Download PDF

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Hypothyroidism wikipedia , lookup

Hyperthyroidism wikipedia , lookup

Transcript
M I N I R E V I E W
Minireview: The Sodium-Iodide Symporter NIS
and Pendrin in Iodide Homeostasis of the Thyroid
Aigerim Bizhanova and Peter Kopp
Division of Endocrinology, Metabolism, and Molecular Medicine, Feinberg School of Medicine, Northwestern
University, Chicago, Illinois 60611
Thyroid hormones are essential for normal development and metabolism. Thyroid hormone biosynthesis requires iodide uptake into the thyrocytes and efflux into the follicular lumen, where it
is organified on selected tyrosyls of thyroglobulin. Uptake of iodide into the thyrocytes is mediated
by an intrinsic membrane glycoprotein, the sodium-iodide symporter (NIS), which actively cotransports two sodium cations per each iodide anion. NIS-mediated transport of iodide is driven by the
electrochemical sodium gradient generated by the Na⫹/K⫹-ATPase. NIS is expressed in the thyroid,
the salivary glands, gastric mucosa, and the lactating mammary gland. TSH and iodide regulate
iodide accumulation by modulating NIS activity via transcriptional and posttranscriptional mechanisms. Biallelic mutations in the NIS gene lead to a congenital iodide transport defect, an autosomal recessive condition characterized by hypothyroidism, goiter, low thyroid iodide uptake, and
a low saliva/plasma iodide ratio. Pendrin is an anion transporter that is predominantly expressed
in the inner ear, the thyroid, and the kidney. Biallelic mutations in the SLC26A4 gene lead to
Pendred syndrome, an autosomal recessive disorder characterized by sensorineural deafness, goiter, and impaired iodide organification. In thyroid follicular cells, pendrin is expressed at the apical
membrane. Functional in vitro data and the impaired iodide organification observed in patients
with Pendred syndrome support a role of pendrin as an apical iodide transporter. (Endocrinology
150: 1084 –1090, 2009)
T
he iodide-containing thyroid hormones T3 and its precursor
T4 are crucial for normal development, growth, and regulation of numerous metabolic pathways. The main function of
the thyroid gland is to concentrate iodide and to make it available
for biosynthesis of thyroid hormones. The significance of this
mechanism is evident in light of the scarcity of iodide in most of
the environment and the fact that insufficient dietary supply of
iodide remains a major public health issue in many parts of the
world (1).
The synthesis of thyroid hormones requires a normally developed thyroid gland, an adequate nutritional intake of iodide,
and a series of sequential biochemical steps. Thyroid hormone
synthesis takes place in the follicles, the functional units of the
gland (2). Each follicle consists of a single layer of thyroid epithelial cells surrounding the follicular lumen. The follicular lumen is filled with colloid, which is predominantly composed of
thyroglobulin, a large glycoprotein that serves as the scaffold for
thyroid hormone synthesis (3).
The synthesis of thyroid hormones requires uptake of iodide
across the basolateral membrane into the thyrocytes, transport
across the cell, and efflux through the apical membrane into the
follicular lumen. Uptake of iodide is mediated by the sodiumiodide symporter (NIS), which cotransports two sodium ions
along with one iodide ion, with the sodium gradient serving as
the driving force (Fig. 1) (1). The energy required to produce the
sodium gradient is provided by the ouabain-sensitive Na⫹/K⫹ATPase (4). The efflux of iodide across the apical membrane is
mediated, at least in part, by pendrin (5). Once iodide reaches the
cell-colloid interface, it is oxidized and rapidly organified by
incorporation into selected tyrosyl residues of thyroglobulin.
This reaction, referred to as organification, is catalyzed by thyroid peroxidase (TPO) in the presence of hydrogen peroxide
and results in the formation of mono- and diiodotyrosines
(MIT and DIT). The generation of hydrogen peroxide is mediated by the calcium-dependent reduced nicotinamide adenine dinucleotide phosphate (NADPH) dual oxidase type 2
ISSN Print 0013-7227 ISSN Online 1945-7170
Printed in U.S.A.
Copyright © 2009 by The Endocrine Society
doi: 10.1210/en.2008-1437 Received October 14, 2008. Accepted January 22, 2009.
First Published Online February 5, 2009
Abbreviations: CICn5, Chloride channel 5; DIT, diiodotyrosine; DUOX2, dual oxidase type
2; ITD, iodide transport defect; MDCK, Madin-Darby canine kidney; MIT, monoiodotyrosine; NIS, sodium-iodide symporter; PKA, protein kinase A; SLC5A, solute carrier 5A; TPO,
thyroperoxidase.
1084
endo.endojournals.org
Endocrinology, March 2009, 150(3):1084 –1090
Downloaded from endo.endojournals.org at Eidgenossische Technische Hochschule Bibliothek on June 12, 2009
Endocrinology, March 2009, 150(3):1084 –1090
Follicular lumen
endo.endojournals.org
I
HO -
CH2
I
DIT
HO -
I
I
O
CH2
T3
I
I
HO HO -
CH2
Organification
CH2
MIT
Coupling
TG
TG
HO -
TPO
CH2
TPO
I
TG
HO -
H2O2
I-
CH2
DIT
I
I
HO -
CH2
DIT
I
H2O2
I-
HO -
I
O
I
CH2
T4
I
I
TG
H2O2
I-
nat
di e
TG
d
I-
Io
TG
Apical membrane
TPO
PDS
DUOX2
DUOXA2
NADP+ NADPH
TG
I-
Ca2+
Hydrolysis
DEHAL1
K+
Na+
Na/K-ATPase
2 Na+
NIS
MIT
DIT
T4
T3
I-
?
Basolateral membrane
Na+
FIG. 1. Main steps in thyroid hormone synthesis. At the basolateral membrane
of thyroid follicular cells, which form the follicles, iodide is transported into
thyrocytes by the NIS. NIS is dependent on the sodium gradient created by the
Na/K-ATPase. At the apical membrane, iodide efflux is, in part, mediated by
pendrin (PDS/SLC26A4). At the cell-colloid interface, iodide is oxidized by TPO in
the presence of H2O2. H2O2 is produced by the calcium- and reduced
nicotinamide adenine dinucleotide phosphate-dependent (NADPH) enzyme
DUOX2. DUOX2 requires a specific maturation factor, DUOXA2. Thyroglobulin
(TG), which is secreted into the follicular lumen, serves as matrix for synthesis of
T4 and T3. First, TPO catalyzes iodination of selected tyrosyl residues
(organification), which results in the formation of MIT and DIT. Subsequently,
two iodotyrosines are coupled to form either T4 or T3 in a reaction that is also
catalyzed by TPO. Iodinated thyroglobulin is stored as colloid in the follicular
lumen. Upon a demand for thyroid hormone secretion, thyroglobulin is
internalized into the follicular cell by pinocytosis and digested in lysosomes, which
generates T4 and T3 that are released into the bloodstream through unknown
mechanisms. The unused MIT and DIT are retained in the cell and deiodinated by the
iodotyrosine dehalogenase 1 (DEHAL1). The released iodide is recycled for thyroid
hormone synthesis.
(DUOX2). TPO also catalyzes the coupling of two iodotyrosines to form either T3 or T4. To release thyroid hormones,
thyroglobulin is engulfed by pinocytosis, digested in lysosomes, and then secreted into the bloodstream at the basolateral membrane (2). The mechanisms of hormone secretion
at the basolateral membrane and the involved channel(s) have
not been characterized.
NIS
NIS is an integral plasma membrane glycoprotein localized at the
basolateral plasma membrane of thyrocytes (6). This protein
plays an essential role in thyroid physiology by mediating uptake
of iodide into the thyrocytes, a key step in thyroid hormone
synthesis. The ability of the thyroid gland to accumulate radioiodine is also a cornerstone in the diagnosis and treatment of
thyroid disorders (7).
1085
The molecular characterization of NIS began in 1996 when
the cDNA encoding rat NIS was isolated by expression-cloning
in Xenopus laevis oocytes (6). Subsequently, the human cDNA
has been cloned by a RT-PCR approach taking advantage of the
homology to rat NIS (8). Rat NIS is predicted to have 618 amino
acids with a relative molecular mass of 65,196 Daltons (6), and
human NIS contains 643 amino acids and exhibits an 84%
amino acid identity and 93% similarity to rat NIS (9). The human NIS gene is located on chromosome 19p12-13.2 and contains 14 introns and 15 exons (9).
NIS (SLC5A5) belongs to the solute carrier family 5A
(SLC5A). All members of this protein family depend on an electrochemical sodium gradient as the driving force for transport of
anions across the plasma membrane (10). The current secondary
structure model predicts that NIS contains 13 transmembrane
domains with the amino terminus located extracellularly and the
carboxy terminus facing the cytosol (1). The mature NIS protein
is approximately 87 kDa in size and has three asparagine-linked
glycosylation sites (1). Glycosylation does not seem to be required for stability, activity, or targeting of the NIS molecule to
the plasma membrane (11). The expression of NIS is differentially regulated and is subjected to various posttranslational
modifications in each tissue in which it is expressed (1, 12).
In addition to thyroid follicular cells, NIS is expressed in several other tissues, including the salivary glands, gastric mucosa,
and the lactating mammary gland, where it mediates active transport of iodide (1). In the lactating mammary gland, NIS plays an
important role by concentrating iodide in the milk, thereby supplying newborns with iodide for thyroid hormone synthesis.
Although NIS has a high affinity for iodide, it is able to mediate transport of other ions (13). Large anions such as thiocyanate and perchlorate can inhibit accumulation of iodide in the
thyroid by competing with iodide (14, 15). Perchlorate is 10 –100
times more potent than thiocyanate in inhibiting iodide accumulation in the thyroid (1). The ability of perchlorate to block
iodide transport has been used in the therapy of hyperthyroidism, and it is used in the perchlorate discharge test, which serves
to detect defects in iodide organification (16). In normal individuals, administration of perchlorate blocks subsequent accumulation of iodide in the thyrocytes but does not cause any discharge of previously accumulated radioiodine because of its
rapid organification. In contrast, in individuals with a total or
partial iodide organification defect, administration of perchlorate results in the rapid release of the unorganified fraction of the
tracer from the thyrocytes (16). Until recently, it has been controversial whether perchlorate acts as a blocker or as a substrate
that is transported by NIS (13, 17, 18). Two recent studies provide evidence that perchlorate is actively transported by NIS (19,
20). Perchlorate transport could be demonstrated in a polarized
cell system (19) as well as by direct measurement with mass
spectrometry (20). Moreover, it has been shown that perchlorate, a widely found pollutant, is transported into the milk (19).
Remarkably, the stoichiometry of the NIS-mediated Na⫹/ClO4⫺
transport is electroneutral, which contrasts with the electrogenic
transport of iodide (two Na⫹ and one I⫺) (19). These findings
indicate that NIS is able to transport different substrates with
distinct stoichiometries (19).
Downloaded from endo.endojournals.org at Eidgenossische Technische Hochschule Bibliothek on June 12, 2009
1086
Bizhanova and Kopp
Minireview
Regulation of NIS protein expression
TSH is the major regulator of thyroid cell proliferation, differentiation, and function, including iodide uptake (21). The effects
of TSH are primarily mediated through the activation of the
cAMP cascade via the GTP-binding protein Gs␣ (22). TSH stimulates iodide accumulation by positively regulating NIS expression at the protein and mRNA level via the cAMP pathway (23).
Hypophysectomized rats with low circulating levels of TSH have
a decreased protein expression of NIS, whereas a single injection
of TSH leads to a prompt increase in NIS expression (24). Rats
maintained on an iodide-deficient diet or treated with propylthiouracil, an agent blocking iodide organification, have high
concentrations of TSH, which correlates with an increase in NIS
protein expression. These findings are in agreement with the
results obtained in human thyroid primary cultures (25, 26) and
the rat thyroid FRTL-5 cell line (23). In FRTL-5 cells, withdrawal
of TSH results in a decrease in intracellular cAMP concentration
and iodide uptake activity (23). Re-addition of TSH increases
NIS mRNA and protein expression and subsequently restores
iodide uptake activity.
Recent studies have shown that TSH not only regulates NIS
transcription and biosynthesis but also mediates NIS activity by
posttranscriptional mechanisms (27). In the presence of TSH,
NIS is active and inserted in the basolateral membrane of thyrocytes (27). Upon TSH withdrawal, NIS protein half-life decreases from 5 to 3 d, and it translocates from the plasma membrane to intracellular compartments (27). As of yet, the
mechanisms regulating the subcellular distribution of NIS are
only partially elucidated. It is known that NIS has several consensus sites for kinases, including protein kinase A (PKA) and
protein kinase C (1). Although it has been shown that NIS is
phosphorylated in vivo (27), the functional significance of this
modification needs further characterization. A more recent study
identified five amino acid residues in NIS that are phosphorylated in vivo, but the phosphorylation status of these amino acid
residues does not affect targeting of NIS to the plasma membrane
(28). NIS contains several sorting sequences that are known to
play a role in targeting, retention, and endocytosis of other membrane proteins (29). For example, the PDZ motif (T/S-X-V/L)
located at the carboxy terminus of NIS is one of the sequences
involved in protein-protein interactions (1). The PDZ motif is
recognized by PDZ-binding proteins that have been implicated in
internalization of other transporters (29). NIS also has a
dileucine motif, L557L558, which is known to interact with the
clathrin-coated system (30). This interaction leads to incorporation of integral membrane proteins into coated vesicles that are
then carried to different destinations within the cell (31).
Iodide is another factor that can regulate iodide accumulation
in the thyroid. In 1948, Wolff and Chaikoff (32) reported that
high doses of iodide block iodide organification in the rat thyroid
in vivo. This phenomenon, known as the acute Wolff-Chaikoff
effect, is a reversible process, because iodide organification resumes when the iodide concentration in the serum decreases. The
mechanisms underlying the Wolff-Chaikoff effect are complex
and involve acute regulation of several key genes and proteins
within the thyrocytes. Several studies have examined the effect of
Endocrinology, March 2009, 150(3):1084 –1090
iodide on NIS mRNA and protein expression in vivo and in vitro
(33–35). In vivo data suggest that high concentrations of iodide
lead to reduction in both NIS mRNA and protein levels, partially
by a transcriptional mechanism. In vitro results suggest that exposure to high doses of iodide results in a decrease in NIS protein
levels that is, at least in part, due to an increase in NIS protein
turnover (33–35).
Congenital iodide transport defect (ITD)
Biallelic mutations in the NIS gene cause a congenital ITD. ITD
is an autosomal recessive condition characterized by hypothyroidism, goiter, reduced or absent thyroid uptake of radioiodide,
and a low saliva/plasma iodide ratio (1, 36). Currently, at least
12 ITD-causing mutations of NIS have been identified (37). Six
of these mutations, namely 226delH, T354P, G395R, Q267E,
G543E, and V59E have been characterized more thoroughly
(38 – 41). The G543E substitution leads to retention of NIS in
intracellular compartments as a result of improper maturation
and trafficking of the protein, the other mutants are still targeted
to the membrane but result in a loss of function (38 – 41). Structural and functional analysis of the T354P mutant protein demonstrated that a hydroxyl group at the ␤-carbon of the residue at
position 354 is crucial for proper NIS function (38). Substitutions of the glycine residue at position 395 residue with several
amino acids indicated that the presence of a small and an uncharged amino acid residue at this position is required for NIS
function (39). Lastly, a recent study revealed that the histidine
residue at position 226 is important for the iodide transport
activity of NIS (37).
Pendrin
Pendrin is a highly hydrophobic membrane protein located at the
apical membrane of thyrocytes (2, 4). In addition to the thyroid,
pendrin is also expressed in the kidney and in the inner ear (42,
43). In the kidney, pendrin plays an important role in acid-base
metabolism as an exchanger of chloride and bicarbonate in ␤-intercalated cells (44). In the inner ear, pendrin is important for
generation of the endocochlear potential (45).
Pendrin belongs to the SLC26A family, which includes several
anion transporters, as well as the motor protein prestin that is
expressed in outer hair cells (46, 47). Pendrin is encoded by the
SLC26A4 gene, which was cloned in 1997 (48). The SLC26A4
gene is located on chromosome 7q21-31 and contains 21 exons
with an open reading frame of 2343 bp (48).
Pendrin is a glycoprotein composed of 780 amino acids (2). It
contains three putative extracellular asparagine-glycosylation
sites (4, 49). Pendrin usually appears as a single protein band
with a molecular mass of 110 –115 kDa when isolated from
human thyroid membranes (49). Pendrin is proposed to have 12
transmembrane domains with both amino and carboxy termini
located inside the cytosol (4, 50). Like other members of the
SLC26A family, pendrin contains a so-called STAS (sulfate
transporter and antisigma factor antagonist) domain (51). The
Downloaded from endo.endojournals.org at Eidgenossische Technische Hochschule Bibliothek on June 12, 2009
Endocrinology, March 2009, 150(3):1084 –1090
exact function of this domain has not been elucidated. Recent
studies, however, suggest that the STAS domain can interact with
the regulatory domain of CFTR (cystic fibrosis transmembrane
conductance regulator) in certain epithelial cells (52–54).
Pendred syndrome
Mutations in the SLC26A4 gene lead to Pendred syndrome (2).
Pendred syndrome is an autosomal recessive disorder characterized by sensorineural deafness, goiter, and a partial defect in
iodide organification (55, 56). Deafness or hearing impairment
is the leading clinical sign of Pendred syndrome (56). In many
patients, hearing loss is prelingual; in some individuals, however,
the hearing loss develops later in childhood (57). Patients with
Pendred syndrome display an enlarged endolymphatic sac and
duct (58 – 60). A subset of patients presents with a so-called
Mondini defect, which is characterized by replacement of the
cochlear turns by a single cavity or a rudimentary cochlea (57,
58). Variability in the hearing loss observed in patients with
SLC26A4 mutations suggests that the phenotype is influenced by
environmental factors and/or genetic modifiers (61).
Goiter usually develops during childhood. There is, however,
a substantial variation within and between families and different
geographic regions (62– 64). Nutritional iodide intake appears
to play an important role as a modifier of the thyroidal phenotype (65, 66). Under conditions of high iodide intake, most individuals have no or only a mild enlargement of the thyroid (60,
67). If the nutritional iodide is scarce, patients with Pendred
syndrome not only develop goiter but may also present with mild
or overt hypothyroidism (68, 69).
Mutations in the SLC26A4 gene, found in patients with Pendred syndrome, are highly heterogeneous (56). Currently, more
than 150 mutations of the SLC26A4 gene have been reported
(56). Most of the mutations are missense mutations, and a
smaller number of mutations consists mainly of nonsense and
intronic mutations (56). Loss of function of some of the mutants
results from the retention of the mutated and misfolded protein
in intracellular compartments, most likely the endoplasmic reticulum (70, 71).
Role of pendrin in the thyroid
Initial functional studies of pendrin in Xenopus oocytes have
demonstrated that pendrin is able to mediate transport of chloride and iodide (5) and that it can act as a chloride/formate
exchanger (72). The ability of pendrin to mediate iodide efflux
(5), the localization of pendrin at the apical membrane of thyrocytes (4, 73), as well as the defect in iodide organification
observed in patients with Pendred syndrome (62, 74), suggested
that pendrin could function as an apical iodide transporter in
thyroid cells (46). The results obtained from a number of independent studies performed in heterologous systems support the
role of pendrin in mediating, at least in part, apical iodide efflux.
Yoshida et al. (75) have demonstrated that iodide efflux is much
higher in nonpolarized Chinese hamster ovary cells expressing
endo.endojournals.org
1087
NIS and pendrin than in cells expressing NIS alone. Electrophysiological studies with transfected COS-7 cells also indicate that
pendrin mediates iodide transport and that it is more efficient at
high extracellular concentrations of chloride (76). In addition,
iodide efflux/chloride influx appears to be more efficient than
chloride efflux/iodide influx (76). These findings are consistent
with results obtained in polarized Madin-Darby canine kidney
(MDCK) cells (50). MDCK cells expressing NIS and pendrin
independently or simultaneously were cultured in a bicameral
system, which allowed measuring iodide uptake at the basolateral membrane and iodide efflux at the apical membrane. Cells
transfected with NIS alone have a significant increase in intracellular iodide uptake compared with untransfected control
cells. In contrast, cells expressing NIS and pendrin show a significant increase in iodide transport into the apical chamber and
consequently a significant decrease in the intracellular iodide
content. These findings support the notion that pendrin may
have a role in facilitating vectorial iodide transport at the apical
membrane (50). The partial organification defect found in patients with Pendred syndrome suggests, however, that iodide can
reach the follicular lumen independently of the presence of
pendrin.
Questions concerning the role of pendrin as
an apical iodide transporter
The traditional concept held that iodide simply crosses the apical
membrane due to the electrochemical gradient that is present
between the cytosol and the follicular lumen (77). However,
autoradiography studies revealed that iodide first accumulates in
the cytosol and subsequently moves to the follicular lumen (78).
This transport of iodide across the apical membrane is rapidly
stimulated by TSH (64, 79). Hence, it has been proposed that
apical iodide efflux is mediated through a specific transporter or
channel. This concept has risen based on several findings. Electrophysiological studies performed with thyroid membrane vesicles suggested the presence of two apical iodide transporters
(80). Functional studies performed in heterologous cells, including polarized cells (50, 75, 76, 81), along with the iodide organification defects found in patients with Pendred syndrome (62,
74), suggested that pendrin mediates apical iodide efflux in thyrocytes. The physiological role of pendrin has, however, been
questioned for several reasons (77). Patients with biallelic mutations in the SLC26A4 gene display a mild or no thyroidal phenotype under conditions of sufficient iodide intake (65). The
pendrin knockout mice, studied under normal iodide intake conditions, do not develop a goiter or abnormal thyroid hormone
levels (45). In addition, it is intriguing that pendrin may have
distinct roles in the thyroid, the inner ear, and the kidney (77).
This led to the proposal that pendrin may be a part of multiprotein complex, the composition of which may vary among different cell types, thereby potentially explaining a variability in anion selectivity of pendrin (77). Other proteins (SLC5A8 and
chloride channel 5 ClCn5) have been proposed to mediate apical
iodide efflux (82, 83). Functional studies performed in Xenopus
oocytes and polarized MDCK cells clearly demonstrate that
Downloaded from endo.endojournals.org at Eidgenossische Technische Hochschule Bibliothek on June 12, 2009
1088
Bizhanova and Kopp
Minireview
SLC5A8, originally designated as human apical iodide transporter (hAIT) (82), does not mediate iodide uptake or efflux (84).
Localization of the ClCn5 protein at the apical membrane of
thyrocytes and a thyroidal phenotype of the ClCn5-deficient
mice that is reminiscent of Pendred syndrome suggest that ClCn5
could be, possibly in conjunction with other chloride channels,
involved in mediating apical iodide efflux or iodide/chloride exchange (83). This possibility has, as of yet, not been corroborated
by further experimental data.
Endocrinology, March 2009, 150(3):1084 –1090
•
•
•
•
Regulation of pendrin expression
TSH stimulates iodide efflux across the apical membrane of thyrocytes (79, 85, 86). After exposure to TSH, iodide efflux is
rapidly stimulated in FRTL-5 cells (85) and in polarized porcine
thyrocytes (79, 86). In polarized porcine thyrocytes grown in a
bicameral system, measurement of iodide transport in both directions demonstrates that TSH up-regulates iodide efflux at the
apical membrane, whereas efflux across the basolateral membrane does not change (79). Treatment of rat thyroid PCCl3 cells
with TSH for a short time results in the rapid translocation of
pendrin from intracellular compartments, specifically endosomes, to the plasma membrane, thus suggesting a role of pendrin in rapid regulation of apical iodide efflux (87). Insertion of
pendrin in the apical membrane correlates with the phosphorylation of pendrin, but it remains unknown whether phosphorylation is necessary or sufficient for this translocation. These
events occur through the PKA pathway and can be inhibited by
H89, a specific PKA inhibitor (87). Interestingly, translocation of
pendrin from the cytosol to the plasma membrane through a
PKC-dependent pathway has been demonstrated after exposure
of cultured rat thyroid cells to insulin for 10, 20, and 40 min (88).
At least in rat FRTL-5 cells, TSH does not significantly modify
SLC26A4 gene expression (4). Interestingly, thyroglobulin has
been shown to up-regulate SLC26A4 mRNA levels in FRTL-5
cells while suppressing expression of several thyroid-specific
genes, including the TSH receptor, NIS, TPO, and TG genes (4).
Treatment with iodide does not affect expression of the
SLC26A4 gene (89). Exposure to thyroglobulin leads to a decreased NIS gene and protein expression and subsequently results in a reduced iodide uptake in vitro (90). In contrast, accumulation of thyroglobulin in the follicular lumen suppresses
iodide uptake in vivo (90). It has been suggested that the inverse
relationship between the concentration of thyroglobulin in the
follicular lumen and iodide uptake in vivo may be important in
the regulation of thyroid function under constant TSH levels (91)
and promote iodide efflux into the follicular lumen (89).
Conclusions
• NIS mediates the active transport of iodide at the basolateral
membrane of thyrocytes.
• Biallelic mutations in NIS cause a congenital iodide transport
defect, an autosomal recessive condition, characterized by hy-
•
pothyroidism, goiter, low thyroid iodide uptake, and low saliva/plasma iodide ratio.
Pendrin is involved in the apical iodide efflux in thyroid cells.
It can also exchange chloride and bicarbonate.
Pendrin is encoded by the SLC26A4 gene. Biallelic mutations
in the SLC26A4 gene cause Pendred syndrome, an autosomal
recessive disorder, characterized by deafness, goiter, and impaired iodide organification.
In the inner ear, pendrin is important for anion and fluid
transport and for maintenance of the endocochlear potential.
In the kidney, pendrin plays a role in acid-base metabolism as
a chloride/bicarbonate exchanger.
In addition to pendrin, other apical iodide channels or transporters may be involved in regulation of apical iodide efflux
in thyrocytes.
Acknowledgments
Address all correspondence and requests for reprints to: Peter Kopp,
M.D., Associate Professor, Director ad interim Center of Genetic Medicine, Division of Endocrinology, Metabolism, and Molecular Medicine,
Northwestern University, Tarry 15, 303 East Chicago Avenue, Chicago,
Illinois 60611. E-mail: [email protected].
Part of this work has been supported by Grant 1R01DK63024-01
from the National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health (to P.K.).
References
1. Dohan O, De la Vieja A, Paroder V, Riedel C, Artani M, Reed M, Ginter CS,
Carrasco N 2003 The sodium/iodide symporter (NIS): characterization, regulation, and medical significance. Endocr Rev 24:48 –77
2. Kopp P 2005 Thyroid hormone synthesis: thyroid iodine metabolism. In:
Braverman L, Utiger R, eds. Werner and Ingbar’s the thyroid: a fundamental
and clinical text. 9th ed. New York: Lippincott Williams Wilkins; 52–76
3. Arvan P, Di Jeso B 2005 Thyroglobulin structure, function, and biosynthesis.
In: Braverman L, Utiger R, eds. Werner and Ingbar’s the thyroid: a fundamental
and clinical text. New York: Lippincott Williams Wilkins; 77–95
4. Royaux IE, Suzuki K, Mori A, Katoh R, Everett LA, Kohn LD, Green ED 2000
Pendrin, the protein encoded by the Pendred syndrome gene (PDS), is an apical
porter of iodide in the thyroid and is regulated by thyroglobulin in FRTL-5
cells. Endocrinology 141:839 – 845
5. Scott DA, Wang R, Kreman TM, Sheffield VC, Karniski LP 1999 The Pendred
syndrome gene encodes a chloride-iodide transport protein. Nat Genet 21:
440 – 443
6. Dai G, Levy O, Carrasco N 1996 Cloning and characterization of the thyroid
iodide transporter. Nature 379:458 – 460
7. Mazaferri EL 2000 Carcinoma of the follicular epithelium. In: Braverman L,
Utiger R, eds. Thyroid: a fundamental and clinical text. 8th ed. New York:
Lippincott Williams & Wilkins; 904 –930
8. Smanik PA, Liu Q, Furminger TL, Ryu K, Xing S, Mazzaferri EL, Jhiang SM
1996 Cloning of the human sodium iodide symporter. Biochem Biophys Res
Commun 226:339 –345
9. Smanik PA, Ryu KY, Theil KS, Mazzaferri EL, Jhiang SM 1997 Expression,
exon-intron organization, and chromosome mapping of the human sodium
iodide symporter. Endocrinology 138:3555–3558
10. Reizer J, Reizer A, Saier Jr MH 1994 A functional superfamily of sodium/solute
symporters. Biochim Biophys Acta 1197:133–166
11. Levy O, De la Vieja A, Ginter CS, Riedel C, Dai G, Carrasco N 1998 N-linked
glycosylation of the thyroid Na⫹/I⫺ symporter (NIS). Implications for its secondary structure model. J Biol Chem 273:22657–22663
12. De La Vieja A, Dohan O, Levy O, Carrasco N 2000 Molecular analysis of the
sodium/iodide symporter: impact on thyroid and extrathyroid pathophysiology. Physiol Rev 80:1083–1105
Downloaded from endo.endojournals.org at Eidgenossische Technische Hochschule Bibliothek on June 12, 2009
Endocrinology, March 2009, 150(3):1084 –1090
13. Eskandari S, Loo DD, Dai G, Levy O, Wright EM, Carrasco N 1997 Thyroid
Na⫹/I⫺ symporter. Mechanism, stoichiometry, and specificity. J Biol Chem
272:27230 –27238
14. Carrasco N 1993 Iodide transport in the thyroid gland. Biochim Biophys Acta
1154:65– 82
15. Wolff J 1964 Transport of iodide and other anions in the thyroid gland. Physiol
Rev 44:45–90
16. Baschieri L, Benedetti G, Deluca F, Negri M 1963 Evaluation and limitations
of the perchlorate test in the study of thyroid function. J Clin Endocrinol Metab
23:786 –791
17. Yoshida A, Sasaki N, Mori A, Taniguchi S, Mitani Y, Ueta Y, Hattori K, Sato
R, Hisatome I, Mori T, Shigemasa C, Kosugi S 1997 Different electrophysiological character of I⫺, ClO4⫺, and SCN⫺ in the transport by Na⫹/I⫺ symporter. Biochem Biophys Res Commun 231:731–734
18. Yoshida A, Sasaki N, Mori A, Taniguchi S, Ueta Y, Hattori K, Tanaka Y,
Igawa O, Tsuboi M, Sugawa H, Sato R, Hisatome I, Shigemasa C, Grollman
EF, Kosugi S 1998 Differences in the electrophysiological response to I⫺ and
the inhibitory anions SCN⫺ and ClO4⫺, studied in FRTL-5 cells. Biochim
Biophys Acta 1414:231–237
19. Dohan O, Portulano C, Basquin C, Reyna-Neyra A, Amzel LM, Carrasco N
2007 The Na⫹/I⫺ symporter (NIS) mediates electroneutral active transport of
the environmental pollutant perchlorate. Proc Natl Acad Sci USA 104:20250 –
20255
20. Tran N, Valentin-Blasini L, Blount BC, McCuistion CG, Fenton MS, Gin E,
Salem A, Hershman JM 2008 Thyroid-stimulating hormone increases active
transport of perchlorate into thyroid cells. Am J Physiol Endocrinol Metab
294:E802–E806
21. Vassart G, Dumont JE 1992 The thyrotropin receptor and the regulation of
thyrocyte function and growth. Endocr Rev 13:596 – 611
22. Laglia G, Zeiger MA, Leipricht A, Caturegli P, Levine MA, Kohn LD, Saji M
1996 Increased cyclic adenosine 3⬘,5⬘-monophosphate inhibits G protein-coupled activation of phospholipase C in rat FRTL-5 thyroid cells. Endocrinology
137:3170 –3176
23. Weiss SJ, Philp NJ, Ambesi-Impiombato FS, Grollman EF 1984 Thyrotropinstimulated iodide transport mediated by adenosine 3⬘,5⬘-monophosphate and
dependent on protein synthesis. Endocrinology 114:1099 –1107
24. Levy O, Dai G, Riedel C, Ginter CS, Paul EM, Lebowitz AN, Carrasco N 1997
Characterization of the thyroid Na⫹/I⫺ symporter with an anti-COOH terminus antibody. Proc Natl Acad Sci USA 94:5568 –5573
25. Saito T, Endo T, Kawaguchi A, Ikeda M, Nakazato M, Kogai T, Onaya T 1997
Increased expression of the Na⫹/I⫺ symporter in cultured human thyroid cells
exposed to thyrotropin and in Graves’ thyroid tissue. J Clin Endocrinol Metab
82:3331–3336
26. Kogai T, Curcio F, Hyman S, Cornford EM, Brent GA, Hershman JM 2000
Induction of follicle formation in long-term cultured normal human thyroid
cells treated with thyrotropin stimulates iodide uptake but not sodium/iodide
symporter messenger RNA and protein expression. J Endocrinol 167:125–135
27. Riedel C, Levy O, Carrasco N 2001 Post-transcriptional regulation of the
sodium/iodide symporter by thyrotropin. J Biol Chem 276:21458 –21463
28. Vadysirisack DD, Chen ES, Zhang Z, Tsai MD, Chang GD, Jhiang SM 2007
Identification of in vivo phosphorylation sites and their functional significance
in the sodium iodide symporter. J Biol Chem 282:36820 –36828
29. Fanning AS, Anderson JM 1999 PDZ domains: fundamental building blocks
in the organization of protein complexes at the plasma membrane. J Clin Invest
103:767–772
30. Tan PK, Waites C, Liu Y, Krantz DE, Edwards RH 1998 A leucine-based motif
mediates the endocytosis of vesicular monoamine and acetylcholine transporters. J Biol Chem 273:17351–17360
31. Marks MS, Ohno H, Kirchnausen T, Bonracino JS 1997 Protein sorting by
tyrosine-based signals: adapting to the Ys and wherefores. Trends Cell Biol
7:124 –128
32. Wolff J, Chaikoff IL 1948 Plasma inorganic iodide as a homeostatic regulator
of thyroid function. J Biol Chem 174:555–564
33. Eng PH, Cardona GR, Fang SL, Previti M, Alex S, Carrasco N, Chin WW,
Braverman LE 1999 Escape from the acute Wolff-Chaikoff effect is associated
with a decrease in thyroid sodium/iodide symporter messenger ribonucleic acid
and protein. Endocrinology 140:3404 –3410
34. Eng PH, Cardona GR, Previti MC, Chin WW, Braverman LE 2001 Regulation
of the sodium iodide symporter by iodide in FRTL-5 cells. Eur J Endocrinol
144:139 –144
35. Spitzweg C, Joba W, Morris JC, Heufelder AE 1999 Regulation of sodium iodide
symporter gene expression in FRTL-5 rat thyroid cells. Thyroid 9:821– 830
36. Wolff J 1983 Congenital goiter with defective iodide transport. Endocr Rev
4:240 –254
37. Wu SL, Ho TY, Liang JA, Hsiang CY 2008 Histidine residue at position 226
endo.endojournals.org
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
1089
is critical for iodide uptake activity of human sodium/iodide symporter. J
Endocrinol 199:213–219
Levy O, Ginter CS, De la Vieja A, Levy D, Carrasco N 1998 Identification of
a structural requirement for thyroid Na⫹/I⫺ symporter (NIS) function from
analysis of a mutation that causes human congenital hypothyroidism. FEBS
Lett 429:36 – 40
Dohan O, Gavrielides MV, Ginter C, Amzel LM, Carrasco N 2002 Na⫹/I⫺
symporter activity requires a small and uncharged amino acid residue at position 395. Mol Endocrinol 16:1893–1902
De La Vieja A, Ginter CS, Carrasco N 2004 The Q267E mutation in the
sodium/iodide symporter (NIS) causes congenital iodide transport defect (ITD)
by decreasing the NIS turnover number. J Cell Sci 117:677– 687
De la Vieja A, Ginter CS, Carrasco N 2005 Molecular analysis of a congenital
iodide transport defect: G543E impairs maturation and trafficking of the
Na⫹/I⫺ symporter. Mol Endocrinol 19:2847–2858
Soleimani M, Greeley T, Petrovic S, Wang Z, Amlal H, Kopp P, Burnham CE
2001 Pendrin: an apical Cl⫺/OH⫺/HCO3⫺ exchanger in the kidney cortex.
Am J Physiol Renal Physiol 280:F356 –F364
Everett LA, Morsli H, Wu DK, Green ED 1999 Expression pattern of the
mouse ortholog of the Pendred’s syndrome gene (Pds) suggests a key role for
pendrin in the inner ear. Proc Natl Acad Sci USA 96:9727–9732
Royaux IE, Wall SM, Karniski LP, Everett LA, Suzuki K, Knepper MA, Green
ED 2001 Pendrin, encoded by the Pendred syndrome gene, resides in the apical
region of renal intercalated cells and mediates bicarbonate secretion. Proc Natl
Acad Sci USA 98:4221– 4226
Everett LA, Belyantseva IA, Noben-Trauth K, Cantos R, Chen A, Thakkar SI,
Hoogstraten-Miller SL, Kachar B, Wu DK, Green ED 2001 Targeted disruption of mouse Pds provides insight about the inner-ear defects encountered in
Pendred syndrome. Hum Mol Genet 10:153–161
Everett LA, Green ED 1999 A family of mammalian anion transporters and
their involvement in human genetic diseases. Hum Mol Genet 8:1883–1891
Zheng J, Shen W, He DZ, Long KB, Madison LD, Dallos P 2000 Prestin is the
motor protein of cochlear outer hair cells. Nature 405:149 –155
Everett LA, Glaser B, Beck JC, Idol JR, Buchs A, Heyman M, Adawi F, Hazani
E, Nassir E, Baxevanis AD, Sheffield VC, Green ED 1997 Pendred syndrome
is caused by mutations in a putative sulphate transporter gene (PDS). Nat Genet
17:411– 422
Porra V, Bernier-Valentin F, Trouttet-Masson S, Berger-Dutrieux N, Peix JL,
Perrin A, Selmi-Ruby S, Rousset B 2002 Characterization and semiquantitative analyses of pendrin expressed in normal and tumoral human thyroid
tissues. J Clin Endocrinol Metab 87:1700 –1707
Gillam MP, Sidhaye AR, Lee EJ, Rutishauser J, Stephan CW, Kopp P 2004
Functional characterization of pendrin in a polarized cell system. Evidence for
pendrin-mediated apical iodide efflux. J Biol Chem 279:13004 –13010
Aravind L, Koonin EV 2000 The STAS domain: a link between anion transporters and antisigma-factor antagonists. Curr Biol 10:R53–R55
Ko SB, Shcheynikov N, Choi JY, Luo X, Ishibashi K, Thomas PJ, Kim JY, Kim
KH, Lee MG, Naruse S, Muallem S 2002 A molecular mechanism for aberrant
CFTR-dependent HCO3⫺ transport in cystic fibrosis. EMBO J 21:5662–5672
Ko SB, Zeng W, Dorwart MR, Luo X, Kim KH, Millen L, Goto H, Naruse S,
Soyombo A, Thomas PJ, Muallem S 2004 Gating of CFTR by the STAS domain
of SLC26 transporters. Nat Cell Biol 6:343–350
Shcheynikov N, Ko SB, Zeng W, Choi JY, Dorwart MR, Thomas PJ, Muallem
S 2006 Regulatory interaction between CFTR and the SLC26 transporters.
Novartis Found Symp 273:177–186; discussion 186 –192, 261–264
Morgans ME, Trotter WR 1958 Association of congenital deafness with goitre; the nature of the thyroid defect. Lancet 1:607– 609
Kopp P, Pesce L, Solis SJ 2008 Pendred syndrome and iodide transport in the
thyroid. Trends Endocrinol Metab 19:260 –268
Reardon W, Coffey R, Phelps PD, Luxon LM, Stephens D, Kendall-Taylor P,
Britton KE, Grossman A, Trembath R 1997 Pendred syndrome: 100 years of
underascertainment? QJM 90:443– 447
Phelps PD, Coffey RA, Trembath RC, Luxon LM, Grossman AB, Britton KE,
Kendall-Taylor P, Graham JM, Cadge BC, Stephens SG, Pembrey ME, Reardon
W 1998 Radiological malformations of the ear in Pendred syndrome. Clin Radiol
53:268 –273
Fugazzola L, Mannavola D, Cerutti N, Maghnie M, Pagella F, Bianchi P,
Weber G, Persani L, Beck-Peccoz P 2000 Molecular analysis of the Pendred’s
syndrome gene and magnetic resonance imaging studies of the inner ear are
essential for the diagnosis of true Pendred’s syndrome. J Clin Endocrinol
Metab 85:2469 –2475
Pryor SP, Madeo AC, Reynolds JC, Sarlis NJ, Arnos KS, Nance WE, Yang Y,
Zalewski CK, Brewer CC, Butman JA, Griffith AJ 2005 SLC26A4/PDS genotype-phenotype correlation in hearing loss with enlargement of the vestibular
Downloaded from endo.endojournals.org at Eidgenossische Technische Hochschule Bibliothek on June 12, 2009
1090
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
Bizhanova and Kopp
Minireview
aqueduct (EVA): evidence that Pendred syndrome and non-syndromic EVA are
distinct clinical and genetic entities. J Med Genet 42:159 –165
Azaiez H, Yang T, Prasad S, Sorensen JL, Nishimura CJ, Kimberling WJ, Smith
RJ 2007 Genotype-phenotype correlations for SLC26A4-related deafness.
Hum Genet 122:451– 457
Fraser GR, Morgans ME, Trotter WR 1960 The syndrome of sporadic goitre
and congenital deafness. Q J Med 29:279 –295
Fraser GR 1965 Association of congenital deafness with goitre (Pendred’s
syndrome): a study of 207 families. Ann Hum Genet 28:201–249
Nilsson LR, Borgfors N, Gamstorp I, Holst HE, Liden G 1964 Nonendemic
goitre and deafness. Acta Paediatr 53:117–131
Sato E, Nakashima T, Miura Y, Furuhashi A, Nakayama A, Mori N,
Murakami H, Naganawa S, Tadokoro M 2001 Phenotypes associated with
replacement of His by Arg in the Pendred syndrome gene. Eur J Endocrinol
145:697–703
Gausden E, Coyle B, Armour JA, Coffey R, Grossman A, Fraser GR, Winter
RM, Pembrey ME, Kendall-Taylor P, Stephens D, Luxon LM, Phelps PD,
Reardon W, Trembath R 1997 Pendred syndrome: evidence for genetic homogeneity and further refinement of linkage. J Med Genet 34:126 –129
Usami S, Abe S, Weston MD, Shinkawa H, Van Camp G, Kimberling WJ 1999
Non-syndromic hearing loss associated with enlarged vestibular aqueduct is
caused by PDS mutations. Hum Genet 104:188 –192
Kopp P, Arseven OK, Sabacan L, Kotlar T, Dupuis J, Cavaliere H, Santos CL,
Jameson JL, Medeiros-Neto G 1999 Phenocopies for deafness and goiter development in a large inbred Brazilian kindred with Pendred’s syndrome associated with a novel mutation in the PDS gene. J Clin Endocrinol Metab 84:
336 –341
Gonzalez Trevino O, Karamanoglu Arseven O, Ceballos CJ, Vives VI, Ramirez
RC, Gomez VV, Medeiros-Neto G, Kopp P 2001 Clinical and molecular analysis of three Mexican families with Pendred’s syndrome. Eur J Endocrinol
144:585–593
Rotman-Pikielny P, Hirschberg K, Maruvada P, Suzuki K, Royaux IE, Green
ED, Kohn LD, Lippincott-Schwartz J, Yen PM 2002 Retention of pendrin in
the endoplasmic reticulum is a major mechanism for Pendred syndrome. Hum
Mol Genet 11:2625–2633
Schnyder S, Chen L, Chan L, Turk A, Gillam MP, Kopp P 2005 Pendrin
mutations that are retained in intracellular compartments induce the IRE1/
XBP1 and the ATF6 unfolded protein response pathways. Thyroid 15(Suppl
1):S-4 –S-5 (Abstract O11)
Scott DA, Karniski LP 2000 Human pendrin expressed in Xenopus laevis
oocytes mediates chloride/formate exchange. Am J Physiol Cell Physiol 278:
C207–C211
Bidart JM, Mian C, Lazar V, Russo D, Filetti S, Caillou B, Schlumberger M
2000 Expression of pendrin and the Pendred syndrome (PDS) gene in human
thyroid tissues. J Clin Endocrinol Metab 85:2028 –2033
Kopp P 1999 Pendred’s syndrome: clinical characteristics and molecular basis.
Current Opin Endocrinol Diabetes 6:261–269
Yoshida A, Taniguchi S, Hisatome I, Royaux IE, Green ED, Kohn LD, Suzuki
K 2002 Pendrin is an iodide-specific apical porter responsible for iodide efflux
from thyroid cells. J Clin Endocrinol Metab 87:3356 –3361
Endocrinology, March 2009, 150(3):1084 –1090
76. Yoshida A, Hisatome I, Taniguchi S, Sasaki N, Yamamoto Y, Miake J, Fukui
H, Shimizu H, Okamura T, Okura T, Igawa O, Shigemasa C, Green ED, Kohn
LD, Suzuki K 2004 Mechanism of iodide/chloride exchange by pendrin. Endocrinology 145:4301– 4308
77. Wolff J 2005 What is the role of pendrin? Thyroid 15:346 –348
78. Andros G, Wollman SH 1967 Autoradiographic localization of radioiodide in
the thyroid gland of the mouse. Am J Physiol 213:198 –208
79. Nilsson M, Bjorkman U, Ekholm R, Ericson LE 1990 Iodide transport in
primary cultured thyroid follicle cells: evidence of a TSH-regulated channel
mediating iodide efflux selectively across the apical domain of the plasma
membrane. Eur J Cell Biol 52:270 –281
80. Golstein P, Abramow M, Dumont JE, Beauwens R 1992 The iodide channel
of the thyroid: a plasma membrane vesicle study. Am J Physiol 263:C590 –
C597
81. Taylor JP, Metcalfe RA, Watson PF, Weetman AP, Trembath RC 2002 Mutations of the PDS gene, encoding pendrin, are associated with protein mislocalization and loss of iodide efflux: implications for thyroid dysfunction in
Pendred syndrome. J Clin Endocrinol Metab 87:1778 –1784
82. Rodriguez AM, Perron B, Lacroix L, Caillou B, Leblanc G, Schlumberger M,
Bidart JM, Pourcher T 2002 Identification and characterization of a putative
human iodide transporter located at the apical membrane of thyrocytes. J Clin
Endocrinol Metab 87:3500 –3503
83. van den Hove MF, Croizet-Berger K, Jouret F, Guggino SE, Guggino WB,
Devuyst O, Courtoy PJ 2006 The loss of the chloride channel, ClC-5, delays
apical iodide efflux and induces a euthyroid goiter in the mouse thyroid gland.
Endocrinology 147:1287–1296
84. Paroder V, Spencer SR, Paroder M, Arango D, Schwartz Jr S, Mariadason JM,
Augenlicht LH, Eskandari S, Carrasco N 2006 Na⫹/monocarboxylate transport (SMCT) protein expression correlates with survival in colon cancer: molecular characterization of SMCT. Proc Natl Acad Sci USA 103:7270 –7275
85. Weiss SJ, Philp NJ, Grollman EF 1984 Effect of thyrotropin on iodide efflux
in FRTL-5 cells mediated by Ca2⫹. Endocrinology 114:1108 –1113
86. Nilsson M, Bjorkman U, Ekholm R, Ericson LE 1992 Polarized efflux of iodide
in porcine thyrocytes occurs via a cAMP-regulated iodide channel in the apical
plasma membrane. Acta Endocrinol 126:67–74
87. Pesce L, Kopp P 2007 Thyrotropin rapidly regulates pendrin membrane abundance via PKA dependent and PKC dependent pathways in rat thyroid cells.
Thyroid 17(Suppl 1):S-136 (Abstract 282)
88. Muscella A, Marsigliante S, Verri T, Urso L, Dimitri C, Botta G, Paulmichl M,
Beck-Peccoz P, Fugazzola L, Storelli C 2008 PKC-␧-dependent cytosol-tomembrane translocation of pendrin in rat thyroid PC Cl3 cells. J Cell Physiol
217:103–112
89. Suzuki K, Kohn LD 2006 Differential regulation of apical and basal iodide
transporters in the thyroid by thyroglobulin. J Endocrinol 189:247–255
90. Suzuki K, Mori A, Saito J, Moriyama E, Ullianich L, Kohn LD 1999 Follicular
thyroglobulin suppresses iodide uptake by suppressing expression of the sodium/iodide symporter gene. Endocrinology 140:5422–5430
91. Kohn LD, Suzuki K, Nakazato M, Royaux I, Green ED 2001 Effects of thyroglobulin and pendrin on iodide flux through the thyrocyte. Trends Endocrinol Metab 12:10 –16
Downloaded from endo.endojournals.org at Eidgenossische Technische Hochschule Bibliothek on June 12, 2009