfused rat
kidney. Clin Sci Mol Med. 54:381, 1978. PMID 639468
262. Stocker R, Yamamoto Y, McDonagh AF, Glazer AN, Ames BN: Bilirubin is an antioxidant of possible
physiological importance. Science. 235:1043, 1987. PMID 3029864
263. Stocker R, Peterhans E: Antioxidant properties of conjugated bilirubin and biliverdin; biologically
relevant scavenging of hypochlorous acid. Free Radic Res Commun. 6:57, 1989. PMID 2542140
264. Malik R, Wrehota EM, Brass CA, Gollan JL: Membrane antioxidant activity of bile pigments is
dependent on the ability to donate hydrogen ions. Hepatology. 12:933, 1990.
265. Wu TW, Carey D, Wu J, Sugiyama H: The cytoprotective effects of bilirubin and biliverdin on rat
hepatocytes and human erythrocytes and the impact of albumin. Biochem Cell Biol. 69:828, 1991.
PMID 1818587
266. Hardy JB, Peeples MO: Serum bilirubin levels in newborn infants. Distributions and associations with
neurological abnormalities during the first year of life. Johns Hopkins Med J. 128:265, 1971. PMID
5556539
267. Gartner LM, Lee K, Vaisman S, Lane D, Zarafu I: Development of bilirubin transport and metabolism
in the newborn Rhesus monkey. J Pediatr. 90:513, 1977. PMID 402463
268. Maisels MJ, Pathak A, Nelson NM, Nathan DG, Smith CA: Endogenous production of carbon
monoxide in normal and erythroblastic newborn infants. J Clin Invest. 50:1, 1971. PMID 5543875
269. Vest M, Strebel L, Hauensiein D: The extent of “shunt” bilirubin and erythrocyte survival in the
newborn infant measured by the administration of ( 15 N) glycine. Biochem J. 95:11c, 1965.
270. Poland RL, Odell GB: Physiologic jaundice The enterohepatic circulation of bilirubin. New Engl J Med.
284:1, 1971. PMID 4922346
271. Clarke CA, Donohoe WTA, Finn R, Lehane D, McConnell RB, Sheppard PM, Towers SH, Woodrow
JC, Bowley CC, Tovey LAD, Bias WB, Krevans JR: Combined study Prevention of Rh hemolytic
disease. : 607, 1971. PMID 4103276
272. Haberman S, Kraft EJ, Leucke PE, Peach RO: ABO isoimmunization The use of the specific Coombs
and best elution tests in the detection of hemolytic disease. J Pediatr. 56:471, 1960.
273. Levi AJ, Gatmaitan Z, Arias IM: Deficiency of hepatic organic anion-binding protein, impaired organic
anion uptake by liver and “physiologic” jaundice in newborn monkeys. N Engl J Med. 283:1136, 1970.
PMID 4990787
274. Brown AK, Zuelzer WW: Studies on the neonatal development of the glucuronide conjugating system.
J Clin Invest. 37:332, 1958.
275. Bosma PJ, Roy Chowdhury J, Bakker C, Gantla S, DeBoer A, Oostra BA, Lindhout D, Tytgat GNJ,
Jansen PLM, Oude Elferink RPJ and Roy Chowdhury N: The genetic basis of the reduced expression of
bilirubin UDP-glucuronosyltransferase 1 in Gilbert’s syndrome. N Engl J Med. 333:1171, 1995. PMID
7565971
276. Bancroft JD, Kreamer B, Gourley GR: Gilbert syndrome accelerates development of neonatal
jaundice. J Pediatr. 132:656, 1998. PMID 9580766

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 80 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

277. Roy Chowdhury N, Deocharan B, Bejjanki HR, Gantla S, Roy Chowdhury J, Koliopoulos C,
Petmezaki S, Valaes T: The presence of a Gilbert-type promoter abnormality increases the level of
neonatal hyperbilirubinemia. Hepatology. 26:370a, 1997.
278. Arthur LJH, Bevan BR, Holton JB: Neonatal hyperbilirubinemia and breast feeding. Dev Med Child
Neurol. 8:279, 1966. PMID 5926432
279. Maisels MJ, Newman TB: Kernicterus in otherwise healthy breast-fed term newborns. Pediatrics.
96:730, 1995. PMID 7567339
280. Arias IM, Gartner LM, Seifter S, Furman M: Prolonged neonatal unconjugated hyperbilirubinemia
associated with breast feeding and a steroid, pregnane-3(alpha), 20(beta)-diol, in maternal milk that
inhibits glucuronide formation in vitro. J Clin Invest. 43:2037, 1964.
281. Holton JB, Lathe GH: Inhibitors of bilirubin conjugation in newborn infant serum and male urine. Clin
Sci. 25:499, 1963.
282. Arias IM, Gartner LM: Production of unconjugated hyperbilirubinemia in full-term newborn infants
following administration of pregnane-3(alpha), 20(beta)-diol. Nature. 203:1292, 1964.
283. Ramos A, Silberberg M, Stern I: Pregnanediols and neonatal hyperbilirubinemia. Am J Dis Child.
111:353, 1966. PMID 5906044
284. Foliot A, Ploussard JP, Housett E, Christoforov B, Luzean R, Odièvre M: Breast milk jaundice In vitro
inhibition of rat liver bilirubin-uridine diphosphate glucuronyl transferase activity and Z
protein-bromosulfophthalein binding by human breast milk. Pediatr Res. 10:594, 1976. PMID 818610
285. Lucey JF, Driscol JJ: Physiological jaundice re-examined, in Sass-Kortsak A (ed) Kernicterus.
Toronto. : University of Toronto Press, 1961, p 29.
286. Arias IM, Wolfson S, Lucey JF, McKay RJ: Transient familial neonatal hyperbilirubinemia. J Clin
Invest. 44:1442, 1965.
287. Hsia DY-T, Patterson P, Allen FH, Diamond LK, Gellis SS: Prolonged obstructive jaundice in infancy
General survey of 156 cases. Pediatrics. 10:243, 1952.
288. Snyder AL, Satterlee W, Robinson SH, Schmid R: Conjugated plasma bilirubin in jaundice caused by
pigment overload. Nature. 213:93, 1967.
289. Robinson S, Vanier T, Desforges JF, Schmid R: Jaundice in thalassemia minor A consequence of
“ineffective erythropoiesis. ” New Engl J Med. 267:512, 1962.
290. Israels LG, Zipursky A: Primary shunt hyperbilirubinemia due to an alternate path of bilirubin
production. Am J Med. 27:693, 1959.
291. Berendsohn S, Lowman J, Sundberg D, Watson CJ: Idiiopathic dyserythropoietic jaundice. Blood.
24:1, 1964.
292. Verwilghen R, Verhaegen H, Waumanns P, Beert J: Ineffective erythropoiesis with morphologically
abnormal erythroblasts and unconjugated hyperbilirubinemia. Br J Haematol. 17:27, 1969. PMID
5807786
293. Verwilghen R, Lewis S, Dacie J, Crookston J, Crookston M: Hempas Congenital dyserythropoietic
anaemia (type II). Q J Med. 42:257, 1973. PMID 4785435
294. Crigler JF, Najjar VA: Congenital familial non-hemolytic jaundice with kernicterus. Pediatrics. 10:169,
1952.
295. Childs B, Sidbury JB, Migeon CJ: Glucuronic acid conjugation by patients with familial non-hemolytic
jaundice and their relatives. Pediatrics. 23:903, 1959.
296. Berk PD, Martin JF, Blaschke TF, Scharschmidt BF, Plotz PH: Unconjugated hyperbilirubinemia
Physiological evaluation and experimental approaches to therapy. Ann Intern Med. 82:552, 1975.
PMID 1119773
297. Sleisenger MG, Kahn I, Barniville H, Rubin W, Benezzer J, Arias IM: Nonhemolytic unconjugated
hyperbilirubinemia with hepatic glucuronyl transferase deficiency A genetic study in four generations.
Trans Assoc Am Physicians. 80:259, 1967. PMID 6082246
298. Szabo L, Ebrey P: Studies on the inheritance of Crigler-Najjar syndrome by the menthol test. Acta
Paediatr Hung. 4:153, 1963.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 81 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

299. Wolkoff AW, Chowdhury JR, Gartner LA, Rose AL, Biempica L, Giblin DR, Fink D, Arias IM:
Crigler-Najjar syndrome (Type I) in an adult male. Gastroenterology. 76:840, 1979. PMID 422012
300. Kapitulnik J, Kaufmann NA, Goitein K, Cividalli G, Blondheim SH: A pigment found in the
Crigler-Najjar syndrome and its similarity to an ultra-filterable photo-derivative of bilirubin. Clin Chim
Acta. 57:231, 1974. PMID 4434643
301. Arias IM, Gartner LM, Cohen M, Benezzer J, Levi AJ: Chronic nonhemolytic unconjugated
hyperbilirubinemia with glucuronyl transferase deficiency Clinical, biochemical, pharmacologic, and
genetic evidence for heterogeneity. Am J Med. 47:395, 1969. PMID 4897277
302. Billing GH, Gray CH, Kulcycka A, Manfield P, Nicholson DC: The metabolism of 14 C-bilirubin in
congenital nonhaemolytic hyperbilirubinaemia. Clin Sci. 27:163, 1964.
303. Blaschke TF, Berk PD, Scharschmidt BF, Guyther JR, Vergalla J, Waggoner JG: Crigler-Najjar
syndrome An unusual course with development of neurologic damage at age eighteen. Pediatr Res.
8:573, 1974. PMID 4825526
304. Novikoff AB, Essner E: The liver cell. Am J Med. 19:102, 1960.
305. Bosma PJ, Roy Chowdhury N, Goldhoorn BG, Hofker MH, Oude Elferink RPJ, Jansen PLM, Roy
Chowdhury J: Sequence of exons and the flanking regions of human
bilirubin-UDP-glucuronosyltransferase gene complex and identification of a genetic mutation in a patient
with Crigler-Najjar syndrome, type I. Hepatology. 15:941, 1992. PMID 1568736
306. Bosma PJ, Roy Chowdhury J, Huang TJ, Lahiri P, Oude Elferink RPJ, Van ES HHG, Lederstein M,
Whitington PF, Jansen PLM, Roy Chowdhury N: Mechanism of inherited deficiencies of multiple
UDP-glucuronosyltransferase isoforms in two patients with Crigler-Najjar syndrome, type I. FASEB J.
6:2859, 1992. PMID 1634050
307. Ritter JK, Yeatman MT, Ferriera P, Owens IS: Identification of a genetic alteration in the code for
bilirubin UDP-glucuronosyltransferase in the UGT1 gene complex of a Crigler-Najjar syndrome, type
I. J Clin Invest. 90:150, 1992. PMID 1634606
308. Kadakol A, Ghosh SS, Sappal BS, Sharma G, Roy Chowdhury J, Roy Chowdhury N: Genetic lesions
of bilirubin uridinediphosphoglucuronate glucuronosyltransferase causing Crigler-Najjar Gilbert’s
syndromes. : Correlation of genotype to phenotype. Hum Mutat.(In Press). PMID 1634606
309. Seppen J, Steenken R, Lindhout D, Bosma PJ, Oude Elferink RP: A mutation which disrupts the
hydrophobic core of the signal peptide of bilirubin UDP-glucuronosyltransferase, an endoplasmic reticulum
membrane protein, causes Crigler-Najjar type II. FEBS Lett. 390:294, 1996 PMID 8706880
310. Gantla S, Bakker CTM, Deocharan B, Thummala NR, Zweiner J, Sinaasappel M, Roy Chowdhury J,
Bosma PJ, Roy Chowdhury N: Splice site mutations A novel genetic mechanism of Crigler-Najjar
syndrome type 1. Am J Hum Genet. 62:585, 1998. PMID 9497253
311. Deocharan B, Gantla S, Morton DH, Rizack L, Roy Chowdhury J, Roy Chowdhury N: Interaction of a
Crigler-Najjar syndrome type I mutation and a Gilbert type promoter defect results in two grades of
hyperbilirubinemia in members of an Amish and a Mennonite kindred of Lancaster County, Pennsylvania.
Gastroenterology. 112:1255A, 1997.
312. Gunn CH: Hereditary acholuric jaundice in a new mutant strain of rats. J Hered. 29:137, 1938.
313. Roy Chowdhury J, Van ES HHG, Roy Chowdhury N: Gunn rat. An animal model of deficiency of
bilirubin conjugation, in Tavoloni N, Berk PD (eds) Hepatic Transport and Bile Secretion. Physiology and
Pathophysiology. New York. : Raven, 1992, p 713.
314. Roy Chowdhury N, Kondapalli R, and Roy Chowdhury J: The Gunn rat An animal model for inherited
deficiency of bilirubin glucuronidation, in Cornelius CE (ed). : Academic, 1993, p 150.
315. Schmid R, Axelrod J, Hammaker L, Swarn RL: Congenital jaundice in rats due to a defective
glucuronide formation. J Clin Invest. 37:1123, 1958.
316. Schutta HS, Johnson L: Bilirubin encephalopathy in the Gunn rat A fine structure study of the
cerebellar cortex. J Neuropathol Exp Neurol. 26:377, 1967. PMID 6027437
317. Rose AL, Johnson A: Bilirubin encephalopathy Neuropathological and histochemical studies in the
Gunn rat model. Neurology. 22:420, 1972.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 82 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

318. Schutta HS, Johnson L: Clinical signs and morphologic abnormalities in Gunn rats treated with
sulfadiethoxine. J Pediatr. 75:1070, 1969. PMID 5389393
319. Call NB, Tisher CC: The urinary concentrating defect in the Gunn strain of rat. Role of bilirubin. J Clin
Invest. 55:319, 1975. PMID 1127102
320. Axelsen RA: Spontaneous renal papillary necrosis in the Gunn rat. Pathology. 5:43, 1973. PMID
4693279
321. Gardner WA, Konigsmark B: Familial nonhemolytic jaundice Bilirubinosis and encephalopathy.
Pediatrics. 43:365, 1969. PMID 5775189
322. Iyanagi T: Molecular basis of multiple UDP-glucuronosyltransferase isoenzyme deficiencies in the
hyperbilirubinemic rat (Gunn rat). J Biol Chem. 266:24048, 1991. PMID 1748678
323. Roy Chowdhury J., Huang TJ, Kasari K, Lederstein M, Arias IM, Roy Chowdhury N: Molecular basis
for the lack of bilirubin-specific and 3-methylcholanthrene-inducible UDP-glucuronosyltransferase activities
in Gunn rats The two isoforms are encoded by distinct mRNA species that share an identical single base
deletion. J Biol Chem. 266:18294, 1991. PMID 1717446
324. Sato H, Aono S, Kashiwamata S, Koiwai O: Genetic defect of bilirubin UDP-glucuronosyltransferase
in the hyperbilirubinemic Gunn rat. Biochem Biophys Res Commun. 177:1161, 1991. PMID 1840486
325. Elawady M, Roy Chowdhury J, Kesari K, Van ES HHG, Lederstein M, Arias IM, Roy Chowdhury N:
Mechanism of the lack of induction of UDP-glucuronosyltransferase activity by 3-methylcholanthrene in
Gunn rats. J Biol Chem. 265:10752, 1990. PMID 2113060
326. Watkins JB, Klassen CD: Induction of UDP-glucuronosyltransferase activities in Gunn, heterozygous
and Wistar rat livers by pregnenolone-16 alpha-carbonitrile. Drug Metab Dispos. 10:590, 1982. PMID
6130905
327. Stevenson IH, Greenwood DT, McEwen J: Hepatic UDP-glucuronosyltransferase in Wistar and Gunn
rats—in vitro activation by diethylnitrosamine. Biochem Biophys Res Commun. 32:866, 1968. PMID
5682310
328. Roy Chowdhury N, Gross F, Moscioni AD, Kram M, Arias IM, Roy Chowdhury J: Isolation and
purification of multiple normal and functionally defective forms of UDP-glucuronosyltransferase from livers
of inbred Gunn rats. J Clin Invest. 79:327, 1987. PMID 3100574
329. Arias IM: Ethereal and N-linked glucuronide formation by normal and Gunn rats in vitro and in vivo.
Biochem Biophys Res Commun. 6:81, 1961.
330. Drucker WD: Glucuronic acid conjugation of tetrahydrocortisone p-nitrophenol in the homozygous
Gunn rats. Proc Soc Exp Biol Med. 129:308, 1968. PMID 5686533
331. Flock EV, Bollman JL, Owen CA, Zollman PE: Conjugation of thyroid hormones and analogues by the
Gunn rat. Endocrinology. 77:303, 1965. PMID 5836170
332. Howan ER, Guarino AM: Biliary excretion of phenol red by Wistar and Gunn rats. Proc Soc Exp Biol
Med. 146:46, 1974. PMID 4827266
333. Van der Veer CN, Schoemaker B, Bakker C, van der Meer R, Jansen PLM, Oude Elferink RP:
Influence of dietary calcium phosphate on the disposition of bilirubin in rats with unconjugated
hyperbilirubinemia. Hepatology. 24:620, 1996. PMID 8781334
334. van der Veere CN, Sinaasappel M, McDonagh AF, Rosenthal P, Labrune P, Odièvre M, Fevery J,
Otte JB, McClean P, Bürk G, Masakowski V, Sperl W, Mowat AP, Vergani GM, Heller K, Wilson JP,
Shepherd R, Jansen PL: Current therapy for Crigler-Najjar syndrome type 1 Report of a world registry.
Hepatology. 24:311, 1996. PMID 8690398
335. Vroemen JPAM, Blankaert N, Buurman WA, Heirwegh KPM, Koostra G: Treatment of enzyme
deficiency by hepatocyte transplantation in rats. J Surg Res. 39:267, 1985. PMID 3928971
336. Gupta S, Aragona E, Vemuru RP, Bhargava KK, Burk RD, Roy Chowdhury J: Permanent engraftment
and functions of hepatocytes delivered to the liver Implications for gene therapy and liver repopulation.
Hepatology. 14:144, 1991. PMID 2066062
337. Ilan Y, Roy Chowdhury N, Prakash R, Jona V, Attavar P, Guha C, Tada K, Roy Chowdhury J:
Massive repopulation of rat liver by transplantation of hepatocytes into specific lobes of the liver and

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 83 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

ligation of portal vein branches to other lobes. Transplantation. 64:8, 1997. PMID 9233693
338. Demetriou AA, Levenson SM, Whiting J, Feldman D, Moscioni AD, Kram M, Roy Chowdhury N, Roy
Chowdhury J: Replacement of hepatic functions in rats by transplantation of microcarrier-attached
hepatocytes. Science. 233:1190, 1986. PMID 2426782
339. Fox IJ, Roy Chowdhury J, Kaufman SS, Goertzen TC, Roy Chowdhury N, Warkentin PI, Dorko BS,
Sauter BV, Strom SC: Treatment of Crigler-Najjar syndrome type I with hepatocyte transplantation.
New Engl J Med. 338:1422, 1998. PMID 9580649
340. Roy Chowdhury J, Strom S, Fox IJ: Human hepatocyte transplantation Gene therapy and more?
Pediatrics. 102:647, 1998. PMID 9738191
341. Lavin A, Sung C, Klibanov AM, Langer R: Enzymatic removal of bilirubin from blood A potential
treatment for neonatal jaundice. Science. 230:543, 1985. PMID 4048947
342. Sugi K, Inoue M, Morino Y: Degradation of plasma bilirubin by a bilirubin oxidase derivative which has
a relatively long half-life in the circulation. Biochim Biophys Acta. 991:405, 1989. PMID 2499360
343. Kapitulnik J: The role of cytochrome P-450 in the alternate pathways of bilirubin metabolism in
congenitally jaundiced Gunn rats and infants with the Crigler-Najjar syndrome, type I. International
Bilirubin Workshop,Trieste, 1992, p 53
344. Roy Chowdhury J, Grossman M, Gupta S, Roy Chowdhury N, Baker JR, Wilson JM: Long term
improvement of hypercholesterolemia after ex vivo gene therapy in LDL-receptor deficient rabbits.
Science. 254:1802, 1991. PMID 1722351
345. Tada K, Roy-Chowdhury N, Prasad V, Kim B-H, Manchikalapudi P, Fox IJ, van Duijvendijk P, Bosma
PJ, Roy-Chowdhury J: Long-term amelioration of bilirubin glucuronidation defect in Gunn rats by
transplanting genetically modified immortalized autologous hepatocytes. Cell Transplant. 7:607, 1998.
PMID 9853589
346. Tada K, Roy Chowdhury N, Neufeld D, Bosma PJ, Heard M, Prasad VR, Roy Chowdhury J:
Long-term reduction of serum bilirubin levels in Gunn rats by retroviral gene transfer in vivo. Liver
Transplant Surg. 4:78, 1998
347. Takahashi M, Ilan Y, Roy Chowdhury N, Guida J, Horwitz MS, Roy Chowdhury J: Long-term
correction of bilirubin UDP-glucuronosyltransferase deficiency in Gunn rats by administration of a
recombinant adenovirus during the neonatal period. J Biol Chem. 271:26536, 1996. PMID 8900123
348. Ilan Y, Attavar P, Takahashi M, Davidson A, Horwitz M, Guida J, Roy Chowdhury N, Roy Chowdhury
J: Induction of central tolerance by intrathymic inoculation of adenoviral antigens into the host thymus
permits long-term gene therapy in Gunn rats. J Clin Invest. 98:2640, 1996. PMID 8958229
349. Ilan Y, Prakash R, Davidson A, Jona V, Droguett G, Horwitz MS, Roy Chowdhury N, Roy Chowdhury
J: Oral tolerization to adenoviral antigens permits long-term gene expression using recombinant
adenoviral vectors. J Clin Invest. 99:1098, 1997. PMID 9062369
350. Ilan Y, Sauter B, Roy Chowdhury N, Reddy B, Thummala NR, Groguett G, Davidson A, Ott M,
Horwitz MS, Roy Chowdhury J: Oral tolerization to adenoviral proteins permits repeated
adenovirus-mediated gene therapy in rats with preexisting immunity to adenovirus. Hepatology.
27:1368, 1998. PMID 9581693
351. Wilson JM, Wu GY, Wu CH, Grossman M, Roy Chowdhury N, Roy Chowdhury J:
Hepatocyte-directed gene transfer in vivo leads to transient improvement of hypercholesterolemia in
LDL-receptor-deficient rabbits. J Biol Chem. 267:963, 1992. PMID 1370472
352. Roy Chowdhury N, Wu CH, Wu GY, Yerneni PC, Bommineni VR, Roy Chowdhury J: Fate of DNA
targeted to the liver by asialoglycoprotein receptor-mediated endocytosis in vivo Prolonged
persistence in cytoplasmic vesicles after partial hepatectomy. J Biol Chem. 268:11265, 1993. PMID
8496181
353. Bommineni VR, Roy Chowdhury N, Wu GY, Wu CH, Franki N, Hays RM, Roy Chowdhury J:
Depolymerization of hepatocellular microtubules after partial hepatectomy. J Biol Chem. 269:25200,
1994. PMID 7929209
354. Kren BT, Bandyopadhyay P, Steer CJ: In vivo site-directed mutagenesis of the factor IX gene by

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 84 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

chimeric RNA/DNA oligonucleotides. Nat Med. 4:274, 1998. PMID 9500596


355. Kren BT, Parashar B, Bandyopadhyay P, Chowdhury NR, Chowdhury JR, Steer CJ: Correction of the
UDP-glucuronosyltransferase gene defect in the Gunn rat model of Crigler-Najjar syndrome type I with a
chimeric oligonucleotide. Proc Natl Acid Sci USA. 96:10349, 1999.
356. Arias IM: Chronic unconjugated hyperbilirubinemia without overt signs of hemolysis in adolescents
and adults. J Clin Invest. 41:2233, 1962.
357. Gollan JL, Huang SM, Billing B, Sherlock S: Prolonged survival in three brothers with severe type II
Crigler-Najjar syndrome. Ultrastructural and metabolic studies. Gastroenterology. 68:1543, 1975.
PMID 805737
358. Gordon ER, Shaffer EA, Sass-Kortsak A: Bilirubin secretion and conjugation in the Crigler-Najjar
syndrome type II. Gastroenterology. 70:761, 1976. PMID 1261771
359. Fevery J, Blanckaert N, Heirwegh KPM, Preaux A-M, Berthelot P: Unconjugated bilirubin and an
increased proportion of bilirubin monoconjugates in the bile of patients with Gilbert’s syndrome and
Crigler-Najjar syndrome. J Clin Invest. 60:970, 1977. PMID 409736
360. Thompson RPH, Pilcher CWT, Robinson J, Strathers GM, McLean AEM, Williams R: Treatment of
unconjugated jaundice with dicophane. Lancet. 2:4, 1969. PMID 4182808
361. Hunter J, Thompson RPH, Rake MO, Williams R: Controlled trial of phetharbital, a non-hypnotic
barbiturate, in unconjugated hyperbilirubinaemia. BMJ. 2:497, 1971. PMID 4931488
362. Orme MLE: Increased glucuronidation of bilirubin in men and rat by administration of antipyrine
(phenazone). Clin Sci Mol Med. 46:511, 1974. PMID 4208404
363. Blaschke TF, Berk PD, Rodkey FL, Scharschmidt BF, Collison HA, Waggoner JG: Effects of
glutethimide and phenobarbital on hepatic bilirubin clearance, plasma bilirubin turnover, and carbon
monoxide production in man. Biochem Pharmacol. 23:2795, 1974. PMID 4429586
364. Gabilan JC, Benattar C, Lindenbaum A: Clofibrate treatment of neonatal jaundice. Pediatrics. 86:647,
1990. PMID 2216638
365. Seppen J, Bosma P, Roy Chowdhury J, Roy Chowdhury N, Jansen PLM, Oude Elferink R:
Discrimination between Crigler-Najjar syndrome type I and II by expression of mutant
bilirubin-UDP-glucuronosyltransferase. J Clin Invest. 94:2385, 1994. PMID 7989595
366. Gilbert A, Lereboullet P: La cholamae simple familiale. Semin Med. 21:241, 1901.
367. Thompson RPH: Genetic transmission of Gilbert’s syndrome, in Okolicsanyi L (ed). : Familial
Hyperbilirubinemia. New York, Wiley, 1981, p 91.
368. Powell LW, Hemingway E, Billing BH, Sherlock S: Idiopathic unconjugated hyperbilirubinemia
(Gilbert’s syndrome) A study of 42 families. New Engl J Med. 277:1108, 1967. PMID 6054997
369. Black M, Billing BH: Hepatic bilirubin UDP glucuronyltransferase activity in liver disease and Gilbert’s
syndrome. New Engl J Med. 280:1266, 1969. PMID 5770050
370. Auclair C, Hakim J, Boivin P, Troube H, Boucherrot J: Bilirubin and paranitrophenol glucuronyl
transferase activity of the liver in patients with Gilbert’s syndrome. Enzyme. 21:97, 1976. PMID
816648
371. Berk PD, Bloomer JR, Howe RB, Berlin NI: Constitutional hepatic dysfunction (Gilbert’s syndrome) A
new definition based on kinetic studies with unconjugated radiobilirubin. Am J Med. 49:296, 1970.
PMID 5455561
372. Goresky CA, Gordon ER, Shaffer EA, Parie P, Carassavas D, Aronoff A: Definition of a conjugation
dysfunction in Gilbert’s syndrome Studies of the handling of bilirubin loads and of the pattern of bilirubin
conjugates secreted in bile. Clin Sci Mol Med. 1:63, 1978.
373. Berk PD, Blaschke TF, Waggoner JG: Defective BSP clearance in patients with constitutional hepatic
dysfunction (Gilbert’s syndrome). Gastroenterology. 63:472, 1972. PMID 5071284
374. Cartel GVM, Chisesi T, Cazzavillian M, Barbui T, Battista R, Dini E: Bromsulphthalein-Ausscheidung
und Hyperbilirubinamia beim Gilbert Syndrome. Dtsch Z Verdau Stoffwechselkr. 35:169, 1975. PMID
1227811
375. Cobelli C, Ruggeri A, Toffolo G, Okolicsanyi L, Venuti M, Orlando R: BSP vs bilirubin kinetics in

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 85 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

Gilbert’s syndrome, in Okolicsanyi L (ed) Familial Hyperbilirubinemia. New York. : Wiley, 1981, p 121.
376. Felsher BF, Rickard D, Redeker AG: The reciprocal relation between caloric intake and the degree of
hyperbilirubinemia in Gilbert’s syndrome. New Engl J Med. 283:170, 1970. PMID 5424007
377. Bloomer JR, Barrett PV, Rodkey FL, Berlin NI: Studies on the mechanisms of fasting
hyperbilirubinemia. Gastroenterology. 61:479, 1971. PMID 5114639
378. Barrett PVD: Hyperbilirubinemia of fasting. JAMA. 217:1349, 1971. PMID 5109641
379. Felsher BF, Carpio NM, Van Couvering K: Effect of fasting and phenobarbital on hepatic
UDP-glucuronic acid formation in the rat. J Lab Clin Med. 93:414, 1979. PMID 107254
380. Gollan JL, Hatt KJ, Billing BH: The influence of diet on unconjugated hyperbilirubinemia in the Gunn
rat. Clin Sci Mol Med. 49:229, 1975. PMID 1175339
381. Cowan RE, Thompson RPH, Kaye JP, Clark GM: The association between fasting
hyperbilirubinaemia and serum non-esterified fatty acids in man. Clin Sci Mol Med. 53:155, 1977.
PMID 891104
382. Fromke VL, Miller D: Constitutional hepatic dysfunction (CHD Gilbert’s disease). : 451, 1972. PMID
4564417
383. Ohkubo H, Musha H, Okuda K: Studies on nicotinic acid interaction with bilirubin metabolism. Dig Dis
Sci. 24:700, 1979. PMID 487925
384. Bosma PJ, Roy Chowdhury J, Bakker C, Gantla S, DeBoer A, Oostra BA, Lindhout D, Tytgat GNJ,
Jansen PLM, Oude Elferink RPJ, Roy Chowdhury N: The genetic basis of the reduced expression of
bilirubin UDP-glucuronosyltransferase 1 in Gilbert’s syndrome. N Engl J Med. 333:1171, 1995. PMID
7565971
385. Chalasani N, Roy Chowdhury N, Roy Chowdhury J, Boyer TD: Kernicterus in an adult who is
heterozygous for Crigler-Najjar syndrome and homozygous for Gilbert-type genetic defect.
Gastroenterology. 112:2099, 1997. PMID 9178703
386. Portman OW, Roy Chowdhury J, Roy Chowdhury N, Alexander M, Cornelius CE, Arias IM: A
non-human primate model for Gilbert’s syndrome. Hepatology. 4:175, 1984. PMID 6706295
387. Portman OW, Alexander M, Roy Chowdhury J, Roy Chowdhury N, Cornelius CE, Arias IM: Effects of
nutrition on hyperbilirubinemia in Bolivian squirrel monkeys. Hepatology. 4:454, 1984. PMID 6724513
388. Sprinz H, Nelson RS: Persistent nonhemolytic hyperbilirubinemia associated with lipochrome-like
pigment in liver cells Report of four cases. Ann Intern Med. 41:952, 1954.
389. Dubin IN: Chronic idiopathic jaundice A review of fifty cases. Am J Med. 23:268, 1958.
390. Cohen L, Lewis C, Arias IM: Pregnancy, oral contraceptives, and chronic familial jaundice with
predominantly conjugated hyperbilirubinemia (Dubin-Johnson syndrome). Gastroenterology. 62:1182,
1972. PMID 5050316
391. Shani M, Seligsohn U, Gilon E, Sheba C, Adam A: Dubin-Johnson syndrome in Israel. I. Clinical,
laboratory, and genetic aspects of 101 cases. West J Med. 39:549, 1970.
392. Levanon M, Rimon S, Shani M, Ramot B, Goldberg E: Active and inactive factor-VII in Dubin-Johnson
syndrome with factor-VII deficiency, hereditary factor-VII deficiency and on coumadin administration.
Br J Haematol. 23:669, 1972. PMID 4675008
393. Javitt NB, Konso T, Kuchiba K: Bile acid excretion in Dubin-Johnson syndrome. Gastroenterology.
75:931, 1978. PMID 700338
394. Kondo T, Yagi R, Kuchiba K: Dubin-Johnson syndrome in a neonate. New Engl J Med. 292:1028,
1975. PMID 1117948
395. Nakata F, Oyanagi K, Fujiwara M, Sogawa H, Minain R, Horino K, Nakao T, Kondo T: Dubin-Johnson
syndrome in a neonate. Eur J Pediatr. 132:299, 1979. PMID 520367
396. Dittrich H, Seifert E: Uber das verhalten des pigmentes sowie der biligrafin auscheidung bei einem
patienten mit Dubin-Johnson syndrom. Acta Hepatosplenol. 9:45, 1962.
397. Morita M, Kihava T: Intravenous cholecystography and metabolism of meglumine iodipamide
(biligrafin) in Dubin-Johnson syndrome. Radiology. 99:57, 1971. PMID 5548689

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 86 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

398. Essner E, Novikoff AB: Human hepatocellular pigments and lysosomes. J Ultrastruct Res. 3:3764,
1960.
399. Ehrlick JC, Novikoff AB, Platt R, Essner E: Hepatocellular lipofuscin and the pigment of chronic
idiopathic jaundice. Bull NY Acad Med. 36:488, 1960.
400. Wegmann R, Rangier M, Eteve J, Charbonnier A, Caroli J: Melanose hepatosplenique avec ictere
chronique a bilirubine directe Maladie de Dubin-Johnson? Etude clinique et biologique de la maladie.
Etudie histochimique, chimique et spectrographique du pigment anormal. Semin Hop Paris. 26:1761,
1960.
401. Arias IM, Bernstein L, Roffler R, Ben Ezzer J: Black liver diseases in Corriedale sheep Metabolism of
tritiated epinephrine and incorporation of isotope into the hepatic pigment in vivo. J Clin Invest.
44:1026, 1065.
402. Swartz HM, Sarna T, Varma RR: On the nature and excretion of the hepatic pigment in the
Dubin-Johnson syndrome. Gastroenterology. 76:958, 1979. PMID 220131
403. Kitamura T, Alroy J, Gatmaitan Z, Inoue M, Mikami T, Kansen PLM, Arias IM: Defective biliary
excretion of epinephrine metabolites in mutant TR rats Relation to the pathogenesis of rat liver in
Dubin-Johnson syndrome and Corriedale sheep with an analogous excretory defect. Hepatology.
15:1154, 1992. PMID 1592353
404. Rubinstein ZJ, Seligson U, Modan M, Shani M: Hepatic computerized tomography in the
Dubin-Johnson syndrome Increased liver density as a diagnostic acid. Comput Radiol. 9:315, 1985.
PMID 4064636
405. Ware A, Eigenbrodt E, Naftalis J, Combes B: Dubin-Johnson syndrome and viral hepatitis.
Gastroenterology. 67:560, 1974. PMID 4852761
406. Billing BH, Williams R, Richards TG: Defects in hepatic transport of bilirubin in congenital
hyperbilirubinaemia. An analysis of plasma bilirubin disappearance curves. Clin Sci. 27:245, 1964.
407. Schoenfield LJ, McGill DB, Hunton DB, Foulk MT, Butt HR: Studies of chronic idiopathic jaundice
(Dubin-Johnson syndrome). I. Demonstration of hepatic excretory defect. Gastroenterology. 44:101,
1963.
408. Erlinger S, Dhumeaux D, Desjeux JF, Benhamou JP: Hepatic handling of unconjugated dyes in the
Dubin-Johnson syndrome. Gastroenterology. 64:106, 1973. PMID 4683844
409. Mandema E, De Fraiture WH, Neiweg HO, Arends A: Familial chronic idiopathic jaundice
(Dubin-Sprinz disease), with a note on bromsulphalein metabolism in this disease. Am J Med. 28:42,
1960.
410. Rodes J, Zubizarreta A, Bruguera M: Metabolism of the bromsulphalein in Dubin-Johnson syndrome.
Diagnostic value of the paradoxical in plasma levels of BSP. Dig Dis. 17:545, 1972.
411. Koskelo P, Toivonen I, Adlercreutz H: Urinary coproporphyrin isomer distribution in Dubin-Johnson
syndrome. Clin Chem. 13:1006, 1967. PMID 6054688
412. Kaplowitz N, Javitt N, Kappas A: Coproporphyrin I and III excretion in bile and urine. J Clin Invest.
51:2895, 1972. PMID 5080415
413. Kondo T, Kuchiba K, Shimizu Y: Coproporphyrin isomers in Dubin-Johnson syndrome.
Gastroenterology. 70:1117, 1976. PMID 1269872
414. Wolkoff AW, Cohen LE, Arias IM: Inheritance of the Dubin-Johnson syndrome. N Engl J Med.
288:113, 1973. PMID 4682034
415. Ben-Ezzer J, Blonder J, Shani M, Seligsohn U, Post CA, Adam A, Szeinberg A: Dubin-Johnson
syndrome. Abnormal excretion of the isomers of urinary coproporphyrin by clinically unaffected family
members. Isr J Med Sci. 9:1431, 1973. PMID 4775130
416. Edwards RH: Inheritance of the Dubin-Johnson-Sprinz syndrome. Gastroenterology. 68:734, 1975.
PMID 1123140
417. Wolkoff AW, Arias IM: Coproporphyrin excretion in amniotic fluid and urine from premature infants A
possible maturation defect. Pediatr Res. 8:591, 1974. PMID 4825527
418. Aziz MA, Schwartz S, Watson CJ: Studies of coproporphyrin. VIII. Reinvestigation of the isomer

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 87 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

distribution in jaundice and liver diseases. J Lab Clin Med. 63:596, 1964.
419. Shimizu Y, Kondo T, Kuchiba K, Urata G: Uroporphyrin III cosynthetase in liver and blood in the
Dubin-Johnson syndrome. J Lab Clin Med. 89:517, 1977. PMID 839109
420. Kondo T, Kuchiba K, Shimizu Y: Metabolic fate of exogenous delta-aminolevulinic acid in
Dubin-Johnson syndrome. J Lab Clin Med. 94:421, 1979. PMID 469378
421. Oude Elferink RPJ, Meijer DKF, Kuipers F, Jansen PLM, Groen AK, Groothuis GMM: Hepatobiliary
secretion of organic compounds; molecular mechanisms of membrane transport. Biochim Biophys
Acta. 1241:215, 1995. PMID 7640297
422. Jansen, PLM, van Klinken JW, van Gelder M, Ottenhoff R, Oude Elferink RPJ: Preserved organic
anion transport in mutant TR rats with a hepatobiliary secretion defect. Am J Physiol. 265:G445
G452, 1993. PMID 8214066
423. Huber M, Guhlmann A, Jansen PLM, Keppler D: Hereditary defect of hepatobiliary cysteinyl
leukotriene elimination in mutant rats with defective hepatic anion excretion. Hepatology. 7:224, 1987.
PMID 3030916
424. Oude Elferink RPJ, Ottenhoff R, Liefting WGM, Schoemaker B, Groen AK, Jansen PLM: ATP
dependent efflux of GSSG and GS conjugate from isolated rat hepatocytes. Am J Physiol. 258:G699,
1990. PMID 2333997
425. Kobayashi K, Sogame Y, Hayashi K, Nicotera P, Orrenius S: ATP stimulates the uptake of S
dinitrophenylglutathione by rat liver plasma membrane vesicles. FEBS Lett. 240:55, 1988. PMID
3191999
426. Akerboom TPM, Narayanaswami V, Kunst M, Sies H: ATP dependent S (2,4 dinitrophenyl)glutathione
transport in canalicular plasma membrane vesicles from rat liver. J Biol Chem. 266:13147, 1991.
PMID 2071597
427. Kobayashi K, Komatsu S, Nishi T, Hara H, Hayashi K: ATP dependent transport for glucuronides in
canalicular plasma membrane vesicles. Biochem Biophys Res Commun. 176:622, 1991. PMID
2025276
428. Oude Elferink, RPJ, de Haan J, Lambert KJ, Hagey LR, Hofmann AF, Jansen PLM: Selective
hepatobiliary transport of nordeoxycholate side chain conjugates in mutant rats with a canalicular
transport defect. Hepatology. 9:861, 1989. PMID 2714736
429. Higgins CF: ABC transporters From microorganisms to man. Annu Rev Cell Biol. 867:1650, 1992.
430. Paulusma CC, Bosma PJ, Zaman GJR, Bakker CTM, Otter M, Scheffer GL, Scheper RJ, Borst P,
Oude Elferink RPJ: Congenital jaundice in rats with a mutation in a multidrug resistance-associated
protein gene. Science. 271:1126, 1996. PMID 8599091
431. Ling V: P-glycoprotein its role in drug resistance. Am J Med 99(suppl). : 31, 1995.
432. Muller M, Meijer C, Zaman GJR, Borst P, Scheper RJ, Mulder NH, de Vires EGE, Jansen PLM:
Overexpression of the multidrug resistance-associated protein results in increased ATP-dependent
glutathione S-conjugate transport. Proc Natl Acad Sci U S A. 91:13033, 1994. PMID 7809167
433. Evers R, Zaman GJR, van Deemter L, Jansen H, Calafat J, Oomen LCJM, Oude Elferink RPJ, Borst
P, Schinkel AH: Basolateral localization and export activity of the human multidrug resistance-associated
protein in polarized pig kidney cells. J Clin Invest. 97:1, 1996. PMID 8550819
434. Paulusma CC, Kool M, Bosma PJ, Scheffer GL, ter Borg F, Scheper RJ, Borst P, Baas F, Oude
Elferink RPJ: A mutation in the human cMOAT gene causes the Dubin Johnson gene. Hepatology.
25:1539, 1997. PMID 9185779
435. Allikmets R, Gerrard B, Hutchinson A, Dean M: Characterization of the human ABC superfamily
Isolation and mapping of 21 new genes using the Expressed Sequence Tags database. Hum Mol
Gen. 5:1649, 1996. PMID 8894702
436. Kartenbeck J, Leuschner U, Mayer R, Keppler D: Absence of the canalicular isoform of the
MRP-gene encoded conjugate export pump from the hepatocytes in Dubin Johnson syndrome.
Hepatology. 23:1061, 1996. PMID 8621134
437. Gourley GR: Bilirubin metaboolism and kernicterus. Adv Pediatr. 44:173, 1997. PMID 9265971

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 88 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

438. Seligsohn U, Shani M, Ramot B, Adam A, Sheba C: Dubin-Johnson Syndrome in Israel. II.
Association with factor-VII deficiency. Q J Med. 39:569, 1970. PMID 5499883
439. de Grouchy J, Dautzenberg M-D, Turleau C, Beguin S, Chavin-Colin F: Regional mapping of clotting
factors VII and X to 13q34. Expression of factor VII through chromosome 8. Hum Genet. 66:230,
1984. PMID 6714981
440. Rotor AB, Manahan L, Florentin A: Familial nonhemolytic jaundice with direct van den Bergh reaction.
Acta Med Phil. 5:37, 1948.
441. Pereira-Lima JE, Utz E, Rosenberg I: Hereditary nonhemolytic conjugated hyperbilirubinemia without
abnormal liver cell pigmentation. A family study. Am J Med. 40:628, 1966.
442. Wolkoff AW, Wolpert E, Pascasio FN, Arias IM: Rotor’s syndrome A distinct inheritable
pathophysiologic entity. Am J Med. 60:173, 1976. PMID 766621
443. Wolpert E, Pascasio FM, Wolkoff AW, Arias IM: Abnormal sulfobromophthalein metabolism in Rotor’s
syndrome and obligate heterozygotes. New Engl J Med. 296:1099, 1977. PMID 850521
444. Kawasaki H, Kinwa N, Irisa T, Hirayama C: Dye clearance studies in Rotor’s syndrome. Am J
Gastroenterol. 71:380, 1979. PMID 453129
445. Dhumeaux D, Berthelot P: Chronic hyperbilirubinemia associated with hepatic uptake and storage
impairment A new syndrome resembling that of the mutant Southdown sheep. Gastroenterology.
69:988, 1975. PMID 1175891
446. Shimizu Y, Naruto H, Ida S, Kohakura M: Urinary coproporphyrin isomers in Rotor’s syndrome. A
study in eight families. Hepatology. 1:173, 1981. PMID 7286897
447. Rapacini GL, Topi GC, Anti M, D’Allasandro GL, Griso D, Amantea A, Devitis I, Fedeli G: Porphyrins
in Rotor syndrome A study on an Italian family. Hepatogastroenterology. 33:11, 1986. PMID 3957223
448. Localio SA, Schwartz MS, Gannon CF: The urinary/fecal coproporphyrin ratio in liver disease. J Clin
Invest. 20:7, 1941.
449. Clayton RJ, Iber FL, Ruebner BH: Byler disease. : 112, 1969. PMID 5762004
450. Bull LN, van Eijk MJ, Pawlikowaska L, DeYoung JA, Juijn JA, Liao M, Klomp LW, Lomri N, Berger R,
Scharshmidt BF, Knisely AS, Houwen RH, Freimer NB: A gene encoding a P-type ATPase mutated in two
forms of hereditary cholestasis. Nat Genet. 18:219, 1998. PMID 9500542
451. Summerskill WHJ, Walshe JM: Benign recurrent intrahepatic obstructive jaundice. Lancet. 2:686,
1959.
452. Williams R, Cartter MA, Sherlock S, Sgheuer PJ, Hill KR: Idiopathic recurrent cholestasis A study of
the functional and pathological lesions in four cases. Q J Med. 33:387, 1964.
453. De Pagter AGF, Van Berge Henegouwen GP, Bokkel-Huinnuk JA, Brandt K-H: Familial benign
recurrent intrahepatic cholestasis. Gastroenterology. 71:202, 1976. PMID 939378
454. Summerskill WHJ: The syndrome of benign recurrent cholestasis. Am J Med. 38:298, 1965.
455. Tygstrup N, Jensen B: Intermittent intrahepatic cholestasis of unknown etiology in five young males
from the Faroe Islands. Acta Med Scand. 185:523, 1969. PMID 5807632
456. Summerfield JA, Scott J, Berman M, Ghent C, Bloomer JR, Berk PD, Sherlock S: Benign recurrent
intrahepatic cholestasis Studies of bilirubin kinetics, bile acids, and cholangiography. Gut. 21:154,
1980. PMID 7380339
457. Biempica L, Gutstein S, Arias IM: Morphological and biochemical studies of benign recurrent
cholestasis. Gastroenterology. 52:521, 1967. PMID 6019971
458. Strautnieks SS, Kagalwalla AF, Tanner MS, Knisley AS, Bull L, Freimer N, Kocoshis SA, Gardiner
RM, Thompson RJ: Identification of a locus for progressive familial intrahepatic cholestasis PFIC2 on
chromosome 2q24. Am J Hum Genet. 61:630, 1997. PMID 9326328
459. De Vree J, Jacquemin E, Strum E: Mutations in the MDR3 gene cause progressive familial
intrahepatic cholestasis. Proc Acad Natl Sci U S A. 25:282, 1998.
460. Zhang N, Arias IM: Deficient MDR3 expression in liver from patients with Navajo neuropathy A human
homologue of mdr2 −/− mice. Hepatology. 26:69A, 1997.

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 89 - Any use is subject to the Terms of Use on www.ommbid.com.
Chapter 125: Hereditary Jaundice and Disorders of Bilirubin Metabolism

461. Oda T, Elkaholoun AG, Meltzer PS, Chandrashekharappa SC: Identification and cloning of the human
homolog (Jag1) of the rat jagged 1 gene from the Alagille syndrome critical region at 20p12.
Genomics. 43:376, 1997. PMID 9268641

Copyright © The McGraw-Hill Companies, Inc. All rights reserved. - 90 - Any use is subject to the Terms of Use on www.ommbid.com.

You might also like