metricas
covid
Buscar en
Clinics
Toda la web
Inicio Clinics The Hirschsprung's–multiple endocrine neoplasia connection
Journal Information
Vol. 67. Issue S1.
Pages 63-67 (April 2012)
Share
Share
Download PDF
More article options
Vol. 67. Issue S1.
Pages 63-67 (April 2012)
REVIEW
Open Access
The Hirschsprung's–multiple endocrine neoplasia connection
Visits
851
Sam W. MooreI,
Corresponding author
swm@sun.ac.za

Tel.: 27 21 9389280
, Monique ZaahlII
I Division of Pediatric Surgery, Faculty of Medicine, University of Stellenbosch, Tygerberg, South Africa.
II Division of Human Genetics, Faculty of Medicine, University of Stellenbosch, Tygerberg, South Africa.
This item has received

Under a Creative Commons license
Article information
Abstract
Full Text
Bibliography
Download PDF
Statistics
Special issue
This article is part of special issue:
Vol. 67. Issue S1
More info

The risk of patients with Hirschsprung's disease later developing multiple endocrine neoplasia remains a matter of concern. The multiple endocrine neoplasia 2–Hirschsprung's disease association has been shown to cosegregate in Hirschsprung's disease patients with both short- and long-segment aganglionosis, although patients with long-segment aganglionosis a to carry the greatest risk. The Hirschsprung's disease–medullary thyroid carcinoma relationship also appears to be bi-directional, and activation or suppression of the rearranged during transfection gene appeared to vary over succeeding generations within the same family. Rearranged during transfection gene variations are associated with both conditions. The cosegregation of Hirschsprung's disease and multiple endocrine neoplasia 2 is particularly interesting as it involves both “switch off” and “switch on” of the rearranged during transfection proto-oncogene in the same patient. This cosegregation mostly relates to the cysteine-rich area on RET-620 (the “Janus gene”). The mechanism whereby rearranged during transfection influences gene activation in multiple endocrine neoplasia 2 is complex, but genetic variations impair the rearranged during transfection tyrosine kinase response to tyrosine kinase activation, thus appearing to dictate downstream signaling cascade responses. Better understanding of the RET-620 relationship allows for a more cost-effective method of identifying those at risk by focusing rearranged during transfection gene testing to this specific area as a “hot spot”. The clinical awareness of possible medullary thyroid carcinoma has led to timely intervention and early treatment of this chemo- and radioresistant tumor with poor prognosis. Establishment of “risk” by genetic testing has become a classic model of molecular medicine being integrated into patient care and offering rearranged during transfection-directed prophylactic surgical management. In addition, novel approaches to treatment based on this genetic knowledge have already shown early promise in randomized clinical trials.

KEYWORDS:
Hirschsprung's disease
Genetics
Novel mutations
RET
Endothelin receptor B gene
Full Text
INTRODUCTION

Hirschsprung's disease (HSCR), or aganglionic megacolon, occurs due to the congenital absence of parasympathetic neuronal ganglia in a segment of the hindgut, resulting in malfunction of the affected segment. The condition presents as a common cause of intestinal obstruction in neonates, with an incidence of 1 in 5,000 (0.02%) live births and an overall risk to siblings of 4% (1–3). A study of HSCR indicates several clear-cut associations that are known or are suspected to be related to an increased risk of HSCR. These include its recurrence in families and the association with the major susceptibility genes points convincingly to genetic factors in its multifactorial pathogenesis. The main susceptibility gene for Hirschsprung's disease is the RET (rearranged during transfection) proto-oncogene tyrosine kinase which is situated at chromosome 10q11.2.

There are early reports of associations between HSCR and a number of associated anomalies related to neural cell development. These include neuroblastoma (1), phaeochromocytoma (1),(4),(5) and multiple endocrine neoplasia (MEN) type 2 A and B syndromes (5–7), among others (1). Reports on this relatively uncommon cosegregation of HSCR and MEN2 in the same patient (6),(8–20) exist because of the common factor of the RET gene being associated with both conditions (HSCR, MEN type 2 and medullary thyroid carcinoma (MTC). This is an extremely interesting observation, as it involves both gain and loss of function of the same gene in the same patient.

Which patients are at risk?

The MEN–HSCR association has been shown to cosegregate in HSCR patients with both short- and long-segment aganglionosis (21). It seems to be particularly associated with long-segment aganglionic segments, and Decker et al. (21) reported a long aganglionic segment (L-HSCR) in seven out of 13 patients (54%). Our experience is of total colonic aganglionosis (TCA) in the index patient in all three families identified in our series (22). It therefore appears that patients with long-segment HSCR carry the highest risk of developing MTC and should have a detailed family history taken: the presence of a long-segment HSCR should be an important selection criterion for gene testing in HSCR.

It is important to explore this concept further in families where HSCR and MTC coexist, as it will yield possible RET gene associations and insights into possible molecular reasons for the phenotypic expression. It is generally accepted that aberrant RET protein synthesis, due to inactivating genetic variations, lead to the congenital malformation of the enteric nervous system (ENS) which we call Hirschsprung's disease. RET gene activation occurs in MEN2A.

The HSCR–MTC relationship also appears to be bi-directional, and RET gene activation or suppression appeared to vary over succeeding generations within the same family (22). Butter et al. (23) reported a 50% incidence of HSCR in 20 patients undergoing a prophylactic thyroidectomy for RET-associated MTC risk (a RET C620 mutation). In one further reported case of familial MTC (24) with a C620S point mutation, the MTC developed 12 years after surgical correction of HSCR in the child plus a maternal MTC 7 years after the child's birth. In our reported series (22), MTC was detected in the parent 5 years after the birth of the affected child.

Hirschsprung's, MEN and the RET proto-oncogene

The RET proto-oncogene [10q11.22] is the major gene involved in the pathogenesis of HSCR with causative loss-of-function mutations being identified in more than 70% of cases (25). In essence, the extracellular domain RET mutations alter the protein and possibly its processing in the endoplasmic reticulum (26). As a result, RET transport to and its expression at the cell surface is decreased.

Multiple endocrine neoplasia also results from autosomally dominant, inherited, highly penetrant germline RET mutations that predispose patients to the development of tumors in cells derived from neural crest origin. Of the more than 25 RET proto-oncogene gene mutations which have been described in association with MEN type 2 syndromes, the most important are associated with the six cysteine alleles of the extracellular portion of the RET proto-oncogene in MEN2A.

RET is a vital gene which directs the migration, proliferation and survival of the enteric neural crest-derived cells of the enteric nervous system (ENS) during embryogenesis. It is also responsible for the development of the autonomic nervous system as well as for controlling kidney development and spermatogenesis, among other functions. Genetic mutation and/or variation may result in RET malfunction, which has been associated with at least four clinical conditions (HSCR, MEN type 2 syndromes (A and B) and familial medullary thyroid carcinoma (FMTC). In HSCR, hypomorphic RET alleles may cause delayed neuroblast migration and non-apoptotic cell death. This causes ENS maldevelopment, resulting in aganglionosis (HSCR) and other related conditions of the ENS.

The presence of a HSCR–MEN2 association raises the question of which HSCR patients are at greatest risk and should therefore be further investigated by RET gene analysis. Hereditary MTC can, with rare exceptions, be traced back in family trees (27) and the HSCR–MEN2 familial association may be missed because the MTC may appear later in life without the connection being made. As a result, the risk of HSCR patients later developing features of an MEN2 syndrome (with MTC) is thought to be under-evaluated. The risk of a future thyroid carcinoma remains a matter of general concern in long-term HSCR management. It is becoming clear that there is a need for screening in order to determine which patients require RET gene analysis as a means of determining risk.

The most frequently identified mutation responsible for MEN2A involves the substitution of the cysteine amino acid at position 634 (Cys634Arg or C634R) by the amino acid arginine. However, point mutations in the cysteine-rich RET area at the 620 position, along with some other RET gene areas (e.g. C609, C611 and C618) have been reported as being mostly associated with the HSCR–MEN2 phenotype. By way of contrast, in >90% of MEN2B patients, the amino acid methionine is replaced by threonine at position 918 (M918T) in the intracellular portion of the gene. There are also a number of other RET variations that are associated with thyroid cancers and familial medullary thyroid carcinoma which may give rise to an overactive abnormal RET protein.

The cosegregation of HSCR and MEN2 is particularly interesting as it involves both a “switch off” and a “switch on” of the gene in the same patient. Some remarkable family lineages have been reported which demonstrate the occurrence of the two conditions in family lineages (17),(22).

Which RET gene variations are important?

Like other receptor tyrosine kinases (RTKs), RET is made up of at least three functional areas (extracellular, transmembrane and intracellular regions) which appear to have specific roles in cellular function (28). It is interesting that, in contrast to other susceptibility areas, the RET mutations which have been identified as carrying both a risk for HSCR as well as MEN2 approximate the transmembrane domain of the gene (mostly C620R and occasionally C620S and rarely C620W) (21). Although the majority of reports connect these two conditions with point mutations in the cysteine-rich RET-area at the 620 position, other RET gene areas (e.g. C609, C611 and C618) have been reported as alternative sites. The high frequency with which the C620 RET mutation occurs in HSCR–MEN patients suggests the concept of the so-called “Janus gene” mutation in this position, which, like the Roman god of doorways, can face in both directions (i.e. activation (MEN/MTC) and inactivation (HSCR).

Our own studies (22) have shown the RET-620 cysteine variations to be present in 2.5% of HSCR families (three out of 118 families investigated), which is in keeping with other reports (21). There was, however, an uncertain prevalence of HSCR–MEN in related family members as not all could be tested.

Studies have also shown that the closer mutations are to the transmembrane domain, the higher their importance in terms of the development of cellular proliferation and tumor activation (29). A strong self-association has been shown in the transmembrane portion of the gene (RET-TM), which has been suggested as a possible explanation of the oncogenic activation due to RET–cysteine mutations in oncogenesis (29). This suggests an area that plays an important role in MTC risk evaluation and may identify a target area for novel therapeutic strategies.

How does RET gene variation result in cellular activation?

Because MEN2A relates mostly to the RET alleles at the six cysteine positions in the extracellular domain of the gene, it appears to induce a different set of downstream signaling genes from that carrying the MEN 2B mutation.

The multidocking intracellular portion of the RET gene where MEN2B RET variations are situated appears to be vital to both tyrosine kinase function and downstream signaling due to the number of signaling molecules that interact there (e.g. Shc, Src, FRS2, IRS1, Gab1/2, and Enigma) (30–32). It is therefore easier to understand how mutations within this tyrosine-kinase rich region (e.g. in 95% of MEN2B patients) alters the substrate specificity of RET tyrosine kinase and results in gene activation.

The mechanism whereby RET influences gene activation in MEN2A is more complex. Genetic variations impair the RET tyrosine kinase response to tyrosine kinase activation, thus appearing to dictate downstream signaling cascade response (33). The induction of a disulfide-linked homodimerization resulting from the cysteine-related mutations within the extracellular portion of the RET proto-oncogene promote dimer formation within the gene as covalent receptor dimers are linked by disulfide bonds between unpaired cysteines (34),(35). It is known that the RET protein targets a number of intracellular signaling cascades, including the RAS-RAF-ERK1/2, PIK3-AKT, and STAT transcription pathways (36). The activation of the MAPK and Ras/ERK molecular signaling cascades involves grb2/mSOS recruitment (37). In addition, recent studies have suggested that the mammalian target of the rapamycin (mTOR) intracellular signaling pathway is functionally activated in MTC (especially in those with germline RET mutations) (38), and that the role of the other downstream RET tyrosine kinase signaling cascades (e.g. BRAF, RAS isoforms and PI3 kinase) is less certain.

How does the RET-620 relationship work?

Oncogenic RET mutations mostly occur de novo in sporadic MTC, and it is now well accepted that RET gene variations produce aberrant proteins which may bypass the normal linking to growth factors outside the cell, most probably triggering cells to grow and divide abnormally and, in turn, promoting tumor formation. Ret activation then promotes Ret-dimer formation and activation of the tyrosine kinase outside of the normal ligand binding system.

However, little is known of the actual mechanisms whereby extracellular domain mutations (e.g. C620S) may result in C-cell proliferation (39). Current understanding of RET gene activation in MTC is that the gene is activated by alterations in autophosphorylation, or alternatively by the modification of the subcellular distribution of the active kinase or the induction of unusual interactions with other proteins (e.g. RET/PTC) (40).

It is possible that the MEN2A–HSCR connection represents a second alternative intracellular mechanism in the pathogenesis of MTC (41). The less common codons C620, C618 and C611 within exon 10 confer a weaker transforming activity in vitro than exon 11 (C634) RET mutations (42). It has been suggested that the MEN2A–HSCR connection represents a second alternative intracellular mechanism in the pathogenesis of MTC involving the endoplasmic reticulum which has been shown to decrease cell surface RET expression in HSCR (26). Changes in polarity in MEN2A–HSCR may result in the unfolding of the RET protein initially failing to achieve homeostasis via the unfolded protein response (UPR) of the endoplasmic reticulum, leading to apoptosis and HSCR. Subsequent upregulation of the UPR may then induce changes which allow gene activation, thus providing a growth advantage to tumor cells.

Oncogenic RET mutations affect protein synthesis and function, which then influence a number of downstream molecular signaling pathways (43). Studies have demonstrated that Ret influences the downstream phosphorylation of the insulin receptor substrate-2 (IRS 2) with subsequent activation of phosphatidylinositol 3 kinase (PI 3-kinase) and protein kinase B (PKB/AKT) signaling pathways (44). In addition, the induction of phosphorylation of SHC results in growth factor receptor binding protein 2 (Grb-2) binding and activation of the mitogen-activating protein (MAP) kinase pathway (45). The resulting MAPK IRS-2, PI 3-kinase and PKB/AKT activation could transduce the oncogenic Ret signal to promote gene transcription, possibly through activation of the jun/AP-1 transcription cascades (44).

Impact on treatment

This paper draws attention to the apparent genotype–phenotype correlation which exists between RET and the MEN2 syndromes (46),(47). Clinical awareness of possible MTC has led to timely intervention and early treatment of this poor prognosis chemo- and radioresistant tumor. Establishment of “risk” by genetic testing has become a classic model of molecular medicine being integrated into patient care. It is therefore fairly clear that RET gene screening is of considerable value in familial screening, and offering RET testing is considered best practice for the clinical management of patients at risk of MEN2A and MEN2B.

Gene mutation in the RET 620 position carries significant risk of developing HSCR and/or MTC. In the interests of cost-effectiveness, it may be part of a targeted investigation of high-risk areas of the gene in HSCR patients at risk, thus facilitating follow-up and future management. Genetic screening is an extremely sensitive method of identifying risk in MEN2 syndromes (±98%) and prophylactic surgical intervention may prevent the onset of the radio- and chemoresistant MTC (48).

Better understanding of the RET-620 relationship would also allow for a more cost-effective method of identifying those at risk by focusing RET gene testing to this specific area as a “hot spot”. In addition, a number of attractive targets are being identified which may be suitable for novel therapeutic signaling pathway-specific drug design (49). Novel approaches to treatment based on this genetic knowledge have already resulted in randomized clinical trials into the effectiveness of multi-kinase inhibitors such as axitinib, sorafenib, motesanib, and XL-184. These have shown early promise in selected cases which may soon change the management of thyroid cancer (50).

REFERENCES
[1]
E Passarge .
The genetics of Hirschsprungs Disease.
N Engl J Med, 276 (1967), pp. 43
[2]
M Angrist , E Kauffman , SA Slaugenhaupt , TC Matise , EG Puffenberger , SS Washington , et al.
A gene for Hirschsprung disease (megacolon) in the pericentromeric region of human chromosome 10.
Nat Genet, 4 (1993), pp. 6
[3]
S Lyonnet , A Bolino , A Pelet , L Abel , C Nihoul-Fekete , M Briard , et al.
A gene for HSCR maps to the proximal long arm of chromosome 10.
Nat Genet, 4 (1993), pp. 50
[4]
E Schocket , HA Telok .
Aganglionic megacolon, phaeochromocytoma, megaloureter and neurofibromatosis.
Am J Dis Child, 94 (1957), pp. 91
[5]
RP Bolande .
The neurocristopathies, a unifying concept of disease arising in neural crest maldevelopment.
Hum Pathol, 5 (1974), pp. 29
[6]
RA Decker , ML Peacock .
Occurrence of MEN 2a in familial Hirschsprung's disease: a new indication for genetic testing of the RET proto-oncogene.
J Pediatr Surg, 33 (1998), pp. 14
[7]
AH Khan , JG Desjardin , H Gregoire , E Seidman .
Gastrointestinal manifestations of the Sipple syndrome in children.
J Pediatr Surg, 22 (1987), pp. 23
[8]
S Dvorakova , K Dvorakova , M Malikova , R Skaba , P Vlcek , B Bendlova .
A novel Czech kindred with familial medullary thyroid carcinoma and Hirschsprung's disease.
J Pediatr Surg, 40 (2005), pp. 6
[9]
LM Mulligan , C Eng , T Attie , S Lyonnet , DJ Marsh , VJ Hyland , et al.
Diverse phenotypes associated with exon 10 mutations of the RET proto-oncogene.
Hum Mol Genet, 3 (1994), pp. 7
[10]
P Caron , T Attie , D David , J Amiel , F Brousset , P Roger , et al.
C618R mutation in exon 10 of the RET proto-oncogene in a kindred with multiple endocrine neoplasia type 2A and Hirschsprung's disease.
J Clin Endocrinol Metab, 81 (1996), pp. 3
[11]
H Peretz , R Luboshitsky , E Baron , A Biton , R Gershoni , S Usher , et al.
Cys 618 Arg mutation in the RET proto-oncogene associated with familial medullary thyroid carcinoma and maternally transmitted Hirschsprung's disease suggesting a role for imprinting.
Hum Mutat, 10 (1997), pp. 9
[12]
G Romeo , I Ceccherini , J Celli , M Priolo , N Betsos , G Bonardi , et al.
Association of multiple endocrine neoplasia type 2 and Hirschsprung disease.
J Intern Med, 243 (1998), pp. 20
[13]
RH Sijmons , RM Hofstra , FA Wijburg , TP Links , RP Zwierstra , A Vermey , et al.
Oncological implications of RET gene mutations in Hirschsprung's disease.
[14]
M Verdy , AM Weber , CC Roy , CL Morin , M Cadotte , P Brochu .
Hirschsprung's disease in a family with multiple endocrine neoplasia type 2.
J Pediatr Gastroenterol Nutr, 1 (1982), pp. 7
[15]
MJ Borst , JM VanCamp , ML Peacock , RA Decker .
Mutational analysis of multiple endocrine neoplasia type 2A associated with Hirschsprung's disease.
Surgery, 117 (1995), pp. 91
[16]
RD Blank , CA Sklar , AB Dimich , MP LaQuaglia , MF Brennan .
Clinical presentations and RET protooncogene mutations in seven multiple endocrine neoplasia type 2 kindreds.
Cancer, 78 (1996), pp. 2003
[17]
S Borrego , C Eng , B Sanchez , ME Saez , E Navarro , G Antinolo .
Molecular analysis of the ret and GDNF genes in a family with multiple endocrine neoplasia type 2A and Hirschsprung disease.
J Clin Endocrinol Metab, 83 (1998), pp. 4
[18]
K Inoue , T Shimotake , K Inoue , K Tokiwa .
Mutational analysis of the RET proto-oncogene in a kindred with multiple endocrine neoplasia type 2A and Hirschsprung's disease.
J Pediatr Surg, 34 (1999), pp. 4
[19]
B Pasini , R Rossi , MR Ambrosio , MC Zatelli , M Gullo , M Gobbo , et al.
RET mutation profile and variable clinical manifestations in a family with multiple endocrine neoplasia type 2A and Hirschsprung's disease.
[20]
SW Moore , MG Zaahl .
Multiple endocrine neoplasia syndromes, children, Hirschsprung's disease and RET.
Pediatr Surg Int, 24 (2008), pp. 30
[21]
RA Decker , ML Peacock , P Watson .
Hirschsprung disease in MEN 2A: increased spectrum of RET exon 10 genotypes and strong genotype-phenotype correlation.
Hum Mol Genet, 7 (1998), pp. 134
[22]
SW Moore , M Zaahl .
Familial associations in medullary thyroid carcinoma with Hirschsprung disease: the role of the RET-C620 “Janus” genetic variation.
J Pediatr Surg, 45 (2010), pp. 6
[23]
A Butter , J Gagne , A Al Jazaeri , MA Emran , C Deal , D St Vil .
Prophylactic thyroidectomy in pediatric carriers of multiple endocrine neoplasia type 2A or familial medullary thyroid carcinoma: mutation in C620 is associated with Hirschsprung's disease.
J Pediatr Surg, 42 (2007), pp. 6
[24]
Y Sasaki , T Shimotake , S Go , N Iwai .
Total thyroidectomy for hereditary medullary thyroid carcinoma 12 years after correction of Hirschsprung's disease.
Eur J Surg, 167 (2001), pp. 9
[25]
RM Hofstra , Y Wu , RP Stulp , P Elfferich , J Osinga , SM Maas , et al.
RET and GDNF gene scanning in Hirschsprung patients using two dual denaturing gel systems.
Hum Mutat, 15 (2005), pp. 29
[26]
S Kjaer , CF Ibanez .
Intrinsic susceptibility to misfolding of a hot-spot for Hirschsprung disease mutations in the ectodomain of RET.
Hum Mol Genet, 12 (2003), pp. 44
[27]
A Machens , H Dralle .
Familial prevalence and age of RET germline mutations: implications for screening.
Clin Endocrinol (Oxf), 69 (2008), pp. 7
[28]
I Ceccherini , R Bocciardi , Y Luo , B Pasini , R Hofstra , M Takahashi , et al.
Exon structure and flanking intronic sequences of the human RET proto-oncogene.
Biochem Biophys Res Commun, 196 (1993), pp. 95
[29]
S Kjaer , K Kurokawa , M Perrinjaquet , C Abrescia , CF Ibanez .
Self-association of the transmembrane domain of RET underlies oncogenic activation by MEN2A mutations.
Oncogene, 25 (2006), pp. 95
[30]
M Santoro , F Carlomagno , A Romano , DP Bottaro , NA Dathan , M Grieco , et al.
Activation of RET as a dominant transforming gene by germline mutations of MEN2A and MEN2B.
Science, 267 (1995), pp. 3
[31]
E Arighi , L Alberti , F Torriti , S Ghizzoni , MG Rizzetti , G Pelicci , et al.
Identification of Shc docking site on Ret tyrosine kinase.
Oncogene, 14 (1997), pp. 82
[32]
MJ Lorenzo , GD Gish , C Houghton , T Stonehouse , T Pawson , B Ponder , et al.
RET alternate splicing influences the interaction of activated RET with the SH2 and PTB domains of Shc, and the SH2 domain of Grb2.
Oncogene, 14 (1997), pp. 71
[33]
CD Andl , AK Rustgi .
No one-way street: cross-talk between e-cadherin and receptor tyrosine kinase (RTK) signaling: a mechanism to regulate RTK activity.
Cancer Biol Ther, 4 (2005), pp. 31
[34]
N Asai , T Iwashita , M Matsuyama , M Takahashi .
Mechanism of activation of the ret proto-oncogene by multiple endocrine neoplasia 2A mutations.
Mol Cell Biol, 15 (1995), pp. 9
[35]
M Santoro , M Grieco , RM Melillo , A Fusco , G Vecchio .
Molecular defects in thyroid carcinomas: role of the RET oncogene in thyroid neoplastic transformation.
Eur J Endocrinol, 133 (1995), pp. 22
[36]
H Prazeres , JP Couto , F Rodrigues , J Vinagre , J Torres , V Trovisco , et al.
In vitro transforming potential, intracellular signaling properties and sensitivity to a kinase inhibitor (Sorafenib) of RET proto-oncogene variants Glu511Lys, Ser649Leu and Arg886Trp.
Endocr Relat Cancer, 18 (2011), pp. 12
[37]
P Edery , S Lyonnet , LM Mulligan , A Pelet , E Dow , L Abel , et al.
Mutations of the RET proto-oncogene in Hirschsprung's disease.
Nature, 367 (1994), pp. 80
[38]
I Rapa , E Saggiorato , D Giachino , N Palestini , F Orlandi , M Papotti , et al.
Mammalian target of rapamycin pathway activation is associated to RET mutation status in medullary thyroid carcinoma.
J Clin Endocrinol Metab, 96 (2011), pp. 53
[39]
SW Moore .
Molecular targets in medullary thyroid carcinoma.
Gene Ther Mol Biol, 12 (2008), pp. 8
[40]
M Santoro , RM Melillo , F Carlomagno , A Fusco , G Vecchio .
Molecular mechanisms of RET activation in human cancer.
Ann N Y Acad Sci, 963 (2002), pp. 21
[41]
O Geneste , C Bidaud , G De Vita , RM Hofstra , S Tartare-Deckert , CH Buys , et al.
Two distinct mutations of the RET receptor causing Hirschsprung's disease impair the binding of signalling effectors to a multifunctional docking site.
Hum Mol Genet, 8 (1999), pp. 99
[42]
A Machens , S Hauptmann , H Dralle .
Modification of multiple endocrine neoplasia 2A phenotype by cell membrane proximity of RET mutations in exon 102.
Endocr Relat Cancer, 16 (2009), pp. 7
[43]
F Carlomagno , G De Vita , MT Berlingieri , V de Franciscis , RM Melillo , V Colantuoni , et al.
Molecular heterogeneity of RET loss of function in Hirschsprung's disease.
EMBO J, 15 (1996), pp. 25
[44]
AM Hennige , R Lammers , D Arlt , W Hoppner , V Strack , G Niederfellner , et al.
Ret oncogene signal transduction via a IRS-2/PI 3-kinase/PKB and a SHC/Grb-2 dependent pathway: possible implication for transforming activity in NIH3T3 cells.
Mol Cell Endocrinol, 167 (2000), pp. 76
[45]
G Pelicci , F Troglio , A Bodini , RM Melillo , V Pettirossi , L Coda , et al.
The neuron-specific Rai (ShcC) adaptor protein inhibits apoptosis by coupling Ret to the phosphatidylinositol 3-kinase/Akt signaling pathway.
Mol Cell Biol, 22 (2002), pp. 63
[46]
ML Brandi , RF Gagel , A Angeli , JP Bilezikian , P Beck-Peccoz , et al.
Guidelines for diagnosis and therapy of MEN type 1 and type 2.
J Clin Endocrinol Metab, 86 (2001), pp. 71
[47]
L Yip , GJ Cote , SE Shapiro , GD Ayers , CE Herzog , RV Sellin , et al.
Multiple endocrine neoplasia type 2: evaluation of the genotype-phenotype relationship 1.
Arch Surg, 138 (2003), pp. 16
[48]
M Wiench , M Kwasniewski , E Gubala , Z Wygoda , A Pawlaczek , M Oczko , et al.
[Proto-oncogene RET somatic mutations in medullary thyroid carcinoma].
Wiad Lek, (2001), pp. 415-421
[49]
X Tong , W Lan , X Zhang , H Wu , M Liu , C Cao .
Solution structure of all parallel G-quadruplex formed by the oncogene RET promoter sequence.
Nucleic Acids Res, 39 (2011), pp. 63
[50]
CA Perez , ES Santos , BA Arango , LE Raez , EE Cohen .
Novel molecular targeted therapies for refractory thyroid cancer.
Nucleic Acids Res, 39 (2011), pp. 63

No potential conflict of interest was reported.

Moore SW is an expert on the genetic connections and other aspects of Hirschsprung's disease and MEN2 syndromes. He wrote the review and was responsible for the clinical component and interpretation of results. Zaahl M is a scientist who collaborated for a long period on this project. She is the head of the Department of Genetics at Stellenbosch University and runs the laboratory where the genetic work was carried out. She personally oversaw most of the laboratory work presented in this study.

Copyright © 2012. CLINICS
Article options
es en pt

¿Es usted profesional sanitario apto para prescribir o dispensar medicamentos?

Are you a health professional able to prescribe or dispense drugs?

Você é um profissional de saúde habilitado a prescrever ou dispensar medicamentos