Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

Contents lists available at ScienceDirect

Molecular and Cellular Endocrinology


journal homepage: www.elsevier.com/locate/mce

Review

Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4)


Rajesh V. Thakker ⇑
Academic Endocrine Unit, Radcliffe Department of Medicine, University of Oxford, Oxford Centre for Diabetes, Endocrinology and Metabolism (OCDEM), Churchill Hospital,
Headington, Oxford OX3 7LJ, United Kingdom

a r t i c l e i n f o a b s t r a c t

Article history: Multiple endocrine neoplasia (MEN) is characterized by the occurrence of tumors involving two or more
Available online xxxx endocrine glands within a single patient. Four major forms of MEN, which are autosomal dominant dis-
orders, are recognized and referred to as: MEN type 1 (MEN1), due to menin mutations; MEN2 (previ-
Keywords: ously MEN2A) due to mutations of a tyrosine kinase receptor encoded by the rearranged during
Parathyroid transfection (RET) protoncogene; MEN3 (previously MEN2B) due to RET mutations; and MEN4 due to
Pancreatic islet cyclin-dependent kinase inhibitor (CDNK1B) mutations. Each MEN type is associated with the occurrence
Pituitary
of specific tumors. Thus, MEN1 is characterized by the occurrence of parathyroid, pancreatic islet and
Tumors
Neuroendocrine tumors
anterior pituitary tumors; MEN2 is characterized by the occurrence of medullary thyroid carcinoma
Histone methylation (MTC) in association with phaeochromocytoma and parathyroid tumors; MEN3 is characterized by the
occurrence of MTC and phaeochromocytoma in association with a marfanoid habitus, mucosal neuromas,
medullated corneal fibers and intestinal autonomic ganglion dysfunction, leading to megacolon; and
MEN4, which is also referred to as MENX, is characterized by the occurrence of parathyroid and anterior
pituitary tumors in possible association with tumors of the adrenals, kidneys, and reproductive organs.
This review will focus on the clinical and molecular details of the MEN1 and MEN4 syndromes. The gene
causing MEN1 is located on chromosome 11q13, and encodes a 610 amino-acid protein, menin, which
has functions in cell division, genome stability, and transcription regulation. Menin, which acts as scaffold
protein, may increase or decrease gene expression by epigenetic regulation of gene expression via histone
methylation. Thus, menin by forming a subunit of the mixed lineage leukemia (MLL) complexes that
trimethylate histone H3 at lysine 4 (H3K4), facilitates activation of transcriptional activity in target genes
such as cyclin-dependent kinase (CDK) inhibitors; and by interacting with the suppressor of variegation
3–9 homolog family protein (SUV39H1) to mediate H3K methylation, thereby silencing transcriptional
activity of target genes. MEN1-associated tumors harbor germline and somatic mutations, consistent
with Knudson’s two-hit hypothesis. Genetic diagnosis to identify individuals with germline MEN1 muta-
tions has facilitated appropriate targeting of clinical, biochemical and radiological screening for this high
risk group of patients for whom earlier implementation of treatments can then be considered. MEN4 is
caused by heterozygous mutations of CDNK1B which encodes the 196 amino-acid CDK1 p27Kip1, which
is activated by H3K4 methylation.
Ó 2013 Elsevier Ireland Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2. Multiple endocrine neoplasia type 1 (MEN1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.1. Clinical features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.2. Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.3. Genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.4. MEN1 mutations in sporadic non-MEN1 endocrine tumors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.5. MEN1 mutations in hereditary endocrine disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.6. Genetic testing and screening in MEN1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.6.1. MEN1 mutational analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.7. Detection of MEN1 tumors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

⇑ Tel.: +44 1865 857501; fax: +44 1865 857502.


E-mail address: rajesh.thakker@ndm.ox.ac.uk

0303-7207/$ - see front matter Ó 2013 Elsevier Ireland Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.mce.2013.08.002

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
2 R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

2.8. MEN1 phenocopies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00


3. Function of menin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1. Menin and epigenetic regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4. Multiple endocrine neoplasia type 4 (MEN4) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

1. Introduction either may be inherited as autosomal-dominant syndromes (Thak-


ker, 1998) or may occur sporadically; that is, without a family his-
Multiple endocrine neoplasia (MEN) is characterized by the tory (Thakker, 2010). However, this distinction between sporadic
occurrence of tumors involving two or more endocrine glands and familial cases may sometimes be difficult, because in some
within a single patient (Thakker, 2010). The disorder has previ- sporadic cases, a family history may be absent because the parent
ously been referred to as multiple endocrine adenopathy or the with the disease may have died before symptoms developed.
pluriglandular syndrome. However, glandular hyperplasia and MEN1, which is also referred to as Wermer’s syndrome (Thakker,
malignancy may also occur in some patients, and the term multiple 1998) is characterized by the combined occurrence of tumors of
endocrine neoplasia is now preferred. Four major forms of MEN are the parathyroid glands, the pancreatic islet cells, and the anterior
recognized and referred to as types 1–4 (MEN1–MEN4), and each pituitary. In addition to these tumors, adrenal cortical tumors, car-
form is characterized by the development of tumors within specific cinoid, facial angiofibromas, collagenomas, and lipomatous tumors
endocrine glands (Table 1) (Thakker, 1998). All these forms of MEN have been described in patients with MEN1 (Trump et al., 1996).

Table 1
Multiple endocrine neoplasia syndromes and their characteristic tumors and associated genetic abnormalities.

Type (chromosome location) Tumors (estimated penetrance) Gene; most frequently mutated codons
MEN1 (11q13) Parathyroid adenoma (90%) MEN1
Entero-pancreatic tumor (30–70%) 83/84, 4-bp del (4%)
– Gastrinoma (40%) 119, 3-bp del (3%)
– Insulinoma (10%) 209–211, 4-bp del (8%)
– Non-functioning & PPoma (20–55%) 418, 3-bp del (4%)
– Glucagonoma (<1%) 514–516, del or ins (7%)
– VIPoma (<1%) Intron 4 ss, (10%)
Pituitary adenoma (30–40%)
– Prolactinoma (20%)
– Somatotrophinoma (10%)
– Corticotrophinoma (<5%)
– Non-functioning (<5%)
Associated Tumors/
– Adrenal cortical tumor (40%)
– Phaeochromocytoma (<1%)
– Brochopulmonary NET (2%)
– Thymic NET (2%)
– Gastric NET (10%)
– Lipomas (30%)
– Angiofibromas (85%)
– Collagenomas (70%)
– Meningiomas (8%)

MEN2 (10 cen-10q11.2)


MEN2A MTC (90%) RET
Phaeochromocytoma (50%) 634, missense
Parathyroid adenoma (20–30%) e.g. Cys ? Arg (85%)
MTC only MTC (100%) RET
618, missense (>50%)
MEN2B (also known as MEN3) MTC (>90%) RET
Phaeochromocytoma (40–50%) 918, Met ? Thr (>95%)
Associated abnormalities (40–50%)
Mucosal neuromas
Marfanoid habitus
Medullated corneal nerve fibers
Megacolon
MEN4 (12p13) Parathyroid adenomaa CDKN1B
Pituitary adenomaa No common mutations identified to date
Reproduction organ tumorsa (e.g. testicular cancer, neuroendocrine cervical carcinoma)
Adrenal + renal tumorsa

Autosomal-dominant inheritance of the MEN syndromes has been established. del, deletion; ins, insertion; PPoma, pancreatic polypeptide–secreting tumor; VIPoma,
vasoactive intestinal polypeptide–secreting tumor; MTC, Medullary Thyroid Cancer. Adapted from Thakker RV: Clinical Practice Guidelines for Multiple Endocrine Neoplasia
Type 1 (MEN1). J Clin Endocrinol Metab 97: 2990–3011, Ó2012. The Endocrine Society (Thakker et al., 2012).
a
Insufficient numbers reported to provide prevalence information.

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx 3

However, in MEN2, which is also called Sipple’s syndrome and was


previously referred to as MEN2A, medullary thyroid carcinoma
(MTC) occurs in association with phaeochromocytoma and para-
thyroid tumors (Thakker, 1998). In MEN3, which was previously
referred to as MEN2B, MTC and pheochromocytoma occur in asso-
ciation with a marfanoid habitus, mucosal neuromas, medullated
corneal fibers, and intestinal autonomic ganglion dysfunction lead-
ing to megacolon (Thakker, 1998), with rare occurrence of parathy-
roid tumors. MTC may also occur as the sole manifestation of the
MEN2 syndrome, and this variant is referred to as MTC-only. In
MEN4, which is also referred to as MENX, patients develop para-
thyroid and anterior pituitary tumors, which may occur in associ-
ation with tumors of the reproductive organs, adrenals and
kidneys. This review will focus on the clinical features, genetics
and epigenetics of the MEN1 and MEN4 syndromes.
Fig. 1. Basis for a diagnosis of MEN1 in individuals. A diagnosis of MEN1 based on
clinical and familial criteria may be confounded by the occurrence of phenocopies.
(Reproduced with permission from Turner et al. (2010).) (Turner et al., 2010).
2. Multiple endocrine neoplasia type 1 (MEN1)

islet tumors, gastrinomas and foregut carcinoids (Dean et al.,


2.1. Clinical features
2000). In addition, these deaths occur at a significantly younger
age than those occurring in unaffected individuals (Dean et al.,
The incidence of MEN1 has been estimated from randomly cho-
2000). Furthermore, a recent multicentre study from France and
sen postmortem studies to be 0.25% and to be 1–18% among pa-
Belgium has suggested that 70% individuals with MEN1 currently
tients with primary hyperparathyroidism, 16–38% among
die of causes directly related to MEN1 (Goudet et al., 2010). In par-
patients with gastrinomas, and less than 3% among patients with
ticular, malignant pancreatic islet tumors and thymic carcinoid tu-
pituitary tumors (Thakker, 2010). The disorder affects all age
mors were associated with a marked increased in risk of death
groups, with a reported age range of 5–81 years, with clinical
(Hazard ratio >3, P < 0.005) (Goudet et al., 2010).
and biochemical manifestations of the disorder having developed
A diagnosis of MEN1 may be established in an individual by one
in 80% and more than 98% of patients, respectively, by the fifth dec-
of three criteria (Fig. 1). Thus, MEN1 may be clinically diagnosed in
ade (Thakker et al., 1989, 2012; Trump et al., 1996). The clinical
an individual on the basis of the occurrence of two or more MEN1-
manifestations of MEN1 are related to the sites of tumors and to
associated endocrine tumors (Fig. 1) (Thakker et al., 2012; Turner
their products of secretion. Parathyroid tumors, resulting in pri-
et al., 2010). In addition, a diagnosis of familial MEN1 is established
mary hyperparathyroidism, are the most common feature of
for an individual who has the occurrence of one of the MEN1-asso-
MEN1 and occur in 95% of MEN1 patients (Benson et al., 1987;
ciated tumors and is a first degree relative of a patient with a clin-
Brandi et al., 2001; Calender et al., 1995; Marx et al., 1998; Trump
ical diagnosis of MEN1. Finally, a genetic diagnosis of MEN1 is
et al., 1996). Pancreatic islet tumors, which are also referred to as
made on identification of a germline MEN1 mutation in an individ-
pancreatic neuroendocrine tumors (NETs) consist of gastrinomas,
ual, who may be asymptomatic and has not yet developed any of
insulinomas, pancreatic polypeptidomas (PPomas), glucagonomas,
the serum biochemical or radiological abnormalities indicative of
vaso-active intestinal polypeptidomas (VIPomas) and non-func-
tumor development (Fig. 2) (Turner et al., 2010).
tioning tumors and these occur in 40% of MEN1 patients; and
anterior pituitary tumors, consisting of prolactinomas, somatotro-
phinomas, corticotrophinomas, and non-functioning adenomas, 2.2. Treatment
occur in 30% of patients (Benson et al., 1987; Calender et al.,
1995; Marx et al., 1998; Trump et al., 1996). In addition, some The treatment for each type of MEN1 associated endocrine tu-
MEN1 patients may also develop adrenocortical tumors, lipomas, mor is generally similar to that for the respective tumors occurring
carcinoid tumors, facial angiofibromas and collagenomas (Bassett in non-MEN1 patients. However, the treatment outcomes of
et al., 1998; Brandi et al., 2001; Calender et al., 1995; Marx et al., MEN1-associated tumors are not as successful as those in non-
1998; Trump et al., 1996). Parathyroid tumors are the first mani- MEN1 patients, for several reasons. First, MEN1-associated tumors,
festation of MEN1 in more than 85% of patients, and in the remain- with the exception of pituitary NETs, are usually multiple, thereby
ing fewer than 15% of patients, the first manifestation may be an making it difficult to achieve a successful surgical cure. For exam-
insulinoma or a prolactinoma (Thakker et al., 1989; Trump et al., ple, MEN1 patients often develop multiple submucosal duodenal
1996). Combinations of these affected glands and their pathologic gastrinomas, thereby reducing surgical cure rates compared with
features (for example, hyperplasia or single or multiple adenomas similar sporadic solitary tumors, such that only 15% of MEN1 pa-
of the parathyroid glands) have been reported to differ in members tients, compared to 45% of non-MEN1 patients, are free of disease
of the same family (Thakker et al., 1989; Trump et al., 1996) and immediately after surgery, and at 5 years this decreased to 5% in
even between identical twins (Flanagan et al., 1996). MEN1 is MEN1 patients, compared to 40% in non-MEN1 patients (Brandi
inherited as an autosomal-dominant disorder in such families, et al., 2001; Imamura et al., 2011; Norton et al., 1999; Norton
but a non-familial (i.e., sporadic) (Trump et al., 1996) form may and Jensen, 2007; Tonelli et al., 2006). MEN1 patients also develop
have developed in 8–14% of patients with MEN1, and molecular ge- multiple parathyroid tumors, and subtotal parathyroidectomy has
netic studies have confirmed the occurrence of de novo mutations resulted in persistent or recurrent hypercalcaemia within 10 years
of the MEN1 gene in approximately 10% of patients with MEN1 in 20–60% of MEN1 patients, as opposed to 4% in non-MEN1 pa-
(Bassett et al., 1998). In the absence of treatment, these tumors tients (Brandi et al., 2001; Schreinemakers et al., 2011; Waldmann
have been observed to be associated with an earlier mortality in et al., 2010). Secondly, occult metastatic disease is more prevalent
patients with MEN1 (Burgess et al., 1998). Studies of MEN1-related in MEN1 patients with NETs than in patients with sporadic endo-
mortality have demonstrated that 28–46% of deaths are directly re- crine tumors. For example, metastases are present in up to 50%
lated to MEN1, most commonly as a result of malignant pancreatic of patients with MEN1-associated insulinomas, whereas <10% of

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
4 R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

Fig. 2. Detection of MEN1 mutation in exon 3 in family 8/89 by restriction enzyme analysis. DNA sequence analysis of individual II.1 revealed a 1-bp deletion at the second
position (GGT) of codon 214 (A). The deletion has caused a frameshift that continues to codon 223 before a stop codon (TGA) is encountered in the new frame. The 1-bp
deletion results in the loss of an MspI restriction enzyme site (C/CGG) from the normal (wild-type, WT) sequence (A), and this finding has facilitated detection of this mutation
in the other affected members (II.4, III.3, and III.4) of this family (B). The mutant (m) polymerase chain reaction product is 190 bp, whereas the WT products are 117 and 73 bp
(C). The affected individuals were heterozygous, and the unaffected members were homozygous for the WT sequence. Individuals III.6 and III.10, who are 40 and 28 years old,
respectively, are mutant gene carriers who are clinically and biochemically normal because of the age-related penetrance of this disorder (Fig. 5). These individuals would still
require screening (Fig. 6) by clinical, biochemical, and radiologic assessment because they still have a residual risk (i.e., 100% – age-related penetrance) of 2% and >13%,
respectively, of tumors developing by age 60 years. Individuals are represented as male (square); female (circle); unaffected (open); affected with parathyroid tumors (solid
upper right quadrant), gastrinoma (solid lower right quadrant), or prolactinoma (solid upper left quadrant); and unaffected mutant gene carriers (dot in the middle of the
open symbol). Individual I.2, who is dead but was known to be affected (tumor details not known), is shown as a solid symbol. The age is indicated below for each individual
at diagnosis or at the time of the last biochemical screening. The standard size marker (S) in the form of the 1-kb ladder is indicated. Cosegregation of this mutation with
MEN1 in family 8/89 and its absence in 110 alleles from 55 unrelated normal individuals (N1–3 shown) indicate that it is not a common DNA sequence polymorphism.
(Adapted from Bassett JHD, Forbes SA, Pannett AAJ, et al.: Characterization of mutations in patients with multiple endocrine neoplasia type 1 (MEN1). Am J Hum Genet 62:
232–244, 1998, with permission.) (Bassett et al., 1998).

non-MEN1 insulinomas metastasise (Akerstrom and Hellman, as menin. Menin is ubiquitously expressed and is predominantly a
2007). Thirdly, MEN1-associated tumors may be larger, more nuclear protein in non-dividing cells. Mutations of the MEN1 gene
aggressive, and resistant to treatment. For example, 85% of ante- (Figs. 2 and 3) have been characterized, and a total of 1336 muta-
rior pituitary tumors in MEN1 patients, as opposed to 64% in non- tions (1133 germline and 203 somatic mutations) were reported in
MEN1 patients are macroadenomas at the time of diagnosis; 30% the first decade following the identification of the gene (Lemos and
of anterior pituitary tumors in MEN1 patients have invaded sur- Thakker, 2008). The 1133 germ-line mutations of the MEN1 gene,
rounding tissue (Hardy classification grades III and IV), compared which consist of 459 different mutations are scattered throughout
to 10% in non-MEN1 patients; and >45% of anterior pituitary NETs the entire 1.830 bp coding region and splice sites of the MEN1 gene
in MEN1 patients had persistent hormonal over-secretion follow- (Agarwal et al., 1997; Bassett et al., 1998; Chandrasekharappa
ing appropriate medical, surgical and radiotherapy treatment, et al., 1997; Lemos and Thakker, 2008; The European Consortium
compared with between 10% and 40% in non-MEN1 patients (Fer- on MEN1, 1997). Approximately 23% are nonsense mutations,
nandez et al., 2010; Trouillas et al., 2008; Verges et al., 2002). Thus, around 41% are frameshift deletions or insertions, 6% are in-frame
the treatment of MEN1-associated tumors is difficult requires a deletions or insertions, 9% are splice-site mutations, and 20% are
multi-disciplinary team of experts. missense mutations, and 1% are whole or particular gene deletions
(Fig. 4) (Lemos and Thakker, 2008). More than 10% of the MEN1
2.3. Genetics mutations arise de novo and may be transmitted to subsequent
generations (Bassett et al., 1998; Lemos and Thakker, 2008; Zable-
The MEN1 gene is located on chromosome 11q13 and consists of wska et al., 2003). It also is important to note that between 5% and
10 exons (Fig. 3) that encode a 610-amino acid protein, referred to 10% of patients with MEN1 may not harbor mutations in the coding

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx 5

Fig. 3. Schematic representation of the genomic organization of the MEN1 gene, its encoded protein (menin) and regions that interact with other proteins. (A) The human
MEN1 gene consists of 10 exons that span more than 9 kb of genomic DNA and encodes a 610-amino acid protein. The 1.83 kb coding region (indicated by shaded region) is
organized into 9 exons (exons 2–10) and 8 introns (indicated by a line but not to scale). The sizes of the exons (boxes) range from 41 to 1297 bp, and that of the introns range
from 80 to 1564 bp. The start (ATG) and stop (TGA) codons in exons 2 and 10, respectively, are indicated. Exon 1, the 50 part of exon 2, and the 30 part of exon 10 are
untranslated (indicated by open boxes). The promoter region is located within a few 100 bp upstream of exon 2. The sites of the nine germline mutations (I–IX) that occur
with a frequency >1.5% (Table 2) are shown and their respective frequencies (scale shown on the right) are indicated by the vertical lines above the gene. These germline
mutations, which collectively represent 20.6% of all reported germline mutations, are: I – c.249_252delGTCT; II – c.292C>T; III – c.358_360delAAG; IV4 – c.628_631delACAG;
V – c.7849G>A; VI – c.1243C>T; VII – c.1378C>T; VIII – c.1546delC; IX – c.1546_1547insC. The locations of the 24 polymorphisms (a–x, Table 4) are illustrated. (B) Menin has
three nuclear localization signals (NLSs) at codons 479–497 (NLS1), 546–572 (NLSa) and 588–608 (NLS2), indicated by closed boxes, and five putative guanosine
triphosphatase (GTPase) sites (G1–G5) indicated by closed bars. (C) Menin regions that have been implicated in the binding to different interacting proteins (Table 6) are
indicated by open boxes. These are JunD (codons 1–40, 139–242, 323–428); nuclear factor-kappa B (NF-jB) (codons 305–381); Smad3 (codons 40–278, 477–610); placenta
and embryonic expression, pem (codons 278–476); NM23H1 (codons 1–486); a subunit of replication protein A (RPA2) (codons 1–40, 286–448); NMHC II-A (codons 154–
306); FANCD2 (codons 219–395); mSin3A (codons 371–387); HDAC1 (codons 145–450); ASK (codons 558–610) and CHES1 (codons 428–610). The regions of menin that
interact with GFAP, vimentin, Smad 1/5, Runx2, MLL-histone methyltransferase complex and estrogen receptor-alpha remain to be determined. (Reproduced from Lemos MC,
Thakker RV: Multiple Endocrine Neoplasia Type 1 (MEN1): Analysis of 1336 mutations reported in the first decade following identification of the gene. Hum Mutat 29: 22–32,
2008, with permission.) (Lemos and Thakker, 2008).

region of the MEN1 gene (Agarwal et al., 1997; Bassett et al., 1998; Thakker, 2008). These mutations at these nine different sites could
Chandrasekharappa et al., 1997; Lemos and Thakker, 2008; The be considered to represent potential ‘‘hot’’ spots (Table 2). Such
European Consortium on MEN1, 1997), and that these individuals deletional and insertional hot spots may be associated with DNA
may have whole gene deletions or mutations in the promoter or sequence repeats, which may consist of long tracts of either single
untranslated regions, which remain to be investigated. One study nucleotides or shorter elements ranging from dinucleotides to
showed that approximately 33% of patients who do not have muta- octanucleotides (Bassett et al., 1998). The DNA sequences in the
tions within the coding region have large deletions involving com- vicinity of codons 83 and 84 in exon 2, and codons 210–211 in
plete exons (Cavaco et al., 2002). Such abnormalities will not be exon 3, contain CT and CA dinucleotide repeats, respectively, flank-
easily detected by DNA sequence analysis. ing the 4-bp deletions; this finding would be consistent with a rep-
Most (75%) of the MEN1 mutations are inactivating (Lemos and lication-slippage model in which misalignment of the dinucleotide
Thakker, 2008) and are consistent with those expected in a tumor- repeat takes place during replication, followed by excision of the 4-
suppressor gene. The mutations not only are diverse in their types bp single-stranded loop (Bassett et al., 1998). A similar replication-
but are also scattered throughout the 1830-bp coding region of the slippage model may also be involved at codons 119 and 120, each
MEN1 gene with no evidence for clustering as observed in MEN2 of which consists of AAG nucleotides encoding a lysine (K) residue.
(Table 1). However, some of the mutations have been observed The deletions and insertions of codon 516 involve a poly(C)7 tract,
to occur several times in unrelated families (Fig. 3 and Table 2). and a slipped-strand mispairing model also is the most likely
Mutations at nine sites in the MEN1 gene accounted for over 20% mechanism to be associated with this mutational hot spot (Bassett
of all the germ-line mutations (Table 2). Of these nine types of et al., 1998; Lemos and Thakker, 2008). Thus, the MEN1 gene ap-
mutations, five are deletional and insertional mutations involving pears to contain DNA sequences that may render it susceptible to
codons 83 and 84 (nt359 del 4), 120 (Lys[K]120 del), 210–211 deletional and insertional mutations.
(nt 738 del 4), and codons 514–516 (nt 1656–7 del or ins C); one Correlations between MEN1 mutations and clinical manifesta-
is a novel acceptor splice site in intron 4, and three are nonsense tions of the disorder appear to be absent. For example, a detailed
mutations (Arg98Stop, Arg415Stop, and Arg460Stop) (Lemos and study of five unrelated families with the same 4-bp deletion in

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
6 R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

50 Table 3
Multiple endocrine neoplasia type 1 – associated tumors in five unrelated families
Germline, n=653 with a 4-bp deletion at codons 210 and 211.
Somatic, n=194 Tumors Family
40
1 2 3 4 5
Parathyroid + + + + +
30 Gastrinoma +  + + +
Percent

Insulinoma  +   
Glucagonoma     +
Prolactinoma  + + + +
20 Carcinoid +    

+, Presence; , Absence of tumors.


Adapted from Thakker RV: Multiple endocrine neoplasia—syndromes of the twen-
10 tieth century. J Clin Endocrinol Metab 83: 2617–2620, Ó1998. The Endocrine Society
(Thakker, 1998).

0
ns fs del/ins sp ms Table 4
Polymorphisms of the MEN1 gene.
ns = nonsense sp = splice site
fs = frame-shift ms = missense Polymorphisma DNA sequence Exon Codon change Allele
del/ins = deletions / insertions *P<0.001 changeb frequencyc
a c.-533T>A 1 – 0.32
Fig. 4. Frequency of germ-line and somatic MEN1 mutations. A total of 1133 germ-
b c.-533T>C 1 – 0.12
line mutations and 203 somatic mutations have been reported (Agarwal et al.,
c c.-39C>G Intron 1 – 0.20
1997), and these are of diverse types (e.g., nonsense, frameshifts, deletions,
d n/a 2 Leu10Leu n/a
insertions, splice-site, and missense mutations). The frequencies of each type of
e c.435C>T 2 Ser145Ser 0.01
mutation in the germ-line and somatic groups are similar, with the exception of the
f c.445+183G>A Intron 2 – 0.05
missense mutations, which are found more frequently in tumors (i.e., the somatic
g c.446127A>T Intron 2 – n/a
group). (Reproduced from Lemos MC, Thakker RV: Multiple Endocrine Neoplasia
h c.44658C>T Intron 2 – 0.01
Type 1 (MEN1): Analysis of 1336 mutations reported in the first decade following
i c.512G>A 3 Arg171Gln 0.01
identification of the gene. Hum Mutat 29: 22–32, 2008, with permission.) (Lemos
j c.768T>C 4 Leu256Leu 0.01
and Thakker, 2008).
k c.824+31T>C Intron 5 – n/a
l c.9133C>G Intron 6 – 0.02
m c.1026G>A 7 Ala342Ala 0.01
Table 2 n c.105092C>T Intron 7 – 0.03
MEN1 germline mutations occurring in over 1.5% of affected families. o c.10503C>G Intron 7 – 0.02
p c.1101A>C 8 Val367Val n/a
Mutationa DNA sequence Exon Codon Predicted Frequency
q c.1254C>T 9 Asp418Asp 0.42
changeb effectc (%)d
r c.1296G>A 9 Leu432Leu 0.01
I c.249_252delGTCT 2 83–84 fs 4.5 s c.1299T>C 9 His433His 0.01
II c.292C>T 2 98 ns, 1.5 t c.1350+103G>C Intron 9 – 0.42
Arg98Stop u c.1434C>T 10 Gly478Gly n/a
III c.358_360delAAG 2 120 if 1.7 v c.1621G>A 10 Ala541Thr 0.04
IV c.628_631delACAG 3 210– fs 2.5 w c.1764G>A 10 Lys588Lys n/a
211 x c.1833305_1833 10 – 0.05
V c.784-9G>A Intron – sp 1.9 307delCTC
4
VI c.1243C>T 9 415 ns, 1.5 n/a – Data not available. (From Lemos M and Thakker RV: Multiple Endocrine
Arg415Stop Neoplasia Type 1 (MEN1): Analysis of 1336 Mutations Reported in the First Decade
VII c.1378C>T 10 460 ns, 2.6 Following Identification of the Gene. Hum Mutat 29: 22–32, 2008, with permission).
Arg460Stop (Lemos and Thakker, 2008).
a
VIII c.1546delC 10 516 fs 1.8 Polymorphism letter as referred to in Fig. 3.
b
IX c.1546_1547insC 10 516 fs 2.7 Polymorphisms are numbered in relation to the MEN1 cDNA reference sequence
(GenBank accession number NM_130799.1), whereby nucleotide +1 corresponds to
a
Mutation number as referred to in Fig. 3. the A of the ATG-translation initiation codon.
b c
Mutations are numbered in relation to the MEN1 cDNA reference sequence Frequency presented in first literature report.
(GenBank accession number NM_130799.1), whereby nucleotide +1 corresponds to
the A of the ATG-translation initiation codon.
c
fs – frameshift mutation; ns – nonsense mutation; sp – splice site mutation; if –
in-frame mutation. of seven unrelated families with the same g ? a novel acceptor
d
Frequencies based on 1133 reported MEN1 independent germline mutations. splice-site mutation in intron 4 revealed a similarly wide range
(From Lemos M and Thakker RV: Multiple Endocrine Neoplasia Type 1 (MEN1): of MEN1-associated tumors and a lack of genotype/phenotype cor-
Analysis of 1336 Mutations Reported in the First Decade Following Identification of relation. The apparent lack of genotype/phenotype correlation,
the Gene. Hum Mutat 29: 22–32, 2008, with permission) (Lemos and Thakker,
which contrasts with the situation in MEN2 (Table 1), together
2008).
with the wide diversity of mutations in the 1830-bp coding region
of the MEN1 gene, makes mutational analysis for diagnostic pur-
poses in MEN1 more difficult than that for MEN2 (Thakker, 1998).
codons 210 and 211 (Table 3) revealed a wide range of MEN1-asso- A total of 24 different polymorphisms (12 in the coding region
ciated tumors (Bassett et al., 1998; Lemos and Thakker, 2008; [10 synonymous and 2 non synonymous], 9 in the introns, and 3
Thakker, 1998); all affected family members had parathyroid tu- in the untranslated regions) of the MEN1 gene have been reported
mors, but members of families 1, 3, 4, and 5 had gastrinomas, (Fig. 3 and Table 4) (Lemos and Thakker, 2008). It is important to
whereas members of family 2 had insulinomas. In addition, prolac- recognize the occurrence of these polymorphisms as they need to
tinomas occurred in members of families 2, 3, 4, and 5 but not in be distinguished from mutations when performing analysis for ge-
family 1, which was affected with carcinoid tumors. Another study netic diagnosis.

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx 7

More than 90% of tumors from MEN1 patients have loss of het- missense mutations, and fewer than 31% are nonsense or frame-
erozygosity (LOH), and this has generally been taken as evidence shift mutations, which would result in a truncated and likely inac-
that the MEN1 gene acts as a tumor-suppressor gene, consistent tivated protein. This contrasts significantly (P < 0.01) with the sit-
with Knudson’s two-hit hypothesis (Friedman et al., 1989; Larsson uation in MEN1 patients in whom more than 65% of the germ-
et al., 1988; Thakker et al., 1989). However, this LOH represents line mutations are protein-truncation and about 23% are missense
only one mechanism by which the second hit may occur, with mutations (Fig. 4). These observations are consistent with a more
the other mechanisms being intragenic deletions and point muta- likely association between missense mutations and the milder
tions. MEN1 tumors (e.g., parathyroids, insulinoma, and lipoma) FIHP variant, but it is important to note that the mutations associ-
that do not have LOH have been shown to harbor different somatic ated with FIHP are also scattered throughout the coding region and
and germ-line mutations of the MEN1 gene (Fig. 4) (Pannett and not clustered, a situation that is similar to that found for germ-line
Thakker, 2001), and this is consistent with the Knudson two-hit MEN1 mutations (Fig. 3). Furthermore, the occurrence of protein-
hypothesis (Knudson et al., 1973). truncating mutations in FIHP patients and particularly deletions,
such as the 4 bp, involving codons 83–84 (Table 2), which are iden-
2.4. MEN1 mutations in sporadic non-MEN1 endocrine tumors tical to those observed in MEN1 patients, makes it difficult to
establish an unequivocal phenotype–genotype correlation. How-
Parathyroid, pancreatic islet cell, and anterior pituitary tumors ever, the sole occurrence of parathyroid tumors in these families
may occur either as part of MEN1 or more commonly as sporadic, that harbor MEN1 mutations that are similar to those found in fam-
nonfamilial tumors. Tumors from patients with MEN1 have been ilies with MEN1 is remarkable, and the mechanisms that deter-
observed to harbor the germ-line mutation together with a somatic mine the altered phenotypic expressions of these mutations
LOH involving chromosome 11q13 (Friedman et al., 1989; Larsson remain to be elucidated. However, nonsense mutations (Tyr312-
et al., 1988; Thakker et al., 1989, 1993), or point mutations, as ex- Stop and Arg460Stop) have been detected (Olufemi et al., 1998)
pected from Knudson’s model (Pannett and Thakker, 2001) and the in MEN1 families with the Burin or prolactinoma variant (Agarwal
proposed role of the MEN1 gene as a tumor suppressor. However, et al., 1997) which is characterized by a high occurrence of prolac-
LOH involving chromosome 11q13, which is the location of tinomas and a low occurrence of gastrinomas (Hao et al., 2004;
MEN1, has also been observed in 5–50% of sporadic endocrine tu- Olufemi et al., 1998). Furthermore, a splice mutation (c.446-
mors, thus implicating the MEN1 gene in the etiology of these tu- 3c ? g) has been detected in an MEN1 kindred from Tasmania, in
mors (Thakker et al., 1993). Approximatley 200 somatic MEN1 whom there was an absence of somatotrophinomas (Burgess
mutations (Fig. 4) have been reported between 1997 and 2007, et al., 2000). However, some other families with isolated acromeg-
and these have occurred in several different endocrine tumors (Le- aly do not have abnormalities of the MEN1 gene (Lemos and Thak-
mos and Thakker, 2008). Thus, these have been detected in 18% of ker, 2008), even though segregation analysis in families with
sporadic parathyroid tumors (total number, n = 452), 38% of gastri- isloated acromegaly and LOH studies of somatotrophinomas indi-
nomas (n = 105) (Wang et al., 1998), 14% of insulinomas (n = 43), cated that the gene was likely to be located on chromosome
57% of VIPomas (n = 7) (Wang et al., 1998), 16% of non-functioning 11q13 (Soares et al., 2005). Interestingly, mutations of the aryl
pancreatic tumors (n = 32), 60% of glucagonomas (n = 5), 2.0% of hydrocarbon receptor interacting protein (AIP) gene, which is also
adrenal cortical tumors (n = 83) (Lemos and Thakker, 2008), 35% located on chromosome 11q13 and 2.7 Mb telomeric to the MEN1
of bronchial carcinoid tumors (n = 26) (Lemos and Thakker, gene, have been identified in some families with isolated acromeg-
2008), 3.5% of anterior pituitary adenomas (n = 167) (Lemos and aly (Vierimaa et al., 2006). However AIP mutations have not been
Thakker, 2008; Prezant et al., 1998), 10% of angiofibromas deteced in patients with MEN1, who do not have MEN1 mutations.
(n = 19), and 28% of lipomas (n = 8). These somatic mutations are
scattered throughout the 1830-bp coding region (Fig. 3), and 18% 2.6. Genetic testing and screening in MEN1
are nonsense mutations, 40% are frameshift deletions or insertions,
6% are in-frame deletions or insertions, 7% are splice-site muta- MEN1 is an uncommon disorder, but because of its autosomal-
tions, and 29% are missense mutations (Fig. 4) (Lemos and Thakker, dominant inheritance, the finding of MEN1 in a patient has impor-
2008). A comparison of the locations of the somatic and germ-line tant implications for other family members; first-degree relatives
mutations revealed a higher frequency (39% [somatic] vs. 23% have about a 50% risk of development of the disease (Trump
[germline]; P < 0.001) of somatic mutations in exon 2, but the sig- et al., 1996). Screening for MEN1 in patients involves the detection
nificance of this observation (Lemos and Thakker, 2008) in the con- of tumors and ascertainment of the germ-line genetic state, that is,
text of the Knudson two-hit hypothesis remains to be elucidated. normal or mutant gene carrier (Fig. 1) (Thakker et al., 2012). Detec-
The tumors harboring a somatic MEN1 mutation had chromosome tion of tumors entails clinical, biochemical, and radiologic investi-
11q13 LOH as the other genetic abnormality, or ‘‘hit’’, consistent gations for MEN1-associated tumors in patients (Trump et al.,
with Knudson’s hypothesis. These studies (Lemos and Thakker, 1996). The characterization of the MEN1 gene (Chandrasekharappa
2008; Prezant et al., 1998; Wang et al., 1998) indicate that et al., 1997; The European Consortium on MEN1, 1997) has facili-
although inactivation of the MEN1 gene may have a role in the eti- tated identification of individuals who have mutations and hence
ology of some sporadic endocrine tumors, the involvement of other a high risk of acquiring the disease (Figs. 2 and 3).
genes, for example, the GNAS1 gene encoding the G protein-stim-
ulatory a subunit (Lyons et al., 1990), with major roles in the eti- 2.6.1. MEN1 mutational analysis
ology of such sporadic endocrine tumors, is highly likely. MEN1 mutational analysis is helpful in clinical practice in sev-
eral ways that include: (1) confirmation of the clinical diagnosis;
2.5. MEN1 mutations in hereditary endocrine disorders (2) identification of family members who harbor the MEN1 muta-
tion and require screening for tumor detection and early treat-
The role of the MEN1 gene in the etiology of other inherited ment; and (3) identification of the 50% of family members who
endocrine disorders, in which either parathyroid or pituitary tu- do not habour the familial germline MEN1 mutation and can there-
mors occur as an isolated endocrinopathy, has been investigated fore be reassured and alleviated of the burden of anxiety of devel-
by mutational analysis. MEN1 mutations have been reported in oping tumors. This latter aspect cannot be over-emphasized as it
42 families with isolated hyperparathyroidism (FIHP) (Hannan not only helps to reduce the personal cost to the individuals and
et al., 2008; Lemos and Thakker, 2008), and 38% of these are their children, but also to the health services in not having to

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
8 R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

Table 5
Summary of biochemical and radiological screening guidelines in individuals at high risk of developing MEN1.

Tumor Age to begin (yr) Biochemical test (annually) Imaging test (every 3 years)
Parathyroid 8 Calcium, PTH None
Pancreatic neuroendocrine
Gastrinoma 20 Gastrin (±gastric acid output) None
Insulinoma 5 Fasting glucose, insulin None
Other enteropancreatic <10 Chromogranin-A; pancreatic polypeptide, glucagon; VIP MRI, CT or EUS (annually)
Anterior pituitary 5 Prolactin, IGF-1 MRI (every 3 years)
Adrenal <10 None, unless symptoms or signs of functioning tumor MRI or CT annually with pancreatic imaging
and/or tumor >1 cm are identified
Foregut carcinoid 20 None CT or MRI (every 1–2 years)

CT, computer tomography; MRI, magnetic resonance imaging; EUS, endoscopic ultrasound.
Reproduced from Thakker RV: Clinical Practice Guidelines for Multiple Endocrine Neoplasia Type 1 (MEN1). J Clin Endocrinol Metab 97: 2990–3011, Ó2012. The Endocrine
Society (Thakker et al., 2012).

undertake unnecessary biochemical and radiological investigations guidelines recommend that these investigations should commence
(Table 5). Thus, MEN1 mutational analysis can be useful in clinical from the age of 5 years to detect anterior pituitary tumors and ins-
practice. ulinomas, the age of 8 years to detect parathyroid tumors, and the
The current guidelines recommend that MEN1 mutational anal- age of 20 years to detect gastrinomas and foregut carcinoid tumors
ysis should be undertaken in: (1) an index case with two or more (Table 5), and be undertaken thereafter on an annual basis (Thak-
MEN1-associated endocrine tumors (i.e. parathyroid, pancreatic ker et al., 2012).
or pituitary tumors); (2) asymptomatic first degree relatives of a
known MEN1 mutation carrier; (3) a first-degree relative of a 2.7. Detection of MEN1 tumors
MEN1 mutation carrier expressing familial MEN1 (i.e. having
symptoms, signs, biochemical or radiological evidence for one or Biochemical screening for the development of MEN1 tumors in
more MEN1-associated tumors) (Thakker et al., 2012). In addition, asymptomatic members (Fig. 2) of families with MEN1 is of great
MEN1 mutational analysis should be considered in patients with importance in as much as earlier diagnosis and treatment of these
suspicious or atypical MEN1. This would include individuals with: tumors may help reduce morbidity and mortality (Thakker et al.,
parathyroid adenomas before the age of 30 years or multigland 2012). The age-related penetrance (i.e., the proportion of gene car-
parathyroid disease; gastrinoma or multiple pancreatic islet cell riers manifesting symptoms or signs of the disease by a given age)
tumors at any age; or individuals who have two or more MEN1- has been ascertained (Bassett et al., 1998) (Fig. 5), and the muta-
associated tumors, which are not part of the classical triad of para- tion appears to be nonpenetrant in those younger than 5 years.
thyroid, pancreatic islet and anterior pituitary tumors (e.g. para- Thereafter, the mutant MEN1 gene has a high penetrance, being
thyroid tumor plus adrenal tumor) (Thakker et al., 2012). greater than 50% penetrant by 20 years of age and greater than
Family members, including asymptomatic individuals who have 95% penetrant by 40 years (Bassett et al., 1998). Screening for
been identified to harbor a MEN1 mutation will require biochemi- MEN1 tumors is difficult because the clinical and biochemical
cal and radiological screening (5). In contrast, those relatives who manifestations in members of any one family are not uniformly
do not harbor the MEN1 mutation will have their risk of developing similar (Thakker et al., 1989; Trump et al., 1996). Attempts to
MEN1-associated endocrine tumors markedly decreased from 1 in screen for the development of MEN1 tumors in the asymptomatic
2 for an autosomal dominant disorder, to 1 in 3000, 1 in 100,000 relatives of an affected individual have depended largely on mea-
and 1 in 1000 which are the respective risks for parathyroid, pan- suring the serum con centrations of calcium, gastrointestinal (g–
creatic islet cell and anterior pituitary tumors for the general pop- i) hormones (e.g., gastrin), and prolactin, as well as on radiologic
ulation; thus, these relatives without the MEN1 mutation will be imaging of the abdomen and pituitary (Thakker et al., 1989; Thak-
freed from the requirement of further repeated clinical investiga- ker, 1998). Parathyroid overactivity causing hypercalcemia is al-
tions (Lemos and Thakker, 2008; Thakker et al., 2012). most invariably the first manifestation of the disorder and has
MEN1 germline mutational analysis should be considered in become a useful and easy biochemical screening investigation
those presenting at an early age with a single, apparently sporadic (Trump et al., 1996). In addition, hyperprolactinemia, which may
MEN1-associated tumor. The occurrence of germline MEN1 muta- be asymptomatic, may represent the first manifestation in fewer
tions in all patients with sporadic, non-familial parathyroid adeno- than 10% of patients and may thus also be a useful and easy bio-
mas is 1%, in gastrinomas is 5%, in prolactinoma is 1%, and in chemical screening investigation (Thakker et al., 2012; Trump
foregut carcinoids is 2% (Thakker et al., 2012). However, the occur- et al., 1996). Pancreatic involvement in asymptomatic individuals
rence of MEN1 mutations tumors in patients developing parathy- has been detected by estimating the fasting plasma concentrations
roid tumors below the age of 40 years has been reported to be of gastrin, PP, and glucagon and by abdominal imaging (Scarsbrook
higher at 5–13%. The occurrence rates of germline MEN1 mutations et al., 2006; Thakker et al., 2012; Trump et al., 1996). However, one
in individuals presenting with a single apparent sporadic pancre- study has shown that a stimulatory meal test is a better method for
atic NET at similarly younger age, has not been established, and detecting pancreatic disease in individuals who have no demon-
the current guidelines recommend that MEN1 mutational analysis strable pancreatic tumors by CT. An exaggerated increase in serum
should be considered in those with gastrinoma or multiple pancre- gastrin or PP or both proved to be a reliable early indicator for the
atic NETS (Thakker et al., 2012). development of pancreatic tumors in these individuals.
Mutational analysis in asymptomatic individuals should be At present, it is suggested that individuals at high risk for MEN1
undertaken at the earliest opportunity, and where possible in the (i.e., mutant gene carriers) undergo biochemical screening (Fig. 6)
first decade of life as tumors have developed in some children by at least once per annum and also have baseline pituitary and
the age of 5 years (Newey et al., 2009; Stratakis et al., 2000; Thak- abdominal imaging (e.g., MRI or CT), which should then be re-
ker et al., 2012). Appropriate biochemical and radiological investi- peated at 1- to 3-year intervals (Table 5) (Thakker et al., 2012).
gations aimed at detecting the development of tumors should then Screening should commence in early childhood because the
be undertaken in these individuals (Tables 1 and 5). The current disease has developed in some individuals by the age of 5 years

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx 9

Cushing’s disease, and visual field loss and the presence of subcu-
taneous lipomas, angiofibromas, and collagenomas (Thakker et al.,
2012; Trump et al., 1996). Biochemical screening should include
estimations of serum calcium, PTH, gastrointestinal hormones
(e.g., gastrin, insulin with a fasting glucose, glucagon, VIP, and
PP), chromogranin A, prolactin, and IGF-1 in all individuals (Thak-
ker et al., 2012), and more specific endocrine-function tests should
be undertaken in individuals who have symptoms or signs sugges-
tive of a clinical syndrome. Radiologic screening should include an
MRI (or CT scanning) of the pancreas, adrenals, and pituitary, ini-
tially as a baseline and then every 1–3 years, as well as imaging
for foregut carcinoids (Scarsbrook et al., 2006; Thakker et al., 2012).

2.8. MEN1 phenocopies

Approximately 5–25% of patients with MEN1 may not have


mutations of the MEN1 gene. This variability in detecting MEN1
mutations may partly be attributable to differences in methods
used to identify the mutations; for example most studies do
not systematically examine for large gene deletions, which may
be found in up to 33% of patients who do not have coding region
mutations (Cavaco et al., 2002). In addition, this variability may
be due to phenotype ascertainment, as some studies have in-
cluded non-familial (i.e. sporadic) patients who may have devel-
oped only two (or fewer) endocrine tumors, and the detection
rate for MEN1 mutations in these patients was found to be
<5% (Tham et al., 2007). Such patients with MEN1-associated tu-
mors but without MEN1 mutations may represent phenocopies,
or have mutations involving other genes. Phenocopy refers to
the development of disease manifestations usually associated
with mutations of a particular gene, but instead are due to an-
other etiology, and the occurrence of phenocopies, has been re-
ported in 5–10% of MEN1 kindreds (Burgess et al., 2000;
Newey and Thakker, 2011; Turner et al., 2010). These phenocop-
ies occurred in two settings. Firstly, in the context of familial
MEN1 (Fig. 1), in which a patient with one MEN1-associated tu-
mor, e.g. a prolactinoma, did not have the familial mutation; and
secondly, in the context of clinical MEN1, in which patients with
two MEN1-associated tumors, who did not have an MEN1 muta-
Fig. 5. Age distributions (A) and age-related penetrance (B) of multiple endocrine
tion, were demonstrated to have involvement of other genes.
neoplasia type 1 (MEN1) determined from an analysis of 174 mutant gene carriers.
A, The age distributions were determined for three groups of MEN1 mutant gene These genes may include: CDC73, which encodes parafibromin,
carriers from 40 families in whom mutations were detected (Bassett et al., 1998). whose mutations result in the Hyperparathyroid–jaw tumor
The 91 members of group A had symptoms, whereas the 40 members of group B (HPT–JT) syndrome; the calcium sensing receptor (CaSR), whose
were asymptomatic and detected by biochemical screening. The 43 members of mutations result in familial benign hypercalciuric hypercalcemia
group C represent individuals who are MEN1 mutant gene carriers (see Fig. 2) and
who remain asymptomatic and biochemically normal. The ages included for
(FBHH) (Turner et al., 2010); and the aryl hydrocarbon receptor-
members of groups A, B, and C are those at the onset of symptoms, at the finding of interacting protein (AIP), a tumor suppressor located on chromo-
the biochemical abnormality, and at the last clinical and biochemical evaluation, some 11q13 whose mutations are associated with familial iso-
respectively. Groups B and C contained members who were significantly younger lated pituitary adenomas (FIPA) (Belar et al., 2012). The
than those in group A (P < 0.001). The younger age of the group C mutant gene
occurrence of MEN1 phenocopies may confound the diagnosis
carriers is consistent with an age-related penetrance for MEN1, which was
calculated (B) for the first five decades. The age-related penetrance (i.e., the of MEN1 (Fig. 1) in an individual, and it therefore appears advis-
proportion of mutant gene carriers with manifestations of the disease by a given able to offer genetic testing to determine the MEN1 mutation
age) increased steadily from 7% in the group younger than 10 years to 52%, 87%, status to symptomatic family members within a MEN1 kindred,
98%, 99%, and 100% by the ages of 20, 30, 40, 50, and 60 years, respectively. The as well as to all index cases (i.e. patient) with two or more
residual risk (100% – age-related penetrance) for the development of MEN1 tumors
in asymptomatic mutant gene carriers who are biochemically normal would then
endocrine tumors. If a MEN1 mutation is not identified in the in-
be 93%, 48%, 13%, 2%, and 1% at the ages of 10, 20, 30, 40, and 50 years, respectively. dex case with two or more endocrine tumors, then clinical and
(From Bassett JHD, Forbes SA, Pannett AAJ, et al.: Characterization of mutations in genetic tests for other disorders such as HPT–JT, FBHH, or FIPA
patients with multiple endocrine neoplasia type 1 (MEN1). Am J Hum Genet 62: should be considered, as these patients may represent phenocop-
232–244, 1998, with permission.) (Bassett et al., 1998).
ies for MEN1.

(Thakker et al., 2012), and it should continue for life because the 3. Function of menin
disease has not developed in some individuals until the eighth
decade. The screening history and physical examination should Initial analysis of the predicted amino acid sequence of menin
be directed toward eliciting the symptoms and signs of hypercalce- did not reveal homologies to any other proteins, sequence motifs,
mia, nephrolithiasis, peptic ulcer disease, neuroglycopenia, hypo- signal peptides, or consensus nuclear localization signal (Chandr-
pituitarism, galactorrhea and amenorrhea in women, acromegaly, asekharappa et al., 1997), and thus the putative function of menin

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
10 R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

Fig. 6. An approach to screening in an asymptomatic relative of a patient with multiple endocrine neoplasia type 1 (MEN1). The relative should have first undergone clinical
evaluation for MEN1-associated tumors to establish that the individual is asymptomatic. Relatives who are symptomatic, who should also have a test for MEN1 mutations,
should proceed to appropriate investigations and management. If mutational analysis for MEN1 is not available, then this protocol could be adapted for first-degree relatives
(Trump et al., 1996). The use of mutational analysis and such screening methods in children is controversial and varies in different countries. It has been suggested that
nonessential genetic testing in a child who is not old enough to make important long-term decisions be deferred. However, the finding that a child from a family with MEN1
does not have any MEN1 mutations removes the burden of repeated clinical, biochemical, and radiologic investigations and enables health resources to be more effectively
directed toward those children who are MEN1 mutant gene carriers. The approaches to genetic testing and screening in MEN1 vary in different countries. PIT, pituitary; PANC,
pancreas; ADR, adrenal; CAR, carcinoid; PTH, parathyroid hormone; PRL, prolactin; IGF-1, insulin-growth-factor-1; CgA, chromogranin A; and g–i, and gastro-intestinal gut-
hormones. (Reproduced with permission from Thakker RV (2010) Multiple Endocrine Neoplasia Type 1. In Endocrinology, sixth ed., Eds: LJ De Groot, JL Jameson.) (Thakker,
2010).

could not be deduced. However, studies based on immunofluores- activation (Heppner et al., 2001); members of the Smad family,
cence, Western blotting of subcellular fractions, and epitope Smad3 and the Smad 1/5 complex, to inhibit the transforming
tagging with enhanced green fluorescent protein revealed that me- growth factor-b (TGF-b) and the bone morphogenetic protein-2
nin was located primarily in the nucleus (Guru et al., 1998). Menin (BMP-2) signaling pathways, respectively; Runx2, also called cbfa1,
has been shown to have at least three nuclear localization signals which is a common target of TGF-b and BMP-2 in differentiating
(NLSs) that are located in the C-terminal quarter of the protein osteoblasts (Sowa et al., 2004); and the mouse placental embryonic
(Guru et al., 1998; La et al., 2006) (Fig. 3). Interestingly, the trun- (pem) expression gene, which encodes a homeobox-containing
cated MEN1 proteins that would result from the nonsense and protein. Recently, the forkhead transcription factor CHES1 has been
frameshift mutations, if expressed, would lack at least one of these shown to be a component of this transcriptional repressor complex
nuclear localization signals (Fig. 3). Menin is predominantly a nu- and to interact with menin in an S-phase checkpoint pathway re-
clear protein in non-dividing cells, but in dividing cells, it is found lated to DNA damage response. Menin uncouples ELK-1, JunD,
in the cytoplasm (Guru et al., 1998). The function of menin still re- and c-Jun phosphorylation from mitogen-activated protein kinase
mains to be elucidated, but it has been shown to interact with a (MAPK) activation and suppresses insulin-induced c-Jun-mediated
number of proteins that are involved in transcriptional regulation transactivation in CHO-1R cells.
(Agarwal et al., 1999; Dreijerink et al., 2006; Heppner et al., Menin has also been reported to directly bind to doubled-
2001; Hughes et al., 2004; Karnik et al., 2005; La et al., 2007; Milne stranded DNA and this is mediated by the positively charged resi-
et al., 2005; Roudaia and Speck, 2008; Scacheri et al., 2006; Yokoy- dues in the nuclear localization signals (NLSs) in the carboxyl ter-
ama and Cleary, 2008; Yokoyama et al., 2004), genome stability, minus of menin. The NLSs appear to be necessary for menin to
cell division and proliferation (Lemos and Thakker, 2008) (Fig. 3 repress the expression of the insulin-like growth factor binding
and Table 6). Thus, in transcriptional regulation, menin has been protein-2 (IGFBP-2) gene by binding to the IGFBP-2 promoter. In
shown to interact with: the activating protein-1 (AP-1), transcrip- addition, each of the NLSs has also been reported to be involved
tion factors JunD (Agarwal et al., 1999) and C-Jun (Yumita et al., in menin-mediated induction of caspase 8 expression (La et al.,
2003), and to suppress Jun-mediated transcriptional activation; 2006, 2007). The NLSs may therefore have roles in controlling gene
members (e.g., p50, p52, and p65) of the NF-RB family of transcrip- transcription as well as targeting menin into the nucleus. Further-
tional regulators to repress NF-RB-mediated transcriptional more, gene expression profile studies, using pituitary and

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx 11

Table 6 in CHO-IR cells also suppressed insulin-induced AP-1 transactiva-


Functions of menin indicated by direct interactions with proteins and other tion, and this was accompanied by an inhibition of C-Fos induction
molecules.
at the transcriptional level (Kim et al., 1999). Furthermore, menin
Functiona Interacting partner expression in Men1-deficient mouse Leydig tumor cell lines in-
Transcription JunD duced cell cycle arrest and apoptosis (Hussein et al., 2007). In con-
regulation trast, depletion of menin in human fibroblasts resulted in their
NF-jB (P50, P52, P65) immortalisation (Lin and Elledge, 2003). Thus, menin appears to
pem
SIN3A
have a large number of functions through interactions with pro-
HDAC teins, and these mediate alterations in cell proliferation.
Smad1
Smad3 3.1. Menin and epigenetic regulation
Smad5
Runx2
MLL histone methyl-transferase complex Menin which acts as a scaffold protein, may increase or de-
ER-Alpha crease gene expression by epigenetic regulation of gene expression
CHES1 via histone methylation or acetylation. For examaple, menin has
Double-stranded DNA been shown to be an integral component of histone methyltrans-
Genome stability RPA2 ferase complexes that contain members from the mixed-lineage
FANCD2 leukemia (MLL) and trithorax protein family (Hughes et al.,
Cell division NMMHC II-A 2004). These can methylate the lysine 4 residue of histone H3
GFAP (H3K4) and H3K4 trimethylation is linked to activation of tran-
Vimentin
scription. Menin, by serving as a molecular adaptor, physically
Cell cycle control NM23 (A) links the MLL histone methyl transferase with the lens epithe-
ASK
lium-derived growth factor (LEDGF), which is a chromatin-associ-
Epigenetic regulation MLL histone methyl-transferase complex ER-Alpha ated protein implicated in the eitology of leukemia, autoimmunity
HDAC
and human immuno-defeciency virus-1 (HIV-1) disease (Roudaia
a
Functions reported include involvement in cell cycle withdrawal, decrease of and Speck, 2008; Yokoyama and Cleary, 2008). Thus, menin has a
cell motility, cell differentiation, apoptosis and DNA repair. (Adapted from Lemos M critical function in the MLL complex and its regulation of multiple
and Thakker RV: Multiple Endocrine Neoplasia Type 1 (MEN1): Analysis of 1336
transcriptional pathways, for example, menin, as a component of
Mutations Reported in the First Decade Following Identification of the Gene. Human
Mutation 29: 22–32, 2008, with permission) (Lemos and Thakker, 2008). this MLL complex, regulates the expression of genes such as the
Hox homeobox genes (Agarwal and Jothi, 2012; Yokoyama et al.,
2004) and the genes for cyclin-dependent kinase inhibitors, p27
pancreatic islet tumors obtained from men1 mouse models, have and p18 (Karnik et al., 2005; Milne et al., 2005). Interestingly, spe-
revealed altered expression of genes involved in transcription, cell cific deregulation of a subset of 23 Hox genes has been reported in
cycle and chromatin remodeling. parathyroid tumors from patients with MEN1 (Shen et al., 2008). In
A role for menin in controlling genome stability has been pro- addition, menin has been shown to directly interact with protein
posed because of its interactions with: a subunit of replication pro- arginine methyltransferase 5 (PRMT5), and to recruit PRMT5 to
tein (RPA2), which is a heterotrimeric protein required for DNA the promoter of the gene encoding the growth arrest-specific pro-
replication, recombination, and repair; and the FANCD2 protein, tein 1 (GAS1), which is an important factor for binidng of Sonic
that is involved in DNA repair and mutations of which result in Hedgehog (Shh) ligand to its cell surface receptor patched (PTCH1)
the inherited cancer-prone syndrome of Fanconi’s anemia. Menin with subsequent activation of the Hedgehog signaling pathway,
also has a role in regulating cell division as it interacts with the which increases repressive histone arginine symmetric dimethyla-
nonmuscle myosin II-A heavy chain (NMHC II-A), which partici- tion (H4R3M2S) and suppresses GAS1 expression (Gurung et al.,
pates in mediating alterations in cytokinesis and cell shape during 2013). GAS1 blocks entry to S phase and prevents cycling of normal
cell division and the glial fibrillary acidic protein (GFAP) and and transformed cells, hereby acting as a tumor suppressor. Fur-
vimentin, which is involved in the intermediate filament network. thermore, menin has been reported to interact with the suppressor
Menin also has a role in cell cycle control as it interacts with: the of varigation 3–9 homolog family protein (SUV39H1) to mediate
tumor metastases suppressor NM23H1/nucleoside diphosphate ki- H3K mehylation, and thereby silence transcrpitional activty of tar-
nase, which induces guanosine triphosphate (GTP)ase activity; and get genes (Yang et al., 2013). Menin has been shown to directly
the activator of S-phase kinase (ASK), which is a component of the interact with the nuclear receptor for estrogen (ERa) and to act
Cdc7/ASK kinase complex that is crucial for cell proliferation. In- as a coactivator for ERa–mediated transcription, linking the acti-
deed, menin has been shown to completely repress ASK-induced vated estrogen receptor to histone H3K4 trimethylation (Dreijerink
cell proliferation (Schnepp et al., 2004). et al., 2006). Additional studies have also shown that the interac-
The functional role of menin as a tumor suppressor also has tion of menin with JunD may be mediated by a histone deacetyl-
been investigated, and studies in human fibroblasts have revealed ase-dependent mechanism, via recruitment of an mSin3A-histone
that menin acts as a repressor of telomerase activity via hTERT (a deacetylase (HDAC) complex to repress JunD transcriptional activ-
protein component of telomerase) (Lin and Elledge, 2003). Further- ity (Kim et al., 2003, 2005). Thus, these studies reveal that menin
more, overexpression of menin in the human endocrine pancreatic can act both as an activator and repressor of transcription by epi-
tumor cell line (BONI) resulted in an inhibition of cell growth (Stal- genetic mechanisms that involve integration of different chroma-
berg et al., 2004) that was accompanied by upregulation of JunD tin modification complexes. For example, menin may mediate
expression but downregulation of d-like protein 1/preadipocyte repression of genes targeted by the transcription factor JunD
factor-1, proliferating cell nuclear antigen, and QM/Jif-1, which is through the mSin3A-histone deacetylse (HDAC) complex, and the
a negative regulator of C-Jun (Stalberg et al., 2004). These findings growth factor pleiotrophin gene through polycomb group com-
of growth suppression by menin also are observed in other cell plex-mediated histone H3K27 methylation (Gao et al., 2009). Me-
types. Thus, expression of menin in RAS-transformed NIH3T3 cells nin may also promote expression of genes, such as those
partially suppressed the RAS-mediated tumor phenotype in vitro encoding the cyclase-dependent-kinase (CDK) inhibitors, p27Kip1
and in vivo (Stalberg et al., 2004), and overexpression of menin and P18INK4c, via interaction with the MLL-HMT complex that

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
12 R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

Fig. 7. Role of menin in regulating transcription by epigenetic mechanisms. Menin may act as an activator or repressor of transcription depending on its inteaction with the
MLL complex, HDAC or SUV39H1. Thus, interaction with MLL results in methylation of histone H3 (H3K4), which in turn regulates suppression of the CDK inhibitors p18 and
p27, and Hox genes, to suppress cell proliferation. However, interaction with HDAC and SUV39H1 results in acetylation of H3 and methylation of H3K9me3, respectively,
which in turn regulate expression of GBX2 and IGFBP2 to promote cell proliferation. Thus, menin plays a role in selective mediation of chromatin remodeling and thereby in
regulating gene expression and cell proliferation.

mediates histone H3K4 methylation (Fig. 7) (Huang et al., 2012; approximatley 3% of these patients with MEN1-associated tumors,
Hughes et al., 2004; Karnik et al., 2005; Milne et al., 2005). Target- such as parathyroid adenomas, pituitary adenomas and pancreatic
ing of these epigenetic mechanisms by use of small molecules is NETs in association with gonadal, adrenal, renal and thyroid tu-
providing important insights for the development of therapeutic mors have CDNK1B mutations. To date, 8 different heterozygous
compounds. For example, the crystal structures of human menin loss-of-function CDNK1B mutations have been identified in pa-
in its free form and in complexes with MLL1 have shown that me- tients with MEN1-like tumors, and this indicates that MEN4 in
nin contains a deep pocket that binds to a short peptide of MLL1 man is an autosomal dominant-disorder, unlike MENX in rats
(Huang et al., 2012). High-throughput screening studies to identify which is autosomal recessive. In addition, germline CDNK1B muta-
compounds that target menin and suppress its interaction with tions may also, rarely, be found in patients with sporadic (i.e. non-
MLL have identified that thienopyrimidine has such functions familial) forms of primary hyperparathyroidism.
(Grembecka et al., 2012; McCarthy, 2012; Shi et al., 2012). Further-
more, their structure activity analyses led to generation of some
5. Conclusions
thienopyrimidine analogs, which bind to wild-type menin but
not menin mutants that involve the interaction site with MLL,
Combined clinical and laboratory investigations of MEN1 have
and others which are capable of inhibiting the MLL-fusion pro-
resulted in an increased understanding of this disorder, which
tein-mediated leukemogenic transformation (Grembecka et al.,
may be inherited as an autosomal-dominant condition. Defining
2012; McCarthy, 2012; Shi et al., 2012). Thus, understanding the
the features of each disease manifestation in MEN1 has improved
role of menin in epigenetic mechanisms has provided a rational
patient management and treatment and has also facilitated a
and molecular basis for the development of menin-MLL inhibitors.
screening protocol to be instituted. Application of the techniques
of molecular biology has enabled identification of the gene causing
4. Multiple endocrine neoplasia type 4 (MEN4) MEN1 and detection of mutations in patients. In addition, these re-
cent advances have facilitated the identification of mutant MEN1
Approximately 5–10% of patients with MEN1 do not have muta- gene carriers who are at high risk for development of this disorder
tions of the MEN1 gene (Lemos and Thakker, 2008; Thakker et al., and thus require regular and biochemical and radiologic screening
2012), and these patients may have mutations involving other to detect the development of endocrine tumors. The function of the
genes. One of these genes is the CDNK1B, which encodes the 196 protein encoded by the MEN1 gene, menin, has been shown to in-
amino acid cyclin-dependent kinase inhibitor (CK1) p27kip1 and volve regulation of transcription, which involves genetic mecha-
was identified to be involved by investigations of a recessive nisms, genome stability, and cell division, but much still remains
MEN-like syndrome in a naturally occuring rat model, referred to to be elucidated. Furthermore, inhibitors of the menin-MLL inter-
as MENX (Pellegata et al., 2006). Rats with MENX were observed action hold promise for provoking new treatments.
to develop parathyroid adenomas, pancreatic islet-cell hyperplasia,
thyroid C-cell hyperplasia, bilateral phaeochromocytomas, para- Acknowledgments
gangliomas and cataracts. The disorder was inherited as an autoso-
mal recessive trait. Genetic mapping studies localized MENX to the I am grateful to the Medical Research Council (UK) for support;
distal part of rat chromosome 4, a region that also contained the and to Mrs. Tracey Walker for typing the manuscript.
putative tumor suppressor CK1P27kip1, which is also refered to
as p27. Mutational analysis of the CDNK1B gene identified, in rats
with MENX, a homozygous frameshifting insertion of 8 bp, at co- References
don 177, that resulted in a missense peptide and termination at co-
Agarwal, S.K., Jothi, R., 2012. Genome-wide characterization of menin-dependent
don 218 (Pellegata et al., 2006). This CDNK1B mutation resulted in H3K4me3 reveals a specific role for menin in the regulation of genes implicated
an absence of p27 protein in the tumor cells (Pellegata et al., 2006). in MEN1-like tumors. PLoS One 7, e37952.
These findings prompted studies in patients with MEN1, who did Agarwal, S.K., Kester, M.B., Debelenko, L.V., Heppner, C., Emmert-Buck, M.R.,
Skarulis, M.C., et al., 1997. Germline mutations of the MEN1 gene in familial
not harbor MEN1 mutations, for abnormalities of CDKN1B which multiple endocrine neoplasia type 1 and related states. Hum. Mol. Genet. 6,
in man is located on chromosome 12p13. This revealed that 1169–1175.

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx 13

Agarwal, S.K., Guru, S.C., Heppner, C., Erdos, M.R., Collins, R.M., Park, S.Y., et al., 1999. Karnik, S.K., Hughes, C.M., Gu, X., Rozenblatt-Rosen, O., McLean, G.W., Xiong, Y.,
Menin interacts with the AP1 transcription factor JunD and represses JunD- et al., 2005. Menin regulates pancreatic islet growth by promoting histone
activated transcription. Cell 96, 143–152. methylation and expression of genes encoding p27Kip1 and p18INK4c. Proc.
Akerstrom, G., Hellman, P., 2007. Surgery on neuroendocrine tumours. Best Pract. Natl. Acad. Sci. USA 102, 14659–14664.
Res. Clin. Endocrinol. Metab. 21, 87–109. Kim, Y.S., Burns, A.L., Goldsmith, P.K., Heppner, C., Park, S.Y., Chandrasekharappa,
Bassett, J.H., Forbes, S.A., Pannett, A.A., Lloyd, S.E., Christie, P.T., Wooding, C., et al., S.C., et al., 1999. Stable overexpression of MEN1 suppresses tumorigenicity of
1998. Characterization of mutations in patients with multiple endocrine RAS. Oncogene 18, 5936–5942.
neoplasia type 1. Am. J. Hum. Genet. 62, 232–244. Kim, H., Lee, J.E., Cho, E.J., Liu, J.O., Youn, H.D., 2003. Menin, a tumor suppressor,
Belar, O., De la Hoz, C., Perez-Nanclares, G., Castano, L., Gaztambide, S., 2012. Novel represses JunD-mediated transcriptional activity by association with an
mutations in MEN1, CDKN1B and AIP genes in patients with multiple endocrine mSin3A-histone deacetylase complex. Cancer Res. 63, 6135–6139.
neoplasia type 1 syndrome in Spain. Clin. Endocrinol. (Oxf.) 76, 719–724. Kim, H., Lee, J.E., Kim, B.Y., Cho, E.J., Kim, S.T., Youn, H.D., 2005. Menin represses
Benson, L., Ljunghall, S., Akerstrom, G., Oberg, K., 1987. Hyperparathyroidism JunD transcriptional activity in protein kinase C theta-mediated Nur77
presenting as the first lesion in multiple endocrine neoplasia type 1. Am. J. Med. expression. Exp. Mol. Med. 37, 466–475.
82, 731–737. Knudson Jr., A.G., Strong, L.C., Anderson, D.E., 1973. Heredity and cancer in man.
Brandi, M.L., Gagel, R.F., Angeli, A., Bilezikian, J.P., Beck-Peccoz, P., Bordi, C., et al., Prog. Med. Genet. 9, 113–158.
2001. Guidelines for diagnosis and therapy of MEN type 1 and type 2. J. Clin. La, P., Desmond, A., Hou, Z., Silva, A.C., Schnepp, R.W., Hua, X., 2006. Tumor
Endocrinol. Metab. 86, 5658–5671. suppressor menin: the essential role of nuclear localization signal domains in
Burgess, J.R., Greenaway, T.M., Shepherd, J.J., 1998. Expression of the MEN-1 gene in coordinating gene expression. Oncogene 25, 3537–3546.
a large kindred with multiple endocrine neoplasia type 1. J. Int. Med. 243, 465– La, P., Yang, Y., Karnik, S.K., Silva, A.C., Schnepp, R.W., Kim, S.K., et al., 2007. Menin-
470. mediated caspase 8 expression in suppressing multiple endocrine neoplasia
Burgess, J.R., Nord, B., David, R., Greenaway, T.M., Parameswaran, V., Larsson, C., type 1. J. Biol. Chem. 282, 31332–31340.
et al., 2000. Phenotype and phenocopy: the relationship between genotype and Larsson, C., Skogseid, B., Oberg, K., Nakamura, Y., Nordenskjold, M., 1988. Multiple
clinical phenotype in a single large family with multiple endocrine neoplasia endocrine neoplasia type 1 gene maps to chromosome 11 and is lost in
type 1 (MEN1). Clin. Endocrinol. (Oxf.) 53, 205–211. insulinoma. Nature 332, 85–87.
Calender, A., Giraud, S., Lenoir, G.M., Cougard, P., Chanson, P., Proye, C., 1995. Lemos, M.C., Thakker, R.V., 2008. Multiple endocrine neoplasia type 1 (MEN1):
[Hereditary multiple endocrine neoplasia. New genetic data and clinical analysis of 1336 mutations reported in the first decade following identification
applications in type 1 multiple endocrine neoplasia]. Presse Med. 24, 542–546. of the gene. Hum. Mutat. 29, 22–32.
Cavaco, B.M., Domingues, R., Bacelar, M.C., Cardoso, H., Barros, L., Gomes, L., et al., Lin, S.Y., Elledge, S.J., 2003. Multiple tumor suppressor pathways negatively regulate
2002. Mutational analysis of Portuguese families with multiple endocrine telomerase. Cell 113, 881–889.
neoplasia type 1 reveals large germline deletions. Clin. Endocrinol. (Oxf.) 56, Lyons, J., Landis, C.A., Harsh, G., Vallar, L., Grunewald, K., Feichtinger, H., et al., 1990.
465–473. Two G protein oncogenes in human endocrine tumors. Science 249, 655–659.
Chandrasekharappa, S.C., Guru, S.C., Manickam, P., Olufemi, S.E., Collins, F.S., Marx, S., Spiegel, A.M., Skarulis, M.C., Doppman, J.L., Collins, F.S., Liotta, L.A., 1998.
Emmert-Buck, M.R., et al., 1997. Positional cloning of the gene for multiple Multiple endocrine neoplasia type 1: clinical and genetic topics. Ann. Int. Med.
endocrine neoplasia-type 1. Science 276, 404–407. 129, 484–494.
Dean, P.G., van Heerden, J.A., Farley, D.R., Thompson, G.B., Grant, C.S., Harmsen, W.S., McCarthy, N., 2012. Anticancer drugs: targeting menin. Nat. Rev. Drug Discov. 11,
et al., 2000. Are patients with multiple endocrine neoplasia type I prone to 190.
premature death? World J. Surg. 24, 1437–1441. Milne, T.A., Hughes, C.M., Lloyd, R., Yang, Z., Rozenblatt-Rosen, O., Dou, Y., et al.,
Dreijerink, K.M., Mulder, K.W., Winkler, G.S., Hoppener, J.W., Lips, C.J., Timmers, 2005. Menin and MLL cooperatively regulate expression of cyclin-dependent
H.T., 2006. Menin links estrogen receptor activation to histone H3K4 kinase inhibitors. Proc. Natl. Acad. Sci. USA 102, 749–754.
trimethylation. Cancer Res. 66, 4929–4935. Newey, P.J., Thakker, R.V., 2011. Role of multiple endocrine neoplasia type 1
Fernandez, A., Karavitaki, N., Wass, J.A., 2010. Prevalence of pituitary adenomas: a mutational analysis in clinical practice. Endocr. Pract. 17 (Suppl. 3), 8–17.
community-based, cross-sectional study in Banbury (Oxfordshire, UK). Clin. Newey, P.J., Jeyabalan, J., Walls, G.V., Christie, P.T., Gleeson, F.V., Gould, S., et al.,
Endocrinol. (Oxf.) 72, 377–382. 2009. Asymptomatic children with multiple endocrine neoplasia type 1
Flanagan, D.E., Armitage, M., Clein, G.P., Thakker, R.V., 1996. Prolactinoma mutations may harbor nonfunctioning pancreatic neuroendocrine tumors. J.
presenting in identical twins with multiple endocrine neoplasia type 1. Clin. Clin. Endocrinol. Metab. 94, 3640–3646.
Endocrinol. (Oxf.) 45, 117–120. Norton, J.A., Jensen, R.T., 2007. Role of surgery in Zollinger–Ellison syndrome. J. Am.
Friedman, E., Sakaguchi, K., Bale, A.E., Falchetti, A., Streeten, E., Zimering, M.B., et al., Coll. Surg. 205, S34–S37.
1989. Clonality of parathyroid tumors in familial multiple endocrine neoplasia Norton, J.A., Fraker, D.L., Alexander, H.R., Venzon, D.J., Doppman, J.L., Serrano, J.,
type 1. N. Engl. J. Med. 321, 213–218. et al., 1999. Surgery to cure the Zollinger–Ellison syndrome. N. Eng. J. Med. 341,
Gao, S.B., Feng, Z.J., Xu, B., Wu, Y., Yin, P., Yang, Y., et al., 2009. Suppression of lung 635–644.
adenocarcinoma through menin and polycomb gene-mediated repression of Olufemi, S.E., Green, J.S., Manickam, P., Guru, S.C., Agarwal, S.K., Kester, M.B., et al.,
growth factor pleiotrophin. Oncogene 28, 4095–4104. 1998. Common ancestral mutation in the MEN1 gene is likely responsible for
Goudet, P., Murat, A., Binquet, C., Cardot-Bauters, C., Costa, A., Ruszniewski, P., et al., the prolactinoma variant of MEN1 (MEN1Burin) in four kindreds from
2010. Risk factors and causes of death in MEN1 disease. A GTE (Groupe d’Etude Newfoundland. Hum. Mutat. 11, 264–269.
des Tumeurs Endocrines) cohort study among 758 patients. World J. Surg. 34, Pannett, A.A., Thakker, R.V., 2001. Somatic mutations in MEN type 1 tumors,
249–255. consistent with the Knudson ‘‘two-hit’’ hypothesis. J. Clin. Endocrinol. Metab.
Grembecka, J., He, S., Shi, A., Purohit, T., Muntean, A.G., Sorenson, R.J., et al., 2012. 86, 4371–4374.
Menin-MLL inhibitors reverse oncogenic activity of MLL fusion proteins in Pellegata, N.S., Quintanilla-Martinez, L., Siggelkow, H., Samson, E., Bink, K., Hofler,
leukemia. Nat. Chem. Biol. 8, 277–284. H., et al., 2006. Germ-line mutations in p27Kip1 cause a multiple endocrine
Guru, S.C., Goldsmith, P.K., Burns, A.L., Marx, S.J., Spiegel, A.M., Collins, F.S., et al., neoplasia syndrome in rats and humans. Proc. Natl. Acad. Sci. USA 103, 15558–
1998. Menin, the product of the MEN1 gene, is a nuclear protein. Proc. Natl. 15563.
Acad. Sci. USA 95, 1630–1634. Prezant, T.R., Levine, J., Melmed, S., 1998. Molecular characterization of the men1
Gurung, B., Feng, Z., Iwamoto, D.V., Thiel, A., Jin, G., Fan, C.M., et al., 2013. Menin tumor suppressor gene in sporadic pituitary tumors. J. Clin. Endocrinol. Metab.
epigenetically represses Hedgehog signaling in MEN1 tumor syndrome. Cancer 83, 1388–1391.
Res. 73, 2650–2658. Roudaia, L., Speck, N.A., 2008. A MENage a Trois in leukemia. Cancer Cell 14, 3–5.
Hannan, F.M., Nesbit, M.A., Christie, P.T., Fratter, C., Dudley, N.E., Sadler, G.P., et al., Scacheri, P.C., Davis, S., Odom, D.T., Crawford, G.E., Perkins, S., Halawi, M.J., et al.,
2008. Familial isolated primary hyperparathyroidism caused by mutations of 2006. Genome-wide analysis of menin binding provides insights into MEN1
the MEN1 gene. Nat. Clin. Pract. Endocrinol. Metab. 4, 53–58. tumorigenesis. PLoS Genet. 2, e51.
Hao, W., Skarulis, M.C., Simonds, W.F., Weinstein, L.S., Agarwal, S.K., Mateo, C., et al., Scarsbrook, A.F., Thakker, R.V., Wass, J.A., Gleeson, F.V., Phillips, R.R., 2006. Multiple
2004. Multiple endocrine neoplasia type 1 variant with frequent prolactinoma endocrine neoplasia: spectrum of radiologic appearances and discussion of a
and rare gastrinoma. J. Clin. Endocrinol. Metab. 89, 3776–3784. multitechnique imaging approach. Radiographics 26, 433–451.
Heppner, C., Bilimoria, K.Y., Agarwal, S.K., Kester, M., Whitty, L.J., Guru, S.C., et al., Schnepp, R.W., Hou, Z., Wang, H., Petersen, C., Silva, A., Masai, H., et al., 2004.
2001. The tumor suppressor protein menin interacts with NF-kappaB proteins Functional interaction between tumor suppressor menin and activator of S-
and inhibits NF-kappaB-mediated transactivation. Oncogene 20, 4917–4925. phase kinase. Cancer Res. 64, 6791–6796.
Huang, J., Gurung, B., Wan, B., Matkar, S., Veniaminova, N.A., Wan, K., et al., 2012. Schreinemakers, J.M., Pieterman, C.R., Scholten, A., Vriens, M.R., Valk, G.D., Rinkes,
The same pocket in menin binds both MLL and JunD but has opposite effects on I.H., 2011. The optimal surgical treatment for primary hyperparathyroidism in
transcription. Nature 482, 542–546. MEN1 patients: a systematic review. World J. Surg. 35, 1993–2005.
Hughes, C.M., Rozenblatt-Rosen, O., Milne, T.A., Copeland, T.D., Levine, S.S., Lee, J.C., Shen, H.C., Rosen, J.E., Yang, L.M., Savage, S.A., Burns, A.L., Mateo, C.M., et al., 2008.
et al., 2004. Menin associates with a trithorax family histone methyltransferase Parathyroid tumor development involves deregulation of homeobox genes.
complex and with the Hoxc8 locus. Mol. Cell 13, 587–597. Endocr. Relat. Cancer 15, 267–275.
Hussein, N., Casse, H., Fontaniere, S., Morera, A.M., Asensio, M.J., Bakeli, S., et al., Shi, A., Murai, M.J., He, S., Lund, G., Hartley, T., Purohit, T., et al., 2012. Structural
2007. Reconstituted expression of menin in Men1-deficient mouse Leydig insights into inhibition of the bivalent menin-MLL interaction by small
tumour cells induces cell cycle arrest and apoptosis. Eur. J. Cancer 43, 402–414. molecules in leukemia. Blood 120, 4461–4469.
Imamura, M., Komoto, I., Ota, S., Hiratsuka, T., Kosugi, S., Doi, R., et al., 2011. Soares, B.S., Eguchi, K., Frohman, L.A., 2005. Tumor deletion mapping on
Biochemically curative surgery for gastrinoma in multiple endocrine neoplasia chromosome 11q13 in eight families with isolated familial somatotropinoma
type 1 patients. World J. Gastroenterol. 17, 1343–1353. and in 15 sporadic somatotropinomas. J. Clin. Endocrinol. Metab. 90, 6580–6587.

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002
14 R.V. Thakker / Molecular and Cellular Endocrinology xxx (2013) xxx–xxx

Sowa, H., Kaji, H., Hendy, G.N., Canaff, L., Komori, T., Sugimoto, T., et al., 2004. Menin neoplasia type 1 syndrome (MEN1): a case-control study in a series of 77
is required for bone morphogenetic protein 2- and transforming growth factor patients versus 2509 non-MEN1 patients. Am. J. Surg. Pathol. 32, 534–543.
beta-regulated osteoblastic differentiation through interaction with Smads and Trump, D., Farren, B., Wooding, C., Pang, J.T., Besser, G.M., Buchanan, K.D., et al.,
Runx2. J. Biol. Chem. 279, 40267–40275. 1996. Clinical studies of multiple endocrine neoplasia type 1 (MEN1). Qjm 89,
Stalberg, P., Grimfjard, P., Santesson, M., Zhou, Y., Lindberg, D., Gobl, A., et al., 2004. 653–669.
Transfection of the multiple endocrine neoplasia type 1 gene to a human Turner, J.J., Christie, P.T., Pearce, S.H., Turnpenny, P.D., Thakker, R.V., 2010.
endocrine pancreatic tumor cell line inhibits cell growth and affects expression Diagnostic challenges due to phenocopies: lessons from Multiple Endocrine
of JunD, delta-like protein 1/preadipocyte factor-1, proliferating cell nuclear Neoplasia type1 (MEN1). Hum. Mutat. 31, E1089–E1101.
antigen, and QM/Jif-1. J. Clin. Endocrinol. Metab. 89, 2326–2337. Verges, B., Boureille, F., Goudet, P., Murat, A., Beckers, A., Sassolas, G., et al., 2002.
Stratakis, C.A., Schussheim, D.H., Freedman, S.M., Keil, M.F., Pack, S.D., Agarwal, S.K., Pituitary disease in MEN type 1 (MEN1): data from the France–Belgium MEN1
et al., 2000. Pituitary macroadenoma in a 5-year-old: an early expression of multicenter study. J. Clin. Endocrinol. Metab. 87, 457–465.
multiple endocrine neoplasia type 1. J. Clin. Endocrinol. Metab. 85, 4776–4780. Vierimaa, O., Georgitsi, M., Lehtonen, R., Vahteristo, P., Kokko, A., Raitila, A., et al.,
Thakker, R.V., 1998. Multiple endocrine neoplasia–syndromes of the twentieth 2006. Pituitary adenoma predisposition caused by germline mutations in the
century. J. Clin. Endocrinol. Metab. 83, 2617–2620. AIP gene. Science 312, 1228–1230.
Thakker, R.V., 2010. Multiple endocrine neoplasia type 1. In: De Groot, L., Jameson, Waldmann, J., Lopez, C.L., Langer, P., Rothmund, M., Bartsch, D.K., 2010. Surgery for
J.L. (Eds.), Endocrinology, sixth ed. Elsevier, Philadelphia, pp. 2719–2741. multiple endocrine neoplasia type 1-associated primary hyperparathyroidism.
Thakker, R.V., Bouloux, P., Wooding, C., Chotai, K., Broad, P.M., Spurr, N.K., et al., Br. J. Surg. 97, 1528–1534.
1989. Association of parathyroid tumors in multiple endocrine neoplasia type 1 Wang, E.H., Ebrahimi, S.A., Wu, A.Y., Kashefi, C., Passaro Jr., E., Sawicki, M.P., 1998.
with loss of alleles on chromosome 11. N. Engl. J. Med. 321, 218–224. Mutation of the MENIN gene in sporadic pancreatic endocrine tumors. Cancer
Thakker, R.V., Pook, M.A., Wooding, C., Boscaro, M., Scanarini, M., Clayton, R.N., Res. 58, 4417–4420.
1993. Association of somatotrophinomas with loss of alleles on chromosome 11 Yang, Y.J., Song, T.Y., Park, J., Lee, J., Lim, J., Jang, H., et al., 2013. Menin mediates
and with gsp mutations. J. Clin. Invest. 91, 2815–2821. epigenetic regulation via histone H3 lysine 9 methylation. Cell Death Dis. 4,
Thakker, R.V., Newey, P.J., Walls, G.V., Bilezikian, J., Dralle, H., Ebeling, P.R., et al., e583.
2012. Clinical practice guidelines for multiple endocrine neoplasia type 1 Yokoyama, A., Cleary, M.L., 2008. Menin critically links MLL proteins with LEDGF on
(MEN1). J. Clin. Endocrinol. Metab. 97, 2990–3011. cancer-associated target genes. Cancer Cell 14, 36–46.
Tham, E., Grandell, U., Lindgren, E., Toss, G., Skogseid, B., Nordenskjold, M., 2007. Yokoyama, A., Wang, Z., Wysocka, J., Sanyal, M., Aufiero, D.J., Kitabayashi, I., et al.,
Clinical testing for mutations in the MEN1 gene in Sweden: a report on 200 2004. Leukemia proto-oncoprotein MLL forms a SET1-like histone
unrelated cases. J. Clin. Endocrinol. Metab. 92, 3389–3395. methyltransferase complex with menin to regulate Hox gene expression. Mol.
The European Consortium on MEN1, 1997. Identification of the multiple endocrine Cell. Biol. 24, 5639–5649.
neoplasia type 1 (MEN1) gene. Hum. Mol. Genet. 6, 1177–1183. Yumita, W., Ikeo, Y., Yamauchi, K., Sakurai, A., Hashizume, K., 2003. Suppression of
Tonelli, F., Fratini, G., Nesi, G., Tommasi, M.S., Batignani, G., Falchetti, A., et al., 2006. insulin-induced AP-1 transactivation by menin accompanies inhibition of c-Fos
Pancreatectomy in multiple endocrine neoplasia type 1-related gastrinomas induction. Int. J. Cancer 103, 738–744.
and pancreatic endocrine neoplasias. Ann. Surg. 244, 61–70. Zablewska, B., Bylund, L., Mandic, S.A., Fromaget, M., Gaudray, P., Weber, G., 2003.
Trouillas, J., Labat-Moleur, F., Sturm, N., Kujas, M., Heymann, M.F., Figarella-Branger, Transcription regulation of the multiple endocrine neoplasia type 1 gene in
D., et al., 2008. Pituitary tumors and hyperplasia in multiple endocrine human and mouse. J. Clin. Endocrinol. Metab. 88, 3845–3851.

Please cite this article in press as: Thakker, R.V. Multiple endocrine neoplasia type 1 (MEN1) and type 4 (MEN4). Molecular and Cellular Endocrinology
(2013), http://dx.doi.org/10.1016/j.mce.2013.08.002

You might also like