Defects in protein folding in congenital hypothyroidism

Defects in protein folding in congenital hypothyroidism

Journal Pre-proof Defects in protein folding in congenital hypothyroidism Héctor M. Targovnik, Karen G. Scheps, Carina M. Rivolta PII: S0303-7207(19)...

2MB Sizes 2 Downloads 39 Views

Journal Pre-proof Defects in protein folding in congenital hypothyroidism Héctor M. Targovnik, Karen G. Scheps, Carina M. Rivolta PII:

S0303-7207(19)30340-5

DOI:

https://doi.org/10.1016/j.mce.2019.110638

Reference:

MCE 110638

To appear in:

Molecular and Cellular Endocrinology

Received Date: 29 July 2019 Revised Date:

21 October 2019

Accepted Date: 1 November 2019

Please cite this article as: Targovnik, Hé.M., Scheps, K.G., Rivolta, C.M., Defects in protein folding in congenital hypothyroidism, Molecular and Cellular Endocrinology (2019), doi: https://doi.org/10.1016/ j.mce.2019.110638. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier B.V.

1

Defects in protein folding in congenital hypothyroidism.

1 2 3

Héctor M. Targovnik1,2 *, Karen G. Scheps1,2, Carina M. Rivolta1,2

4 5 1

6

Universidad de Buenos Aires. Facultad de Farmacia y Bioquímica. Departamento de

7

Microbiología, Inmunología, Biotecnología y Genética/Cátedra de Genética. Buenos Aires,

8

Argentina. 2

9 10

CONICET-Universidad de Buenos Aires. Instituto de Inmunología, Genética y Metabolismo (INIGEM). Buenos Aires, Argentina.

11 12

Short title: Protein folding in congenital hypothyroidism

13

Keywords: Thyroid genes, Protein folding, ERAD, Congenital hypothyroidism.

14

Word count of the full article:

15 16

*Address correspondence and requests for reprints to:

17

Dr. Héctor M. Targovnik

18

CONICET-Universidad de Buenos Aires. Instituto de Inmunología, Genética y Metabolismo

19

(INIGEM). Hospital de Clínicas “José de San Martín”, Av. Córdoba 2351, Cuarto Piso, Sala

20

5, C1120AAR - Buenos Aires, Argentina.

21

Tel.: 54-11-5950-8805 E-mail: [email protected], [email protected]

22

2

1

Abstract

2

Primary congenital hypothyroidism (CH) is the most common endocrine disease in children and

3

one of the most common preventable causes of both cognitive and motor deficits. CH is a

4

heterogeneous group of thyroid disorders in which inadequate production of thyroid hormone

5

occurs due to defects in proteins involved in the gland organogenesis (dysembryogenesis) or in

6

multiple steps of thyroid hormone biosynthesis (dyshormonogenesis). Dysembryogenesis is

7

associated with genes responsible for the development or growth of thyroid cells: such as NKX2-1,

8

FOXE1, PAX8, NKX2-5, TSHR, TBX1, CDCA8, HOXD3 and HOXB3 resulting in agenesis,

9

hypoplasia or ectopia of thyroid gland. Nevertheless, the etiology of the dysembryogenesis remains

10

unknown for most cases. In contrast, the majority of patients with dyshormonogenesis has been

11

linked to mutations in the SLC5A5, SLC26A4, SLC26A7, TPO, DUOX1, DUOX2, DUOXA1,

12

DUOXA2, IYD or TG genes, which usually originate goiter.

13

About 800 genetic mutations have been reported to cause CH in patients so far, including

14

missense, nonsense, in-frame deletion and splice-site variations. Many of these mutations are

15

implicated in specific domains, cysteine residues or glycosylation sites, affecting the maturation of

16

nascent proteins that go through the secretory pathway. Consequently, misfolded proteins are

17

permanently entrapped in the endoplasmic reticulum (ER) and are translocated to the cytosol for

18

proteasomal degradation by the ER- associated degradation (ERAD) machinery.

19

Despite of all these remarkable advances in the field of the CH pathogenesis, several points on

20

the development of this disease remain to be elucidated. The continuous study of thyroid gene

21

mutations with the application of new technologies will be useful for the understanding of the

22

intrinsic mechanisms related to CH. In this review we summarize the present status of knowledge

23

on the disorders in the protein folding caused by thyroid genes mutations.

24 25 26

3

1

1. Introduction

2

The thyroid hormone (TH) is essential for the growth and neural maturation, as well as

3

cardiovascular and renal functions. This hormone also regulates the lipid, carbohydrate and protein

4

metabolism [Mullur et al., 2014]. The thyroid produces predominantly the prohormone 3,5,3´,5´-

5

tetraiodothyronine (T4) and minority the 3,5,3´-triiodothyronine (T3), T3 is obtained by deiodination

6

of the outer phenolic ring of the T4 by means of type 1 iodothyronine deiodinases (D1) and type 2

7

iodothyronine deiodinases (D2) [Mullur et al., 2014]. TH action is mainly mediated by binding of

8

T3 to its nuclear receptor, a member of the superfamily of hormone-responsive nuclear transcription

9

factors [Cheng et al., 2010; Samuels et al., 1974].

10

The biosynthesis of TH takes place at the interface of the apical cell plasma membrane and the

11

colloid in polarized follicular thyroid cells and requires the presence of iodide, thyroid peroxidase

12

(TPO), a supply of hydrogen peroxide, an iodine acceptor protein, thyroglobulin (TG), proteolytic

13

enzymes and the rescue and recycling of iodide by action of iodotyrosine deiodinase-1 (IYD-1)

14

(Figure 1) [Degroot and Niepomniszcze, 1977; Gnidehou et al., 2004; Targovnik et al. 2010, 2011].

15

After the iodide is uptake in the basolateral membrane by the Na+/I- symporter (NIS) [Bizhanova

16

and Kopp, 2009; Dohan et al., 2003], it is transported at the apical membrane into the follicular

17

lumen by an anion exchange transmembrane protein called pendrin (Figure 1) [Bizhanova and

18

Kopp, 2009; Royaux et al., 2000; Yoshida et al., 2002]. Na+/I- symporter (NIS) cotransports one

19

iodide ion with two sodium ions against an electrochemical gradient. This sodium gradient is

20

generated by Na+/K+-ATPase. Roepke et al. [2009] and Purtell et al. [2012] demonstrated that

21

KCNQ1-KCNE2 potassium channel is required for NIS-mediated thyroid cell uptake of I- (Figure

22

1). It was later described that the Ca2+-activated chloride channel TMEM16A or anoctamin-1

23

(encoded by ANO1 gene) [Siveira and Kopp, 2015; Twyffels et al. 2014] and SLC26A7 also

24

mediate the iodide efflux through the apical membrane of thyroid cells [Ishii et al., 2019]. The

25

oxidized iodide is incorporated into selected tyrosyl residues of TG which results in formation of

26

monoiodotyrosines (MID) and diiodotyrosines (DIT). Coupling between either two DIT, or between

4

1

a DIT and a MIT, generates the formation of T4 or T3 respectively (Figure 1). The oxidition of I−

2

into I+, iodination and coupling reactions are mediated by TPO in presence of H2O2 [Degroot and

3

Niepomniszcze, 1977]. Hydrogen peroxide is generated by a metabolic pathway, involving two

4

members of the NADPH oxidase family: dual oxidase 1 (DUOX1) and dual oxidase 2 (DUOX2)

5

and two ER-resident proteins: DUOX maturation factor 1 (DUOXA1) and DUOX maturation factor

6

2 (DUOXA2) [De Deken et al., 2000, 2002; De Deken and Miot, 2019; Dupuy et al, 1999;

7

Grasberger and Refetoff, 2006]. TPO and the DUOX system are in close proximity, forming a

8

complex at the level of the apical plasma membrane of the follicular cell (Figure 1) [Song et al.,

9

2010].

10

The thyroid cells produce free T4 and T3 by endocytic internalization and proteolytic lysosomal

11

cleavage (cathepsin proteases and glutamate carboxypeptidase) of the TG, which are delivered to

12

the blood circulation for action at their peripheral target tissues [Jordans et al., 2009; Linke et al.,

13

2002; Suban et al., 2012]. The monocarboxylate transporter 8 (MCT8, encoded by SLC16A2 gene)

14

facilitates the efflux of TH from the thyrocytes (Figure 1) [Di Cosmo et al., 2010]. From TG, free

15

MIT and DIT are also released by the thyroid cells. Therefore, the rescue and recycling into thyroid

16

cell of iodide from MIT to DIT by the action of IYD-1 are essential to prevent loss of this rare

17

element, especially in iodine-deficient areas.

18

The thyroid stimulating hormone (TSH) and its receptor (TSHR), a member of the G- protein-

19

coupled receptor superfamily (GPCRs), constitute the main regulatory pathway of the thyroid

20

(Figure 1) [Vassart and Dumont, 1992]. The cAMP and Gq/phospholipase C cascade pathways

21

mediate most of its effects, including iodide uptake, expression of thyroid genes, biosynthesis of

22

TH, TPO activity, thyroid H2O2 generating system, endocytosis, proteolysis and hormone release.

23

[Vassart and Dumont, 1992].

24

In order to fulfill their functions, the specialized thyroid proteins described above must be

25

exported to the colloid or located at the basolateral and apical membranes through a series of way-

5

1

stations. Exported proteins enter at the endoplasmic reticulum (ER), the first station of this pathway.

2

The proper folding and assembly of new synthesized exportable proteins occur within the ER

3

compartment. Then, secretory proteins are translocated into the Golgi complex by carrier vesicles

4

and finally, they are exported to their final destination. Exportable protein folding is enhanced by

5

the actions of compartment-specific molecular chaperones and enzymes. Remarkably, unfolding or

6

misfolding of exportable proteins cause retention within the ER, induction of the synthesis of ER

7

chaperones and removal of misfolded proteins from the ER lumen by an anterograde traffic out of

8

the ER to be catabolized in the proteasomes, a phenomenon known as ER-associated degradation

9

(ERAD) mechanism. ERAD prevents the accumulation or secretion of the abnormal protein [Oi et

10

al., 2017].

11

Here, we review the latest contributions in the field of protein folding disorders in congenital

12

hypothyroidism caused by mutations in the main proteins that participate in the biosynthesis of TH

13

and in thyroid embryogenesis. We also provide a large references list of original publications.

14 15

2. Congenital hypothyroidism

16

Congenital hypothyroidism (CH) is the most frequent endocrine disease in infants, affects about

17

1 in 2,000-3000 newborns and is characterized by elevated levels of thyroid-stimulating hormone

18

(TSH) and low TH serum levels [Kwak, 2018; Park and Chatterjee, 2005; Rastogi and LaFranchi,

19

2010]. It is also one of the most common preventable causes of cognitive and motor deficits [Kwak,

20

2018, Park and Chatterjee, 2005; Rastogi and LaFranchi, 2010]. The genetic classification divides

21

CH into two main categories according to their causes: (a) disorders of thyroid gland development

22

(dysembriogenesis or thyroid dysgenesis group) or (b) defects in any of the steps of thyroid

23

hormone synthesis (dyshormonogenesis group) [Park and Chatterjee, 2005; Rastogi and LaFranchi,

24

2010; Targovnik et al. 2010, 2011, 2017]. The dysembryogenesis or thyroid dysgenesis group,

25

which accounts for the 80-85 % of the cases, results from a thyroid gland that is completely absent

6

1

in orthotopic or ectopic location (agenesis or athyreosis), severely reduced in size but in the proper

2

position in the neck (orthotopic hypoplasia) or located in an unusual position (thyroid ectopy) at the

3

base of the tongue or along the thyroglossal tract [Abu-Khudir et al., 2017; Targovnik et al. 2010,

4

2011, 2017]. In few patients, the CH has been associated with mutations in genes responsible for

5

the development or growth of thyroid cells: NKX2-1 (also known as TTF1), FOXE1 (also known as

6

TTF2 or FKHL15), PAX8 and NKX2-5 genes [Grasberger and Refetoff, 2011; Kwak, 2018;

7

Targovnik et al. 2010, 2011, 2017]. Recently, exome analysis identified DUOX2 mutations in

8

patients with ectopic thyroid suggesting that DUOX2 could also be involved in thyroid

9

embryogenesis [Kizys et al., 2017]. More recently, Zou et al. [2018] identified mutations in the

10

TBX1, CDCA8, HOXD3 and HOXB3 genes associated with dysgenesis. Loss-of-function mutations

11

in the TSHR gene have been shown to cause thyroid hypoplasia and TSH resistance in humans

12

[Grasberger and Refetoff, 2017]. In the majority of patients with a defect in thyroid organogenesis,

13

the genetic mechanism responsible remains to be elucidated.

14

Dyshormonogenesis, which accounts for the remaining 15-20% of the cases, has been linked to

15

mutations in the SLC5A5 [Santos-Silva et al., 2019; Spitzweg and Morris, 2010; Targovnik et al.,

16

2017], SLC26A4 [Bizhanova and Kopp, 2010; Chem et al., 2018; Dossena et al., 2009, 2011;

17

Makretskaya et al., 2018; Sun et al., 2018; Wémeau and Kopp, 2017, Zou et al., 2018], SLC26A7

18

[Cangul, et al., 2018; Ishii et al., 2019; Zou et al., 2018], TPO [Long et al., 2018; Makretskaya et

19

al., 2018; Ris-Stalpers and Bikker, 2010; Santos-Silva et al., 2019; Sun et al., 2018; Targovnik et

20

al., 2017; Zou et al., 2018], DUOX1 [Aycan, et al., 2017; Liu et al., 2019; Watanabe et al., 2019;

21

Zou et al., 2018], DUOX2 [Belforte et al., 2016a, 2016b; Chem et al., 2018; Grasberger, 2010; Long

22

et al., 2018; Makretskaya et al., 2018; Muzza and Fugazzola, 2017; Sun et al., 2018; Yu et al., 2018;

23

Zou et al., 2018], DUOXA1 [Liu et al., 2019], DUOXA2 [Chem et al., 2018; Muzza and Fugazzola,

24

2017 Sun et al., 2018], IYD [Moreno and Visser, 2010; Sun et al., 2018; Targovnik et al., 2017] and

25

TG [Chem et al., 2018; Citterio et al., 2019; Di Jeso and Arvan, 2016; Long et al., 2018;

7

1

Makretskaya et al., 2018; Santos-Silva et al., 2019; Sun et al., 2018; Targovnik et al., 2010, 2011,

2

2016, 2017; Yu et al., 2018; Zou et al., 2018] genes. In the last three decades about eight hundred

3

variants have been described in unrelated CH patients. These mutations produce a heterogeneous

4

spectrum of congenital hypothyroidism, characterized usually by the presence of congenital goiter

5

or goiter appearing shortly after birth with an autosomal recessive inheritance. Therefore, the

6

patients are typically homozygous or compound heterozygous for the gene mutations and the

7

parents, carriers of one mutation.

8

Iodide organification defects (IOD) are linked with inactivating mutations in the TPO, DUOX2,

9

DUOXA2 or SLC26A4 genes and probably associated with mutations in the DUOX1 and DUOXA1

10

genes. IOD is characterized by high levels of serum TG and a positive Perchlorate Discharge Test

11

(PDT) [El-Desouki et al., 1995], indicating that the iodide is taken up by thyroid cells but it is not

12

incorporated into the TG when the oxidation of iodine, iodination and coupling mechanisms fail.

13

PDT is used to distinguish Total Iodide Organification Defect (TIOD) from Partial Iodide

14

Organification Defect (PIOD, between 10 and 90%) [Kwak. 2018]. In patients with PIOD usually a

15

single allele is affected, whereas homozygous or compound-heterozygous mutations are associated

16

with TIOD. Most cases of CH associated with alterations in DUOX2 are caused by either biallelic or

17

monoallelic mutations which lead to extremely complex correlation between DUOX2 genotypes

18

and clinical phenotypes [De Marco et al., 2011; Moreno and Visser, 2007; Wang et al., 2014]. Both

19

biallelic and monoallelic DUOX2 mutations could be associated with transient CH (TCH) or

20

permanent CH (PCH) [Fu et al., 2016a; Tan et al., 2016]. Mutations in SLC26A4 gene cause

21

Pendred syndrome characterized by congenital sensorineural hearing loss, goiter with or without

22

hypothyroidism and usually PIOD [Bizhanova and Kopp, 2010; Wémeau and Kopp, 2017]. In

23

patients with IYD-1 deficiency, the organification process is not affected whereas the serum TG

24

levels are elevated [Moreno and Visser, 2010; Targovnik et al., 2017]. In contrast, the presence of

25

low or undetectable levels of TG and also negative PDT in a goitrous individual suggest a TG

8

1

defect [Citterio et al., 2019; Di Jeso and Arvan, 2016; Targovnik et al., 2010, 2011, 2017]. Patients

2

with an iodide transport defect by mutations in SLC5A5 gene have a normal-sized or somewhat

3

enlarged thyroid gland, elevated plasma TG levels and no radio-iodide uptake [Spitzweg and

4

Morris, 2010; Targovnik et al., 2017].

5

Recent technological advances in DNA sequencing through the implementation of next-

6

generation sequencing (NGS) platforms that allow the massive identification of sequence variants

7

have led to the identification of new mutations in the thyroid genes. Interestingly, Fu et al. [2016a,

8

2016b, 2016c] using NGS detected that most of the cases of CH with one or two DUOX2 gene

9

mutations are associated with subclinical or TCH, whereas patients with three or four (one in

10

homozygous) or five (two in homozygous) DUOX2 gene mutations are mostly associated with

11

PCH. The new technologies also showed the coexistence of multiple mutations in different thyroid

12

genes in the same patient with CH, for instances, mutations in DUOX2 associated with mutations in

13

DUOXA2 or TPO or TG or TSHR or SLC26A4 genes [Fu et al., 2016a, 2016b; Jiang 2016; Park et

14

al., 2016; Sun et al., 2018]. The identification of simultaneous mutations in the same gene or in

15

different thyroid specific genes could contribute to the accurate diagnosis and classification of the

16

defects.

17 18

3. Defects in protein folding of thyroglobulin

19

TG represents a highly specialized homodimeric glycoprotein which acts as a substrate for the

20

synthesis of T4 and T3 as well as the storage of the inactive forms of iodine. [Citterio et al., 2019;

21

Targovnik 2013; Targovnik et al., 2017]. Human TG gene is a single copy of 268 Kb long that maps

22

on chromosome 8 (8:132,866,958-133,134,903; GRCh38 assembly) and contains an 8,455 nt of

23

mRNA sequence (with a coding sequences of 8304, NCBI Accesion Number: NM_003235.5)

24

divided into 48 exons [Malthiéry and Lissitzky, 1987; Mendive et al., 2001; Mercken et al., 1985;

25

van de Graaf et al., 2001].

9

1

The monomeric human TG preprotein has a leader peptide of 19 amino acids followed by a

2

2749-amino-acid polypeptide (Figure 2). The most important T4-forming site couples donor DIT149

3

to acceptor DIT24 [Dunn et al., 1998; Lamas et al., 1989; Palumbo et al., 1990]; whereas the main

4

T3-forming site couples an MIT2766 at the antepenultimate residue of one TG monomer with the

5

antepenultimate DIT2766 in the opposite monomer of the TG dimer [Citterio et al., 2017].

6

The internal protein organization makes TG an example of gene evolution by intragenic

7

duplication events and gene fusions [Parma et al., 1987]. The N-terminal and the central parts of the

8

monomer include three types of repetitive motifs, called TG type 1, TG type 2, and TG type 3,

9

comprising cysteine-rich repeats covalently bound by disulfide bonds (Figure 2). The TG type 1

10

motif derives from an ancestral gene that was duplicated and modified during evolution by

11

mutations [Holzer et al., 2016; Malthiery and Lissitzky, 1987; Mercken et al., 1985; Molina et al.,

12

1996, Parma et al., 1987, van de Graaf et al., 2001] whereas TG type 2 repeats are distant members

13

of the GCC2/GCC3 domain superfamily [Lee et al, 2011] and TG type 3 repeats exhibit only

14

internal homology. Tg monomer contains eleven TG type 1 elements located between positions 31 -

15

358, 605 - 1210 and between 1511 - 1565; three TG type 2 elements located between amino acids

16

1456 and 1503, and five TG type 3 elements between residues 1603 and 2187 (Figure 2) [Holzer et

17

al., 2016; Malthiery and Lissitzky, 1987; Mercken et al., 1985; Molina et al., 1996, Parma et al.,

18

1987, van de Graaf et al., 2001]. TG is organized in four structural regions (I, II, III and IV) (Figure

19

2). Region I comprises 10 of the 11 TG type 1 repeats, two segments called linker (residues 359 to

20

604) and hinge (residues 1211 to 1455) plus the N-terminal T4 forming site. Region II contains 3

21

TG type 2 repeats and the eleventh TG type 1 repeat, whereas region III contains the five TG type 3

22

repeats. The fourth region is composed of the cholinesterase-like (ChEL) domain, a member of the

23

α/β-hydrolase fold family, located between residues 2211 to 2735 of the TG [Park and Arvan, 2004;

24

Swillens et al., 1986] and the C-terminal T3 forming site. The transitions between repetitive motifs

25

TG type 2-3 and TG type 1-11, between TG type 1-11 and TG type 3-a1, and between TG type 3-a3

10

1

and the ChEL domain correspond to non-repetitive sequences that could be called spacer 1 (residues

2

1504 to 1510), spacer 2 (1566 to 1602) and spacer 3 (2188 to 2210) segments, respectively (Figure

3

2).

4

During translation/translocation in the lumen of the ER, the newly synthesised TG starts to fold

5

and acquires its final three-dimensional glycoprotein structure as it passes through the Golgi

6

complex. TG is transferred to the follicle lumen as a noncovalent dimer, concentrated into secretory

7

vesicles and stored as colloid. Intensive post translational modifications take place in the ER, which

8

include glycosylation and formation of intrachain disulfide bonds.

9

All glycoproteins go through similar trimming in the ER and acquire their final glycoprotein

10

structure as they pass through the Golgi [Bedard et al, 2004]. The addition of carbohydrates

11

stabilizes the protein and increases its solubility. TG acquires the N-glycans during the process of

12

translocation and elongation of the polypeptide chain into the ER. First, two N-acetylglucosamines

13

(GlcNAc) and nine mannoses with three terminal glucose residues are assembled onto a core

14

oligosaccharide, which is then en bloc transferred to the consensus N-glycosylation sites

15

(Asparagine-X-Serine/Threonine, where X is any amino acid residue except proline) on nascent TG

16

chains by action of an oligosaccharyltransferase (Figure 3). Next, the three terminal glucoses of this

17

core are trimmed by sequential action of I and II α-glucosidases. Finally, a terminal mannose is

18

trimmed by α-mannosidase I with the formation of high-mannose type oligosaccharide [Helenius

19

and Aebi, 2004] (Figure 3).

20

Correct folding is assisted by interaction with chaperones and enzymes of the ER [Bedard et al.,

21

2005], such as calnexin (CNX), calreticulin (CRT,), GRP94, BiP, Protein Disulfide Isomerase

22

(PDI), ERp29, ERp57, ERp72 [Baryshev et al, 2004; Di Jeso et al., 2003, 2005; Kim et al., 1992;

23

Kim and Arvan, 1995; Kuznetsov et al., 1994; Menon et al., 2007; Muresan and Arvan, 1997;

24

Sargsyan et al., 2002]. Such interactions may also serve to prevent premature export of unfolded or

25

misfolded secretory proteins from the ER by a process known as ER-quality control [Bedard et al.,

11

1

2005; Qi el al., 2017], an adaptive cellular reaction that maintains ER homeostasis avoiding

2

perturbation by the accumulation of incorrectly folded proteins. GRP94 is associated with the

3

nascent TG [Muresan and Arvan, 1997], whereas BiP binds to unfolded polypeptides and prevents

4

protein aggregation through noncovalent associations [Chung et al, 2002]. The oxidative folding of

5

TG proceeds with the aid of two chaperone-oxidoreductase complexes, Bip/PDI and

6

CNX/CRT/ERp57 [Meunier et al., 2002]. The formation of intradomain disulfide bonds can be

7

catalyzed near BiP bound sites by PDI and near CNX/CRT bound sites by the ERp57

8

oxidoreductase [Di Jeso et al., 2005].

9

The last stage of TG conformational maturation in the ER is the non-covalent TG

10

homodimerization, without any interchain disulfide bridge, and only after TG has dissociated from

11

CNX/CRT system. ChEL domain is required for protein dimerization. Whereas correctly folded

12

proteins are exported from the ER to Golgi, misfolded proteins are retained in the ER and

13

retrotranslocated to the cytosol for proteasomal degradation by the ERAD machinery [Kim an

14

Arvan, 1998; Qi et al., 2017].

15

The next step in the post translational modifications of TG N-glycans occurs within the Golgi

16

and follows presumably the same general process described for other proteins. In brief, after

17

removal

18

glycosyltransferases (GnT) catalyze the branching and elongation of the carbohydrate chains,

19

producing hybrid-type or complex-type oligosaccharide [Vagin et al., 2009] (Figure 3).

of

mannose

residues

by

Golgi

mannosidases,

several

N-acetylglucosamine

20

The cDNA sequence of human TG contains 20 putative N-linked glycosylation sites, 16 of them

21

being glycosylated in the mature protein [Yang et al., 1996]. Eight sites (asparagines residues 76,

22

484, 529, 748, 816, 1716, 1774 and 2250) are linked to complex type oligosaccharide units

23

containing fucose and galactose in addition to mannose and glucosamine [Yang et al., 1996]. The

24

sites in positions 1220, 1349, 2013, 2295 and 2582 contain high-mannose type units (mannose and

25

glucosamine) and the sites in positions 198 and 1365 are linked to oligosaccharide units containing

12

1

galactose in addition to mannose and glucosamine but no fucose and may be either hybrid-type or

2

complex-type. [Yang et al., 1996]. Finally, very different oligosaccharide composition types were

3

found associated with position 947 (complex or high mannose) [Yang et al., 1996]. The human TG

4

also contains sulfated O-linked glycosaminoglycans [Fogelfeld and Schneider, 1990; Schneider et

5

al., 1988; Spiro and Bhoyroo, 1988]. Interestingly, non-glycosylated thyroglobulin loses the ability

6

to synthesize thyroid hormones [Mallet et al., 1995].

7

Phosphorylation of TG occurs in Golgi apparatus within carbohydrates and serine and tyrosine

8

residues [Consiglio et al., 1987], whereas sulfation is a late post translational modification step that

9

takes place in the trans-Golgi and contributes to the negative charge of the TG [Cauvi et al., 2003].

10

In mammals, newly synthesized TG is firstly glycosylated and folded via the formation of

11

numerous disulfide bonds [Citterio et al., 2019; Targovnik, 2013]. As indicated above, TG type 1,

12

TG type 2 and TG type 3 are cysteine-rich repeats that are covalently bound by intramolecular

13

disulfide bonds, which are involved in the conformation of the tertiary structure of TG and in the

14

hormonogenesis. The formation of disulfide bonds is catalyzed by endogenous ER oxidoreductases.

15

The TG type 1 domains are grouped into two subgroups, type 1A and type 1B. Type 1A repeats

16

(TG type 1–1 to TG type 1–8 and TG type 1–10) have a total of six Cys residues, whereas type 1B

17

repeats (TG type 1–9 and TG 1–11) contain only four Cys residues [Molina et al., 1996].

18

Conservative intradomain disulfide bonds patterns Cys1–Cys2, Cys3–Cys4, and Cys5–Cys6 prevail

19

in all type 1A repeats [Molina et al., 1996]. TG type 1B, TG type 2 and TG type 3 repeats and the

20

six cysteine residues of the ChEL domain include also intradomain disulfide bonding.

21

TG containing missense mutations in the cysteine residues or in the ChEL domain and splice site

22

mutations causing exon skipping could represent targets for intracellular retention and degradation

23

by ERAD pathway.

24

Sequencing analysis of the TG gene revealed missense mutations that involved cysteine residues:

25

p.C160S, p.C183Y, p.C194G, p.C726Y, p.C1077R, p.C1264R, p.C1281Y, p.C1476R, p.C1493Y,

13

1

p.C1510F, p.C1607F, p.C1897Y, p.C1904G, p.C1996S, p.C2000W, p.C2006Y and p.C2154Y

2

(Figure 2, Table 1) [Baryshev et al, 2004; Caputo et al., 2007; Citterio et al., 2013; Hishinuma et al,

3

1999, 2005, 2006; Hu et al., 2016; Kanou et al., 2007; Kitanaka et al., 2006; Narumi et al., 2011;

4

Nicholas et al., 2016; Zou et al. 2018]. The loss of cysteine residues can eliminate disulfide bonds

5

and alter the normal conformational structure of the TG, possibly preventing the interaction of

6

hormonogenic acceptor and donor sites, and causing a defect in the intracellular transport of the

7

protein. A detailed study of two unrelated patients with CH and two siblings with adenomatous

8

goiter revealed replacement of two conserved cysteines by arginine (C1264R) and serine (C1996S),

9

which resulted in the retention and aggregation of TG in the ER [Hishinuma et al. 1999]. Thyroid

10

tissue extracted from all patients exhibited higher expression levels of ER chaperones, BiP, GRP94,

11

PDI, ERp29 and ERp72 [Baryshev et al, 2004].

12

The ChEL domain is essential for the intracellular transport of TG to the site of its

13

hormonogenesis, via the secretory pathway [Lee et al., 2008, 2009, 2011; Lee and Arvan, 2011;

14

Park and Arvan, 2004; Wang et al., 2010]. It functions as an intramolecular chaperone and as a

15

molecular escort for TG regions I, II, and III [Lee et al., 2008]. The ChEL domain is also required

16

for protein dimerization and consequently plays a critical structural and functional role in the TG

17

protein. It is well documented that pathogenic missense mutations in the ChEL domain can result in

18

the intracellular retention of TG. The missense mutations located in this domain may cause TG

19

retention in the ER and premature degradation. Several missense mutations associated with CH

20

were reported to be present in the ChEL domain: p.A2234D [Caputo et al., 2007; Citterio et al.,

21

2013; Machiavelli et al., 2010; Pardo et al. 2008; 2009, Santos-Silva et al., 2019], p.R2242H [Caron

22

et al., 2003; Machiavelli et al., 2010; Raef et al., 2010, Zou et al., 2018], p.N2285D [Makretskaya et

23

al., 2018], p.G2319D [Hishinuma et al., 2006], p.R2336Q [Hishinuma et al., 2006; Kitanaka et al.,

24

2006 ; Siffo et al., 2018], p.G2341S [Watanabe et al. 2018], p.A2362P [Siffo et al., 2018],

25

p.W2365R [Siffo et al., 2018], p.G2374V [Hishinuma et al., 2006], p.G2375R [Hishinuma et al.,

14

1

2005; 2006; Kanou et al.; 2007; Long et al., 2018], p.I2394M [Sun et al., 2018], p.R2455H [Hu et

2

al., 2016; Long et al., 2018, Yu et al., 2018], p.A2471S [Sun et al., 2018], p.L2547Q [Nicholas et

3

al., 2016], p.R2585W [Hu et al., 2016, Jiang et al., 2016 ; Sun et al., 2018] p.N2616I [Hu et al.,

4

2016; Long et al., 2018, Sun et al. 2018; Yu et al., 2018], p.W2685L [Nicholas et al., 2016], and

5

p.A2713T [Long et al., 2018] (Figure 2, Table 2). Functional analysis suggests that the p.A2215D

6

mutation results in retention of the TG protein inside the ER and degradation via the proteasome

7

system [Pardo et al., 2009], as already observed in the cog/cog congenital goiter mouse [Kim et al,

8

1998] and the rdw/rdw non-goitrous CH rat [Kim et al, 2000]. The cog/cog mouse harboring the

9

p.L2263P TG mutation [Kim et al., 1998] and the WIC-rdw rat harboring the p.G2300R TG

10

mutation [Hishinuma et al., 2000; Kim et al., 2000] exhibit full-length synthesis of TG but impaired

11

transport from the ER.

12

Exon skipping in the TG gene can be caused by nucleotide substitutions or deletion in acceptor

13

or donor splice sites involving the -3/-2/-1 (c.275-3C>G, c.5042-2A>G, c.6563-2A>G, c.2762-

14

1G>A, c.6200-1G>C, c.7998-1G>A) or +1/+2/+3/+4/+5/+6 position (c.638+1G>A, c.745+1G>A,

15

c.2176+1G>A, c.4932+1G>C, c.5686+1G>T, c.5686+1G>A, c.5686+1G>C, c.6262+1delG,

16

c.6876+1delG, c.274+2T>G, c.5401+2T>C, c.7036+2T>A, c.7862+2T>A, c.4159+3_+4delAT;

17

c.3433+3_+6delGAGT, c.638+5G>A), respectively [Abdul-Hassan et al., 2013; Alzahrani et al.,

18

2006; Chen et al., 2018; Citterio et al., 2015, Fu et al., 2016c, Gutnisky et al., 2004; Hermanns et

19

al., 2013, Hishinuma et al., 2006; Hu et al., 2016, Ieiri et al. 1991; Makretskaya et al., 2018;

20

Medeiros-Neto et al., 1996; Narumi et al., 2011; Nicholas et al., 2016; Niu et al., 2009; Pardo et al.,

21

2008, 2009; Peteiro-Gonzalez et al., 2010; Rubio et al., 2008; Targovnik et al., 1995, 2001, 2012;

22

Watanabe et al., 2019; Zou et al., 2018]. Recently, two exonic cryptic 5' splicing sites in exons 6

23

(c.745+1G>A) [Citterio et al., 2015] and 19 (c.4159+3_4delAT) [Targovnik et al., 2012] of the TG

24

gene have been identified. The elimination of exons in the TG gene by aberrant splicing results in

25

an altered ability to transfer an iodophenoxyl group from the donor site to the acceptor iodotyrosine.

15

1

Interestingly, two affected patients with the c.5686+1G>T mutation which leads to the skipping of

2

exon 30 [Targovnik et al., 2001], showed a marked increase of the ER chaperones, GRP94 and BiP

3

[Medeiros-Neto et al., 1996]. Immunohistochemical localization of TG showed a notable decrease

4

of immunoreaction product in the follicular lumen of thyroid sections from affected patients. In

5

contrast, accumulation of intracellular immunoreactive TG was clearly demonstrable in the thyroid

6

epithelium, suggesting a defect in intracellular TG transport [Medeiros-Neto et al., 1996].

7 8

4. Defects in protein folding of thyroid peroxidase

9

TPO is a thyroid-specific enzyme that plays a key role in thyroid hormones biosynthesis and is

10

the major autoantigen in Hashimoto’s disease, the most common organ-specific autoimmune

11

disease (Czarnocka et al., 1985; Portmann et al., 1985). TPO is a member of the peroxidase family

12

which employs a heme group as part of their catalytic unit. The human TPO gene is located on

13

chromosome 2 (2:1,374,066-1,543,711; GRCh38 assembly). It comprises 17 exons, spands 170 kb

14

of genomic DNA and encodes a glycosylated hemoprotein of 933 amino acids (Figure 4). The

15

mRNA is 3,152 nucleotides long (NCBI Reference Sequence NM_000547.5) [Kimura et al., 1989]

16

and the pre-protein is composed of a putative 18 amino acids signal peptide followed by a 915

17

amino acids polypeptide which codifies a large extracellular domain (residues 19-846), a

18

transmembrane domain (residues 847-871), and a short intracellular tail (residues 872-933) (Figure

19

4) [https://www.uniprot.org/uniprot/P07202]. Baker et al. (1994) and Nishikawa et al. (1994)

20

suggested that human TPO may be a disulfide-linked dimer. TPO shows 42% sequence homology

21

with the myeloperoxidase (MPO) (Libert et al., 1987) and is part of the animal haem peroxidase

22

superfamily. The exons 13 and 14 encode the complement control protein (CCP)-like (residues 740-

23

795) and calcium-binding epidermal growth factor (EGF-Ca2+ binding)-like (residues 796-839)

24

domains, respectively (Figure 4) [https://www.uniprot.org/uniprot/P07202, Ruff and Carayon,

25

2006]. EGF-Ca2+ shows structural characteristics of a Sushi domain.

16

1

The TPO enzyme activity depends on both proper folding and membrane insertion, and an intact

2

catalytic site. After its synthesis, human TPO undergoes various postranslational modifications such

3

as N- and O-linked glycosylation and heme fixation [Ruff and Carayon, 2006]. Five potential

4

glycosylation sites are present at asparagine residues 129, 307, 342, 478, and 569 [Ruff and

5

Carayon, 2006]. N-glycans play an essential role in the correct folding, intracellular trafficking and

6

activity of TPO [Fayadat et al., 1998]. Exons 8, 9 and 10 encode the catalytic center of the TPO

7

protein (heme-binding region) [Rodrigues et al., 2005] which is crucial for the enzymatic activity

8

and contain the proximal (His239) and distal (His494) putative heme binding histidine residues

9

(Figure 4). According to the studies with MPO and lactoperoxidase, the heme group in TPO is a

10

bis-hydroxylated heme b covalently bound via ester linkages to Glu399 and Asp238 of the apoprotein

11

(Figure 4) [Ruff and Carayon, 2006].

12

The first description of a human TPO mutation was conducted in a boy with iodide

13

organification defect, who presented with hypothyroidism at the age of 4 months and developed a

14

compressive goiter at the age of 12 years [Abramowicz et al, 1992]. A homozygous duplication of a

15

tetranucleotide -GGCC- in exon 8 of the TPO gene was identified (c.1186_1187insGGCC or

16

c.1183_1186dupGGCC, also reported as 1187_1188insGCCG or c.1184_1187dup GCCG), 152 bp

17

upstream from the junction with intron 8. The resulting frameshift generates a stop codon in exon 9

18

(p.A397Pfs*76, Figure 4), which would result in a truncated protein with mutation of the proximal

19

and deletion of the distal putative heme binding histidine residues where by TPO activity is

20

expected to be absent. However, alternative splicing restores the normal reading frame

21

[Abramowicz et al, 1992]. The result is a nearly full-length protein with a 91-amino acid segment

22

replaced by a 51 residue unrelated segment. This translation product is expected to have an

23

unchanged distal putative heme binding His494, but to lack the proximal putative heme binding

24

His239 [Abramowicz et al, 1992]. The duplication GGCC in exon 8 is a common mutation of the

25

TPO gene in Caucasian population [Abramowicz et al., 1992; Altmann et al, 2013; Bakker et al.,

17

1

2000; Belforte et al., 2012; Bikker et al, 1995; Cangul et al., 2013; Cangul et al., 2015; Fugazzola et

2

al., 2005; Gruters et al.,1996; Nascimento et al., 2003, Neves et al., 2010; Rivolta et al 2003; Santos

3

et al., 1999].

4

Six cystein residues responsible for the formation of 3 disulfide bridges: Cys800-Cys814, Cys808-

5

Cys823 and Cys825-Cys838 are located in the EGF-Ca2+ binding domain of human TPO. The

6

substitution of cysteine for other amino acids affects the formation of the disulfide bond and

7

therefore can alter the tertiary structure of the EGF-Ca2+ binding domain of TPO [Rodrigues, et al.,

8

2005]. Disruption of the disulfide bonds, as result of p.C800R [Bӧrgel et al., 2005], p.C808R

9

[Rivolta et al., 2003] or p.C838S [Rodrigues, et al., 2005] mutations (Figure 4), would generate

10

significant conformational changes which could partially retain the TPO on the ER on the one hand,

11

and on the other hand, the proteins with conformational defects responsible for a lower net activity

12

of the enzyme might reach the membrane.

13

Interestingly, the p.S131P mutation (Figure 4) disrupts the potential glycosylation site at Asn129

14

residue [Rodrigues, et al., 2005]. The severity of this mutation is evident, the N-X-S/T site is

15

modified to N-X-P and consequently is very likely to induce major defects in the intracellular traffic

16

of the TPO.

17

Umeki et al. (2004) analyzed five mutated TPOs with alteration in expression probably due to

18

incorrect folding, four of them were missense mutations (p.R175Q, p.G533C, p.R665W, and

19

p.G771R, Figure 4) [Kotani et al., 2003, 2004; Umeki et al., 2002] and one corresponded to the

20

elimination of two contiguous amino acids (p.D574-L575del, Figure 4). The five mutated TPOs had

21

a pathogenic mutation on the extracellular domain but had no abnormality on the transmembrane

22

and cytoplasmic domains. The mutated mRNAs were transfected in Cho-Ki and NRK cells and then

23

they were studied by indirect immunofluorescence and flow cytometry to determine the intracellular

24

localization. In permeabilized condition the 5 mutated proteins were located in intracellular

25

structures like the nuclear envelope and ER in both cells line, whereas in nonpermeabilized

18

1

condition only p.D574-L575del was observed on the cell surface of Cho-KI and p.D574-L575del in

2

addition to p.G771R, on the cell surfase of NRK cells [Umeki et al. (2004]. The analysis of the

3

TPO protein expression on the cell surface by flow cytometry showed a positive fluorescense peak

4

in p.G533C, p.D574-L575del and p.G771R with less intensity compared with wild-type TPO. In

5

contrast, p.R175Q and p.R665W were not expressed on the cell surface. It has been reported that

6

calnexin and calreticulin binding to TPO are necessary for the initial folding steps [Fayadat et al.,

7

2000]. In immunoprecipitation by anti-TPO antibodies followed by Western blot using anti-TPO,

8

anti-calnexin and anti-calreticulin, p.G533C, p.D574-L575del and p.G771R exhibited increasing

9

interaction with calnexin. This suggests that some p.G533C, p.D574-L575del and p.G771R mutated

10

proteins would be fold properly in the calnexin cycle and secreted via Golgi to apical membrane of

11

the thyrocyte, while the remaining of these misfolded TPOs would be degraded in the ERAD

12

pathway. In p.R175Q and p.R665W missense mutations, an arginine was substituted for glutamine

13

or tryptophan, respectively. Arginine is a basic amino acid with positive charge, whereas glutamine

14

is uncharged and tryptophan is a hydrophobic amino acid. The loss of an arginine could cause

15

structural changes and as a consequence p.R175Q and p.R665W mutations would follow the

16

anterograde pathway and would be degraded in proteasomes.

17 18

5. Defects in protein folding of DUOX system

19

The DUOX1 and DUOX2 genes are located in opposite transcriptional orientation in an operon-

20

like unit on chromosome 15, encoding similar proteins that are inserted in the apical membrane of

21

thyroid cells [De Deken et al. 2000, 2002; De Deken and Miot, 2019; Dupuy et al, 1999;

22

Grasberger, 2010; Muzza and Fugazzola, 2017, Ohye and Sugawara, 2010]. The DUOX2 gene

23

spans 22 Kb of genomic DNA (15:45,092,650-45,114,344; reverse strand, GRCh38 assembly)

24

which includes 34 exons, being the first one non-coding. The mRNA (NCBI Reference Sequence:

25

NM_014080.4) is 6,428 nucleotides long and the preprotein is composed of a putative 21 amino

19

1

acids signal peptide followed by a 1,527 amino acids polypeptide (Figure 5) [De Deken et al. 2000,

2

2002; Dupuy et al, 1999]. DUOX1 gene encodes a homologous protein displaying 83 % sequence

3

similarity. The 36 kb DUOX1 gene (15:45,129,933-45,165,576; forward strand, GRCh38 assembly)

4

consists of 35 exons (being the first two noncoding) encoding a protein of 1551 amino acids

5

(mRNA: 5675 nucleotides long; NCBI Reference Sequence: NM_017434.5) in which the first 21

6

amino acids correspond to the putative signal peptide [De Deken et al. 2000, 2002; Dupuy et al,

7

1999]. The structure of DUOX1 and DUOX2 proteins contains three regions: 1) peroxidase

8

homology ectodomain (peroxidase-like domain): an N-terminal extracellular region, which has 43%

9

sequence homology with hemoperoxidases, which distinguishes DUOX from other family

10

NOX/DUOX, 2) a central region encompassing an intracellular loop containing potentially two EF-

11

hand calcium binding motifs, which would be involved in regulating the conformation and activity

12

of NADPH oxidase; 3) gp91phox/NOX2-like domain: a C-terminal region, comprising the last six

13

transmembrane domains (Figure 5) [De Deken et al., 2000; Dupuy et al., 1999; Morand et al., 2004;

14

Ohye and Sugawara, 2010]. The C-terminal domain includes four conserved histidines, which are

15

coordination sites for two heme prosthetic group [Muzza and Fugazzola, 2017]. The C-terminal

16

cytosolic orientation presents a binding site for FAD and NADPH (Figure 5) [De Deken et al.,

17

2000; Dupuy et al., 1999; Morand et al., 2004; Ohye and Sugawara, 2010]. It is believed that this

18

region is responsible for the transfer of electrons from NADPH through the membrane. DUOX2

19

exhibits higher expression and is also more efficient in the production of peroxide than DUOX1

20

[Muzza and Fugazzola, 2017]. DUOX1 is regulated by protein kinase A through Gs-PKA pathway,

21

while DUOX2 is stimulated by protein kinase C through Gq-phospholipase C pathway [Rigutto et

22

al., 2009]. Although both enzymes are expressed in the thyroid, DUOX1 is also expressed in the

23

respiratory and tongue epithelia, testis and cerebellum, and DUOX2, in respiratory epithelia, uterus,

24

gallbladder, pancreatic islets and digestive tract where these enzymes could provide H2O2 for

25

peroxidase-mediated antimicrobial defense processes [El Hassani et al., 2005; Forteza et al., 2005;

20

1

Geiszt et al, 2003]. DUOX1 and DUOX2 present two N-glycosylated states, the high mannose

2

glycosylated immature form of 180 kDa expressed in ER and the fully glycosylated mature form of

3

190 kDa which is translocated at the plasma membrane and generates H2O2 [Morand et al., 2004;

4

Muzza and Fugazzola, 2017]. Partially glycosylated DUOX2 remains in the ER [Grasberger and

5

Refetoff, 2006].

6

DUOXA1 and DUOXA2 are essential for DUOX maturation [Grasberger and Refetoff, 2006,

7

Muzza and Fugazzola, 2017]. The DUOXA1 and DUOXA2 genes are oriented head to head in the

8

intergenic region of 16 kb (DUOXA2: 15:45,114,321-45,118,421, forward strand; DUOXA1:

9

15:45,117,367-45,129,938, reverse strand, GRCh38 assembly) between both DUOX genes. The

10

DUOXA2 open reading frame spans 6 exons encoding a protein of 320 amino acids (mRNA: 1755

11

nucleotides long; NCBI Reference Sequence: NM_207581.4), whereas the DUOXA1 spans 11

12

exons encoding a protein of 483 amino acids (mRNA: 2021 nucleotides long; NCBI Reference

13

Sequence: NM_144565.4). The two DUOXA genes encode a N-terminal ectodomain, five

14

transmembrane domains and a C-terminal cytoplasmic region. The first extracellular loop presents

15

three N-glycosylation sites [Grasberger and Refetoff, 2006; Muzza and Fugazzola, 2017].

16

DUOXA1 and DUOXA2 enable the ER-to-Golgi transition, maturation, and surface expression of

17

functional DUOX1 and DUOX2 in a heterologous cell system (Figure 6) [De Deken and Miot,

18

2019; Grasberger and Refetoff, 2006]. A stable complex between DUOX2 and DUOXA2 is then

19

required for their enzymatic activity (Figure 6). DUOXA2 would be part of the DUOX2 quality

20

control system, indicating that DUOXA2 displays a chaperone-like function with respect to its partner.

21

[Carré et al., 2015; Grasberger et al. 2007].

22

DUOX2 alterations were first described in 2002 [Moreno et al.]. In the literature, Sanger

23

sequencing has been commonly used for DUOX2 mutations identification in both research and

24

clinical studies. Recently, high throughput sequencing has been reported to be a rapid and effective

25

tool for largest DUOX2 mutation screening for patients with or without family history [Chem et al.,

21

1

2018; Long et al., 2018; Makretskaya et al., 2018; Sun et al., 2018; Yu et al., 2018; Zou et al.,

2

2018]. The mutational spectrum shows heterogenous alterations dispersed in the exons and introns

3

of the entire gene (Figure 5) [Belforte et al., 2016a; De Deken and Miot, 2019]. To date, molecular

4

research has shown more than one hundred fifty pathogenic variations have been found to be

5

associated with TCH, PCH or subclinical congenital hypothyroidism [Belforte et al., 2016a; De

6

Deken and Miot, 2019; Fu et al., 2016a].

7

mutations were found to be altered proteins with a single amino acid substitution, followed by

8

deletions/insertion and nonsense mutations causing premature truncation of the DUOX2 protein

9

[Belforte et al., 2016a; Muzza and Fugazzola, 2017]. Premature termination codons can delete the

10

gp91phox/NOX2-like domain and the peroxidase-like domain. The functional consequence could

11

be the complete impossibility to generate H202 from the DUOX2 system. Only six DUOX2 splice

12

site mutations have been reported, c.514-2A>G [Tan et al., 2016], c.2335-1G>C [Belforte et al.,

13

2016b; Muzza et al., 2014], c.2655-2A>C [Varela et al., 2006], c.3516-1G>A [Tan et al., 2016]

14

c.3693+1G>T [Tan et al., 2016] and c.4240-1G>C [Matsuo et al., 2016]. The p.K530*, p.E879K,

15

p.R1110Q, p.L1160del, p.R1334W and p.L1343F mutations have been the most frequently

16

identified DUOX2 mutations in Asian population (Figure 5) [Belforte et al., 2016a; Muzza and

17

Fugazzola, 2017], whereas the frequent p.F965Sfs*29 (c.2895_2898delGTTC) mutation (Figure 5)

18

has proved to be the most frequent in the Caucasian population [Belforte et al., 2016a; Muzza and

19

Fugazzola, 2017].

The most frequent events of deleterious DUOX2

20

Grasberger et al. [2007] analyzed three missense mutations in DUOX2 (p.Q36H [Varela et al.,

21

2006], p.R376W [Cortinovis et al., 2008; Lü et al., 2011; Vigone et al., 2005] and p.D506N

22

[Cortinovis et al., 2008; Pfarr et al., 2006] involved in congenital hypothyroidism (Figure 5), in

23

terms of its impact on the generation of H2O2, folding, intracellular traffic and DUOXA2

24

interaction, in a reconstituted heterologous system. The three missense mutations displayed either

25

complete (p.Q36H and p.R376W) or partial loss (p.D506N) of H2O2-generating activity. The

22

1

mutations p.Q36H and p.R376W completely prevent the transit of DUOX2 to the cell surface and

2

showed a diminished oxidized-to-reduced ratio [Grasberger et al. 2007]. The mutant proteins are

3

present predominantly as core N-glycosylated (a thiol-reduced folding intermediate) and are

4

retained by the quality control system within the ER as indicated by a greater complexation with the

5

lectin calnexin. This interaction may be related to retrotranslocation and proteasomal degradation.

6

On the contrary, the p.D506N mutation showed a similar ratio of oxidized-to-reduced forms and a

7

diminished expression at the plasma membrane [Grasberger et al. 2007]. This showed a successful

8

route to the plasma membrane through of Golgi complex but with an abnormal maturation.

9

Oxidative folding of DUOX2 in the ER appears to be the rate-limiting step in the maturation of

10

DUOX2. ER-localized DUOX2 is almost exclusively in the reduced form (presumably inactive)

11

[Grasberger et al. 2007]. The oxidized form (presumably active) does not accumulate in the ER due

12

to rapid, DUOXA2-mediated export (Figure 6). DUOXA2 allows rapid ER exit of DUOX2 folded

13

and also enhances the degradation of DUOX2 protein mutants not suitable for ER output

14

[Grasberger et al. 2007]. However, DUOXA2 does not affect the reduced-to-oxidized transition,

15

this observation would indicate that DUOXA2 would act after the maturation of DUOX2, once

16

DUOX2 has acquired a stable disulfide-bonded conformation. Intramolecular (DUOX2) and

17

intermolecular (between DUOX2 and DUOXA2) disulfide bonds which are promoted and

18

maintained by the oxidizing environment of the ER seem to play an important role in the protein

19

folding, stability, function and protein-protein interaction. Carré et al. [2015] reported an

20

intramolecular bond in the DUOX2 between cysteine residues 124 (located in the N-terminal ectodomain)

21

and 1162 (located in the second extracellular loop), that provides structural support for the formation of

22

interdisulfide bridges between DUOX2 and DUOXA2 (Figure 6). The intermolecular bonds between

23

DUOX2 and DUOXA2 involve the N-terminal cysteine residues 568 and 582 of DUOX2 respectively and

24

cysteine residues 167 of the first extacellular loop and 233 of the second extacellular loop of DUOXA2

25

(Figure 6) [Carré et al., 2015]. It is importan to note that mutations of cysteine residues in the N-

26

terminal domain of human DUOX2 protein affect both targeting and extracellular H2O2 production

23

1

by human DUOX2 [Fortunato et al., 2010; Morand et al., 2004]. Recently, the p.C124* (dbSNP:

2

rs1197021807) and p.C568R (rs770603845) mutations (Figure 5) was reported in gnomAD data set

3

[https://gnomad.broadinstitute.org/gene/ENSG00000140279] in a heterozygous state, with an allele

4

frequency of 2/249972 (0,000008001) and 4/188376 (0.00002123) respectively.

5 6

6. Defects in protein folding of NIS

7

The SLC5A5 gene that encodes NIS, is located on chromosome 19, spanning 23 Kb

8

(19:17,871,945-17,895,174; GRCh38 assembly). The full-length human mRNA (NCBI Reference

9

Sequence: NM_000453.3) consists of 15 exons (3604 nucleotides) and encodes a 643 amino acids

10

peptide with 13 transmembrane segments, an extracelular N-terminal extreme and an intracellular

11

C-terminus (Figure 7) [De la Vieja et al.,2004, 2005; Spitzweg and Morris, 2010; Targovnik et al.,

12

2017]. This transporter has three N-glycosilation sites (Figure 7) [Spitzweg and Morris, 2010],

13

although their glycosilation is not essential for its function as iodine transporter [Levy et al., 1998]

14

or for reaching and inserting into the plasma membrane [Li et al., 2013].

15

The expression of NIS is regulated primarily by TSH levels [Carrasco, 2013], I- concentration in

16

plasma [Portulano et al., 2014] and by TG [Vono-Toniolo and Kopp, 2004]. Its trafficking to the

17

cell membrane is also regulated by the interaction with other proteins, such as pituitary tumor–

18

transforming gene (PTTG) binding factor (PBF) which promotes its internalization [Stratford et al.

19

2005].

20

The analysis of the effect of reported mutations affecting SLC5A5 and of different mutagenesis

21

studies allowed understanding the key role of certain residues for the folding and trafficking of the

22

NIS. Several loss-of-function mutations leading to I- transport defect (ITD) have been reported

23

[Martín et al., 2018a]; five of them lead to the intracellular retention of this transporter: p.R124H,

24

p.V270E, p.A439_P443del, p.S509Rfs*7 and p.G543E (Figure 7). The first three inhibit also the

25

transporter activity [Ravera et al., 2017]. The interaction between the δ-amino group at 124 and the

24

1

thiol in the Cys440, lost in p.R124H, is crucial for the folding of NIS and targeting to the plasma

2

membrane [Paroder et al, 2013]. The deleterious effect of p.A439_P443del is due to the loss of

3

Asn441 of the side chain which interacts with Gly444, capping the α-helix of transmembrane segment

4

XII [Li , et al., 2013]. The frameshift mutation p.S509Rfs*7 [Pohlenz, et al., 1998; 2000] leads to

5

the loss of the C-terminal portion of this transporter. Different studies [Martín et al., 2018b; Martín

6

et al., 2018c] have demonstrated the importance of this region (between residued Ile546 and Lys618),

7

which is subjected to post translational modifications, such as phosporilation for the migration to

8

the plasma membrane. It was shown that a putative tryptophan-acidic motif (Trp565-Asp566) is

9

required for leaving the ER and reaching the membrane. Furthermore, the region between Val580

10

and Lys618 is crucial for its basolateral location. Only alanine or serine are tolerated at position 543,

11

other substitutions as p.G543E lead to the intracellular retention of the altered peptide [De la Vieja

12

et al., 2005]. The mutation p.V270E alters the positive charge in the surface of the molecule, which

13

would alter the interaction with the proteins involved in the trafficking of NIS [Nicola et al, 2015].

14

The systematic assessment of sorting motifs in NIS, carried out by Darrouzet et al. [2016] revealed

15

an internal noncanonical PDZ-binding motif, that involves the residues Tyr118, Glu119, Tyr120 and

16

Leu121, necessary to reach the plasma membrane. This site was predicted bioinformatically and

17

assessed with functional assays, including directed mutagenesis which validated the role played by

18

Leu121.

19 20

7. Defects in protein folding of Pendrin

21

The passive transport of the iodine to the follicular lumen is mostly facilitated by pendrin [Bidart

22

et al., 2000], encoded by the SLC26A4 gene, that maps in chromosome 7 (7:107,660,828-

23

107,717,782; GRCh38 assembly). It comprises 21 exons (the first one is noncoding), covers 57 kb

24

of genomic DNA and encodes 780 amino acids (Figure 8). The mRNA is 4,737 nucleotides long

25

(NCBI Reference Sequence: NM_000441.2) and the canonical isoform codifies for a classic

25

1

predicted topology of 12 transmembrane domains with intracellular N- and C-terminal tails (Figure

2

8). Dossena et al., [2009] proposed another model of pendrin consisting of 15 transmembrane

3

segments, an extracelular N-terminal extreme and an intracellular C-terminus. It is interesting to

4

note that there is no experimental data proving any of the suggested models. Pendrin has a STAS

5

(sulfate transporter and antisigma factor antagonist) domain (Figure 8) [Aravind and Koonin, 2000].

6

Pendrin has affinity for chloride, iodide and bicarbonate, among other anions. The peptide is N-

7

glycosylated exclusively at Asn167 and Asn172 which are located in the second putative extracellular

8

loop of pendrin (Figure 8). N-glycosylation of membrane proteins plays a role in their correct

9

folding and targeting to the membrane [Azroyan et al., 2011].

10

As was previously mentioned, mutations in this gene can lead to nonsyndromic hearing loss

11

associated with enlarged vestibular aqueduct (EVA) or Pendred syndrome, an autosomal recessive

12

disorder characterized by sensorineural hearing impairment, thyroid goiter and a partial

13

organification problem detectable by a positive perchlorate test [Bizhanova and Kopp, 2009]. The

14

main molecular mechanism proposed for Pendred Syndrome has been SLC26A4 mutations causing

15

the retention of the mutants in the ER [Rotman-Pikielny et al, 2002]. Retention of misfolded

16

pendrin in the ER and subsequent degradation is a major pathogenic mechanism for Pendred

17

syndrome. There are more than 170 sequence variants associated with this phenotype [Campbell, et

18

al., 2001; Dossena et al., 2009, 2011; Tsukamoto et al., 2003], 64% of these are single nucleotide

19

changes leading to amino acid substitutions and at least eight of them (p.G102R; p.V138F; p.T193I;

20

p.L236P; p.T410M; p.L445W; p.Q446R and p.Y556C) [Coyle et al., 1998; de Moraes et al., 2016;

21

Taylor et al., 2002] (Figure 8) lead to diminished or abolished expression of the mutant pendrin in

22

the apical membrane. Although the specific molecular chaperones involved in the folding of this

23

protein have yet to be identified, it has been reported that Hsp90 is not involved, and furthermore,

24

its inhibitor 17-allylamino-17-demethoxygeldanamycin (17-AAG) increases the stability of Pendrin

25

by inducing the transcription of heat shock transcription factor 1 (Hsf1)-dependent chaperones [Lee,

26

1

et al., 2012]. The same authors also demonstrated that the E3 ubiquitin ligase Rma1 is involved in

2

ERAD Pendrin degradation.

3

Jung et al [2006] reported that the cytosolic heat shock cognate protein 70 (Hsc70) and the co-

4

chaperone DNAJC14 are involved in the unconventional Golgi-bypass trafficking pathway that can

5

be stimulated by ER stress-inducing agents. This can lead to cell surface expression of misfolded

6

variants, making them potential targets for treating Pendred syndrome.

7 8

8. Defects in protein folding of SLC26A7

9

In the recent years, the role of pendrin as the main apical iodine transporter has been challenged,

10

as consequence of the lack of a complete correlation between deleterious mutations affecting this

11

transporter and hypothyroidism [Bizhanova and Kopp, 2011] and the fact that apical iodide efflux is

12

also possible in the absence of pendrin [Silveira and Kopp, 2015]. Interestingly, 50% of patients

13

with Pendred syndrome are euthyroid. In the last two years, recessive loss-of-function mutations in

14

another transporter named SLC26A7 were identified associated with goitrous CH [Cangul, et al.,

15

2018; Ishii et al., 2019; Zou et al., 2018]. The SLC26A7 gene is located on chromosome 8, spanning

16

149 Kb (8:91,249,319-91,398,155; GRCh38 assembly) and including 19 exons, being the first one

17

non-coding. The full-length human mRNA (NCBI Reference Sequence:

18

of 5,237 nucleotides. Sequence analysis revealed an open reading frame of 1,971 nucleotides

19

encoding a 656-amino acid peptide [Lohi et al, 2002]. The structure of SLC26A7 is characterized

20

by citoplasmatic N- and C-terminal domains, 12 putative transmembrane segments and a C-terminal

21

cytoplasmic STAS consensus domain (Figure 9) [Alper and Sharma, 2013; Lohi et al, 2002]. In

22

addition, it contains two predicted N-glycosilation sites at Asn125 and Asn131 (Figure 9) [Lohi et al,

23

2002]. The SLC26A7 was first discovered as a chloride/anion exchanger in the kidney (expressed in

24

the basolateral membrane of the outer medullary collecting duct intercalated cells) and in the

25

stomach (expressed in the basolateral membrane of the gastric parietal cells) [Petrovic et al., 2003;

NM_052832.4) consists

27

1

2004]. Recently, immunohistochemical and RT-PCR analyses confirmed the expression of

2

SLC26A7 in human thyroid follicular cells, located predominantly on the apical side facing the

3

follicular lumen [Ishii et al., 2019].

4

In 2018, a homozygous mutation in the SLC26A7 gene (c.927_930delCACT, c.932_933delAA

5

[p.I309Mfs*9,p.E311Gfs*28] or c.927_933delinsG [p.I309_E311delinsM]) was firstly reported in a

6

family of Saudi origin with two CH patients [Zou et al., 2018] (Figure 9). The unaffected mother,

7

father and a sibling carried the mutation in a heterozygous state. Almost simultaneously, Cangul et

8

al. [2018] identified novel homozygous mutations in the SLC26A7 gene in six unrelated families

9

with goitrous CH. By whole exome sequencing and subsequent Sanger sequencing, it was identified

10

in three consanguineous families of Pakistani or Turkish origin a homozygote c.679C>T transition,

11

which replaces the arginine residue in position 227 with a stop codon (p.R227*) [Cangul et al.,

12

2018] (Figure 9). In the same paper, in other three non-consanguineous families of Finnish origin,

13

the authors reported a homozygous frameshift mutation (c.1893delT, p.F631Lfs*8) (Figure 9).

14

More recently, Ishii et al. [2019] described two Japanese siblings with goitrous CH caused by a

15

homozygous nonsense mutation (c.1498C>T; p.Q500*) (Figure 9). In vitro studies of the mutant

16

evidence a complete loss of iodide transport activity. The p.Q500* mutation, located in the STAS

17

domain, impaired the localization of SLC26A7 in the cell membrane. It is well documented that the

18

STAS domain is required for membrane localization and exchanger function [Dorwart et al., 2008;

19

Sharma,et al., 2011].

20

Remarkably, hypothyroidism caused by SLC26A7 mutants cannot be rescued by wild-type

21

pendrin, suggesting that SLC26A7 plays a critical function in thyroid hormone synthesis. [Zou et

22

al., 2018]. Ishii et al. [2019] hypothesized that the SLC26A7 may be expressed more abundantly

23

than pendrin in the human thyroid gland and that SLC26A7 may have another function in the

24

biosynthesis of thyroid hormones in addition to being an iodide transporter. These hypotheses

25

require further investigation.

28

1

9. Defects in protein folding of the proteins involved in disembryogenesis

2

In only 5-10% of the thyroid dysgenesis, the underlying genetic cause can be unequivocally

3

determined and linked to a specific sequence variant. As was mentioned earlier, the mutations

4

described affect mainly genes involved in the development, growth and differentiation of the

5

thyroid cells (NKX2-1, FOXE1, PAX8, NKX2-5, TSHR, TBX1, CDCA8, HOXD3 and HOXB3).

6

The TSHR is member of the superfamily of G Protein-coupled receptors [Vassart and Dumont,

7

1992, Tao and Conn, 2014]. The canonical isoform of 764 amino acids consists of seven

8

transmembrane domains, with an N-terminal extracellular segment that interacts with the TSH and a

9

cytoplasmatic C-terminal domain, capable of interacting with the G protein [Corvilain, et al., 2001].

10

As other glycosylated transmembrane proteins, TSHR undergoes post translational modifications

11

and folding in the ER lumen where it interacts with the resident chaperones. Studies by Siffroi-

12

Fernandez et al., [2002] and Mizrachi and Segaloff [2004] showed the interaction of this nascent

13

peptide with calnexin and calreticulin. Siffroi-Fernandez et al., [2002] were also able to detect the

14

binding of BiP in K652 cells, whereas Mizrachi and Segaloff [2004] could not find this association

15

and demonstrated the interaction with the protein disulfide isomerase (PDI) in HEK293 cells.

16

Loss-of-function germline mutations in TSHR are expected to cause congenital hypothyroidism

17

with a syndrome of resistance to TSH [Corvilain et al., 2001]. Most of these variants lead to

18

misfolding of the receptor and its intracellular retention in the ER [Mizrachi and Segaloff, 2004].

19

Familial cases of congenital hypothiroidim due to mutations in NKX2-1 and PAX8, often show

20

incomplete penetrance, suggesting the presence of additional modifiers [De Felice and Di Lauro,

21

2004]. Molecular analyses of DHTP mice (Double Heterozygous for both Titf1- and Pax8-null

22

mutations) that presented thyroid hypoplasia during embryogenesis and high incidence of thyroid

23

hemiagenesis as adults [Amendola, et al., 2005], revealed the presence of a strain specific mutation

24

in DNAJC17 [Amendola, et al., 2010], making it a candidate locus for hypothyroidism

25

predisposition. The DNAJC17 human gene encodes 304 amino acids and is a member of the heat-

29

1

shock-protein-40 family (Hsp40), acting usually as cofactor of the Hsp70 chaperone, regulating

2

their ATPase activity, through the J domain that all members display. [Kampinga and Craig, 2010].

3

Nevertheless, many of the functions that this co-chaperon exercises are still unknown. Pascarella, et

4

al [2018] demonstrated its co-localization in the nuclear speckles (also known as interchromatin

5

granule clusters) and its interaction with spliceosomal components, suggesting a participation in

6

splicing processes. Germline loss-of-function mutations in a homozygous state in this gene were

7

also associated with the recessive inheritance of retinal dystrophy [Patel, et al., 2016] This gene was

8

screened in 89 patients with congenital hypothyroidism, but only a possible damaging rare variant

9

c.610G>C was identified in a single patient [Nettore, et al., 2018], showing that DNAJC17 is not an

10

initial candidate gene to be analyzed in thyroid dysgenesis.

11

In the recent years, it has been described that mutations affecting different genes lead to

12

congenital hypothirodism associated to other syndromes, such as the Krüppel-like zinc finger

13

transcription factor GLI-similar 3 (GLIS3) which causes neonatal diabetes, a large spectrum of

14

abnormalities and clinical presentations of congenital hypothyroidism, from thyroid agenesia to

15

dyshormogenesis [Dimitri, et al., 2015]. This transcription factor mediates the TSH/TSHR

16

proliferation of thyrocites and is required for the biosynthesis of thyroid hormones, since

17

experiment in mice showed it regulates the expression of NIS and pendrin [Kang, et al., 2017]. On

18

the other hand, loss-of-function mutations that affect the C-terminus of the immunoglobulin

19

superfamily member 1 and impair its trafficking to the plasma membrane, lead to X-linked

20

congenital central hypothyroidism, with prolactin and GH deficiency, delayed pubertal development

21

and testicular enlargement [Sun, et al., 2012; Turgeon, et al, 2017]. Deficient levels of

22

immunoglobulin superfamily member 1, decrease the pituitary response to TRH because the

23

expression of its receptor is downregulated.

24 25

10. Conclusions and future perspectives

30

1

Congenital hypothyroidism (CH) describes a hormonal state of insufficient thyroid hormone

2

secretion detected at birth or postnatally. Over the last 20 years, mutations in the candidate genes

3

were identified and functionally characterized providing an overwhelming wealth of new insights

4

into basic molecular mechanisms.

5

Mutations affecting the proteins involved in the cascade of biosynthesis of TH, thyroid

6

organogenesis and its regulating pathways have been classified as missense and nonsense

7

mutations, frameshift mutations by single or multiple nucleotide deletion or insertion and splicing

8

mutations in the exonic or intronic consensus sites. The mutations are located in specific domains or

9

cysteine residues or glycosylation sites affecting a crucial function or the formation of disulfide

10

bridges or conformational maturation, respectively. In the dishormonogenesis the thyroid gland

11

develops hypertrophy and hyperplasia by the proliferative effect as a consequence of the chronic

12

elevation of TSH. In contrast, the dysembryogenesis results agenesis or hypoplasia or ectopia.

13

Many of the mutations cause protein misfolding and accumulation in the ER, affecting the secretory

14

capacity of the thyroid cell. The retention of misfolded proteins alters the ER homeostasis and

15

activates the ERAD mechanism, responsible for their cytosolic degradation.

16

Despite all these remarkable advances numerous challenging aspects remain to be solved. The

17

recent generation of new potent approaches including new sequencing technology and the

18

generation of thyroid organoids provides additional tools that might be of great help in such studies.

19 20 21 22 23 24 25

31

1

Declaration of interest

2

The authors declare that there is no conflict of interest that could be perceived as prejudicing the

3

impartiality of the research reported.

4 5

Funding

6

This study was funded by grants from the FONCyT-ANPCyT-MINCyT (PICT 2014-1193 to CMR

7

and PICT 2015-1811 to HMT), CONICET (PIP 2015-11220150100499 to CMR) and Universidad

8

de Buenos Aires (UBACyT 2016-20020150100099BA to CMR).

9 10

Acknowledgements

11

C.M. Rivolta and H.M. Targovnik are established investigators of the Consejo Nacional de

12

Investigaciones Científicas y Técnicas (CONICET).

13

K.G. Scheps is a postdoctoral research fellow of the CONICET.

14

32

1

Legends of Figures

2

Figure 1. Biosynthesis of thyroid hormones and iodine recycling in the polarized follicular

3

thyroid cells. The major steps in iodide (I−) metabolism (uptake, efflux, organification, and

4

recycling) in the different compartments and thyroid proteins involved are shown. After I− uptake

5

at the basolateral membrane of thyrocytes by the Na+/I- symporter (NIS), it is transported at the

6

apical membrane into the follicular lumen through the pendrin, anoctamin-1 and SLC26A7

7

transporters. NIS is dependent on the sodium gradient created by the Na/K-ATPase and KCNQ1-

8

KCNE2 potassium channel is required for NIS-mediated thyroid cell uptake of I-. In the cell-colloid

9

interface, the I- is oxidized to iodinium (I+) and rapidly organified by incorporation into selected

10

tyrosyl residues of thyroglobulin (TG), which results in the formation of two intermediaries, 3-

11

monoiodotyrosine (MIT) and 3,5-diiodotyrosine (DIT). Subsecuently, iodotyrosine coupling results

12

in the synthesis of 3,5,3'-triiodothyronine (T3) and 3,5,3',5'-tretraiodothyronine (T4). The oxidizing,

13

iodination and coupling reactions are mediated by thyroid peroxidase (TPO) in presence of H2O2

14

generated by dual oxidase 1 and 2/dual oxidase maturation factor 1 and 2 (DUOX1 and

15

2/DUOXA1 and DUOXA2) complex. Iodinated TG is stored as colloid in the follicular lumen.

16

Once endocytosed into the thyrocytes, the TG is digested by lysosomal proteases, thus releasing

17

attached MIT, DIT, T4 and T3. T4 and T3 are then transported out of the follicular cells and into the

18

bloodstream, through the monocarboxylate transporter 8 (MCT8). Meanwhile, MITs and DITs are

19

deiodinated by the iodotyrosine deiodinase 1 (IYD-1), present at the apical plasma membrane and

20

in endocytic vesicles of the thyrocytes and the iodine atoms as I- are transported to the

21

intrathyroidal iodide pool. Simultaneously, free amino acids (aa) generated during proteolysis of Tg

22

are conducted to the intrathyroidal aa pool. The transcription of specific thyroid genes and thyroid

23

function is under the control of thyrotropin (TSH) via its protein G-coupled receptor (TSHR).

24

GDP: guanosine diphosphate.

33

1

Figure 2. Structural organization of the thyroglobulin homodimer. The repetitive units of

2

thyroglobulin (TG) TG type 1, TG type 2 and TG type 3 and the cholinesterase-like (ChEL)

3

domain, drawn to scale, are represented by boxes. N-terminal T4 (coupling of a donor DIT149 with

4

the acceptor DIT24) and N-terminal T3 (coupling of a MIT2766 at the antepenultimate residue of one

5

TG monomer with the antepenultimate DIT2766 in the opposite monomer) forming sites are shown.

6

The amino acid sequences (represented by single-letter code) of TG are indicated below of the

7

respective protein schematic diagrams. The mutated residues are shown in red and double

8

underlined (see Table 1 and 2). The amino acid position is designated according to the reference

9

sequences reported in NCBI accession number: NM_003235.5 The amino acid (aa) positions are

10

numbered including the 19 amino acid of the signal peptide (SP).

11 12

Figure 3. Schematic model of the N-glycosylation pathway of thyroglobulin. First, the nascent

13

polypeptide acquires the N-glycans by action of an oligosaccharyltransferase that transfers Glc3

14

(□), Man9 (○), GlcNAc2 (■) from dolichol phosphate precursor to the consensus N-glycosylation

15

sites (Asn, asparagine). Next, I and II α-glucosidases trim two glucose residues creating a

16

monoglucosylated N-glycan, later the third glucose is removed by α-glucosidase II. After removal

17

of one mannose residue in the ER, α-mannosidase I, the thyroglobulin is exported to the Golgi.

18

Finally, Golgi I and II α-mannosidases remove mannose residues and several N-acetylglucosamine

19

glycosyltransferases (GnT) may catalyze branching and elongation of the carbohydrate chains,

20

resulting in hybrid-type or complex-type N-glycans. N-glycans are elongated by addition of

21

galactose, fucose and sialic acid residues catalyzed by galactosyltransferases, fucosyltransferases

22

and sialyltransferases, respectively.

23 24

Figure 4. Structural organization of the thyroid peroxidase. The signal peptide (SP),

25

complement control protein (CCP) domain, calcium-binding EGF (EGF-Ca2+) domain, animal

34

1

haem peroxidase (An peroxidase) region and heme-binding sites (proximal histidine (H494), distal

2

histidine (H239), aspartic acid (D238), glutamic acid (E399), drawn to scale, are shown. The arrows

3

indicate where the mutated residues are located. The amino acid sequences (represented by single-

4

letter code) of thyroid peroxidase protein are indicated below of the respective protein schematic

5

diagrams. The mutated residues are shown in red and double underlined. The amino acid position is

6

designated according to the reference sequences reported in NCBI accession number:

7

NM_000547.5. The amino acid positions are numbered including the 18 amino acid of the signal

8

peptide (SP). ECD: extracellular domain, TM: transmembrane domain, ICD: intracellular domain,

9

aa: amino acid.

10 11

Figure 5. Structural organization of the DUOX2. The peroxidase-like and gp91phox/NOX2-like

12

domains, glycosilation sites (NXS/T site), EF-hand calcium binding motifs and FAD and NADPH

13

binding sites, drawn to scale, are shown. The seven alpha helice transmembrane domains are

14

represented by boxes (I, II, III, IV, V, VI and VII). The arrows indicate where the mutated residues

15

are located. The amino acid sequences (represented by single-letter code) of DUOX2 are indicated

16

below of the respective protein schematic diagrams. The mutated residues are shown in red and

17

double underlined. The amino acid position is designated according to the reference sequences

18

reported in NCBI accession number: NM_014080.4. The amino acid (aa) positions are numbered

19

including the 21 amino acid of the signal peptide (SP).

20 21

Figure 6. Folding model of DUOX2-DUOXA2. Formation of intermolecular disulfide bonds

22

between DUOX2 and DUOXA2 and trafficking of the DUOX2-DUOXA2 heterodimer to the

23

apical plasma membrane of the thyrocyte are shown. DUOX2 requires a DUOXA2 to exit from the

24

endoplasmic reticulum (ER) and reach the apical plasma membrane, through the Golgi apparatus.

25

The oxidative folding of DUOX2 and DUOXA2 takes place in the ER, generating an

35

1

intramolecular disulfide bridge in DUOX2 (between cysteine124 and cysteine1162) and

2

intermolecular disulfide bridges between the DUOX2 (cysteine568 and cysteine582) and DUOXA2

3

(cysteine167 and cysteine233). Both proteins form a stable heterodimer complex at the cell surface

4

which is fundamental for the reactive oxygen species production. N : N-glycosylation site,

5

cysteine, NADPH: binding site for the NADPH, FAD: binding site for the FAD.

:

6 7

Figure 7. Thirteen-TMS human Na+/I- symporter secondary structure model. Schematic

8

representation adapted from De la Vieja et al. [2004, 2005] and Spitzweg and Morris [2010] is

9

showed. The arrows indicate where the mutated residues are located. The amino acid sequences

10

(represented by single-letter code) of Na+/I- symporter are indicated below of the respective protein

11

schematic diagrams. The mutated residues are shown in red and double underlined.

12

glycosylation site, TM: transmembrane domain.

N : N-

13 14

Figure 8. Twelve-TMS human pendrin secondary structure model. Schematic representation

15

adapted from Dossena et al. [2009] is showed. The arrows indicate where the mutated residues are

16

located. The amino acid sequences (represented by single-letter code) of pendrin are indicated

17

below of the respective protein schematic diagrams. The mutated residues are shown in red and

18

double underlined. STAS (sulfate transporter and antisigma factor antagonist) domain is

19

underlined. N : N-glycosylation site, TM: transmembrane domain.

20 21

Figure 9. Twelve-TMS human SLC26A7 secondary structure model. Schematic representation

22

adapted from Ishii et al. [2019] is showed. The arrows indicate where the mutated residues are

23

located. The amino acid sequences (represented by single-letter code) of the SLC26A7 are

24

indicated below of the respective protein schematic diagrams. The mutated residues are shown in

36

1

red and double underlined. The STAS (sulfate transporter and antisigma factor antagonist) domain

2

is underlined. N : N-glycosylation site, TM: transmembrane domain.

3

37

1

References

2

Abramowicz, M.J., Targovnik, H.M., Varela, V., Cochaux, P., Krawiec, L., Pisarev, M.A., Propato,

3

F.V.E., Juvenal, G., Chester, H.A., Vassart, G., 1992. Identification of a mutation in the coding

4

sequence of the human thyroid peroxidase gene causing congenital goiter. J. Clin. Invest. 90,

5

1200–1204.

6

Abdul-Hassan, I.A., AL-Ramahi, I.J., AL-Faisal, A.H.M., 2013. Detection of T.G. and TO genes

7

compound mutations associated with thyroid carcinoma, toxic goiter and hypothyroidism in Iraqi

8

patients. J. Med. Sci. 13, 676-683.

9 10 11 12

Abu-Khudir, R., Larrivée-Vanier, S., Wasserman, J.D., Deladoëy, J., 2017. Disorders of thyroid morphogenesis. Best Pract. Res. Clin. Endocrinol. Metab. 31, 143-159. Alper, S.L., Sharma, A.K., 2013. The SLC26 gene family of anion transporters and channels. Mol. Aspects Med. 34, 494–515.

13

Altmann, K., Hermanns, P., Mühlenberg, R., Fricke-Otto, S., Wentzell, R., Pohlenz, J., 2013.

14

Congenital goitrous primary hypothyroidism in two German families caused by novel thyroid

15

peroxidase (TPO) gene mutations. Exp. Clin. Endocrinol. Diabetes 121, 343-346.

16

Alzahrani, A.S., Baitei, E.Y., Zou, M., Shi, Y., 2006. Metastatic thyroid follicular carcinoma arising

17

from congenital goiter due to a novel splice donor site mutation in the thyroglobulin gene. J.

18

Clin. Endocrinol. Metab. 91, 740-746.

19

Amendola, E., De Luca, P., Macchia, P.E., Terracciano, D., Rosica, A., Chiappetta, G., Kimura

20

S, Mansouri A, Affuso A, Arra C, Macchia V, Di Lauro R, De Felice M., 2005. A mouse model

21

demonstrates a multigenic origin of congenital hypothyroidism. Endocrinology 146, 5038-5047.

22

Amendola, E., Sanges, R., Galvan, A., Dathan, N., Manenti, G., Ferrandino, G., Alvino, F.M., Di

23

Palma, T., Scarfò, M., Zannini, M., Dragani, T.A., De Felice, M., Di Lauro, R., 2010. A locus on

24

mouse chromosome 2 is involved in susceptibility to congenital hypothyroidism and contains an

25

essential gene expressed in thyroid. Endocrinology 151, 1948-1958.

38

1 2

Aravind, L., Koonin, E.V., 2000. The STAS domain-a link between anion transporters and antisigma-factor antagonists. Curr. Biol. 10, R53–R55.

3

Aycan, Z., Cangul, H., Muzza, M., Bas, V.N., Fugazzola, L., Chatterjee, V.K., Persani, L.,

4

Schoenmakers, N., 2017. Digenic DUOX1 and DUOX2 mutations in cases with congenital

5

hypothyroidism. J. Clin. Endocrinol. Metabol. 102, 3085–3090.

6 7 8 9

Azroyan, A., Laghmani, K., Crambert, G., Mordasini, D., Doucet, A., Edwards, A., 2011. Regulation of pendrin by pH: dependence on glycosylation. Biochem. J. 434, 61-72. Baker, J. R., Arscott, P., Johnsoon, J., 1994. An analysis of the structure and antigenicity of different forms of human thyroid peroxidase. Thyroid 4, 173–178.

10

Bakker, B., Bikker, H., Vulsma, T., de Randamie, J.S.E., Wiedijk, B.M., de Vijlder, J.J.M. 2000

11

Two decades of screening for congenital hypothyroidism in the Netherlands: TPO gene

12

mutations in total iodide organification defect (an update). J. Clin. Endocrinol. Metab. 85, 3708-

13

3712.

14

Baryshev, M., Sargsyan, E., Wallin, G., Lejnieks, A., Furudate, S., Hishinuma, A., Mkrtchian, S.,

15

2004. Unfolded protein response is involved in the pathology of human congenital hypothyroid

16

goiter and rat non-goitrous congenital hypothyroidism. J. Mol. Endocrinol. 32, 903–920.

17 18

Bedard, K., Szabo, E., Michalak, M., Opas, M., 2005. Cellular functions of endoplasmic reticulum chaperones calreticulin, calnexin, and ERp57. Int. Rev. Cytol. 245, 91-121.

19

Belforte, F.S., Miras, M.B., Olcese, M.C., Sobrero, G., Testa, G., Muñoz, L., Gruñeiro-Papendieck,

20

L., Chiesa, A., González-Sarmiento, R., Targovnik, H.M., Rivolta, C.M., 2012. Congenital

21

goitrous hypothyroidism: Mutation analysis in the thyroid peroxidase gene. Clin. Endocrinol.

22

(Oxf) 76, 568-576.

23 24

Belforte, F.S., Olcese, M.C., Siffo, S., Papendieck, P., Enacan, R., Gruñeiro-Papendieck, L., Chiesa, A.,

Targovnik,

H.M.,

Rivolta,

C.M.,

2016a.

Biallelic

stop

codon

mutations

39

1

(p.F353Pfs*36/p.Y425X) in DUOX2 gene associated with transient congenital hypothyroidism:

2

Report of a family and literature review. Annals Thyroid Res. 2, 69-78.

3

Belforte, F.S., Cintia, E., Citterio, C.E., Testa, G., Olcese, M.C., Sobrero, G., Miras, M.B.,

4

Targovnik, H.M., Rivolta, C.M., 2016b. Compound heterozygous DUOX2 gene mutations

5

(c.2335-1G>C/c.3264_3267delCAGC)

6

Characterization of complex cryptic splice sites by minigene analysis. Mol. Cell. Endocrinol.

7

419, 172-184.

associated

with

congenital

hypothyroidism.

8

Bidart, J.M., Mian, C., Lazar, V., Russo, D., Filetti, S., Caillou, B., Schlumberger, M., 2000.

9

Expression of pendrin and the Pendred syndrome (PDS) gene in human thyroid tissues. J. Clin.

10

Endocrinol. Metab. 85, 2028-2033.

11

Bikker, H., Vulsma, T., Baas, F., Vijlder, J.J.M., 1995. Identification of five novel inactivating

12

mutations in the human thyroid peroxidase gene by denaturating gradient gel electrophoresis.

13

Hum. Genet. 6, 9-16.

14 15 16 17 18 19

Bizhanova, A., Kopp, P., 2009. Minireview: The sodium-iodide symporter NIS and pendrin in iodide homeostasis of the thyroid. Endocrinology 150,1084-1090. Bizhanova, A., Kopp, P., 2010. Genetics and phenomics of Pendred syndrome. Mol. Cell. Endocrinol. 322, 83-90. Bizhanova, A., Kopp, P., 2011. Controversies concerning the role of pendrin as an apical iodide transporter in thyroid follicular cells. Cell. Physiol. Biochem. 28, 485-490.

20

Bӧrgel, K., Pohlenz, J., Holzgreve, W., Bramswig, J.H., 2005. Intrauterine therapy of goitrous

21

hypothyroidism in a boy with a new compound heterozygous mutation (Y453D and C800R) in

22

the thyroid peroxidase gene. A long-term follow-up. Am. J. Obstet. Gynecol. 193, 857-858.

23 24

Campbell,

C., Cucci,

R.A., Prasad,

S., Green,

G.E., Edeal,

J.B., Galer,

C.E., Karniski,

L.P., Sheffield, V.C., Smith, R.J.H., 2001. Pendred Syndrome, DFNB4, and PDS/SLC26A4

40

1

Identification of eight novel mutations and possible genotype-phenotype correlations. Hum.

2

Mutat. 17: 403-411.

3

Cangul, H., Aycan, Z., Olivera-Nappa, A., Saglam, H., Schoenmakers, N.A., Boelaert, K.,

4

Cetinkaya, S., Tarim, O., Bober, E., Darendeliler, F., Bas, V., Demir, K., Aydin, B.K., Kendall,

5

M., Cole, T., Hogler, W., Chatterjee, V.K.K., Barrett, T.G., Maher, E.R., 2013. Thyroid

6

dyshormonogenesis is mainly caused by TPO mutations in consanguineous community. Clin.

7

Endocrinol. (Oxf) 79, 275-281.

8

Cangül, H., Demir, K., Babayiğit, H.Ö., Abacı, A., Böber, E., 2015. The missense alteration A5T of

9

the thyroid peroxidase gene is pathogenic and associated with mild congenital hypothyroidism. J.

10

Clin. Res. Pediatr. Endocrinol. 7, 238-241.

11

Cangul, H., Liao, X.-H., Schoenmakers, E., Kero, J., Barone, S., Srichomkwun, P., Iwayama, H.,

12

Serra, E.G., Saglam, H., Eren, E., Tarim, O., Nicholas, A.K., Zvetkova, I., Anderson, C.A.,

13

Frankl, F.E.K., Boelaert, K., Ojaniemi, M., Jääskeläinen, J., Patyra, K., Löf, C., Williams, E.D.,

14

UK10K Consortium, Soleimani, M., Barrett, T., Maher, E.R., Chatterjee, V.K., Refetoff, S.,

15

Schoenmakers, N., 2018. Homozygous loss-of-function mutations in SLC26A7 cause goitrous

16

congenital hypothyroidism. JCI insight 3, e99631.

17

Caputo, M., Rivolta, C.M., Esperante, S.A., Gruñeiro-Papendieck, L., Chiesa, A., Pellizas, C.G.,

18

González-Sarmiento, R., Targovnik, H.M., 2007. Congenital hypothyroidism with goitre caused

19

by new mutations in the thyroglobulin gene. Clin. Endocrinol. (Oxf) 67, 351-357.

20

Carré, A., Louzada, R. A. N., Fortunato, R. S., Ameziane-El-Hassani, R., Morand, S., Ogryzko, V.,

21

de Carvalho, D.P., Grasberger, H., Leto, T.L., Dupuy, C., 2015. When an intramolecular

22

disulfide bridge governs the interaction of DUOX2 with its partner DUOXA2. Antioxid. Redox

23

Signal. 23, 724–733.

24

Caron, P., Moya, C.M., Malet, D., Gutnisky, V.J., Chabardes, B., Rivolta, C.M., Targovnik, H.M.,

25

2003. Compound heterozygous mutations in the thyroglobulin gene (1143delC and

41

1

6725G>A[R2223H]) resulting in fetal goitrous hypothyroidism. J. Clin. Endocrinol. Metab. 88,

2

3546-3553.

3

Carrasco, N., 2013. Thyroid hormones synthesis: thyroid iodide transport. Werner and Ingbar’s The

4

Thyroide: A Fundamental and Clinical Text, Tenth Edition, 32-47. Editors: L.E. Braverman,

5

D.S. Cooper, Lippincott Williams & Wilkins, Philadelphia, USA.

6

Cauvi, D., Venot, N., Nlend, M.-C., Chabaud, O. M. 2003. Thyrotropin and iodide regulate sulfate

7

concentration in thyroid cells. Relationship to thyroglobulin sulfation. Can. J. Physiol.

8

Pharmacol. 81, 1131–1138.

9

Chen, X., Kong, X., Zhu, J., Zhang, T., Li, Y., Ding, G., Wang, H., 2018. Mutational spectrum

10

analysis of seven genes associated with thyroid dyshormonogenesis. Int. J. Endocrinol.

11

2018:8986475

12 13

Cheng, S.Y., Leonard, J.L., Davis, P.J., 2010. Molecular aspects of thyroid hormone actions. Endocr. Rev. 31, 139–170.

14

Citterio, C.E., Machiavelli, G.A., Miras, M.B., Gruñeiro-Papendieck, L., Lachlan, K., Sobrero, G.,

15

Chiesa, A., Walker, J., Muñoz, L., Testa, G., Belforte, F.S., Gonzalez-Sarmiento, R., Rivolta,

16

C.M., Targovnik, H.M., 2013. New insights into thyroglobulin gene: molecular analysis of seven

17

novel mutations associated with goiter and hypothyroidism. Mol. Cell. Endocrinol. 365, 277-

18

291.

19

Citterio, C.E., Morales, C.M., Bouhours-Nouet, N., Machiavelli, G.A., Bueno, E., Gatelais, F.,

20

Coutant, R., Gonzalez-Sarmiento, R., Rivolta, C.M., Targovnik, H.M., 2015. Novel compound

21

heterozygous thyroglobulin mutations c.745+1G>A/c.7036+2T>A associated with congenital

22

goiter and hypothyroidism in a Vietnamese family. Identification of a new cryptic 5′ splice site

23

in the exon 6. Mol. Cell. Endocrinol. 404, 102-112.

24

Citterio, C.E., Veluswamy, B., Morgan, S.J., Galton, V.A., Banga, J.P., Atkins, S., Morishita, Y.,

25

Neumann, S., Latif, R., Gershengorn, M.C., Smith, T.J., Arvan, P., 2017. De novo

42

1

triiodothyronine formation from thyrocytes activated by thyroid stimulating hormone. J. Biol.

2

Chem. 292, 15434-15444.

3 4

Citterio, C.E., Targovnik, H.M., Arvan P., 2019. The role of thyroglobulin in thyroid hormonogenesis. Nat. Rev. Endocrinol. 15, 323-338

5

Consiglio, E., Acquaviva AM, Formisano S, Liguoro D, Gallo A, Vittorio T, Santisteban P, De

6

Luca M, Shifrin S, Yeh HJC, Kohn L., 1987. Characterization of phosphate residues on

7

thyroglobulin. J. Biol. Chem. 262, 10304–10314.

8

Cortinovis, F., Zamproni, I., Persani, L., Vigone, M.C., Mora, S., Weber, G., Chiumello, G., 2008.

9

Prevalence of DUOX2 mutations among children affected by congenital hypothyroidism and

10 11 12

dyshormonogenesis. Horm. Res. 70 (suppl 1), 40. Corvilain, B., Van Sande, J., Dumont, J.E., Vassart, G. 2001. Somatic and germline mutations of the TSH receptor and thyroid diseases. Clin. Endocrinol. (Oxf), 55, 143-158.

13

Coyle, B., Reardon, W., Herbrick, J.-A., Tsui, L.-C., Gausden, E., Lee, J., Coffey, R., Grueters,

14

A., Grossman, A., Phelps, P.D., Luxon, L., Kendall-Taylor, P., Scherer, S.W., Trembath, R.C.,

15

1998. Molecular analysis of the PDS gene in Pendred syndrome (sensorineural hearing loss and

16

goitre). Hum. Mol. Genet. 7, 1105-1112.

17

Czarnocka, B, Ruf, J., Ferrand, M., Carayon, P., Lissitzky, S. 1985 Purification of the human

18

thyroid peroxidase and its identification as the microsomal antigen involved in autoimmune

19

thyroid diseases. FEBS Lett. 190, 147-152.

20

Darrouzet, E., Graslin, F., Marcellin, D., Tcheremisinova, I., Marchetti, C., Salleron, L., Pognonec,

21

P., Pourcher, T., 2016. A systematic evaluation of sorting motifs in the sodium-iodide symporter

22

(NIS). Biochem. J. 473, 919–928

23

De Deken, X., Wang, D., Many, M.C., Costagliola, S., Libert, F., Vassart, G., Dumont, J.E., Miot,

24

F., 2000. Cloning of two human thyroid cDNAs encoding new members of the NADPH oxidase

25

family. J. Biol. Chem. 275, 23227-23233.

43

1 2 3 4 5 6 7 8

De Deken, X., Wang, D., Dumont, J.E., Miot, F., 2002. Characterization of ThOX proteins as components of the thyroid H(2)O(2)-generating system. Exp. Cell Res. 15, 187-196. De Deken, X., Miot, F., 2019. DUOX defects and their roles in congenital hypothyroidism. Methods Mol Biol 1982, 667-693. De Felice, M., Di Lauro, R. 2004 Thyroid development and its disorders: genetics and molecular mechanisms. Endocr. Rev. 25, 722-746. Degroot, L.J., Niepomniszcze, H., 1977. Biosynthesis of thyroid hormone: basic and clinical aspects. Metabolism 26, 665-718.

9

De la Vieja, A., Ginter, C.S., Carrasco, N., 2005. Molecular analysis of a congenital iodide transport

10

defect: G543E impairs maturation and trafficking of the Na+/I− symporter. Mol. Endocrinol. 19,

11

2847-2858.

12

De Marco, G., Agretti, P., Montanelli, L., Di Cosmo, C., Bagattini, B., De Servi, M., Ferrarini,

13

E., Dimida, A., Freitas Ferreira, A.C., Molinaro, A., Ceccarelli, C., Brozzi, F., Pinchera,

14

A., Vitti, P., Tonacchera, M., 2011. Identification and functional analysis of novel dual oxidase 2

15

(DUOX2) mutations in children with congenital or subclinical hypothyroidism. J. Clin.

16

Endocrinol. Metab. 96, 1335–1339.

17

de Moraes, V. C.S., Bernardinelli, E., Zocal, N., Fernandez, J.A., Nofziger, C., Castilho, A.M.,

18

Sartorato, E.L., Paulmichl, M., Dossena, S. 2016. Reduction of cellular expression levels is a

19

common feature of functionally affected pendrin (SLC26A4) protein variants. Mol. Med. 22, 41-

20

53.

21

Di Cosmo, C., Liao, X.H., Dumitrescu, A.M., Philp, N.J., Weiss, R.E., Refetoff, S., 2010. Mice

22

deficient in MCT8 reveal a mechanism regulating thyroid hormone secretion. J. Clin.

23

Invest. 120, 3377-3388

44

1

Di Jeso, B., Ulianich, L., Pacifico, F., Leonardi, A., Vito, P., Consiglio, E., Formisano, S., Arvan,

2

P., 2003. The folding of thyroglobulin in the calnexin/calreticulin pathway and its alteration by a

3

loss of Ca2_ from the endoplasmic reticulum. Biochem. J. 370, 449–458.

4

Di Jeso, B., Park, Y.-N., Ulianich, L., Treglia, A.S., Urbanas, M.L., High, S., Arvan, P., 2005.

5

Mixed-disulfide folding intermediates between thyroglobulin and endoplasmic reticulum

6

resident oxidoreductases ERp57 and protein disulfide isomerase. Mol. Cell. Biol. 25, 9793-9805.

7

Di Jeso, B., Arvan, P., 2016. Thyroglobulin from molecular and cellular biology to clinical

8

endocrinology. Endocr. Rev. 37, 2-36.

9

Dimitri, P., Habeb, A.M., Garbuz, F., Millward, A., Wallis, S., Moussa, K., Akcay, T., Taha,

10

D., Hogue, J., Slavotinek, A., Wales, J.K.H., Shetty, A., Hawkes, D., Hattersley, A.T., Ellard,

11

S., De Franco, E., 2015. Expanding the clinical spectrum associated with GLIS3 mutations. J.

12

Clin. Endocrinol. Metab. 100, E1362-E1369.

13

Dohan, O., De la Vieja, A., Paroder, V., Riedel, C., Artani, M., Reed, M., Ginter, C.S., Carrasco,

14

N., 2003. The sodium/iodide symporter (NIS): characterization, regulation, and medical

15

significance. Endocr. Rev. 24, 48–77.

16

Dorwart, M. R., Shcheynikov, N., Baker, J.M.R., Forman-Kay, J.D., Muallem, S., Thomas P.J. 2008.

17

Congenital chloride-losing diarrhea causing mutations in the STAS domain result in misfolding

18

and mistrafficking of SLC26A3. J. Biol. Chem. 283, 8711–8722.

19

Dossena, S., Rodighiero, S., Vezzoli, V., Nofziger, C., Salvioni, E., Boccazzi, M., Grabmayer,

20

E., Bottà, G., Meyer, G., Fugazzola, L., Beck-Peccoz, P., Paulmichl, M., 2009. Functional

21

characterization of wild-type and mutated pendrin (SLC26A4), the anion transporter involved in

22

Pendred síndrome. J. Mol Endocrinol 43, 93-103.

23 24

Dossena, S., Nofziger, C., Tamma, G., Bernardinelli, E., Vanoni, S., Nowak, C., Grabmayer, E., Kössler, S., Stephan, S., Patsch, W., Paulmichl M., 2011.

Molecular and Functional

45

1

Characterization of Human Pendrin and its Allelic Variants. Cell. Physiol. Biochem. 28, 451-

2

466

3 4

Dunn, A.D., Corsi, C.M., Myers, H.E., Dunn, J.T., 1998. Tyrosine 130 is an important outer ring donor for thyroxine formation in thyroglobulin. J. Biol. Chem. 273, 25223-25229.

5

Dupuy, C., Ohayon, R., Valent, A., Noël-Hudson, M.-S., Dème, D., Virion, A., 1999. Purification

6

of a novel flavoprotein involved in the thyroid NADPH oxidase. Cloning of the porcine and

7

human cdnas. J. Biol. Chem. 274, 37265-37269.

8

El-Desouki, M., Al-Jurayyan, N., Al-Nuaim, A., Al-Herbish, A., Abo-Bakr, A., Al-Mazrou, Y., Al-

9

Swailem, A., 1995. Thyroid scintigraphy and perchlorate discharge test in the diagnosis of

10

congenital hypothyroidism. Eur. J. Nucl. Med. 22, 1005-1008.

11

El Hassani, R.A., Benfares, N., Caillou, B., Talbot, M., Sabourin, J.C., Belotte, V., Morand,

12

S., Gnidehou, S., Agnandji, D., Ohayon, R., Kaniewski, J., Noël-Hudson, M.S., Bidart, J.M-

13

, Schlumberger, M., Virion, A., Dupuy, C., 2005. Dual oxidase2 is expressed all along the

14

digestive tract. Am. J. Physiol. Gastrointest. Liver Physiol. 288, G933-G942.

15

Fayadat, L., Niccoli-Sire, P., Lanet, J., Franc, J.L., 1998. Human thyroperoxidase is largely retained

16

and rapidly degraded in the endoplasmic reticulum. Its N-glycans are required for folding and

17

intracelular trafficking. Endocrinology 139, 4277–4285.

18

Fayadat, L., Siffroi-Fernandez, S., Lanet, J., Franc, J.-L., 2000. Calnexin and calreticulin binding to

19

human thyroperoxidase is required for its first folding step(s) but is not sufficient to promote

20

efficient cell Surface expression. Endocrinology 141, 959–966.

21 22

Fogelfeld, L., Schneider, A.B., 1990. Inhibition of chondroitin sulfate incorporation into human thyroglobulin by p-nitrophenyl-b-D-xylopyranoside. Endocrinology 126, 1064-1069.

23

Forteza, R., Salathe, M., Miot, F., Forteza, R., Conner, G.E., 2005. Regulated hydrogen peroxide

24

production by Duox in human airway epithelial cells. Am. J. Respir. Cell Mol. Biol. 32, 462-469.

46

1

Fortunato, R.S., Lima de Souza, E.C., Ameziane-el Hassani, R., Boufraqech, M., Weyemi, U.,

2

Talbot, M., Lagente-Chevallier, O., de Carvalho, D.P., Bidart, J.M., Schlumberger, M., Dupuy,

3

C., 2010. Functional consequences of dual oxidase-thyroperoxidase interaction at the plasma

4

membrane. J. Clin. Endocrinol. Metab. 95, 5403-5411.

5

Fu, C., Luo, S., Zhang, S., Wang, J., Zheng, H., Yang, Q., Xie, B., Hu, X., Fan, X., Luo, J., Chen,

6

R., Su, J., Shen, Y., Gu, X., Chen, S., 2016a. Next-generation sequencing analysis of DUOX2 in

7

192 Chinese subclinical congenital hypothyroidism (SCH) and CH patients. Clin. Chim. Acta.

8

458, 30-34.

9

Fu, C., Xie, B., Zhang, S., Wang, J., Luo, S., Zheng, H., Su, J., Hu, X., Chen, R., Fan, X., Luo, J.,

10

Gu, X., Chen, S., 2016b. Mutation screening of the TPO gene in a cohort of 192 Chinese patients

11

with congenital hypothyroidism. BMJ Open 6: e010719.

12

Fu, C., Wang, J., Luo, S., Yang, Q., Li, Q., Zheng, H., Hu, X., Su, J., Zhang, S., Chen, R., Luo, J.,

13

Zhang, Y., Shen, Y., Wei, H., Meng, D., Gui, B., Zeng, Z., Fan, X., Chen, S, 2006c. Next-

14

generation sequencing analysis of TSHR in 384 Chinese subclinical congenital hypothyroidism

15

(CH) and CH patients. Clin. Chim. Acta 462, 127-132.

16

Fugazzola, L., Mannavola, D., Vigone, M.C., Cirello, V., Weber, G., Beck-Peccoz, P., Persani, L.,

17

2005. Total iodide organification defect: clinical and molecular characterization of an Italian

18

family. Thyroid 15, 1085-1088.

19

Geiszt, M., Witta, J., Baffi, J., Lekstrom, K., Leto, T.L., 2003. Dual oxidases represent novel

20

hydrogen peroxide sources supporting mucosal surface host defense. FASEB J. 17, 1502-1504.

21

Gnidehou, S., Caillou, B., Talbot, M., Ohayon, R., Kaniewski, J., Noël-Hudson, M.S., Morand, S.,

22

Agnangji, D., Sezan, A., Courtin, F., Virion, A., Dupuy, C., 2004. Iodotyrosine dehalogenase 1

23

(DEHAL1) is a transmembrane protein involved in the recycling of iodide close to the

24

thyroglobulin iodination site. FASEB J. 18, 1574-1576.

47

1 2

Grasberger, H., Refetoff, S., 2006. Identification of the maturation factor for dual oxidase. Evolution of an eukaryotic operon equivalent. J. Biol. Chem. 281, 18269-18272.

3

Grasberger, H., De Deken, X., Miot, F., Pohlenz, J., Refetoff, S., 2007. Missense mutations of dual

4

oxidase 2 (DUOX2) implicated in congenital hypothyroidism have impaired trafficking in cells

5

reconstituted with DUOX2 maturation factor. Mol. Endocrinol. 21, 1408-1421.

6 7 8 9 10 11

Grasberger, H., 2010. Defects of thyroidal hydrogen peroxide generation in congenital hypothyroidism. Mol. Cell. Endocrinol. 322, 99-106. Grasberger, H., Refetoff, S., 2011. Genetic causes of congenital hypothyroidism due to dyshormonogenesis. Curr. Opin. Pediatr. 23, 421-428. Grasberger, H., Refetoff, S., 2017. Resistance to thyrotropin. Best Pract. Res. Clin. Endocrinol. Metab. 31, 183-194.

12

Gruters, A., Kohler, B., Wolf, A., Soling, A., de Vijlder, L., Krude, H., Biebermann, H., 1996.

13

Screening for mutations of the human thyroid peroxidase gene in patients with congenital

14

hypothyroidism. Exp. Clin. Endocrinol. Diabetes 104 (Suppl 4), 121-123.

15

Gutnisky, V.J., Moya. C.M., Rivolta, C.M., Domené, S., Varela, V., Toniolo, J.V., Medeiros-Neto,

16

G., Targovnik, H.M., 2004. Two

17

IVS34-1G>C and R277X / R1511X) in the thyroglobulin (TG) gene in affected individuals of a

18

brazilian kindred with congenital goiter and defective TG synthesis. J. Clin. Endocrinol. Metab.

19

89, 646-657.

20 21

distinct

compound

heterozygous constellation (R277X /

Helenius, A., Aebi, M., 2004. Roles of N‐linked glycans in the endoplasmic reticulum. Annu. Rev. Biochem. 73, 1019–1049.

22

Hermanns, P., Refetoff, S., Sriphrapradang, C., Pohlenz, J., Okamato, J., Slyper, L., Slyper, A.H.,

23

2013. A clinically euthyroid child with a large goiter due to a thyroglobulin gene defect: clinical

24

features and genetic studies. J. Pediatr. Endocrinol. Metab. 26, 119-123.

48

1

Hishinuma, A., Takamatsu, J., Ohyama, Y., Yokozawa, T., Kanno, Y., Kuma, K., Yoshida, S.,

2

Matsuura, N., Ieiri, T., 1999. Two novel cysteine substitutions (C1263R and C1995S) of

3

thyroglobulin cause a defect in intracellular transport of thyroglobulin in patients with congenital

4

goiter and the variant type of adenomatous goiter. J. Clin. Endocrinol. Metab. 84, 1438-1444.

5

Hishinuma, A., Furudate, S.-I., Oh-Ishi, M., Nagakubo, N., Namatame, T., Ieiri, T., 2000. A novel

6

missense mutation (G2320R) in thyroglobulin causes hypothyroidism in rdw rats. Endocrinology

7

141, 4050-4055.

8 9

Hishinuma, A., Fukata, S., Kakudo, K., Murata, Y., Ieiri, T., 2005. High incidence of thyroid cancer in long-standing goiters with thyroglobulin mutations. Thyroid 15, 1079-1084.

10

Hishinuma, A., Fukata, S., Nishiyama, S., Nishi, Y., Oh-Ishi, M., Murata, Y., Ohyama, Y.,

11

Matsuura, N., Kasai, K., Harada, S., Kitanaka, S., Takamatsu, J., Kiwaki, K., Ohye, H., Uruno,

12

T., Tomoda, C., Tajima, T., Kuma, K., Miyauchi, A., Ieiri, T., 2006. Haplotype analysis reveals

13

founder effects of thyroglobulin gene mutations C1058R and C1977S in Japan. J. Clin.

14

Endocrinol. Metab. 91, 3100-3104.

15

Holzer, G., Morishita, Y., Fini, J.-B., Lorin, T., Gillet, B., Hughes, S., Tohmé, M., Deléage, G.,

16

Demeneix, B., Arvan, P., Laudet, V., 2016. Thyroglobulin represents a novel molecular

17

architecture of vertebrates. J. Biol. Chem. 291, 16553–16566.

18

Hu, X., Chen, R., Fu, C., Fan, X., Wang, J., Qian, J., Yi, S., Li, C., Luo, J., Su, J., Zhang, S., Xie,

19

B., Zheng, H., Lai, Y., Chen, Y., Li, H., Gu, X., Chen, S., Shen, Y., 2016. Thyroglobulin gene

20

mutations in Chinese patients with congenital hypothyroidism. Mol. Cell. Endocrinol. 423, 60-

21

66.

22

Ieiri, T., Cochaux, P., Targovnik, H.M., Suzuki, M., Shimoda, S.-I., Perret, J., Vassart, G., 1991. A

23

3' splice site mutation in the thyroglobulin gene responsible for congenital goiter with

24

hypothyroidism. J. Clin. Invest. 88, 1901-1905.

49

1

Ishii, J., Suzuki, A., Kimura, T., Tateyama, M., Tanaka, T., Yazawa, T., Arimasu, Y., Chen, I-

2

S., Aoyama, K., Kubo, Y., Saitoh, S., Mizuno, H., Kamma, H., 2019. Congenital goitrous

3

hypothyroidism is caused by dysfunction of the iodide transporter SLC26A7. Commun. Biol. 2,

4

270.

5

Jiang, H., Wu, J., Ke, S., Hu, Y., Fei, A., Zhen, Y., Yu, J., Zhu, K., 2016. High prevalence of

6

DUOX2 gene mutations among children with congenital hypothyroidism in central China. Eur. J.

7

Med. Genet. 59, 526-531.

8

Jordans, S., Jenko-Kokalj, S., Kühl, N.M., Tedelind, S., Sendt, W., Brömme, D., Turk, D., Brix, K.,

9

2009. Monitoring compartment-specific substrate cleavage by cathepsins B, K, L, and S at

10

physiological pH and redox conditions. BMC Biochem. 10, 23.

11

Jung, J., Kim, J., Roh, S.H., Jun, I., Sampson, R.D., Gee, H.Y., Choi, J.Y., Lee, M.G., 2016. The

12

HSP70 co-chaperone DNAJC14 targets misfolded pendrin for unconventional protein secretion.

13

Nat. Commun. 7, 11386.

14 15

Kampinga, H.H., Craig, E.A., 2010. The HSP70 chaperone machinery: J proteins as drivers of functional specificity. Nat. Rev. Mol. Cell Biol. 11, 579-592.

16

Kang, H.S., Kumar, D., Liao, G., Lichti-Kaiser, K., Gerrish, K., Liao, X.H., Refetoff, S., Jothi, R.,

17

Jetten, A.M., 2017. GLIS3 is indispensable for TSH/TSHR-dependent thyroid hormone

18

biosynthesis and follicular cell proliferation. J. Clin. Invest. 127, 4326-4337.

19

Kanou, Y., Hishinuma, A., Tsunekawa, K., Seki, K., Mizuno, Y., Fujisawa, H., Imai, T., Miura, Y.,

20

Nagasaka, T., Yamada, C., Ieiri, T., Murakami, M., Murata, Y., 2007. Thyroglobulin gene

21

mutations producing defective intracellular transport of thyroglobulin are associated with

22

increased thyroidal type 2 iodothyronine deiodinase activity. J. Clin. Endocrinol. Metab. 92,

23

1451-1457.

24

Kim, P., Bole, D., Arvan, P., 1992. Transient aggregation of nascent thyroglobulin in the

25

endoplasmic reticulum: relationship to the molecular chaperone, BiP. J Cell Biol 118, 541–549.

50

1 2

Kim, P.S., Arvan, P., 1995. Calnexin and BiP act as sequential molecular chaperones during thyroglobulin folding in the endoplasmic reticulum. J. Cell Biol. 128, 29–38.

3

Kim, P.S., Arvan, P., 1998. Endocrinopathies in the family of endoplasmic reticulum (ER) storage

4

diseases: Disorders of protein trafficking and the role of ER molecular chaperones. Endocr. Rev.

5

19, 173–202

6

Kim, P.S., Hossain, S.A., Park, Y.-N., Lee, I., Yoo, S.-E., Arvan, P., 1998. A single amino acid

7

change in the acetylcholinesterase-like domain of thyroglobulin causes congenital goiter with

8

hypothyroidism in the cog/cog mouse: A model of human endoplasmic reticulum storage

9

diseases. Proc. Natl. Acad. Sci. USA 95, 9909-9913.

10

Kim, P.S., Ding, M., Menon, S., Jing, C.-G., Cheng, J.-M., Miyamoto, T., Li, B., Furudate, S.-i.,

11

Agui, T., 2000. A missense mutation G2320R in the thyroglobulin gene causes non-goitrous

12

congenital primary hypothyroidism in the WIC-rdw rat. Mol. Endocrinol. 14, 1944-1953.

13

Kimura, S., Hong, Y.-S., Kotani, T., Othaki, S. Kikkawa, F., 1989. Structure of the human thyroid

14

peroxidase gene: comparison and relationship to the human myeloperoxidase gene. Biochemistry

15

28, 4481-4489.

16

Kitanaka, S., Takeda, A., Sato, U., Miki, Y., Hishinuma, A., Ieiri, T., Igarashi, T., 2006. A novel

17

compound heterozygous mutation in the thyroglobulin gene resulting in congenital goitrous

18

hypothyroidism with high serum triiodothyronine levels. J. Hum. Genet. 51, 379-382.

19

Kizys, M.M.L., Louzada, R.A., Mitne-Neto, M., Jara, J.R., Furuzawa, G.K., de Carvalho, D.P.,

20

Dias-da-Silva, M.R., Nesi-França, S., Dupuy, C., Maciel, R.M.B., 2017. DUOX2 mutations are

21

associated with congenital hypothyroidism with ectopic thyroid gland. J. Clin. Endocrinol.

22

Metab. 102, 4060-4071.

23

Kotani, T., Umeki, K., Kawano, J., Suganuma, T., Hishinuma, A., Ieiri, T., Harada, S., 2003.

24

Partial iodide organification defect caused by a novel mutation of the thyroid peroxidase gene in

25

three siblings. Clin. Endocrinol. (Oxf) 59, 198–206.

51

1

Kotani T., Umeki K., Kawano J-i., Suganuma T., Yamamoto I., Aratake Y., Ichiba Y., Furujo M.

2

A., 2004. Novel missense mutation in the thyroid peroxidase gene, R175Q, resulting in

3

insufficient cell surface enzyme in two siblings. Clin. Pediatr. Endocrinol. 13, 37-46.

4

Kuznetsov, G., Chen, L.B., Nigam, S.K., 1994. Several endoplasmic reticulum stress proteins,

5

including ERp72, interact with thyroglobulin during its maturation. J. Biol. Chem. 269, 22990–

6

22995.

7 8 9 10 11 12

Kwak, M. J., 2018. Clinical genetics of defects in thyroid hormone synthesis Ann. Pediatr. Endocrinol. Metab. 23,169-175. Lamas, L., Anderson, P.C., Fox, J.W., Dunn, J.T., 1989. Consensus sequences for early iodination and hormonogenesis in human thyroglobulin. J. Biol. Chem. 264, 13541-13545. Lee, J., Di Jeso, B., Arvan, P., 2008. The cholinesterase-like domain of thyroglobulin functions as an intramolecular chaperone. J. Clin. Invest. 118, 2950–2958.

13

Lee, J., Wang, X., Di Jeso, B., Arvan, P., 2009. The cholinesterase-like domain, essential in

14

thyroglobulin trafficking for thyroid hormone synthesis, is required for protein dimerization. J.

15

Biol. Chem. 284, 12752-12761.

16 17 18 19 20

Lee, J., Di Jeso, B., Arvan, P., 2011. Maturation of thyroglobulin protein región I. J. Biol. Chem. 286, 33045-33052. Lee, J., Arvan, P., 2011. Repeat motif-containing regions within thyroglobulin J. Biol. Chem. 286, 26327-26333. Lee, K., Hong, T. J., Hahn, J. S., 2012. Roles of 17‐AAG‐induced molecular chaperones and

21

Rma1 E3 ubiquitin ligase in folding and degradation of Pendrin. FEBS Lett. 586, 2535-2541.

22

Levy, O., De la Vieja, A., Ginter, C.S., Riedel, C., Dai, G., Carrasco, N., 1998. N-linked

23

glycosylation of the thyroid Na+/I− symporter (NIS). Implications for its secondary structure

24

model. J. Biol. Chem. 273, 22657–63

52

1 2

Li, W., Nicola, J.P., Amzel, L.M., Carrasco, N., 2013. Asn441 plays a key role in folding and function of the Na+/I− symporter (NIS). FASEB J. 27, 3229–38

3

Libert, F., Ruell, J., Ludgate, M., Swillens, S., Alexander, N., Vassart, G., Dinsart, C., 1987.

4

Thyroperoxidase, an auto-antigen with a mosaic structure made of nuclear and mitochondrial

5

gene modules. EMBO J. 6, 4193-4196.

6

Linke, M., Jordans, S., Mach, L., Herzog, V. Brix, K. 2002. Thyroid stimulating hormone

7

upregulates secretion of cathepsin B from thyroid epithelial cells. Biol. Chem. 383, 773–784.

8

Liu, S., Han, W., Zang, Y., Zang, H., Wang, F., Jiang, P., Wei, H., Liu, X., Wang, Y., Ma, X., Ge,

9

Y., 2019. Identification of two missense mutations in DUOX1 (p.R1307Q) and DUOXA1

10

(p.R56W) that can cause congenital hypothyroidism through impairing H2O2 generation. Front.

11

Endocrinol. 10, 526.

12

Lohi, H., Kujala, M., Mäkelä, S., Lehtonen, E., Kestilä, M., Saarialho-Kere, U., Markovich, D.,

13

Kere, J., 2002. Functional characterization of three novel tissue-specific anion exchangers

14

SLC26A7, -A8, and -A9*. J. Biol. Chem. 277, 14246–14254.

15

Long, W., Lu, G., Zhou, W., Yang, Y., Zhang, B., Zhou, H., Jiang, L., Yu, B., 2018. Targeted next-

16

generation sequencing of thirteen causative genes in Chinese patients with congenital

17

hypothyroidism. Endocr. J. 65,1019-1028

18 19

Lü, Z.P., Li, G.H., Li, W.J., Liu, S.G., 2011. DUOX2 gene mutation in patients with congenital goiter with hypothyroidism. Chin. J. Pediatr. 49, 943-946.

20

Machiavelli, G.A., Caputo, M., Rivolta, C.M., Olcese, M.C., Gruñeiro-Papendieck, L., Chiesa, A.,

21

González-Sarmiento, R., Targovnik, H.M., 2010. Molecular analysis of congenital goiters with

22

hypothyroidism caused by defective thyroglobulin synthesis. Identification of a novel

23

c.7006C>T [p.R2317X] mutation and expression of minigenes containing nonsense mutations in

24

exon 7. Clin. Endocrinol. (Oxf) 72, 112-121.

53

1

Makretskaya, N., Bezlepkina, O., Kolodkina, A., Kiyaev, A., Vasilyev, E.V., Petrov, V.,

2

Kalinenkova, S., Malievsky, O., Dedov, I.I., Tiulpakov, A., 2018. High frequency

3

of mutations in 'dyshormonogenesis genes' in severe congenital hypothyroidism. Plos One 13,

4

e0204323.

5 6 7 8

Mallet, B., Lejeune, P., Baudry, N., Niccoli, P., Carayon, P., 1995. N-glycans modulate in vivo and in vitro thyroid hormone synthesis. J. Biol. Chem. 270, 29881–29888. Malthièry, Y., Lissitzky, S., 1987. Primary structure of human thyroglobulin deduced from the sequence of its 8448-base complementary DNA. Eur. J. Biochem. 165, 491-498.

9

Martín, M., Geysels, R. C., Peyret, V., Bernal Barquero, C.E., Masini-Repiso, A.M., Nicola, J.P.

10

2018a Implications of Na+/I-symporter transport to the plasma membrane for thyroid

11

hormonogenesis and radioiodide therapy. J. Endocr. Soc. 3, 222-234.

12

Martín, M., Modenutti, C. P., Peyret, V., Geysels, R. C., Darrouzet, E., Pourcher, T., Masini-

13

Repiso, A.M., Martí, M.A., Carrasco, N., Nicola, J. P., 2018b. A carboxy-terminal monoleucine-

14

based motif participates in the basolateral targeting of the Na+/I− symporter. Endocrinology 160,

15

156-168.

16

Martín, M., Modenutti, C.P., Geysels, R.C., Peyret, V., Signorino, M., Testa, G., Masini-Repiso,

17

A.M., Miras, M., Carrasco, N., Martín, M.A., Nicola, J.P., 2008c. A novel iodide transport

18

defect-causing Na1/I2 Symporter (NIS) carboxy-terminus mutant uncovers a critical tryptophan-

19

acid motif required for plasma membrane transport. In: 88th Annual Meeting of the American

20

Thyroid Association; Washington, DC. Abstract BO7.

21

Matsuo, K., Tanahashi, Y., Mukai, T., Suzuki, S., Tajima, T., Azuma, H., Fujieda, K., 2016. High

22

prevalence of DUOX2 mutations in Japanese patients with permanent congenital hypothyroidism

23

or transient hypothyroidism. J. Pediatr. Endocrinol. Metab. 29, 807-812.

24

Medeiros-Neto, G., Kim, P. S., Yoo, S. E., Vono, J., Targovnik, H. M., Camargo, R., Hossain, S.

25

A., Arvan, P., 1996. Congenital hypothyroid goiter with deficient thyroglobulin. Identification of

54

1

an endoplasmic reticulum storage disease with induction of molecular chaperones. J. Clin.

2

Invest. 98, 2338-2844.

3

Mendive, F.M., Rivolta, C.M., Moya, C.M., Vassart, G., Targovnik, H.M., 2001. Genomic

4

organization of the human thyroglobulin gene: the complete intron-exon structure. Eur. J.

5

Endocrinol. 145, 485-496.

6

Menon, S., Lee, J., Abplanalp, W.A., Yoo, S.E., Agui, T., Furudate, S., Kim, P.S., Arvan, P., 2007.

7

Oxidoreductase interactions include a role for ERp72 engagement with mutant thyroglobulin

8

from the rdw/rdw rat dwarf. J. Biol. Chem. 282, 6183-6191.

9

Mercken, L., Simons, M.J., Swillens, S., Massaer, M., Vassart, G., 1985. Primary structure of

10

bovine thyroglobulin deduced from the sequence of its 8,431-base complementary DNA. Nature

11

316, 647-651.

12

Meunier, L., Usherwood, Y.K., Chung, K.T., Hendershot, L.M., 2002. A subset of chaperones and

13

folding enzymes form multiprotein complexes in endoplasmic reticulum to bind nascent

14

proteins. Mol. Biol. Cell 13, 4456–4469.

15

Mizrachi, D., Segaloff, D. L., 2004. Intracellularly located misfolded glycoprotein hormone

16

receptors associate with different chaperone proteins than their cognate wild-type receptors. Mol.

17

Endocrinol. 18, 1768-1777.

18

Molina, F., Bouanani, M., Pau, B., Granier, C., 1996. Characterization of the type-1 repeat from

19

thyroglobulin, a cysteine-rich module found in proteins from different families. Eur. J. Biochem.

20

240, 125-133.

21

Morand, S., Agnandji, D., Noel-Hudson, M.S., Nicolas, V., Buisson, S., Macon-Lemaitre, L.,

22

Gnidehou, S., Kaniewski, J., Ohayon, R., Virion, A., Dupuy, C., 2004. Targeting of the dual

23

oxidase 2 N-terminal region to the plasma membrane. J. Biol. Chem. 279, 30244-30251.

55

1

Moreno, J.C., Bikker, H., Kempers, M.J.E., van Trotsenburg, A.S.P., Baas, F., de Vijlder, J.J.M.,

2

Vulsma, T., Ris-Stalpers, C., 2002. Inactivating mutations in the gene for thyroid oxidase 2

3

(THOX2) and congenital hypothyroidism. N. Engl. J. Med. 347, 95-102.

4 5 6 7 8 9

Moreno, J.C., Visser, T.J., 2007. New phenotypes in thyroid dyshormonogenesis: Hypothyroidism due to DUOX2 mutations. Endocr. Dev. 10, 99-117. Moreno, J.C., Visser, T.J., 2010. Genetics and phenomics of hypothyroidism and goiter due to iodotyrosine deiodinase (DEHAL1) gene mutations. Mol. Cell. Endocrinol. 322, 91-98. Mullur, R., Liu, Y.-Y., Brent, G. A., 2014. Thyroid hormone regulation of metabolism. Physiol. Rev. 94, 355–382.

10

Muresan, Z., Arvan, P., 1997. Thyroglobulin transport along the secretory pathway. Investigation of

11

the role of molecular chaperone, GRP94, in protein export from the endoplasmic reticulum. J.

12

Biol. Chem. 272, 26095–26102.

13

Muzza, M., Rabbiosi, S., Vigone, M.C., Zamproni, I., Cirello, V., Maffini, M.A., Maruca K.,

14

Schoenmakers N., Beccaria L., Gallo F., Park S.-M., Beck-Peccoz P., Persani L., Weber G.,

15

Fugazzola L., 2014. The clinical and molecular characterization of patients with

16

dyshormonogenic congenital hypothyroidism reveals specific diagnostic clues for DUOX2

17

defects. J. Clin. Endocrinol. Metab. 99, 544-553.

18 19

Muzza, M., Fugazzola, L., 2017. Disorders of H2O2 generation. Best Pract. Res. Clin. Endocrinol. Metab. 31, 225-240.

20

Narumi, S., Muroya, K., Asakura, Y., Aachi, M., Hasegawa, T., 2011. Molecular basis of thyroid

21

dyshormonogenesis: genetic screening in population-based japanese patients. J. Clin. Endocrinol.

22

Metab. 96, E1838-E1842.

23

Nascimento, A.C., Guedes, D.R., Santos, C.S., Knobel, M., Rubio, I.G.S., Medeiros-Neto, G., 2003.

24

Thyroperoxidase gene mutations in congenital goitrous hypothyroidism with total and partial

25

iodide organification defect. Thyroid 13, 1145-1151.

56

1

Nettore, I.C., Desiderio, S., De Nisco, E., Cacace, V., Albano, L., Improda, N., Ungaro, P., Salerno,

2

M., Colao, A., Macchia, P.E., 2018. High-resolution melting analysis (HRM) for mutational

3

screening of Dnajc17 gene in patients affected by thyroid dysgenesis. J. Endocrinol Invest. 41,

4

711-717.

5

Neves, S.C., Mezalira, P.R., Dias, V.M.A. Chagas, A.J., Viana, M, Targovnik, H., Knobel,

6

M., Medeiros-Neto, G., Rubio, I.G.S., 2010. Monoallelic thyroid peroxidase gene mutation in a

7

patient with congenital hypothyroidism with total iodide organification defect. Arq. Bras.

8

Endocrinol. Metabol. 54, 732–737.

9

Nicholas, A.K., Serra, E.G., Cangul, H., Alyaarubi, S., Ullah, I., Schoenmakers, E., Deeb, A.,

10

Habeb, A.M., AlMaghamsi, M., Peters, C., Nathwani, N., Aycan, Z., Saglam, H., Bober, E.,

11

Dattani, M., Shenoy, S., Murray, P.G., Babiker, A., Willemsen, R., Thankamony, A., Lyons, G.,

12

Irwin, R., Padidela, R., Tharian, K., Davies, J.H., Puthi, V., Park, S.M., Massoud, A.F., Gregory,

13

J.W., Albanese, A., Pease-Gevers, E., Martin, H., Brugger, K., Maher, E.R., Chatterjee, K.,

14

Anderson, C.A., Schoenmakers, N., 2016. Comprehensive screening of eight causatives genes in

15

congenital hypothyroidism with gland-in-situ. J. Clin. Endocrinol. Metab. 101, 4521-4531.

16

Nicola, J.P., Reyna-Neyra, A., Saenger, P., Rodriguez-Buritica, D.F., Gamez Godoy, J.D.,

17

Muzumdar, R., Amzel L.M., Carrasco, N., 2015. Sodium/iodide symporter mutant V270E causes

18

stunted growth but no cognitive deficiency. J. Clin. Endocrinol. Metab. 100, E1353-E1361.

19

Nishikawa, T., Rapoport, B., McLachlan, S.M. 1994. Exclusion of two major areas on thyroid

20

peroxidase from the immunodominant region containing the conformational epitopes recognized

21

by human autoantibodies. J. Clin. Endocrinol. Metab. 79, 1648–1654.

22

Niu, D.M., Hsu, J.H., Chong, K.W., Huang, C.H., Lu, Y.H., Kao, C.H., Yu, H.C., Lo, M.Y., Jap,

23

T.S., 2009. Six new mutations of the thyroglobulin gene discovered in taiwanese children

24

presenting with thyroid dyshormonogenesis. J. Clin. Endocrinol. Metab. 94, 5045-5052.

57

1 2 3 4

Ohye, H., Sugawara, M., 2010. Dual oxidase, hydrogen peroxide and thyroid diseases. Exp. Biol. Med. (Maywood) 235, 424-433. Palumbo, G., Gentile, F., Condorelli, G.L., Salvatore, G., 1990. The earliest site of iodination in thyroglobulin is residue number 5. J. Biol. Chem. 265, 8887-8892.

5

Pardo, V., Rubio, I.G., Knobel, M., Aguiar-Oliveira, M.H., Santos, M.M., Gomes, S.A., Oliveira,

6

C.R., Targovnik, H.M., Medeiros-Neto, G., 2008. Phenotypic variation among four family

7

members with congenital hypothyroidism caused by two distinct thyroglobulin gene mutations.

8

Thyroid 18, 783-786.

9

Pardo, V., Vono-Toniolo, J., Rubio, I.G., Knobel, M., Possato, R.F., Targovnik, H.M., Kopp, P.,

10

Medeiros-Neto, G., 2009. The p.A2215D thyroglobulin gene mutation leads to deficient

11

synthesis and secretion of the mutated protein and congenital hypothyroidism with wide

12

phenotype variation. J. Clin. Endocrinol. Metab. 94, 2938-2944.

13

Park, K.J., Park, H.K., Kim, Y.J., Lee, K.R., Park, J.H., Park, J.H., Park, H.D., Lee, S.Y., Kim,

14

J.W., 2016. DUOX2 mutations are frequently associated with congenital hypothyroidism in the

15

Korean population. Ann. Lab. Med. 36, 145-153.

16 17 18 19

Park, S.M., Chatterjee, V.K.K., 2005. Genetics of congenital hypothyroidism. J. Med. Genet. 42, 379-389. Park, Y.N., Arvan, P., 2004. The acetylcholinesterase homology region is essential for normal conformational maturation and secretion of thyroglobulin. J. Biol. Chem. 279, 17085-17089.

20

Parma, J., Christophe, D., Pohl, V., Vassart, G., 1987. Structural organization of the 5' region of the

21

thyroglobulin gene. Evidence for intron loss and "exonization" during evolution. J. Mol. Biol.

22

196, 769-779.

23

Paroder, V., Nicola, J. P., Ginter, C. S., Carrasco, N. 2013 The iodide-transport-defect-causing

24

mutation R124H: a δ-amino group at position 124 is critical for maturation and trafficking of the

25

Na+/I− symporter. J. Cell Sci. 126, 3305-3313.

58

1

Pascarella, A., Ferrandino, G., Credendino, S.C., Moccia, C., D’Angelo, F., Miranda, B.,

2

D'Ambrosio, C., Bielli, P., Spadaro, O., Ceccarelli, M., Scaloni, A., Sette, C., De Felice, M., De

3

Vita, G., Amendola, E., 2018. DNAJC17 is localized in nuclear speckles and interacts with

4

splicing machinery components. Sci. Rep. 8, 7794.

5

Patel, N., Aldahmesh, M. A., Alkuraya, H., Anazi, S., Alsharif, H., Khan, A. O., Sunker, A., Al-

6

Mohsen, S., Abboud, E.B., Nowilaty, S.R., Alowain, M., Al-Zaidan, H., Al-Saud, B., Alasmari,

7

A., Abdel-Salam, G.M., Abouelhoda, M., Abdulwahab, F.M., Ibrahim, N., Naim, E., Al-Younes,

8

B., AlMostafa, A.E., AlIssa, A., Hashem, M., Buzovetsky, O., Xiong, Y., Monies, D., Altassan,

9

N., Shaheen, R., Al-Hazzaa, S.A., Alkuraya, F.S., 2016. Expanding the clinical, allelic, and locus

10

heterogeneity of retinal dystrophies. Genet. Med. 18, 554-562.

11

Peteiro-Gonzalez, D., Lee, J., Rodriguez-Fontan, J., Castro-Piedras, I., Cameselle-Teijeiro, J.,

12

Beiras, A., Bravo, S.B., Alvarez, C.V., Hardy, D.M., Targovnik, H.M., Arvan, P., Lado-Abeal,

13

J., 2010. New insights into thyroglobulin pathophysiology revealed by the study of a family with

14

congenital Goiter. J. Clin. Endocrinol. Metab. 95, 3522-3526.

15

Petrovic, S., Ju, X., Barone, S., Seidler, U., Alper, S.L., Lohi, H., Kere, J., Soleimani, M. 2003.

16

Identification of a basolateral Cl-/HCO3- exchanger specific to gastric parietal cells. Am. J.

17

Physiol. Gastrointest. Liver Physiol. 284, G1093–G1103.

18 19

Petrovic, S. et al. 2004 SLC26A7: a basolateral Cl-/HCO3- exchanger specific to intercalated cells of the outer medullary collecting duct. Am. J. Physiol. Ren. Physiol. 286, F161–F169.

20

Pfarr, N., Korsch, E., Kaspers, S., Herbst, A., Stach, A., Zimmer, C., Pohlenz J., 2006. Congenital

21

hypothyroidism caused by new mutations in the thyroid oxidase 2 (THOX2) gene. Clin.

22

Endocrinol. (Oxf). 65, 810-815.

23

Pohlenz, J., Rosenthal, I.M., Weiss, R.E., Jhiang, S.M., Burant, C., Refetoff, S., 1998. Congenital

24

hypothyroidism due to mutations in the sodium/iodide symporter. Identification of a nonsense

25

mutation producing a downstream cryptic 3' splice site. J. Clin. Invest. 101, 1028–1035.

59

1

Pohlenz, J., Duprez, L., Weiss, R.E., Vassart, G., Refetoff, S., Costagliola, S., 2000. Failure of

2

membrane targeting causes the functional defect of two mutant sodium iodide symporters. J.

3

Clin. Endocrinol. Metab. 85, 2366–2369.

4

Portmann, L, Hamada, N., Heinrich, G., Degroot, L.J., 1985. Anti-thyroid peroxidase antibody in

5

patients with autoimmune thyroid disease: possible identity with anti-microsomal antibody. J.

6

Clin. Endocrinol. Metab. 61, 1001-1003.

7 8

Portulano, C., Paroder-Belenitsky, M., Carrasco, N., 2014. The Na+/I− symporter (NIS): mechanism and medical impact. Endocr. Rev. 35, 106–49

9

Purtell, K., Paroder-Belenitsky, M., Reyna-Neyra, A., Nicola, J.P., Koba, W., Fine, E., Carrasco,

10

N., Abbott, G.W. 2012. The KCNQ1-KCNE2 K+ channel is required for adequate thyroid I−

11

uptake. FASEB J. 26, 3252–3259.

12 13

Qi, L., Tsai, B., Arvan, P., 2017. New insights into the physiological role of endoplasmic reticulumassociated degradation. Trends Cell Biol. 27, 430-440.

14

Raef, H., Al-Rijjal, R., Al-Shehri, S., Zou, M., Al-Mana, H., Baitei, E.Y., Parhar, R.S., Al-

15

Mohanna, F.A., Shi, Y., 2010. Biallelic p.R2223H mutation in the thyroglobulin gene causes

16

thyroglobulin retention and severe hypothyroidism with subsequent development of thyroid

17

carcinoma. J. Clin. Endocrinol. Metab. 95, 1000-1006.

18

Rastogi, M.V., LaFranchi, S.H., 2010. Congenital hypothyroidism. Orphanet J. Rare Dis. 5, 17.

19

Ravera, S., Reyna-Neyra, A., Ferrandino, G., Amzel, L. M., Carrasco, N., 2017. The sodium/iodide

20

symporter (NIS): molecular physiology and preclinical and clinical applications. Annu. Rev.

21

Physiol. 79, 261-289.

22

Rigutto, S., Hoste, C., Grasberger, H., Milenkovic, M., Communi, D., Dumont, J.E., Corvilain,

23

B., Miot, F., De Deken, X., 2009. Activation of dual oxidases Duox1 and Duox2: differential

24

regulation mediated by camp-dependent protein kinase and protein kinase C-dependent

25

phosphorylation. J. Biol. Chem. 284, 6725-6734.

60

1 2

Ris-Stalpers, C., Bikker, H., 2010. Genetics and phenomics of hypothyroidism and goiter due to TPO mutations. Mol. Cell. Endocrinol. 322, 38-43.

3

Rivolta, C.M., Esperante, S.A., Gruñeiro-Papendick, L., Chiesa, A., Moya, C.M., Domené, S.,

4

Varela, V., Targovnik, H.M., 2003. Five novel inactivating mutations in the human thyroid

5

peroxidase gene responsible for congenital goiter and iodide organification defect. Hum. Genet.

6

22, 259.

7

Rodrigues, C., Jorge, P., Soares, J.P., Santos, I., Salomão, R., Madeira, M., Osório, R.V. Santos, R.,

8

2005. Mutation screening of the thyroid peroxidase gene in a cohort of 55 Portuguese patients

9

with congenital hypothyroidism. Eur. J. Endocrinol., 152, 193–198.

10

Roepke, T.K., King, E.C., Reyna-Neyra, A., Paroder, M., Purtell, K., Koba, W., Fine, E., Lerner,

11

D.J., Carrasco, N., Abbott, G.W., 2009. Kcne2 deletion uncovers its crucial role in thyroid

12

hormone biosynthesis. Nat. Med. 15, 1186–1194

13

Rotman-Pikielny, P., Hirschberg, K., Maruvada, P., Suzuki, K., Royaux, I. E., Green, E.D., Kohn,

14

L.D., Lippincott-Schwartz, J., Yen, P.M., 2002. Retention of pendrin in the endoplasmic

15

reticulum is a major mechanism for Pendred syndrome. Hum. Mol. Genet. 11, 2625-2633.

16

Royaux, I.E., Suzuki, K., Mori, A., Katoh, R., Everett, L.A., Kohn, L.D., Green, E.D., 2000.

17

Pendrin, the protein encoded by the Pendred syndrome gene (PDS), is an apical porter of iodide

18

in the thyroid and is regulated by thyroglobulin in FRTL-5 cells. Endocrinology 141, 839–845.

19

Rubio, I.G., Galrao, A.L., Pardo, V., Knobel, M., Possato, R.F., Camargo, R.R., Ferreira, M.A.,

20

Kanamura, C.T., Gomes, S.A., Medeiros-Neto, G., 2008. A molecular analysis and long-term

21

follow-up of two siblings with severe congenital hypothyroidism carrying the IVS30+1G>T

22

intronic thyroglobulin mutation. Arq. Bras. Endocrinol. Metabol. 52, 1337–1344.

23 24

Ruff, J., Carayon, P., 2006 Structural and functional aspects of thyroid peroxidase. Arch. Biochem. Biophys, 445, 269–277.

61

1 2

Samuels, H.H., Tsai, J.S., Casanova, J., 1974. Thyroid hormone action: in vitro demonstration of putative receptors in isolated nuclei and soluble nuclear extracts. Science 184, 1188–1191.

3

Santos, C.L.S., Bikker, H., Rego, K.G.M., Nascimento, A.C., Tambascia, M., de Vijlder, J.J.M.,

4

Medeiros-Neto, G., 1999. A novel mutation in the TPO gene in goitrous hypothyroid patients

5

with iodide organification defect. Clin. Endocrinol. (Oxf) 51, 165-172.

6

Santos-Silva, R., Rosário, M., Grangeia, A., Costa, C., Castro-Correia, C., Alonso, I., Leão, M.,

7

Fontoura, M., 2019. Genetic analyses in a cohort of Portuguese pediatric patients with congenital

8

hypothyroidism. J. Pediatr. Endocrinol. Metab. doi: 10.1515/jpem-2019-0047.

9

Sargsyan, E., Baryshev, M., Szekely, L., Sharipo, A., Mkrtchian, S., 2002. Identification of ERp29,

10

an endoplasmic reticulum lumenal protein, as a new member of the thyroglobulin folding

11

complex. J. Biol. Chem. 277, 17009-17015.

12

Schneider, A.B., McCurdy, A., Chang, T., Dudlak, D., Magner, J., 1988. Metabolic labeling of

13

human thyroglobulin with (35S) sulfate: incorporation into chrondroitin 6-sulfate and

14

endoglycosidase-F-susceptible carbohydrate units. Endocrinology 122, 2428-2435.

15 16

Sharma, A.K., Rigby, A.C., Alper, S.L., 2011. STAS domain structure and function. Cell. Physiol. Biochem. 28, 407–422.

17

Siffo, S., Adrover, E., Citterio, C.E., Miras, M.B., Balbi, V.A., Chiesa, A., Weill, J., Sobrero, G.,

18

González, V.G., Papendieck, P., Martinez, E.B., Gonzalez-Sarmiento, R., Rivolta, C.M.,

19

Targovnik, H.M., 2018. Molecular analysis of thyroglobulin mutations found in patients with

20

goiter and hypothyroidism. Mol. Cell. Endocrinol. 473, 1-16.

21

Siffroi‐Fernandez, S., Giraud, A., Lanet, J., Franc, J. L., 2002. Association of the thyrotropin

22

receptor with calnexin, calreticulin and BiP: Effects on the maturation of the receptor. Eur. J.

23

Biochem. 269, 4930-4937.

24 25

Silveira, J. C., Kopp, P. A., 2015. Pendrin and anoctamin as mediators of apical iodide efflux in thyroid cells. Curr. Opin. Endocrinol. Diabetes Obes. 22, 374–380.

62

1

Song, Y., Ruf. J., Lothaire, P., Dequanter, D., Andry, G., Willemse, E., Dumont, J.E., Van Sande,

2

J., De Deken, X., 2010. Association of duoxes with thyroid peroxidase and its regulation in

3

thyrocytes. J. Clin. Endocrinol. Metabol. 95, 375-382.

4

Spiro, R.G., Bhoyroo, V.D., 1988. Occurrence of sulfate in the asparagine-linked complex

5

carbohydrate units of thyroglobulin. Identification and localization of galactose 3-sulfate and N-

6

acetylglucosamine 6-sulfate residues in the human and calf proteins. J. Biol. Chem. 263, 14351–

7

14358.

8 9

Spitzweg, C., Morris, J.C., 2010. Genetics and phenomics of hypothyroidism and goiter due to NIS mutations. Mol. Cell. Endocrinol. 322, 56-63.

10

Stratford, A.L., Boelaert, K., Tannahill, L.A., Kim, D.S., Warfield, A., Eggo, M.C., Gittoes,

11

N.J.L., Young, L.S., Franklyn, J.A., McCabe, C.J., 2005. Pituitary tumor transforming gene

12

binding factor: a novel transforming gene in thyroid tumorigenesis. J. Clin. Endocrinol. Metab.

13

90, 4341–49.

14

Suban, D., Zajc, T., Renko, M., Turk, B., Turk, V., Dolenc, I., 2012. Cathepsin C and plasma

15

glutamate carboxypeptidase secreted from Fischer rat thyroid cells liberate thyroxin from the N-

16

terminus of thyroglobulin. Biochimie 94, 719–726.

17

Sun, F., Zhang, J.-X., Yang, C.-Y., Gao, G.-Q., Zhu, W.-B., Han, B., Zhang, L.-L., Wan, Y.-Y., Ye

18

X.-P, Ma Y.-R, Zhang M.-M, Yang L, Zhang Q.-Y, Liu W, Guo C.-C, Chen G, Zhao S.-X, Song

19

K.-Y, Song H.-D., 2018. The genetic characteristics of congenital hypothyroidism in China by

20

comprehensive screening of 21 candidate genes. Eur. J. Endocrinol. 178, 623-633.

21

Sun, Y., Bak, B., Schoenmakers, N., van Trotsenburg, A.S.P., Oostdijk, W., Voshol, P., Cambridge,

22

E., White, J.K., le Tissier, P., Gharavy, S.N.M., Martinez-Barbera, J.P., Stokvis-Brantsma, W.H.,

23

Vulsma, T., Kempers, M.J., Persani, L., Campi, I., Bonomi, M., Beck-Peccoz, P., Zhu, H., Davis,

24

T.M.E., Hokken-Koelega, A.C.S., Del Blanco, D.G., Rangasami, J.J., Ruivenkamp, C.A.L.,

25

Laros, J.F.J., Kriek, M., Kant, S.G., Bosch, C.A.J., Biermasz, N.R., Appelman-Dijkstra, N.M.,

63

1

Corssmit, E.P., Hovens, G.C.J., Pereira, A.M., den Dunnen, J.T., Wade, M.G., Breuning, M.H.,

2

Hennekam, R.C., Chatterjee, K., Dattani, M.T., Wit, J.M., Bernard, D.J., 2012. Loss-of-function

3

mutations in IGSF1 cause an X-linked syndrome of central hypothyroidism and testicular

4

enlargement. Nat. Genet. 44, 1375-1381.

5

Tan, M., Huang, Y., Jiang, X., Li, P., Tang, C., Jia, X., Chen, Q., Chen, W., Sheng, H., Feng,

6

Y., Wu, D., Liu, L., 2016. The prevalence, clinical, and molecular characteristics of congenital

7

hypothyroidism caused by DUOX2 mutations: a population-based cohort study in Guangzhou.

8

Horm. Metab. Res. 48, 581-588.

9 10

Tao, Y. X., Conn, P. M., 2014. Chaperoning G protein-coupled receptors: from cell biology to therapeutics. Endocr. Rev. 35, 602-647.

11

Targovnik, H., Vono, J., Billerbeck, A.E.C., Cerrone, G.E., Varela, V., Mendive, F., Wajchenberg,

12

B.L., Medeiros-Neto, G., 1995. A 138-nucleotide deletion in the thyroglobulin ribonucleic acid

13

messenger in a congenital goiter with defective thyroglobulin synthesis. J. Clin. Endocrinol.

14

Metab. 80, 3356-3360.

15

Targovnik, H.M., Rivolta, C.M., Mendive, F.M., Moya, C.M., Medeiros-Neto, G., 2001. Congenital

16

goiter with hypothyroidism caused by a 5' splice site mutation in the thyroglobulin gene. Thyroid

17

11, 685-690.

18 19 20 21

Targovnik, H.M., Esperante, S.A., Rivolta, C.M., 2010. Genetics and phenomics of hypothyroidism and goiter due to thyroglobulin mutations. Mol. Cell. Endocrinol. 322, 44-55. Targovnik, H.M., Citterio, C.E., Rivolta, C.M., 2011. Thyroglobulin gene mutations in congenital hypothyroidism. Horm. Res. Pædiatr. 75, 311-321.

22

Targovnik, H.M., Edouard, T., Varela, V., Tauber, M., Citterio, C.E., González-Sarmiento, R.,

23

Rivolta, C.M., 2012. Two Novel Mutations in the Thyroglobulin Gene as Cause of Congenital

24

Hypothyroidism. Identification a Cryptic Donor Splice Site in the Exon 19. Mol. Cell.

25

Endocrinol. 348, 313-321.

64

1

Targovnik, H.M., 2013. Thyroglobulin structure, function and biosynthesis. Werner and Ingbar’s

2

The Thyroid: A Fundamental and Clinical Text, Tenth Edition, 74-92. Editors: L.E. Braverman,

3

D.S. Cooper, Lippincott Williams & Wilkins, Philadelphia, USA.

4

Targovnik, H.M., Citterio, C.E., Siffo, S., Rivolta, C.M., 2016. Advances and perspectives in

5

genetics of congenital thyroid disorders associated with thyroglobulin gene mutations. Peertechz

6

J. Biol. Res. Dev. 1, 062-070.

7 8

Targovnik, H.M., Citterio, C.E., Rivolta, C.M., 2017. Iodide handling disorders (NIS, TPO, TG, IYD). Best Pract. Res. Clin. Endocrinol. Metab. 31, 195-212.

9

Taylor, J.P., Metcalfe, R.A., Watson, P.F., Weetman, A.P., Trembath, R.C., 2002. Mutations of the

10

PDS gene, encoding pendrin, are associated with protein mislocalization and loss of iodide

11

efflux: implications for thyroid dysfunction in Pendred syndrome. J. Clin. Endocrinol.

12

Metab. 87, 1778-1784.

13

Tsukamoto, K., Suzuki, H., Harada, D., Namba, A., Abe, S., Usami, S.-i., 2003. Distribution and

14

frequencies of PDS (SLC26A4) mutations in Pendred syndrome and nonsyndromic hearing loss

15

associated with enlarged vestibular aqueduct: a unique spectrum of mutations in Japanese. Eur. J

16

Hum. Genet. 11, 916–922.

17

Turgeon, M.-O., Silander, T.L., Doycheva, D., Liao, X.-H., Rigden, M., Ongaro, L., Zhou,

18

X., Joustra, S.D., Wit, J.M., Wade, M.G., Heuer, H., Refetoff, S., Bernard, D.J., 2017. TRH

19

action is impaired in pituitaries of male IGSF1-deficient mice. Endocrinology 158, 815-830.

20

Twyffels, L., Strickaert, A., Virreira, M., Massart, C., Van Sande, J., Wauquier, C., Beauwens, R.,

21

Dumont, J. E., Galietta, L. J., Boom, A., Kruys, V., 2014. Anoctamin-1/TMEM16A is the major

22

apical iodide channel of the thyrocyte. Am. J. Physiol. Cell Physiol. 307, C1102–C1112.

23

Umeki, K., Kotani, T., Kawano, J.-I., Suganuma, T., Yamamoto, I., Aratake, Y., Furujo, M., Ichiba,

24

Y., 2002. Two novel missense mutations in the thyroid peroxidase gene: R665W and G771R

25

result in a localization defect and cause congenital hypothyroidism. Eur. J. Endocrinol. 146, 491-

65

1

498.

2

Umeki, K., Kawano, J.-I., Yamamoto, I., Yatsuki, A, Kotani, T., 2004. Comparative analysis and

3

characterization of mutated thyroid peroxidases with disturbance expressed on the cell surface.

4

Mol. Cell. Endocrinol. 223, 77-84.

5 6 7 8

Vagin, O., Kraut, J.K., Sachs G., 2009. Role of N-glycosylation in trafficking of apical membrane proteins in epithelia. Am. J. Physiol. Renal Physiol. 296, F459–F469. van de Graaf, S.A.R., Ris-Stalpers, C., Pauws, E., Mendive, F.M., Targovnik, H.M., de Vijlder, J.J.M., 2001. Up to date with human thyroglobulin. J. Endocrinol. 170, 307–321.

9

Varela, V., Rivolta, C.M., Esperante, S.A., Gruñeiro-Papendieck, L., Chiesa, A., Targovnik, H.M.,

10

2006. Three mutations (p.Q36H, p.G418fsX482, and g.IVS19-2A>C) in the dual oxidase 2 gene

11

responsible for congenital goiter and iodide organification defect. Clin. Chem. 52, 182-191.

12 13

Vassart, G., Dumont, J. E., 1992. The thyrotropin receptor and the regulation of thyrocyte function and growth. Endocr. Rev. 13, 596-611.

14

Vigone, M.C., Fugazzola, L., Zamproni, I., Passoni, A., Di Candia, S., Chiumello, G., Persani L.,

15

Weber, G., 2005. Persistent mild hypothyroidism associated with novel sequence variants of the

16

DUOX2 gene in two siblings. Hum. Mutat. 26; 395.

17 18

Vono-Toniolo, J., Kopp, P., 2004. Thyroglobulin gene mutations and other genetic defects associated with congenital hypothyroidism. Arq. Bras. Endocrinol. Metabol. 48, 70–82

19

Wang, F., Lu, K., Yang, Z., Zhang, S., Lu, W., Zhang, L., Liu, S., Yan, S., 2014. Genotypes and

20

phenotypes of congenital goitre and hypothyroidism caused by mutations in dual oxidase 2

21

genes. Clin. Endocrinol. (Oxf) 81, 452-457.

66

1

Wang, X., Lee, J., Di Jeso, B., Treglia, A.S., Comoletti, D., Dubi, N., Taylor, P., Arvan, P., 2010.

2

Cis and trans actions of the cholinesterase-like domain within the thyroglobulin dimer. J. Biol.

3

Chem. 285, 17564-17573.

4

Watanabe, Y., Sharwood, E., Goodwin, B., Creech, M.K., Hassan, H.Y., Netea, M.G., Jaeger, M.,

5

Dumitrescu, A., Refetoff, S., Huynh, T., Weiss, R.E., 2018. A novel mutation in the TG gene

6

(G2322S) causing congenital hypothyroidism in a Sudanese family: A case report. BMC Med.

7

Genet.19, 69.

8

Watanabe, Y., Bruellman, R.J., Ebrhim, R.S., Abdullah, M.A., Dumitrescu, A.M., Refetoff, S.,

9

Weiss, R.E., 2019. Congenital Hypothyroidism due to Oligogenic Mutations in Two Sudanese

10 11 12

Families. Thyroid 29, 302-304. Wémeau, J.-L., Kopp, P., 2017. Pendred sindrome. Best Pract. Res. Clin. Endocrinol. Metab. 31, 143-159.

13

Yang, S.-X., Pollock, H.G., Rawitch, A.B., 1996. Glycosylation in human thyroglobulin: location

14

of the N-linked oligosaccharide units and comparison with bovine thyroglobulin. Arch.

15

Biochem. Biophys. 327, 61-70.

16

Yoshida, A., Taniguchi, S., Hisatome, I., Royaux, I.E., Green, E.D., Kohn, L.D., Suzuki, K., 2002.

17

Pendrin is an iodide-specific apical porter responsible for iodide efflux from thyroid cells. J.

18

Clin. Endocrinol. Metab. 87, 3356–3361.

19

Yu,

B., Long,

W., Yang,

Y., Wang,

Y., Jiang,

L., Cai,

Z., Wang,

20

Newborn screening and molecular profile of congenital hypothyroidism in

21

Front Genet. 9, 509.

H.,

2018.

chinese population.

22

Zou, M., Alzahrani, A.S., Al-Odaib, A., Alqahtani, M.A., Babiker, O., Al-Rijjal, R.A., BinEssa,

23

H.A., Kattan, W.E., Al-Enezi, A.F., Al Qarni, A., Al-Faham, M.S.A., Baitei, E.Y., Alsagheir, A.,

24

Meyer, B.F., Shi, Y., 2018. Molecular analysis of congenital hypothyroidism in Saudi Arabia:

67

1

SLC26A7 mutation is a novel defect in thyroid dyshormonogenesis. J. Clin. Endocrinol. Metab.

2

103, 1889-1898.

Table 1. Thyroglobulin missense mutations involved in the wild type cysteine residues Exon position

Nucleotide position

Amino acid position

References

Exon 5 Exon 5 Exon 5 Exon 10 Exon 14 Exon 17

c.479G>C c.548G>A c.580T>G c.2177G>A c.3229T>C c.3790T>C

p.C160S p.C183Y p.C194G p.C726Y p.C1077R p.C1264R

Exon 17 Exon 21 Exon 21 Exon 22 Exon 24 Exon 31 Exon 31 Exon 33

c.3842G>A c.4426T>C c.4478G>A c.4529G>T c.4820G>T c.5690G>A c.5986T>A

p.C1281Y p.C1476R p.C1493Y p.C1510F p.C1607F p.C1897Y p.C1904G p.C1996S

Exon 33 Exon 33 Exon 37

c.6000C>G c.6017G>A c.6461G>A

p.C2000W p.C2006Y p.C2154Y

Nicholas et al., 2016 Caputo et al., 2007 Hishinuma et al., 2006 Nicholas et al., 2016 Hishinuma et al., 2005; Hishinuma et al., 2006 Baryshev et al., 2004 ; Hishinuma et al., 1999, 2005, 2006; Kanou et al., 2007; Narumi et al., 2011 Citterio et al., 2013 Zou et al., 2018 Nicholas et al., 2016 Narumi et al., 2011 Hishinuma et al., 2006 Hishinuma et al., 2006; Kitanaka et al., 2006 Hu et al., 2016 Baryshev et al., 2004; Hishinuma et al., 1999, 2005, 2006 Citterio et al., 2013 Hishinuma et al., 2006 Hishinuma et al., 2006

The nucleotide position is designated according to TG mRNA reference sequences reported in NCBI Accession Number: NM_ 003235.5. The A of the ATG of the initiator methionine codon is denoted nucleotide +1. The amino acid positions are numbered including the 19 amino acid of the signal peptide.

Table 2. Thyroglobulin missense mutations in the ChEL domain Exon position

Nucleotide position

Amino acid position

References

Exon 38

c.6701C>A

p.A2234D

Exon 38

c.6725G>A

p.R2242H

Exon 39 Exon 40 Exon 40

c.6853A>G c.6956G>A c.7007G>A

p.N2285D p.G2319D p.R2336Q

Exon 40 Exon 41 Exon 41 Exon 41 Exon 41

c.7021G>A c.7084G>C c.7093T>C c.7121G>T c.7123G>A

p.G2341S p.A2362P p.W2365R p.G2374V p.G2375R

Exon 41 Exon 42

c.7182C>G c.7364G>A

p.I2394M p.R2455H

Exon 43 Exon 44 Exon 44

c.7411G>T c.7640T>A c.7753C>T

p.A2471S p.L2547Q p.R2585W

Exon 45

c.7847A>T

p.N2616I

Exon 47 Exon 47

c.8054G>T c.8137G>A

p.W2685L p.A2713T

Caputo et al., 2007; Citterio et al., 2013; Machiavelli et al., 2010; Pardo et al. 2008, 2009; Santos-Silva et al., 2019 Caron et al., 2003; Machiavelli et al., 2010; Raef et al., 2010; Zou et al., 2018 Makretskaya et al., 2018 Hishinuma et al., 2006 Hishinuma et al., 2006; Kitanaka et al., 2006; Siffo et al., 2018 Watanabe et al. 2018 Siffo et al., 2018 Siffo et al., 2018 Hishinuma et al., 2006 Hishinuma et al., 2005, 2006; Kanou et al.; 2007; Long et al., 2018 Sun et al., 2018 Hu et al., 2016; Long et al., 2018, Yu et al., 2018 Sun et al., 2018 Nicholas et al., 2016 Hu et al., 2016; Jiang et al., 2016; Sun et al., 2018 Hu et al., 2016; Long et al., 2018, Sun et al. 2018, Yu et al., 2018 Nicholas et al., 2016 Long et al., 2018

The nucleotide position is designated according to TG mRNA reference sequences reported in NCBI Accession Number: NM_ 003235.5. The A of the ATG of the initiator methionine codon is denoted nucleotide +1. The amino acid positions are numbered including the 19 amino acid of the signal peptide.

I-

TSH T3 T4 T T4 4 T4

TSHR Basolateral membrane

G protein

MCT8 T3 T

4

GDP

T4 T4 T4

T3 Tg

NIS IYD-1

KCNQ1-KCNE2 K+ channel

Thyroid transcription factors

T4

Na+/K+ ATPase

MIT

I-

DIT T4

Transcription

aa mRNA

T4

IYD-1

Traduction

Iodide recycled

I-

aa pool

TG

TG

Anoctamin-1

Pendrin TPO DUOX1/2

SLC26A7

Apical membrane

Folicular lumen

TG

I+

Figure 1

Iodination of tyrosines Coupling of iodotyrosines

Oxidation of

I-

IH2O2

DUOXA1/2

2130 2188 2211

1996

1893

1724

2736 2768 aa

Region IV

Region III

1511 1566 1603

1146 1211

1074

922

605 659 727

298

359

161

1 31 93

1456 1470 1487

Region II

Region I

ChEL domain Hinge

Linker

NH2

COOH

SP 1-1 1-2

1-3

1-4

1-5 1-6

1-7

1-8

1-9 1-10

TG type 1

2-1-2-3 1-11

3-a1

3-b1

3-a2

3-b2 3-a3

TG type 3

TG type 2

Spacer 3

Spacer 1

NH2

2736 2768

2130 2188 2211

1996

1893

1724

1511 1566 1603

1456 1470 1487

1146 1211

1074

922

605 659 727

298

359

161

1 31 93

Spacer 2

COOH

SP 1-1 1-2

1-3

1-4

1-5 1-6

1-7

1-8

SP

Figure 2

1 61 121 181 241 301 361 421 481 541 601 661 721 781 841 901 961 1021 1081 1141 1201 1261 1321 1381 1441 1501 1561 1621 1681 1741 1801 1861 1921 1981 2041 2101 2161 2221 2281 2341 2401 2461 2521 2581 2641 2701 2761

1-9 1-10

2-1-2-3 1-11

3-a1

3-b1

3-a2

3-b2 3-a3

T4

MALVLEIFTLLASICWVSANIFEYQVDAQPLRPCELQRETAFLKQADYVPQCAEDGSFQT VQCQNDGRSCWCVGANGSEVLGSRQPGRPVACLSFCQLQKQQILLSGYINSTDTSYLPQC QDSGDYAPVQCDVQQVQCWCVDAEGMEVYGTRQLGRPKRCPRSCEIRNRRLLHGVGDKSP PQCSAEGEFMPVQCKFVNTTDMMIFDLVHSYNRFPDAFVTFSSFQRRFPEVSGYCHCADS QGRELAETGLELLLDEIYDTIFAGLDLPSTFTETTLYRILQRRFLAVQSVISGRFRCPTK CEVERFTATSFGHPYVPSCRRNGDYQAVQCQTEGPCWCVDAQGKEMHGTRQQGEPPSCAE GQSCASERQQALSRLYFGTSGYFSQHDLFSSPEKRWASPRVARFATSCPPTIKELFVDSG LLRPMVEGQSQQFSVSENLLKEAIRAIFPSRGLARLALQFTTNPKRLQQNLFGGKFLVNV GQFNLSGALGTRGTFNFSQFFQQLGLASFLNGGRQEDLAKPLSVGLDSNSSTGTPEAAKK DGTMNKPTVGSFGFEINLQENQNALKFLASLLELPEFLLFLQHAISVPEDVARDLGDVME TVLSSQTCEQTPERLFVPSCTTEGSYEDVQCFSGECWCVNSWGKELPGSRVRGGQPRCPT DCEKQRARMQSLMGSQPAGSTLFVPACTSEGHFLPVQCFNSECYCVDAEGQAIPGTRSAI GKPKKCPTPCQLQSEQAFLRTVQALLSNSSMLPTLSDTYIPQCSTDGQWRQVQCNGPPEQ VFELYQRWEAQNKGQDLTPAKLLVKIMSYREAASGNFSLFIQSLYEAGQQDVFPVLSQYP SLQDVPLAALEGKRPQPRENILLEPYLFWQILNGQLSQYPGSYSDFSTPLAHFDLRNCWC VDEAGQELEGMRSEPSKLPTCPGSCEEAKLRVLQFIRETEEIVSASNSSRFPLGESFLVA KGIRLRNEDLGLPPLFPPREAFAEQFLRGSDYAIRLAAQSTLSFYQRRRFSPDDSAGASA LLRSGPYMPQCDAFGSWEPVQCHAGTGHCWCVDEKGGFIPGSLTARSLQIPQCPTTCEKS RTSGLLSSWKQARSQENPSPKDLFVPACLETGEYARLQASGAGTWCVDPASGEELRPGSS SSAQCPSLCNVLKSGVLSRRVSPGYVPACRAEDGGFSPVQCDQAQGSCWCVMDSGEEVPG TRVTGGQPACESPRCPLPFNASEVVGGTILCETISGPTGSAMQQCQLLCRQGSWSVFPPG PLICSLESGRWESQLPQPRACQRPQLWQTIQTQGHFQLQLPPGKMCSADYADLLQTFQVF ILDELTARGFCQIQVKTFGTLVSIPVCNNSSVQVGCLTRERLGVNVTWKSRLEDIPVASL PDLHDIERALVGKDLLGRFTDLIQSGSFQLHLDSKTFPAETIRFLQGDHFGTSPRTWFGC SEGFYQVLTSEASQDGLGCVKCPEGSYSQDEECIPCPVGFYQEQAGSLACVPCPVGRTTI SAGAFSQTHCVTDCQRNEAGLQCDQNGQYRASQKDRGSGKAFCVDGEGRRLPWWETEAPL EDSQCLMMQKFEKVPESKVIFDANAPVAVRSKVPDSEFPVMQCLTDCTEDEACSFFTVST TEPEISCDFYAWTSDNVACMTSDQKRDALGNSKATSFGSLRCQVKVRSHGQDSPAVYLKK GQGSTTTLQKRFEPTGFQNMLSGLYNPIVFSASGANLTDAHLFCLLACDRDLCCDGFVLT QVQGGAIICGLLSSPSVLLCNVKDWMDPSEAWANATCPGVTYDQESHQVILRLGDQEFIK SLTPLEGTQDTFTNFQQVYLWKDSDMGSRPESMGCRKDTVPRPASPTEAGLTTELFSPVD LNQVIVNGNQSLSSQKHWLFKHLFSAQQANLWCLSRCVQEHSFCQLAEITESASLYFTCT LYPEAQVCDDIMESNAQGCRLILPQMPKALFRKKVILEDKVKNFYTRLPFQKLMGISIRN KVPMSEKSISNGFFECERRCDADPCCTGFGFLNVSQLKGGEVTCLTLNSLGIQMCSEENG GAWRILDCGSPDIEVHTYPFGWYQKPIAQNNAPSFCPLVVLPSLTEKVSLDSWQSLALSS VVVDPSIRHFDVAHVSTAATSNFSAVRDLCLSECSQHEACLITTLQTQPGAVRCMFYADT QSCTHSLQGQNCRLLLREEATHIYRKPGISLLSYEASVPSVPISTHGRLLGRSQAIQVGT SWKQVDQFLGVPYAAPPLAERRFQAPEPLNWTGSWDASKPRASCWQPGTRTSTSPGVSED CLYLNVFIPQNVAPNASVLVFFHNTMDREESEGWPAIDGSFLAAVGNLIVVTASYRVGVF GFLSSGSGEVSGNWGLLDQVAALTWVQTHIRGFGGDPRRVSLAADRGGADVASIHLLTAR ATNSQLFRRAVLMGGSALSPAAVISHERAQQQAIALAKEVSCPMSSSQEVVSCLRQKPAN VLNDAQTKLLAVSGPFHYWGPVIDGHFLREPPARALKRSLWVEVDLLIGSSQDDGLINRA KAVKQFEESRGRTSSKTAFYQALQNSLGGEDSDARVEAAATWYYSLEHSTDDYASFSRAL ENATRDYFIICPIIDMASAWAKRARGNVFMYHAPENYGHGSLELLADVQFALGLPFYPAY EGQFSLEEKSLSLKIMQYFSHFIRSGNPNYPYEFSRKVPTFATPWPDFVPRAGGENYKEF SELLPNRQGLKKADCSFWSKYISSLKTSADGAKGGQSAESEEEELTAGSGLREDLLSLQE PGSKTYSK

T3

ChEL domain

Asn

GnT-II

NH2

NH2

α-mannosidase II

α-mannosidase I

NH2

SH

Asn

Asn

GnT-I

Complex-type

SH

SH

Asn

Asn

α-mannosidase I

NH2

Asn

α-glucosidase II

NH2

SH

SH

NH2

SH

Asn

Hybrid-type

NH2

High-mannose type

α-glucosidase II

GnT-IV

α-glucosidase I

GnT-IV

GnT-V GnT-VI

SH

Figure 3

Asn

Asn

NH2

NH2

NH2

Asn

SH

p.R665W p.D574-L575del

p.C800R

p.C838S

p.R175Q

933 aa

pH 494

dH 239

839 847 872

740

1 19

p.G771R

796

p.G533C

p.A397Pfs*76

p.S131P

NH2

COOH SP

D 238

E 399 ECD

An peroxidase

CCP EGF-Ca2+

TM

SP 1 61 121 181 241 301 361 421 481 541 601 661 721 781 841 901

MRALAVLSVTLVMACTEAFFPFISRGKELLWGKPEESRVSSVLEESKRLVDTAMYATMQR NLKKRGILSPAQLLSFSKLPEPTSGVIARAAEIMETSIQAMKRKVNLKTQQSQHPTDALS EDLLSIIANMSGCLPYMLPPKCPNTCLANKYRPITGACNNRDHPRWGASNTALARWLPPV YEDGFSQPRGWNPGFLYNGFPLPPVREVTRHVIQVSNEVVTDDDRYSDLLMAWGQYIDHD IAFTPQSTSKAAFGGGADCQMTCENQNPCFPIQLPEEARPAAGTACLPFYRSSAACGTGD QGALFGNLSTANPRQQMNGLTSFLDASTVYGSSPALERQLRNWTSAEGLLRVHARLRDSG RAYLPFVPPRAPAACAPEPGIPGETRGPCFLAGDGRASEVPSLTALHTLWLREHNRLAAA LKALNAHWSADAVYQEARKVVGALHQIITLRDYIPRILGPEAFQQYVGPYEGYDSTANPT VSNVFSTAAFRFGHATIHPLVRRLDASFQEHPDLPGLWLHQAFFSPWTLLRGGGLDPLIR GLLARPAKLQVQDQLMNEELTERLFVLSNSSTLDLASINLQRGRDHGLPGYNEWREFCGL PRLETPADLSTAIASRSVADKILDLYKHPDNIDVWLGGLAENFLPRARTGPLFACLIGKQ MKALRDGDWFWWENSHVFTDAQRRELEKHSLSRVICDNTGLTRVPMDAFQVGKFPEDFES CDSITGMNLEAWRETFPQDDKCGFPESVENGDFVHCEESGRRVLVYSCRHGYELQGREQL TCTQEGWDFQPPLCKDVNECADGAHPPCHASARCRNTKGGFQCLCADPYELGDDGRTCVD SGRLPRVTWISMSLAALLIGGFAGLTSTVICRWTRTGTKSTLPISETGGGTPELRCGKHQ AVGTSPQRAAAQDSEQESAGMEGRDTHRLPRAL

Figure 4

ICD

SP

I

Peroxidase-like domain

EF-hand calcium biding motifs

heme

II III

1548 aa

1129 1186 1224 1245

855

899

1042 1077

heme

Ca2

Ca2

NH2

Ca Ca22

819

p.C568R 602

NXS/T site 537

NXS/T site 455

NXS/T site 100

1 22

p.Q36H

p.D506N p.K530* p.R1110Q p.L1160del p.R1334W p.F966Sfs*29 p.R376W p.L1343F p.E879K NXS/T site 348 NXS/T site 382

p.C124*

COOH

FAD NADPH IV V VI VII binding binding site site

Alpha helice transmembrane domain gp91phox/NOX2-like domain

SP 1 61 121 181 241 301 361 421 481 541 601 661 721 781 841 901 961 1021 1081 1141 1201 1261 1321 1381 1441 1501

MLRARPEALMLLGALLTGSLGPSGNQDALSLPWEVQRYDGWFNNLRHHERGAVGCRLQRR VPANYADGVYQALEEPQLPNPRRLSNAATRGIAGLPSLHNRTVLGVFFGYHVLSDVVSVE TPGCPAEFLNIRIPPGDPVFDPDQRGDVVLPFQRSRWDPETGRSPSNPRDLANQVTGWLD GSAIYGSSHSWSDALRSFSGGQLASGPDPAFPRDSQNPLLMWAAPDPATGQNGPRGLYAF GAERGNREPFLQALGLLWFRYHNLWAQRLARQHPDWEDEELFQHARKRVIATYQNIAVYE WLPSFLQKTLPEYTGYRPFLDPSISPEFVVASEQFFSTMVPPGVYMRNASCHFRKVLNKG FQSSQALRVCNNYWIRENPNLNSTQEVNELLLGMASQISELEDNIVVEDLRDYWPGPGKF SRTDYVASSIQRGRDMGLPSYSQALLAFGLDIPRNWSDLNPNVDPQVLEATAALYNQDLS QLELLLGGLLESHGDPGPLFSAIVLDQFVRLRDGDRYWFENTRNGLFSKKEIEDIRNTTL RDVLVAVINIDPSALQPNVFVWHKGAPCPQPKQLTTDGLPQCAPLTVLDFFEGSSPGFAI TIIALCCLPLVSLLLSGVVAYFRGREHKKLQKKLKESVKKEAAKDGVPAMEWPGPKERSS PIIIQLLSDRCLQVLNRHLTVLRVVQLQPLQQVNLILSNNRGCRTLLLKIPKEYDLVLLF SSEEERGAFVQQLWDFCVRWALGLHVAEMSEKELFRKAVTKQQRERILEIFFRHLFAQVL DINQADAGTLPLDSSQKVREALTCELSRAEFAESLGLKPQDMFVESMFSLADKDGNGYLS FREFLDILVVFMKGSPEDKSRLMFTMYDLDENGFLSKDEFFTMMRSFIEISNNCLSKAQL AEVVESMFRESGFQDKEELTWEDFHFMLRDHDSELRFTQLCVKGGGGGGNGIRDIFKQNI SCRVSFITRTPGERSHPQGLGPPAPEAPELGGPGLKKRFGKKAAVPTPRLYTEALQEKMQ RGFLAQKLQQYKRFVENYRRHIVCVAIFSAICVGVFADRAYYYGFASPPSDIAQTTLVGI ILSRGTAASVSFMFSYILLTMCRNLITFLRETFLNRYVPFDAAVDFHRWIAMAAVVLAIL HSAGHAVNVYIFSVSPLSLLACIFPNVFVNDGSKLPQKFYWWFFQTVPGMTGVLLLLVLA IMYVFASHHFRRRSFRGFWLTHHLYILLYALLIIHGSYALIQLPTFHIYFLVPAIIYGGD KLVSLSRKKVEISVVKAELLPSGVTYLQFQRPQGFEYKSGQWVRIACLALGTTEYHPFTL TSAPHEDTLSLHIRAVGPWTTRLREIYSSPKGNGCAGYPKLYLDGPFGEGHQEWHKFEVS VLVGGGIGVTPFASILKDLVFKSSLGSQMLCKKIYFIWVTRTQRQFEWLADIIQEVEEND HQDLVSVHIYVTQLAEKFDLRTTMLYICERHFQKVLNRSLFTGLRSITHFGRPPFEPFFN SLQEVHPQVRKIGVFSCGPPGMTKNVEKACQLVNRQDRAHFMHHYENF

Figure 5

N

Disulfide bridge

.

2 O 2-

2 O2

N

N

N

N

NH2

N NH2

Apical membrane

N

H 2 O2

N

Colloid

Cytosol COOH

COOH

DUOXA2/ DUOX2

NADP+ + H +

NADPH

N

N

NH2

Disulfide bridge

N

N

N

N

NH2

N N

Golgi Apparatus

DUOXA2/ DUOX2 COOH

COOH

N

N

NH2

Disulfide bridge

N

N

N

N

NH2

N N

Endoplasmic Reticulum COOH

COOH

N

N

NH2

N

N

N

NH2

N

N

N

DUOXA2

Figure 6

COOH

DUOX2

COOH

N N Extracellular N

NH2 1

p.S509Rfs*7 11

TM1

79

TM2

86

TM3

157

163

TM4

TM5

217

TM6

241

TM7

308

340

TM8

TM9

411

417

468

522

TM10 TM11 TM12

TM13

p.G543E 37

54

111

136

182

191

260

286

368

388

438

p.V270E

444 550

intracellular p.A439_P443del

p.R124H

COOH 643

1 61 121 181 241 301 361 421 481 541 601

MEAVETGERPTFGAWDYGVFALMLLVSTGIGLWVGLARGGQRSAEDFFTGGRRLAALPVG LSLSASFMSAVQVLGVPSEAYRYGLKFLWMCLGQLLNSVLTALLFMPVFYRLGLTSTYEY LEMRFSRAVRLCGTLQYIVATMLYTGIVIYAPALILNQVTGLDIWASLLSTGIICTFYTA VGGMKAVVWTDVFQVVVMLSGFWVVLARGVMLVGGPRQVLTLAQNHSRINLMDFNPDPRS RYTFWTFVVGGTLVWLSMYGVNQAQVQRYVACRTEKQAKLALLINQVGLFLIVSSAACCG IVMFVFYTDCDPLLLGRISAPDQYMPLLVLDIFEDLPGVPGLFLACAYSGTLSTASTSIN AMAAVTVEDLIKPRLRSLAPRKLVIISKGLSLIYGSACLTVAALSSLLGGGVLQGSFTVM GVISGPLLGAFILGMFLPACNTPGVLAGLGAGLALSLWVALGATLYPPSEQTMRVLPSSA ARCVALSVNASGLLDPALLPANDSSRAPSSGMDASRPALADSFYAISYLYYGALGTLTTV LCGALISCLTGPTKRSTLAPGLLWWDLARQTASVAPKEEVAILDDNLVKGPEELPTGNKK PPGFLPTNEDRLFFLGQKELEGAGSWTPCVGHDGGRDQQETNL

Figure 7

p.L236P p.T410M

p.V138F

N

N Extracellular

107

113 154

185

234

268

318

348

408

421

466

p.G102R TM1 87

TM2 131

TM3 136

TM4 207

TM5 213

TM6 286

TM7 295

TM8 368

TM9 384

478

TM10 TM11 TM12 443

448

504

intracellular 534

p.Y556C 1

NH2

p.T193I

p.L445W p.Q446R

STAS domain COOH 780

1 61 121 181 241 301 361 421 481 541 601 661 721

730

MAAPGGRSEPPQLPEYSCSYMVSRPVYSELAFQQQHERRLQERKTLRESLAKCCSCSRKR AFGVLKTLVPILEWLPKYRVKEWLLSDVISGVSTGLVATLQGMAYALLAAVPVGYGLYSA FFPILTYFIFGTSRHISVGPFPVVSLMVGSVVLSMAPDEHFLVSSSNGTVLNTTMIDTAA RDTARVLIASALTLLVGIIQLIFGGLQIGFIVRYLADPLVGGFTTAAAFQVLVSQLKIVL NVSTKNYNGVLSIIYTLVEIFQNIGDTNLADFTAGLLTIVVCMAVKELNDRFRHKIPVPI PIEVIVTIIATAISYGANLEKNYNAGIVKSIPRGFLPPELPPVSLFSEMLAASFSIAVVA YAIAVSVGKVYATKYDYTIDGNQEFIAFGISNIFSGFFSCFVATTALSRTAVQESTGGKT QVAGIISAAIVMIAILALGKLLEPLQKSVLAAVVIANLKGMFMQLCDIPRLWRQNKIDAV IWVFTCIVSIILGLDLGLLAGLIFGLLTVVLRVQFPSWNGLGSIPSTDIYKSTKNYKNIE EPQGVKILRFSSPIFYGNVDGFKKCIKSTVGFDAIRVYNKRLKALRKIQKLIKSGQLRAT KNGIISDAVSTNNAFEPDEDIEDLEELDIPTKEIEIQVDWNSELPVKVNVPKVPIHSLVL DCGAISFLDVVGVRSLRVIVKEFQRIDVNVYFASLQDYVIEKLEQCGFFDDNIRKDTFFL TVHDAILYLQNQVKSQEGQGSILETITLIQDCKDTLELIETELTEEELDVQDEAMRTLAS

Figure 8

p.I309_E311delinsM p.R227*

N 66

71 121

TM1

TM2

46

91

TM3 101

Extracellular

N 142

TM4 162

198

223

TM5 178

TM6 243

279

TM7 259

307

TM8 327

363

TM9 343

375

427

448

TM10 TM11 TM12 395

407

468

intracellular 1

NH2

p.Q500*

495

STAS domain

p.F631Lfs*8

631 656

1 61 121 181 241 301 361 421 481 541 601

COOH

MTGAKRKKKSMLWSKMHTPQCEDIIQWCRRRLPILDWAPHYNLKENLLPDTVSGIMLAVQ QVTQGLAFAVLSSVHPVFGLYGSLFPAIIYAIFGMGHHVATGTFALTSLISANAVERIVP QNMQNLTTQSNTSVLGLSDFEMQRIHVAAAVSFLGGVIQVAMFVLQLGSATFVVTEPVIS AMTTGAATHVVTSQVKYLLGMKMPYISGPLGFFYIYAYVFENIKSVRLEALLLSLLSIVV LVLVKELNEQFKRKIKVVLPVDLVLIIAASFACYCTNMENTYGLEVVGHIPQGIPSPRAP PMNILSAVITEAFGVALVGYVASLALAQGSAKKFKYSIDDNQEFLAHGLSNIVSSFFFCI PSAAAMGRTAGLYSTGAKTQVACLISCIFVLIVIYAIGPLLYWLPMCVLASIIVVGLKGM LIQFRDLKKYWNVDKIDWGIWVSTYVFTICFAANVGLLFGVVCTIAIVIGRFPRAMTVSI KNMKEMEFKVKTEMDSETLQQVKIISINNPLVFLNAKKFYTDLMNMIQKENACNQPLDDI SKCEQNTLLNSLSNGNCNEEASQSCPNEKCYLILDCSGFTFFDYSGVSMLVEVYMDCKGR SVDVLLAHCTASLIKAMTYYGNLDSEKPIFFESVSAAISHIHSNKNLSKLSDHSEV

Figure 9

Highlights

• Congenital hypothyroidism is the most frequent endocrine disease in children. • Mutations in the candidate genes were identified and functionally characterized. • Protein folding disorders are caused by mutations in thyroid genes. • Misfolded of thyroid proteins causes retention within the endoplasmic reticulum. • The retention of misfolded proteins activates the ERAD mechanism.