Subscribe to RSS
DOI: 10.1055/a-2337-2265
An Update on the Genetic Drivers of Corticotroph Tumorigenesis
- Abstract
- Introduction
- Somatic genetic variants associated with sporadic corticotroph tumours
- Syndromes of endocrine neoplasia
- Familial cancer predisposition
- Phakomatoses
- Familial isolated corticotrophinoma
- Germline defects with no evidence of familial disease
- “Feedback tumours”
- Conclusions
- References
Abstract
The genetic landscape of corticotroph tumours of the pituitary gland has dramatically changed over the last 10 years. Somatic changes in the USP8 gene account for the most common genetic defect in corticotrophinomas, especially in females, while variants in TP53 or ATRX are associated with a subset of aggressive tumours. Germline defects have also been identified in patients with Cushing’s disease: some are well-established (MEN1, CDKN1B, DICER1), while others are rare and could represent coincidences. In this review, we summarise the current knowledge on the genetic drivers of corticotroph tumorigenesis, their molecular consequences, and their impact on the clinical presentation and prognosis.
#
Keywords
corticotrophinoma - corticotroph tumour - Cushing’s disease - genetic testing - multiple endocrine neoplasia - pituitary adenoma - next generation sequencing - pituitary neuroendocrine tumour - USP8Introduction
Corticotrophinomas, also known as corticotroph tumours, are infrequent and usually benign pituitary neuroendocrine tumours (PitNETs). Derived from TPIT-expressing progenitors, most corticotrophinomas secrete adrenocorticotropic hormone (ACTH) excessively and lead to Cushing’s disease (CD), although some are clinically silent [1]. The untreated CD is associated with higher morbidity and mortality, and therefore, timely diagnosis and efficient treatment are critical [2] [3] [4] [5]; diagnosis is delayed by an average of 3.2 years [6]. Furthermore, the chronic complications of hypercortisolaemia result in an impaired quality of life that might persist even after successful treatment [7] [8]. The differential diagnosis of Cushing’s syndrome entails a complex series of biochemical and imaging studies [9], although CD is the most common cause of endogenous cortisol excess. Transsphenoidal surgery (TSS) is the first line of treatment and is successful in most cases, but therapeutic options are limited – and often ineffective – in case of postoperative disease persistence or recurrence [10] [11].
A great interest in the genetic defects driving corticotroph tumorigenesis has arisen in the last few years, particularly after the discovery that a large proportion of sporadic corticotrophinomas are driven by somatic variants in a single gene, USP8 [12] [13]. Pathogenic variants in other genes are less common and can be associated with specific phenotypes and clinical courses. Here, we review the current knowledge on the genetic drivers of corticotroph tumorigenesis ([Fig. 1]), their molecular consequences, and their impact on the clinical presentation and prognosis.
#
Somatic genetic variants associated with sporadic corticotroph tumours
Next-generation sequencing (NGS) techniques have allowed the identification of somatic pathogenic variants in a group of genes ([Table 1]). About 40–70% of sporadic cases of CD are due to recurrent somatic variants in hotspot regions of USP8 (15q21.2) or USP48 (1p36.12) [12] [13] [14] [15] [16]. Variants in BRAF (7q34) and NR3C1 (5q31.3) have been described in a few cases, and their actual prevalence is still unknown. Co-occurrence of USP8, USP48, BRAF and NC3R1 variants has never been reported. Additionally, variants in TP53 (17p13.1), ATRX (Xq21.1), and DAXX (6p21.32) are rare in benign adenomas but are overrepresented in corticotrophinomas with aggressive features [16] [17] [18]. Furthermore, other somatic defects, with or without concomitant variants in more prominent genes (such as USP8, USP48, ATRX, NR3C1 and TP53), have also been described; however, their relevance needs to be clarified in further studies (GNAS [19] [20], FAT1, HCFC1, FBL, ANKRD27, STAG2, PKHD1, MXRA5, MAST4, AKAP6 and CHD2 [21]). Somatic MEN1 variants have rarely been identified in corticotroph tumours.
Gene |
Defects in corticotrophinoma* |
Genotyped cases (total screened and prevalence range, if available) |
Clinical features |
---|---|---|---|
USP8 |
Somatic hotspot |
~600 cases (screened: ~2000, prevalence: 0–65%) [12] [13] [16] [18] [22] [23] [24] [25] [27] [28] [29] [30] [31] [33] [34] [37] [44] [49] [50] [51] |
Sex: |
USP48 |
Somatic hotspot |
Fifty-six cases (screened: 577, prevalence: 0–23%) [15] [16] [31] [33] [34] [50] |
|
NR3C1 |
Disperse |
Nine cases (screened: 116, prevalence: 0–6.1%) [13] [15] [16] [33] [35] [56] [57] |
Unknown |
BRAF |
Somatic hotspot (Val600Glu) GOF variant |
Sixteen cases (screened: 517, prevalence: 0–16.5%) [13] [15] [16] [33] [35] [56] [57] |
Unknown |
TP53 |
Somatic LOF variants |
Twenty-one cases (screened: 170, prevalence: 0–33%) [26] [35] [64] [65] [66] [67] [68] [69] |
Higher number of interventions, higher OR for invasion, and shorter overall survival [18] |
ATRX |
Somatic LOF variants |
Ten cases (screened: 41, prevalence: unknown) [16] [17] [26] |
Aggressive and metastatic tumours [17] |
DAXX |
Somatic LOF variants |
Two cases (screened: 18, prevalence: unknown) [16] |
Unknown |
MEN1 |
Germline or somatic LOF variants |
At least 34 cases with germline variants [33] [75] [78] [79] [92] [93] [94] [95] [98] [99] [100] [101] [102] [103] [104] (2.9% out of 245 paediatric cases in one cohort [33]); somatic variant in one case [213] |
MEN1 (might debut as isolated paediatric CD) |
CDKN1B |
Germline LOF variants |
Five cases [125] (1.2% out of 245 paediatric cases in one cohort [33]) |
MEN4, isolated paediatric CD |
SDHA |
Germline LOF variant |
One case (1.2% out of 245 paediatric cases in one cohort [29]) |
Isolated paediatric CD, possibly: 3PAs |
PRKAR1A |
Germline LOF variants |
Three cases [156] [157] [158] (0.4% out of 245 paediatric cases in one cohort [33]) |
CNC |
RET |
Germline GOF variants |
Two cases [142] |
3PAs, MEN2B |
DICER1 |
Germline LOF variants |
DICER1 syndrome, PitB |
|
MSH2 |
Germline LOF variants |
LS, aggressive corticotrophinoma, ACTH-producing metastatic PitNET |
|
MLH1 |
Germline LOF variant |
One case [213] |
LS, aggressive corticotrophinoma |
TSC2** |
Germline or somatic LOF variants |
Four cases, three with germline and one with somatic variants (1.6% out of 245 paediatric cases in one cohort [29]) |
TSC, isolated paediatric CD (somatic variant) |
CABLES1** |
Germline LOF variants |
Five cases (1.2% out of 245 paediatric cases in one cohort [29]) |
Young-onset CD, large corticotrophinoma |
*Only pathogenic and likely pathogenic variants were accounted for. ** Requires further evidence of causality. 3PAs, phaeochromocytoma, paraganglioma and pituitary neuroendocrine tumour; CD, Cushing’s disease; CNC, Carney complex; GOF, gain-of-function; LDDST, low-dose dexamethasone suppression test; LOF, loss-of-function; LS, Lynch syndrome; MEN1, multiple endocrine neoplasia type 1; MEN4, multiple endocrine neoplasia type 4; OR, odds ratio; PitB, pituitary blastoma.
USP8
USP8 codes for a protein with deubiquitinase activity (UniProt P40818) that regulates membrane trafficking and protein turnover through deubiquitination of its substrate proteins. About 40% of sporadic CD tumours carry USP8 variants. The prevalence of USP8 variants in sporadic corticotrophinomas varies between 0–65% depending on the published series, with an average of ~40% [12] [13] [14] [16] [18] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34]. Among ~600 positive CD cases,>99% of the variants are located within a hotspot region coding for the amino acids 713–723 [12] [13] [23] [25] [29] [31] [35], which includes the 14–3–3 binding motif Arg715-Ser716-Tyr717-Ser718-Ser719-Pro720 [36]. Most variants target residues Ser718, Pro720, or both, the most common events being the substitutions p.Ser718Pro, p.Ser719Pro, p.Pro720Arg and p.Pro720Gln, as well as the in-frame deletion p.Ser718del [12] [13] [27] [29] [37]. Variants are present in heterozygosity and only in the tumour DNA, with the exception of a single female patient who carried the USP8 p.Ser719Pro variant as a de novo heterozygous germline defect [38]. This individual developed paediatric CD in the context of a complex syndrome, including developmental delay, dysmorphic features, ichthyosiform hyperkeratosis, chronic lung and kidney disease, dilated cardiomyopathy, hyperinsulinaemia and partial growth hormone (GH) deficiency.
Interaction between USP8 and 14–3–3 is essential for the fine control of USP8 activity, as 14-3-3 retains USP8 in an inactive state in the cytosol [36]. Disruption of that interaction leads to a sustained DUB activity and prevents degradation of its cargo proteins [39]. Hotspot variants also promote cleavage and translocation of USP8 into the nucleus [12] [13]. In addition, a somatic variant upstream of the hotspot (p.Gly664Arg) has been reported in one case. In vitro experiments showed that it reduces USP8/14-3-3 interaction and results in sustained activation of USP8, similar to the hotspot mutants [31]. In corticotroph tumours, the best-characterized mechanism of action of USP8 involves EGFR signalling [13]. EGFR ubiquitination after its activation promotes EGFR endocytosis and lysosomal degradation [39]. USP8 variants reduce EGFR ubiquitination, increase the pool of EGFR in the plasma membrane and enable sustained ERK1/2 signalling, which in turn induces POMC transcription and ACTH secretion [13].
USP8 variants are very specific to corticotrophinomas, as they are absent in ectopic ACTH-producing tumours and other pituitary entities [12] [13] [14] [40] [41] [42]. They are also much more frequent in functional corticotroph tumours than in silent tumours [26] [43] [44]. Some studies suggest that USP8 variants define a specific molecular subgroup of corticotrophinoma, with differential expression of proteins relevant for the corticotroph physiology (cyclin-dependent kinase inhibitor 1B, CDKN1B; cyclin-dependent kinase 5 and Abl enzyme substrate 1, CABLES1; GR, glucocorticoid receptors; somatostatin receptor 5, SST5), proliferation, cell-to-cell junction, and epithelial-to-mesenchymal transition [24] [25] [28] [29] [35] [44] [45] [46]. Tumours with USP8 variants seem to partially retain the original corticotroph physiology, as they exhibit larger suppression of pro-opiomelanocortin (POMC) by dexamethasone test in vitro [29] and better response to desmopressin stimulation test in vivo than the wild type [34]. In vitro experiments have shown that they also respond better to the somatostatin receptor ligand pasireotide [37] [47], although the translation into clinical practice needs further evaluation [44].
USP8 variants are much more common in female patients (range 27–68%) than in males (range 3.5–38%) [12] [13] [14] [16] [22] [23] [27] [28] [29] [43] [48]. It is unclear whether their prevalence is similar in paediatric (range 0%–31%) [23] [30] [33] and adult cases (range 12%–5%) [12] [18] [22] [24] [25] [27] [28] [29] [31] [34] [44] [49] [50], although they seem to be more frequent in teenagers diagnosed at older age [23] [33] and adults at younger age [16] [24] [28] [34]. The association of USP8 mutational status with hormone levels is inconsistent and, in some cases, contradictory. Some authors reported associations with lower plasma ACTH [13] [22] [34], higher ACTH/tumour size ratio [12], lower [49] or higher preoperative 24-hour urinary-free cortisol (UFC) [24], or higher postoperative UFC [43]. Nevertheless, most studies could not show significant differences between wild-type tumours and those with USP8 variants. Several publications suggest that tumours with USP8 variants tend to have a more benign phenotype, such as smaller size [12] [16] [22], lower invasion rates into the cavernous sinus [12] [18] or higher chances of achieving postoperative remission [26] [27]. Other studies have found evidence for a higher risk of invasion [51] or higher risk of recurrence after remission [23] [24] [31] and have reported USP8 variants in difficult-to-manage cases, including few patients with aggressive tumours and pituitary carcinomas [32] [48].
#
USP48
Variants in USP48, which encodes another deubiquitinase (UniProt Q86UV5), were discovered by NGS in USP8 wild-type corticotroph tumours [15] [16]. USP48 variants have been described in 1.5–23% of cases, all USP8 wild type [15] [16] [31] [33] [34] [50]. Similar to USP8, USP48 variants are more frequent in female patients [15] [16]. One study found tumours with USP48 variants to be significantly smaller [16], while another reported a higher rate of cavernous sinus invasion compared to wild-type tumours [50]. All somatic variants identified affect a single amino acid (Met415) and confer higher deubiquitinase activity to the mutant USP48 protein. A potential mechanism of action in corticotroph tumour cells may be through the stabilizing effect of USP48 on GLI1, a downstream transcription factor effector of the hedgehog pathway [52], which contributes to corticotroph pathophysiology via its crosstalk with CRH signalling [53] [54]. In a corticotroph tumour cell model, overexpression of a USP48 mutant enhanced CRH-induced (but not basal) POMC promoter activity in a GLI1-dependent manner, suggesting that it may promote corticotroph tumorigenesis by amplifying the trophic action of CRH [16].
#
NR3C1
Variants in the NR3C1 (nuclear receptor subfamily 3 group C member 1) gene that encodes for the glucocorticoid receptor (UniProt P04150) were initially described as rare events [55]. NGS studies showed somatic NR3C1 variants in ~6% of cases, but no clinical correlations have been established [13] [15] [16] [33] [35] [56] [57]. NR3C1 loss-of-function (LOF) in the corticotroph cells should impair the response to the negative adrenal feedback, although the contribution of this specific genetic defect to corticotroph tumorigenesis is unclear, given its low prevalence [56]. In addition, at the germline level, this genetic defect causes generalized glucocorticoid resistance [58]. Interestingly, germline LOF NR3C1 variants have been documented in two cases of CD, including a young man with generalized glucocorticoid resistance and a dominant-negative de novo variant, as well as a paediatric female patient carrying a germline likely pathogenic variant and no additional clinical features [33] [59].
#
BRAF
The oncogenic BRAF variant p.Val600Glu was reported in 16% of USP8 wild-type corticotrophinomas in one cohort [15], but not in other Asian, American or European cohorts [16] [31] [32] [33] [34] [50]. This hotspot variant is a frequent gain-of-function (GOF) defect in cancer that results in a constitutively active RAS-RAF-MEK-ERK pathway, and it has been exploited as a therapeutic target in various neoplasms [60] [61] [62] [63]. The evidence so far indicates that the BRAF p.Val600Glu variant is extremely rare or absent in the majority of adult and paediatric CD patient cohorts.
#
TP53
Variants in the TP53 tumour suppressor gene were initially considered rare events described only in isolated cases of aggressive and metastatic corticotrophinomas [64] [65] [66] [67] [68] [69]. NGS on selected cohorts of USP8 wild-type, aggressive and/or metastatic corticotroph tumours revealed somatic TP53 variants in up to 33% of cases [16] [17]. A large multicentre study on functional corticotrophinomas reported TP53 variants in 14% of all functional macroadenomas and in 24% of invasive cases [18]. TP53 variants correlated with higher Knosp grades and parasellar invasion, more therapeutic procedures and increased disease-specific mortality. In a case of progressive corticotroph tumour growth after bilateral adrenalectomy (Nelson syndrome), a TP53 variant was detected in the specimen obtained after radiation, but not in the pre-radiated surgical specimens, and it was attributed to the mutagenic action of radiation [67]. However, it should be noted that TP53 variants have also been detected in specimens obtained prior to radiation [18]. The majority of cases report TP53 variants already at first surgery and in both primary tumour and metastases, suggesting an early event in pituitary tumour progression [18] [69]. In addition, a single case of CD due to a microadenoma in a female paediatric patient carrying a germline likely pathogenic TP53 variant with a Li-Fraumeni-like family history has recently been described [33].
#
ATRX and DAXX
ATRX (alpha thalassemia/mental retardation syndrome X-linked, UniProt P46100) and DAXX (death-domain-associated protein, UniProt Q9UER7) are involved in chromatin remodelling and alternative lengthening of telomeres [70]. Inactivating ATRX and DAXX variants are commonly found in pancreatic neuroendocrine neoplasms (PanNENs) and are mutually exclusive [71]. ATRX and DAXX bind to TP53 and in several cancers, ATRX/DAXX variants can be found concomitantly with TP53 variants. ATRX variants were reported in aggressive and metastatic corticotrophinoma [17] [26] [69] [72]. Their prevalence in metastatic corticotroph tumours is estimated at ~30% [17]. In almost half of the reported cases, ATRX variants were accompanied by TP53 (or TP53 plus PTEN/NF1/2) variants. DAXX variants were found in two USP8 wild type aggressive corticotroph tumours, which also had TP53 variants [16]. In some reports, they were described in both primary tumour and metastasis, while in others, only in metastasis [17] [69] [72]. Strong ATRX immunopositivity was found in the majority of corticotroph tumours, and a lack of nuclear ATRX immunostaining was observed in several cases of corticotroph tumours carrying ATRX variants [17] [72] [73].
#
#
Syndromes of endocrine neoplasia
Multiple endocrine neoplasia type 1
This entity (MIM #131100) consists of the association of various endocrine and nonendocrine neoplasms, mainly primary hyperparathyroidism (PHPT, 75–100% of patients), gastrointestinal and PanNENs (41–75%), and PitNETs (30–65%) [74] [75] [76] [77] [78] [79] [80]. Less constant associations include adrenocortical tumours, phaeochromocytomas, bronchopulmonary and thymic NENs, lipomas, angiofibromas, collagenomas, meningiomas, leiomyomas, hibernomas ependymomas, and breast cancer [81]. Half of cases display autosomal dominant (AD) inheritance, of which 70–93% are caused by LOF germline and rarely mosaic, heterozygous MEN1 variants (11q13.1) [79] [80] [82] [83] [84] [85] [86] [87] [88]. The same genetic defect underlies only one-third of sporadic cases, and de novo variants are rare [79] [84] [85] [89]. MEN1 (UniProt O00255–2) is a 610 amino acid predominantly nuclear scaffold protein that regulates gene transcription, genome stability, and cell proliferation, thus acting as a classic tumour suppressor in multiple tissues [90] [91].
A pituitary tumour is the first manifestation of MEN1 in 12–32% of patients, with prolactinomas accounting for two-thirds of these lesions [75] [76] [77] [79] [92] [93] [94] [95]. According to most studies, MEN1 LOF-associated PitNETs present at younger ages, are more often plurihormonal and behave more aggressively than non-MEN1 tumours [92] [93] [95] [96] [97]. Only ̴4% of MEN1-associated pituitary tumours are corticotrophinomas, yet MEN1 LOF is the most common germline genetic cause of CD (2.9% of cases in a paediatric cohort), with at least 34 genotyped cases reported in the literature [33] [75] [78] [79] [92] [93] [94] [95] [98] [99] [100] [101] [102] [103] [104]. Similar to humans, Men1 knockout mice develop pituitary tumours with clear predominance of prolactin and prolactin/GH-expressing tumours [105]. In one study, however, 3% of heterozygous Men1 exon 1–2 deletion mice developed ACTH-expressing pituitary tumours, accounting for 10% of all PitNETs. In contrast, that histotype was not observed in eight previously reported knockout lines [106].
The documented MEN1-associated CD cases include an equal proportion of males and females (n=17 each) [33] [92] [93] [94] [95] [98] [99] [100] [101] [102] [103] [104]. Out of the cases with available data, 61.8% (n=21) were caused by microadenomas, 17.6% (n=6) were due to macroadenomas, and multiple tumours were found in 11.8% (n=4). Of the four patients with multiple tumours, three developed microprolactinomas, and one had a nonfunctioning gonadotroph macroadenoma [99] [101] [103]. A single case of CD caused by a plurihormonal macroadenoma expressing ACTH, prolactin (PRL), GH, and luteinizing hormone (LH) has been published [102]. In most other cases with available data, tumours were monohormonal, displayed no particular imaging or histopathological features and responded well to surgery. CD was diagnosed in childhood or adolescence in 34.4% of cases (n=11) and was the first manifestation of MEN1 in 35.3% (n=12) [33] [92] [93] [98] [100] [103]. Cases presenting during adulthood (n=8 with available data) were diagnosed at age 40.9±10.1 years [93] [94] [99] [101] [102]. Because simplex MEN1 cases have been identified among patients with paediatric CD, genetic testing could be helpful for risk identification, early diagnosis of other disease components, and genetic counselling. Hypercortisolaemia in the setting of MEN1 is usually due to a corticotroph tumour but is caused by primary adrenal disease in one-fifth of cases and is rarely secondary to ACTH secretion from other NENs [101].
#
Multiple endocrine neoplasia type 4
Caused by germline LOF germline heterozygous CDKN1B (12p13.1) variants, this syndrome (MIM #610755) explains ̴2% of MEN1 cases negative for MEN1 defects [107] [108] [109]. Manifestations highly vary among CDKN1B variant carriers, but AD inheritance with incomplete penetrance is well-proven [110]. The most common component of multiple endocrine neoplasia type 4 (MEN4) is PHPT, although renal angiomyolipomas, adrenal nonfunctioning tumours, uterine fibroids, gastrinomas, gastric carcinomas, gastrointestinal and PanNENs, neuroendocrine cervical carcinomas, bronchial neuroendocrine tumours (NEN), and papillary thyroid carcinoma have also been described [107] [108] [109] [111] [112] [113] [114] [115] [116]. CDKN1B (UniProt P46527) is a negative regulator of multiple cyclin-dependent kinase/cyclin complexes that inhibits the progression from the G1 to the S phase of the cell cycle [117] [118].
Low CDKN1B expression is a common feature of human corticotroph tumours, and Cdkn1b knockout mice develop fully penetrant ACTH-secreting hyperplasia or tumours of the pituitary pars intermedia, but no other PitNETs [119] [120] [121] [122] [123] [124]. In contrast, corticotroph tumours account for only 36.8% (n=7) out of the 19 pituitary tumours associated with germline CDKN1B variants in the literature (reviewed in [125]). Presentation was sporadic in 85.7% (n=6) of these cases, and an equal proportion of cases was explained by tumours<10 mm. Onset in adulthood and additional MEN4-associated tumours were documented in 28.6% of cases (n=2), while 71.4% (n=5) presented with isolated paediatric CD [108] [110] [126]. CDKN1B variants associated with these cases were classified as pathogenic or likely pathogenic, except for two missense defects that were variants of uncertain significance (VUS). Functional analyses rendered mixed results for both VUS; one of them was a previously known germline change in familial isolated pituitary adenoma (FIPA) and familial cancer predisposition, and somatic defect in other neoplasms [127] [128] [129].
#
Phaeochromocytoma and paraganglioma syndromes
An infrequent phenotype recently termed as “three P association” (3PAs) consists of the association of a pituitary tumour with a phaeochromocytoma or paraganglioma (PPGL) in a single patient or in different members of the same family [130] [131]. Germline defects in the PPGL-associated genes SDHA, SDHB, SDHC, SDHD, and SDHAF2 (namely SDHx genes), explain ̴40% of the genetically tested 3PAs, as well as rare cases of isolated pituitary tumours [33] [132] [133]. These genes encode the four subunits and the assembly factor 2 of succinate dehydrogenase (SDH), a mitochondrial enzyme that participates in the Krebs cycle and in oxidative phosphorylation [134]. SDH dysfunction leads to tumorigenesis via pseudohypoxia, reactive oxygen species, and defective apoptosis and DNA methylation [134] [135] [136] [137].
At least 25 cases of PitNETs associated with germline pathogenic or likely pathogenic SDHx variants have been documented [33] [138]. The affected gene was SDHB in 12, SDHD in six, SDHA in five, and SDHC in two cases. Sixteen tumours were prolactinomas, four were somatotrophinomas, two were nonfunctioning pituitary tumours, one was a metastatic PitNET of gonadotroph origin, one was a corticotrophinoma, and the tumour type was not reported for one case. The presentation was 3PAs in 20 cases (80%) and isolated in the rest. Loss-of-heterozygosity (LOH) was detected in seven out of eleven cases where it was analysed, and two macroprolactinomas due to an SDHB defect occurred in a single family [130]. SDHx-driven pituitary tumours display a particular histopathological pattern characterized by cytoplasmic vacuoles of variable size [130] [139]. Concordantly, Sdhb +/- mice develop hyperplasia of GH and prolactin-producing cells with nuclear inclusions or pseudoinclusions [131]. The nature of these features remains unclear but could be explained by the engulfment of SDHx-associated dysfunctional mitochondria [140].
Four cases of CD presenting with a 3PAs phenotype have been published. One patient carried a pathogenic RET variant, two had negative genetic tests (for SDHx in one case and SDHx, MEN1, and VHL in the other), and one was not tested but had a family history compatible with MEN2A [131] [141] [142] [143]. More recently, a case of isolated paediatric CD (age 8 years) caused by a pathogenic germline SDHA frameshift variant was reported. The tumour was a 7 mm corticotrophinoma with reduced SDHA, SDHB and SDHD immunostaining and LOH at the variant locus. The variant was of maternal origin, but there were no other cases of pituitary tumours or PPGLs in the family [33]. In addition, three germline SDHA VUS were found in three different patients with isolated paediatric CD. One germline SDHD VUS was found in two other paediatric CD patients; one of them had a family history compatible with FIPA, but the cosegregation was not proven [33] [131]. Functional studies are required to assess the pathogenic potential of these VUS.
#
Carney complex (CNC)
CNC is a rare syndrome with endocrine and cardiocutaneous manifestations [144]. LOF PRKAR1A (coding for protein kinase A, PKA, regulatory subunit-1-alpha 17q24.2, MIM #160980) variants are identified in three-quarters of cases, either as de novo defects or inherited in an AD manner [145] [146]. While one of the most prominent and frequent manifestations of the disease is adrenal Cushing’s syndrome, CNC has only recently been associated with CD [147] [148]. In the pituitary, the cAMP/PKA pathway is involved in cell differentiation and in response to secretagogues and trophic factors in most cell types [149], but alterations most often affect the somatotroph lineage. Single micro or macroadenomas or multiple microadenomas of somatotroph, lactotroph or mammosomatotroph origin, often with surrounding hyperplastic areas, with or without LOH, are characteristic of CNC [150] [151] [152]. PRKAR1A variants have not been detected as somatic changes in sporadic PitNETs [145] [153] [154] [155].
Only three cases of CD in patients with CNC have been reported in the literature, and thus, LOF PRKAR1A variants remain a very rare cause of CD. The three individuals had apparently sporadic presentation but carried germline frameshift PRKAR1A variants and developed additional manifestations of CNC, including PPNAD in one case [156] [157] [158]. The de novo origin of the variant was confirmed in one individual [158]. In one female patient, CD was highly suspected at age three years, but a corticotroph tumour was not proven [156]. The other two patients were males diagnosed with CD at ages 17 and 31 years due to<10 mm corticotrophinomas with LOH at the variant locus [157] [158]. The scarcity of reported cases, however, could be misleading since the diagnosis of CD in this context is challenging, given the high penetrance of PPNAD [159]. Together with McCune-Albright syndrome, MEN1, and, very rarely, MEN type 2 (MEN2, a possible coincidence), CNC could be a cause for “double Cushing’s” [157].
#
Multiple endocrine neoplasia type 2
GOF variants in RET (10q11.2), coding for a transmembrane receptor tyrosine kinase, are the underlying cause of all MEN2 phenotypes: MEN2A, MEN2B and familial medullary cancer (95–98% of cases) both as inherited or de novo forms [160] [161] [162] [163] [164]. Although pituitary tumours are not considered a component of MEN2, the RET protein, as a dependence receptor, plays an important role in the maintenance of somatotroph cell numbers in normal circumstances [165]. The occurrence of pituitary tumours in germline RET variant carriers has only been documented in four cases in the literature [142] [166] [167] [168], which may represent coincidences. Corticotroph tumours accounted for two of these cases and in both individuals, CD preceded typical MEN2 manifestations. The first patient was a 68-year-old male with MEN2A due to a pathogenic RET variant who also developed a phaeochromocytoma, PHPT and MTC [142]. He had a biochemical diagnosis of CD and achieved remission after TSS, but a corticotroph tumour was neither observed by magnetic resonance imaging nor histologically confirmed. A second surgery was performed 15 years later due to recurrence, again with no histological confirmation [142]. The second patient was a male who carried a pathogenic RET variant and was diagnosed with CD at age 13 years. On examination, coarse facial features, mucosal neuromas, and a marfanoid habitus were noticed, and he ultimately developed MTC. Of note, a corticotrophinoma was not observed by imaging studies and was only confirmed after a third TSS, after which he finally achieved remission [168]. Patients with MEN2 phenotypes without genetic testing or negative for RET variants have also been reported [132]. Other causes of hypercortisolaemia are rare among MEN2 patients, although ectopic ACTH secretion from an MTC may occur [169].
#
#
Familial cancer predisposition
DICER1 syndrome
The DICER1 syndrome, DICER1 tumour predisposition syndrome, DICER1 pleuropulmonary blastoma (PPB) familial tumour predisposition syndrome, or PPB familial tumour and dysplasia syndrome (MIM #601200) is an AD entity including MEN and nonendocrine cancer [170]. The characteristic features of the syndrome include dysembryonic tumours that are extremely rare outside this setting, such as PPB, cystic nephroma, ovarian sex cord-stromal tumour, nasal chondromesenchymal hamartoma, ciliary body and cerebral medulloepithelioma, anaplastic kidney sarcoma, pineoblastoma, embryonal rhabdomyosarcoma, and pituitary blastoma (PitB) [170]. Wilms tumour, juvenile hamartomatous intestinal polyps, and pulmonary cysts can also occur, but the most common manifestation of DICER syndrome is thyroid disease, in the form of differentiated thyroid carcinoma, or childhood-onset multinodular goitre. Some of the manifestations have a characteristic age of onset [171] [172].
Around 70% of cases are due to heterozygous germline LOF DICER1 (14q32.13) variants appearing de novo in 10–20% of patients, while somatic mosaic variants have been detected in a further 10% [171] [172] [173] [174]. The latter defects are associated with an extremely rare phenotype known as GLOW syndrome (global developmental delay, lung cysts, overgrowth, and Wilms tumour) [174] [175] [176]. Germline DICER1 variants are most often accompanied by somatic second hits; somatic defects also occur in sporadic tumours [176] [177] [178] [179] [180] [181] [182] [183] [184] [185] [186] [187]. Germline variants are not clustered in hotspots, but mosaic and somatic defects are missense variants that are usually located within the RNase IIIb domain of the protein [170] [176] [184] [188] [189]. DICER1 participates in the processing of mature microRNAs (miRNAs) and small interfering RNAs (siRNAs); its inactivation leads to reduced expression of 5′-derived mature miRNAs, which greatly alters the expression of miRNA targets [187] [189] [190] [191].
PitB, a poorly differentiated ACTH-expressing anterior pituitary neoplasm with an oncofetal molecular signature, has a very low penetrance (<1%) [184] [192] [193]. Most PitBs are diagnosed in neonates or infants as CD or silent pituitary tumours, but two cases have been reported in childhood and young adulthood [194] [195] [196]. Nineteen out of the twenty PitBs so far genotyped were due to LOF DICER1 variants, but it is not clear if any of them were caused by somatic defects [172] [194] [195] [196]. In most cases, PitB was the first manifestation of the syndrome, and nine of these patients died during infancy or childhood [195] [196]. Because PitB is considered a pathognomonic lesion of the DICER1 syndrome, its diagnosis should prompt germline DICER1 screening [172]. These tumours overexpress PRAME, which has been exploited as a target for immunotherapy in other neoplasms [197]. Aside from PitB, a causal association of DICER1 variants with sporadic PitNETs has not been proven [198].
#
Lynch syndrome
Also known as hereditary non-polyposis colorectal cancer (CRC), this AD syndrome is defined by an increased risk for developing CRC, endometrium, ovary, gastric, small bowel, urinary tract, biliary tract, brain, skin, pancreas, and prostate cancer [199]. Germline heterozygous LOF defects in the DNA mismatch repair (MMR) genes MLH1 (3p22.2, MIM #609310), MSH2 (2p21-p16.3, MIM #120435), MSH6 (2p16.3, #614350), and PMS2 (7p22.1, MIM #614337), usually accompanied by somatic second hits, as well as deletions of the last exons of EPCAM (2p21, MIM #613244, causing MSH2 silencing) underlie this phenotype [200] [201] [202] [203] [204] [205] [206] [207] [208]. The MMR system excises base-base mismatches to increase the fidelity of DNA replication and is involved in the response to mechanisms of DNA damage; its failure leads to increased spontaneous mutagenesis and microsatellite instability [209] [210].
Eight cases of pituitary tumours presenting in carriers of germline MMR defects have been documented in the literature, affecting patients within the fourth and seventh decades of life [133] [211] [212] [213] [214] [215]. These cases include one macroprolactinoma, one aggressive prolactinoma, one aggressive nonfunctioning PitNET, one undifferentiated sellar carcinoma, and three ACTH-producing PitNETs. The tumour type was not specified in one case reported in a register-based study [211]. Four patients had MSH2 variants, one carried an MSH6 variant, another one had both genes affected, and individual cases were due to PMS2 and MLH1 variants. Out of the three ACTH-producing tumours, one was a metastatic PitNET and two were aggressive corticotrophinomas [133] [212] [213]. One of the latter tumours displayed LOH at the variant locus (MSH2), while somatic MEN1 and MSH6 variants were detected in the other case [133] [213]. Interestingly, in one of these patients the corticotroph tumour was the first manifestation of LS [133]. Pituitary tumours seem to be causally associated with MMR defects, and although they remain rare LS-associated neoplasms, their unusually aggressive potential in this context should be kept in mind.
#
#
Phakomatoses
Tuberous sclerosis complex (TSC)
The occurrence of hamartomatous lesions affecting brain, skin, heart, lungs, and kidneys, as well as neurological complications such as seizures and behavioural, psychiatric, intellectual, academic, neuropsychological, and psychosocial difficulties characterizes TSC, with a range of tumorous and cutaneous manifestations [216] [217]. TSC is classified as TSC1 (MIM #191100) or TSC2 (MIM #613254), depending on its causative genetic defect: LOF heterozygous variants in TSC1 (9q34.13) or TSC2 (16p13.3), respectively [218] [219]. An AD pattern of inheritance is present in one-third of patients, while the rest are simplex cases due to de novo variants [216]. Germline TSC1 or TSC2 variants are found in around 70–90% of all familial or simplex TSC patients, while germline or somatic mosaic defects cause a small proportion of cases [217] [220] [221] [222]. TSC1 and TSC2 are part of a protein complex that negatively regulates the serine/threonine kinase MTOR and its LOF leads to an overactive MTORC1. [223] [224] [225] [226] [227]. Aside from their role in TSC, somatic defects in these genes have been observed in NENs and urothelial, bladder, and renal cancer [227].
PitNETs have been reported only in eight patients with TSC, while, interestingly, in a rat model with a naturally occurring Tsc2 LOF variant, pituitary tumours are observed with a frequency of 58%, with LOH in one out of three tumours analysed [228]. The human cases include one silent gonadotrophinoma, two somatotrophinomas, one suspected but unconfirmed prolactinoma, and four corticotrophinomas [229] [230] [231] [232] [233] [234]. The CD cases included one young adult (macroadenoma) and three paediatric patients (microadenomas) [33] [232] [233]. Evidence of a pathogenic or likely pathogenic TSC2 variant was available for three cases, and LOH was documented in one out of two cases where it was investigated [33] [233]. Two individuals carrying likely pathogenic variants had no family or personal history of TSC at ages 10 and 15 years. In addition, a pathogenic variant was reported as a somatic change in an 8 mm corticotrophinoma from a paediatric individual with CD [33].
#
PTEN hamartoma tumour syndrome
The PTEN hamartoma tumour syndrome encompasses a spectrum of phenotypes, including Cowden syndrome, MIM #158350), Lhermitte-Duclos disease, Bannayan-Riley-Ruvalcaba syndrome, PTEN-related Proteus syndrome, and PTEN-related Proteus-like syndrome [235]. Cowden syndrome is a rare AD disorder characterized by dermal trichilemmomas and papillomatous papules, bowel, breast and thyroid hamartomatosis, macrocephaly, and high lifelong risk for breast, endometrial, renal colorectal, and non-medullary thyroid cancer [236].
At least four reports of pituitary tumours in patients with CS exist in the literature [237] [238] [239] [240]. The causative defect was determined in only one case a 29-year-old female patient with a macroprolactinoma who carried a germline pathogenic PTEN variant [239]. Another patient developed a microprolactinoma and a paraganglioma, but genetic testing failed to identify a causative defect in PTEN, SDHB, SDHC, and SDHD [237]. The tumour type was not specified in one individual [238]. The most recent report was of an ACTH-producing metastatic pituitary tumour in a 52-year-old female patient with Cowden syndrome without clinical signs of CD [240]. She was concurrently diagnosed with meningiomas, and had a history of macrocephaly, multiple mucocutaneous lesions, multinodular goitre and a breast fibroadenoma, but no genetic test was conducted in this individual. The clinical association of corticotrophinomas with phakomatoses remains a rare occurrence and additional research is required to assess causality.
#
#
Familial isolated corticotrophinoma
The diagnosis of pituitary tumours in two or more members of the same family in the absence of other clinical features is termed FIPA (MIM #102200) and accounts for 2–4% of all PitNETs [241]. Three-quarters of FIPA-associated pituitary tumours are somatotrophinomas, prolactinomas, or mixed GH/prolactin-producing tumours that arise at a younger age compared with sporadic cases [242] [243] [244]. Corticotroph tumours occur in only 7.5% of FIPA families and kindreds with only this type of tumours are extremely rare (1.4%) [245]. The genetic cause of FIPA seems to imply multiple genes and remains unknown in most cases.
Germline LOF AIP (11q13.2) variants underlie 15–20% of all FIPA cases but 25–50% of pituitary tumours in families with exclusively somatotrophinomas, displaying AD inheritance and incomplete penetrance [243] [245] [246]. AIP defects also explain one-third of all cases of gigantism and 6–20% of young-onset pituitary tumours [245] [247] [248]. AIP is a co-chaperone protein involved in pathways relevant for pituitary tumorigenesis, such as cAMP/PKA and RET-dependent apoptosis [249] [250] [251] [252] [253]. Its role in neoplasia is complex and entails both tumour suppression and oncogenesis, depending on the cellular context [254] [255]. Most pituitary tumours associated with AIP defects are sparsely granulated macroadenomas with increased macrophage infiltration, presenting clinically as GH excess, with prolactinomas representing 10% of cases [243] [245] [256]. Five cases of CD associated with germline AIP variants (reference sequence NM_003977.4) have been published, but none of these cases supports the role of AIP in pure corticotroph tumorigenesis [245] [257] [258] [259] [260] [261]. All cases have apparently sporadic presentations. Three variants, p.Lys103Arg in one paediatric case [100], as well as p.Arg304Gln [257] (with allele frequency 0.1%, and four homozygotes in gnomAD v4.0) and p.Arg16His [262] (allele frequency 0.2% and five homozygotes in gnomAD v4.0), found in adults, are VUS or likely benign variants [245]. A fourth variant (p.Lys58Asn, classified as VUS) has been identified in a patient with an apparently multihormonal pituitary tumour secreting both prolactin and ACTH. This patient responded well to dopamine analogue treatment with normalised prolactin and cortisol levels and reduction in tumour size over 14.4 years of follow-up [261]. A fifth case (p.Leu251Argfs*52) had a similar clinical course (prolactin and ACTH staining, response to cabergoline) [258]. Furthermore, a rodent model of biallelic Aip deficiency did not develop corticotroph adenomas, while PIT1 lineage tumours (GH, PRL and thyroid stimulating hormone) were all present [263].
Tandem microduplications including GPR101 (Xq26.3), are the cause of X-linked acrogigantism (MIM #300942), an infrequent form of GH excess with onset in infancy or early childhood [264]. These copy number variants are detected constitutively, or as somatic mosaicism, and although most X-LAG cases occur sporadically, FIPA has been documented [265]. GPR101 defects have not been implicated in CD [266].
Germline heterozygous LOF CDH23 (10q22.1) variants are another reported underlying cause of FIPA (MIM #617540), accounting for one-third of FIPA families and 12% of sporadic pituitary tumours in one cohort [267]. This study reported four rare germline CDH23 VUS in an equal number of individuals with CD, but no experimental validation was performed. This genetic defect requires further confirmation.
Another molecular defect possibly contributing to FIPA was recently proposed, after finding germline missense PAM variants in three first-degree relatives with gigantism and in sporadic patients with functioning pituitary tumours [232]. This gene encodes an enzyme catalysing the C-terminal amidation of secreted peptides; LOF was demonstrated in vitro for multiple variants [268] [269]. Among these cases, two patients with isolated paediatric CD carried a frameshift and a 5′-UTR variant that displayed deleterious effects. A frameshift variant carrier family member displayed elevated midnight serum cortisol but no clinical features of CD. A second study found additional deleterious variants in sporadic patients with functioning PitNETs, including two corticotrophinomas causing cyclical CD [270]. PAM variants seem to result in pituitary hypersecretion, but their role in pituitary tumorigenesis requires further clarification.
#
Germline defects with no evidence of familial disease
CABLES1 variants
CABLES1 (18q11.2) encodes a negative regulator of cell cycle progression that facilitates the interaction of tyrosine kinases with their substrates, thereby modulating crucial phosphorylation cascades [271] [272] [273]. CABLES1 also stabilizes and prevents the degradation of cell-cycle regulators and interacts with TP53 and TP73 to trigger apoptosis [274]. Inactivation of the mouse orthologue Cables1 promotes tumour formation as well as cell proliferation and survival in vitro [ 275]. Although knockout mice are prone to developing colon and endometrial cancer, Cables1 might not be a cancer driver per se [276] [277]. In corticotroph cells, Cables1 is one of the main transactivated genes in response to glucocorticoids, making it a possible mediator of the regulatory adrenal-pituitary feedback loop [278].
The above-mentioned data, plus the observation that CABLES1 immunostaining was often reduced in human corticotrophinomas, prompted the search for CABLES1 genetic defects [278]. Five germline heterozygous missense CABLES1 variants have been identified in five sporadic corticotroph tumours [33] [279]. In vitro, these variants displayed either reduced capacity to block cell proliferation under dexamethasone treatment (four variants) or reduced half-life (one variant) [33] [279] [280]. Three variants were confirmed to be germline, one with proven inheritance, but there was no somatic LOH in any of these cases. No coexisting hotspot somatic drivers of corticotrophinomas were detected either. Interestingly, all patients (three adults and two children) developed aggressive, either functioning (three cases) or silent (two cases) corticotroph tumours. Further research is required to fully define this recently proposed association.
#
#
“Feedback tumours”
In patients with primary adrenal insufficiency, ACTH levels rise due to a lack of adrenal-pituitary negative feedback. This process may, on rare occasions, lead to symptomatic pituitary hyperplasia and even potentially corticotroph tumour formation, as it has been documented in patients with Addison’s disease, the most common cause of primary adrenal insufficiency [281] [282] [283] [284] [285] [286] [287] [288] [289] [290]. While in the past, Addison’s disease was often caused by tuberculosis or Waterhouse–Friderichsen syndrome (adrenal haemorrhage due to severe infection), today, this condition is usually due to destruction of the adrenal cortex by the adaptive immune system, which in most cases is mediated by autoantibodies against CYP21A2 [291]. A complex genetic background underlies this entity, and it is unknown if any of its predisposition loci is also involved in pituitary tumorigenesis. Corticotrophinomas may also occur in patients with monogenic causes of adrenal insufficiency, such as congenital adrenal hyperplasia (CAH) or X-linked adrenal hypoplasia congenita.
Congenital adrenal hyperplasia
This diagnosis includes multiple entities characterized by enzymatic deficiencies involving adrenal steroidogenesis. Over 95% of the cases are due to LOF CYP21A2 (6p21.33, MIM #201910) variants with autosomal recessive inheritance and rarely occurring de novo, while the rest of the cases are due to CYP11B1, HSD3B2, CYP17A1, CYP11A1, POR, and STAR defects [292]. The clinical presentation ranges from the classic salt-wasting and simple-virilising forms in patients with severe CYP21A2 deficiency to the non-classic presentation, characterized by milder degrees of hypocortisolaemia and hyperandrogenaemia. At least four cases of corticotrophinomas presenting in patients with CAH have been documented, all affecting women within the third and fourth decades of life [293] [294] [295]. One patient was homozygous for a CYP21A2 variant associated with non-classic CAH, another one was a compound heterozygous for variants associated with classic and non-classic CAH, and a third one carried a heterozygous variant associated with severe CYP21A2 deficiency. In these three cases, the diagnosis of CAH was established during the diagnostic workup for CD and all had pituitary microadenomas, with histological confirmation in two cases [294] [295]. The patient with no genetic test had a history of ambiguous genitalia and presented with compressive symptoms of a pituitary macroadenoma, confirmed as a corticotrophinoma [293]. The clinical presentation of CD in these patients was characterized by pronounced stigmata of hyperandrogenaemia and variable degrees of hypercortisolaemia, depending on the degree of CYP21A2 deficiency; interestingly, the patient with virilising CAH developed the largest tumour. Notwithstanding these data, it is likely that additional molecular disruptions have contributed to driving corticotroph tumorigenesis in these rare cases.
#
X-linked adrenal hypoplasia congenita (X-linked AHC)
X-linked AHC (MIM #300200) consists of adrenal hypoplasia with primary adrenal insufficiency associated with hypogonadotropic hypogonadism [296]. Two-thirds of patients develop acute infantile adrenal insufficiency, while in the rest, the adrenal insufficiency appears in childhood and, rarely, in early adulthood [297] [298] [299]. Central hypogonadism is diagnosed most often due to delayed or arrested puberty [300] [301] [302]. Most patients are males carrying hemizygous LOF variants in NR0B1 (Xp21.2); heterozygous females are rarely affected [303] [304] [305]. NR0B1 encodes an orphan nuclear receptor with a role in the development of the adrenal glands and the hypothalamic-pituitary-gonadal axis, mainly via repressing steroidogenesis [306]. A single case of a corticotroph tumour associated with a germline frameshift NR0B1 variant has been reported [307]. This male patient had pre-existing adrenal insufficiency, primary hypothyroidism, and hypogonadotropic hypogonadism. He was diagnosed with an invasive corticotrophinoma causing CD at age 33 years, for which three TSS and radiotherapy were required to achieve remission. Maternal inheritance of the genetic defect was proven, but no other affected family members were identified. The causative role of the primary adrenal insufficiency and, therefore, the theoretically heightened stimulatory effect on the central regulation of ACTH is unclear.
#
#
Conclusions
A large proportion of cases of CD are explained by a single somatic driver, while the remaining are due to a heterogeneous genetic background involving multiple germline and somatic defects. Most of our knowledge on the genetic basis of this complex entity derives from research conducted over the last decade. For USP8 variants, most of the available data point towards a relatively benign and treatment-responsive clinical phenotype. In contrast, TP53, ATRX, and likely Lynch syndrome-related gene defects translate into aggressive behaviour. Defining the impact of other individual genetic defects on the pathogenesis of corticotrophinomas remains a pending matter. For instance, it is unclear whether germline defects lead to specific clinical features in CD, other than younger disease onset and the risk of developing other neoplasms in some cases. Further research efforts should focus on developing evidence-based guidelines for genetic testing in CD, as well as on identifying and characterising potential pharmacological targets. For now, the risk of pathogenic/likely pathogenic germline variants is low in patients with corticotrophinomas if no other syndromic manifestation is present. Therefore, germline genetic testing is only suggested in children. Regarding somatic testing, in aggressive macroadenomas, ATRX immunostaining could be informative, but genetic assessment of somatic DNA of the tumour is not recommended outside of clinical research settings in a recent publication, Lin et al. identified TP53 variants or extensive LOH in 9/14 treatment-refractory corticotroph tumours, six of them carrying concomitant ATRX (four) or DAXX (two) alterations [309].
#
#
Conflict of Interest
The authors declare that they have no conflict of interest.
-
References
- 1 Mete O, Grossman A, Yamada S. et al. Pituitary Tumours: Anterior Pituitary Neuroendocrine Tumours (PitNETs)/ PitNETs of TPIT Lineage/ Corticotroph PitNET/Adenoma. In: Osamura RY, Asa SL, eds. WHO Classification of Tumours Editorial Board Endocrine and neuroendocrine tumours. 5 edn. Lyon (France): International Agency for Research on Cancer; 2022
- 2 Etxabe J, Vázquez JA. Morbidity and mortality in Cushing's disease: An epidemiological approach. Clin Endocrinol (Oxf) 1994; 40: 479-484
- 3 Lindholm J, Juul S, Jorgensen JO. et al. Incidence and late prognosis of Cushing's syndrome: A population-based study. J Clin Endocrinol Metab 2001; 86: 117-123
- 4 Clayton RN, Jones PW, Reulen RC. et al. Mortality in patients with Cushing's disease more than 10 years after remission: A multicentre, multinational, retrospective cohort study. Lancet Diabetes Endocrinol 2016; 4: 569-576
- 5 Ragnarsson O, Olsson DS, Papakokkinou E. et al. Overall and disease-specific mortality in patients with Cushing disease: A Swedish nationwide study. J Clin Endocrinol Metab 2019; 104: 2375-2384
- 6 Rubinstein G, Osswald A, Hoster E. et al. Time to diagnosis in Cushing's syndrome: A meta-analysis based on 5367 patients. J Clin Endocrinol Metab 2020; 105: e12-e22
- 7 Lindsay JR, Nansel T, Baid S. et al. Long-term impaired quality of life in Cushing's syndrome despite initial improvement after surgical remission. J Clin Endocrinol Metab 2006; 91: 447-453
- 8 Espinosa-de-Los-Monteros AL, Sosa E, Martinez N. et al. Persistence of Cushing's disease symptoms and comorbidities after surgical cure: A long-term, integral evaluation. Endocr Pract 2013; 19: 252-258
- 9 Fleseriu M, Auchus R, Bancos I. et al. Consensus on diagnosis and management of Cushing's disease: A guideline update. Lancet Diabetes Endocrinol 2021; 9: 847-875
- 10 Feelders RA, Newell-Price J, Pivonello R. et al. Advances in the medical treatment of Cushing's syndrome. Lancet Diabetes Endocrinol 2019; 7: 300-312
- 11 Fleseriu M, Varlamov EV, Hinojosa-Amaya JM. et al. An individualized approach to the management of Cushing disease. Nat Rev Endocrinol 2023; 19: 581-599
- 12 Ma ZY, Song ZJ, Chen JH. et al. Recurrent gain-of-function USP8 mutations in Cushing's disease. Cell Res 2015; 25: 306-317
- 13 Reincke M, Sbiera S, Hayakawa A. et al. Mutations in the deubiquitinase gene USP8 cause Cushing's disease. Nat Genet 2015; 47: 31-38
- 14 Song ZJ, Reitman ZJ, Ma ZY. et al. The genome-wide mutational landscape of pituitary adenomas. Cell Res 2016; 26: 1255-1259
- 15 Chen J, Jian X, Deng S. et al. Identification of recurrent USP48 and BRAF mutations in Cushing's disease. Nat Commun 2018; 9: 3171
- 16 Sbiera S, Pérez-Rivas LG, Taranets L. et al. Driver mutations in USP8 wild-type Cushing's disease. Neuro Oncol 2019; 21: 1273-1283
- 17 Casar-Borota O, Boldt HB, Engstrom BE. et al. Corticotroph aggressive pituitary tumors and carcinomas frequently harbor ATRX mutations. J Clin Endocrinol Metab 2021; 106: 1183-1194
- 18 Pérez-Rivas LG, Simon J, Albani A. et al. TP53 mutations in functional corticotroph tumors are linked to invasion and worse clinical outcome. Acta Neuropathol Commun 2022; 10: 139
- 19 Williamson EA, Ince PG, Harrison D. et al. G-protein mutations in human pituitary adrenocorticotrophic hormone-secreting adenomas. Eur J Clin Invest 1995; 25: 128-131
- 20 Riminucci M, Collins MT, Lala R. et al. An R201H activating mutation of the GNAS1 (Gsalpha) gene in a corticotroph pituitary adenoma. Mol Pathol 2002; 55: 58-60
- 21 Sbiera S, Kunz M, Weigand I. et al. The new genetic landscape of Cushing's disease: Deubiquitinases in the spotlight. Cancers (Basel) 2019; 11
- 22 Hayashi K, Inoshita N, Kawaguchi K. et al. The USP8 mutational status may predict drug susceptibility in corticotroph adenomas of Cushing's disease. Eur J Endocrinol 2016; 174: 213-226
- 23 Faucz FR, Tirosh A, Tatsi C. et al. Somatic USP8 gene mutations are a common cause of pediatric Cushing disease. J Clin Endocrinol Metab 2017; 102: 2836-2843
- 24 Albani A, Pérez-Rivas LG, Dimopoulou C. et al. The USP8 mutational status may predict long-term remission in patients with Cushing's disease. Clin Endocrinol (Oxf) 2018; 89: 454-458
- 25 Ballmann C, Thiel A, Korah HE. et al. USP8 mutations in pituitary Cushing adenomas-targeted analysis by next-generation sequencing. J Endocr Soc 2018; 2: 266-278
- 26 Bujko M, Kober P, Boresowicz J. et al. USP8 mutations in corticotroph adenomas determine a distinct gene expression profile irrespective of functional tumour status. Eur J Endocrinol 2019; 181: 615-627
- 27 Losa M, Mortini P, Pagnano A. et al. Clinical characteristics and surgical outcome in USP8-mutated human adrenocorticotropic hormone-secreting pituitary adenomas. Endocrine 2019; 63: 240-246
- 28 Weigand I, Knobloch L, Flitsch J. et al. Impact of USP8 gene mutations on protein deregulation in Cushing disease. J Clin Endocrinol Metab 2019; 104: 2535-2546
- 29 Sesta A, Cassarino MF, Terreni M. et al. Ubiquitin-specific protease =8 mutant corticotrope adenomas present unique secretory and molecular features and shed light on the role of ubiquitylation on ACTH processing. Neuroendocrinology 2020; 110: 119-129
- 30 Pasternak-Pietrzak K, Faucz FR, Stratakis CA. et al. Is there a common cause for paediatric Cushing's disease?. Endokrynol Pol 2021; 72: 104-107
- 31 Treppiedi D, Barbieri AM, Di Muro G. et al. Genetic profiling of a cohort of Italian patients with ACTH-secreting pituitary tumors and characterization of a novel USP8 gene variant. Cancers (Basel) 2021; 13: 4022
- 32 Andonegui-Elguera S, Silva-Roman G, Pena-Martinez E. et al. The genomic landscape of corticotroph tumors: From silent adenomas to ACTH-secreting carcinomas. Int J Mol Sci 2022; 23: 4861
- 33 Hernández-Ramírez LC, Pankratz N, Lane J. et al. Genetic drivers of Cushing's disease: Frequency and associated phenotypes. Genet Med 2022; 24: 2516-2525
- 34 Shichi H, Fukuoka H, Kanzawa M. et al. Responsiveness to DDAVP in Cushing's disease is associated with USP8 mutations through enhancing AVPR1B promoter activity. Pituitary 2022; 25: 496-507
- 35 Neou M, Villa C, Armignacco R. et al. Pangenomic classification of pituitary neuroendocrine tumors. Cancer Cell 2020; 37: 123-134.e125
- 36 Mizuno E, Kitamura N, Komada M. 14-3-3-dependent inhibition of the deubiquitinating activity of UBPY and its cancellation in the M phase. Exp Cell Res 2007; 313: 3624-3634
- 37 Albani A, Pérez-Rivas LG, Tang S. et al. Improved pasireotide response in USP8 mutant corticotroph tumours in vitro. Endocr Relat Cancer 2022; 29: 503-511
- 38 Cohen M, Persky R, Stegemann R. et al. Germline USP8 mutation associated with pediatric Cushing disease and other clinical features: A new syndrome. J Clin Endocrinol Metab 2019; 104: 4676-4682
- 39 Mizuno E, Iura T, Mukai A. et al. Regulation of epidermal growth factor receptor down-regulation by UBPY-mediated deubiquitination at endosomes. Mol Biol Cell 2005; 16: 5163-5174
- 40 Ronchi CL, Peverelli E, Herterich S. et al. Landscape of somatic mutations in sporadic GH-secreting pituitary adenomas. Eur J Endocrinol 2016; 174: 363-372
- 41 Bi WL, Horowitz P, Greenwald NF. et al. Landscape of genomic alterations in pituitary adenomas. Clin Cancer Res 2017; 23: 1841-1851
- 42 Pérez-Rivas LG, Osswald A, Knosel T. et al. Expression and mutational status of USP8 in tumors causing ectopic ACTH secretion syndrome. Endocr Relat Cancer 2017; 24: L73-L77
- 43 Pérez-Rivas LG, Theodoropoulou M, Ferrau F. et al. The gene of the ubiquitin-specific protease 8 is frequently mutated in adenomas causing Cushing's disease. J Clin Endocrinol Metab 2015; 100: E997-E1004
- 44 Castellnou S, Vasiljevic A, Lapras V. et al. SST5 expression and USP8 mutation in functioning and silent corticotroph pituitary tumors. Endocr Connect 2020; 9: 243-253
- 45 Mossakowska BJ, Rusetska N, Konopinski R. et al. The expression of cell cycle-related genes in USP8-mutated corticotroph neuroendocrine pituitary tumors and their possible role in cell cycle-targeting treatment. Cancers (Basel) 2022; 14: 5594
- 46 Kober P, Rusetska N, Mossakowska BJ. et al. The expression of glucocorticoid and mineralocorticoid receptors in pituitary tumors causing Cushing's disease and silent corticotroph tumors. Front Endocrinol (Lausanne) 2023; 14: 1124646
- 47 Treppiedi D, Marra G, Di Muro G. et al. P720R USP8 mutation is associated with a better responsiveness to pasireotide in ACTH-secreting PitNETs. Cancers (Basel) 2022; 14: 2455
- 48 Pérez-Rivas LG, Theodoropoulou M, Puar TH. et al. Somatic USP8 mutations are frequent events in corticotroph tumor progression causing Nelson's tumor. Eur J Endocrinol 2018; 178: 59-65
- 49 Wanichi IQ, de Paula Mariani BM, Frassetto FP. et al. Cushing's disease due to somatic USP8 mutations: A systematic review and meta-analysis. Pituitary 2019; 22: 435-442
- 50 Abraham AP, Pai R, Beno DL. et al. USP8, USP48, and BRAF mutations differ in their genotype-phenotype correlation in Asian Indian patients with Cushing's disease. Endocrine 2022; 75: 549-559
- 51 Tatsi C, Pankratz N, Lane J. et al. Large genomic aberrations in corticotropinomas are associated with greater aggressiveness. J Cli Endocrinol Metab 2019; 104: 1792-1801
- 52 Zhou A, Lin K, Zhang S. et al. Gli1-induced deubiquitinase USP48 aids glioblastoma tumorigenesis by stabilizing Gli1. EMBO Rep 2017; 18: 1318-1330
- 53 Vila G, Papazoglou M, Stalla J. et al. Sonic hedgehog regulates CRH signal transduction in the adult pituitary. FASEB J 2005; 19: 281-283
- 54 Vila G, Theodoropoulou M, Stalla J. et al. Expression and function of sonic hedgehog pathway components in pituitary adenomas: Evidence for a direct role in hormone secretion and cell proliferation. J Clin Endocrinol Metab 2005; 90: 6687-6694
- 55 Karl M, Von Wichert G, Kempter E. et al. Nelson's syndrome associated with a somatic frame shift mutation in the glucocorticoid receptor gene. J Clin Endocrinol Metab 1996; 81: 124-129
- 56 Miao H, Liu Y, Lu L. et al. Effect of 3 NR3C1 mutations in the pathogenesis of pituitary ACTH adenoma. Endocrinology 2021; 162: bqab167
- 57 Theodoropoulou M. Glucocorticoid receptors are making a comeback in corticotroph tumorigenesis. Endocrinology 2022; 163: bqab257
- 58 Hurley DM, Accili D, Stratakis CA. et al. Point mutation causing a single amino acid substitution in the hormone binding domain of the glucocorticoid receptor in familial glucocorticoid resistance. J Clin Invest 1991; 87: 680-686
- 59 Karl M, Lamberts SW, Koper JW. et al. Cushing's disease preceded by generalized glucocorticoid resistance: Clinical consequences of a novel, dominant-negative glucocorticoid receptor mutation. Proc Assoc Am Physicians 1996; 108: 296-307
- 60 Davies H, Bignell GR, Cox C. et al. Mutations of the BRAF gene in human cancer. Nature 2002; 417: 949-954
- 61 Ikenoue T, Hikiba Y, Kanai F. et al. Functional analysis of mutations within the kinase activation segment of B-Raf in human colorectal tumors. Cancer Res 2003; 63: 8132-8137
- 62 Wellbrock C, Ogilvie L, Hedley D. et al. V599EB-RAF is an oncogene in melanocytes. Cancer Res 2004; 64: 2338-2342
- 63 Proietti I, Skroza N, Michelini S. et al. BRAF inhibitors: Molecular targeting and immunomodulatory actions. Cancers (Basel) 2020; 12: 1823
- 64 Levy A, Hall L, Yeudall WA. et al. p53 gene mutations in pituitary adenomas: Rare events. Clin Endocrinol (Oxf) 1994; 41: 809-814
- 65 Tanizaki Y, Jin L, Scheithauer BW. et al P53 gene mutations in pituitary carcinomas. Endocr Pathol 2007; 18: 217-222 [doi]
- 66 Kawashima ST, Usui T, Sano T. et al P53 gene mutation in an atypical corticotroph adenoma with Cushing's disease. Clin Endocrinol (Oxf) 2009; 70: 656-657 CEN3404 [pii] [doi]
- 67 Pinto EM, Siqueira SA, Cukier P. et al. Possible role of a radiation-induced p53 mutation in a Nelson's syndrome patient with a fatal outcome. Pituitary 2011; 14: 400-404
- 68 Saeger W, Mawrin C, Meinhardt M. et al. Two pituitary neuroendocrine tumors (PitNETs) with very high proliferation and TP53 mutation – High-grade PitNET or PitNEC?. Endocr Pathol 2022; 33: 257-262
- 69 Sumislawski P, Rotermund R, Klose S. et al. ACTH-secreting pituitary carcinoma with TP53, NF1, ATRX and PTEN mutations Case report and review of the literature. Endocrine 2022; 76: 228-236
- 70 Dyer MA, Qadeer ZA, Valle-Garcia D. et al. ATRX and DAXX: Mechanisms and mutations. Cold Spring Harb Perspect Med 2017; 7: a026567
- 71 Jiao Y, Shi C, Edil BH. et al. DAXX/ATRX, MEN1, and mTOR pathway genes are frequently altered in pancreatic neuroendocrine tumors. Science 2011; 331: 1199-1203
- 72 Casar-Borota O, Botling J, Granberg D. et al. Serotonin, ATRX, and DAXX expression in pituitary adenomas: Markers in the differential diagnosis of neuroendocrine tumors of the sellar region. Am J Surg Pathol 2017; 41: 1238-1246
- 73 Alzoubi H, Minasi S, Gianno F. et al. Alternative lengthening of telomeres (ALT) and telomerase reverse transcriptase promoter methylation in recurrent adult and primary pediatric pituitary neuroendocrine tumors. Endocr Pathol 2022; 33: 494-505
- 74 Benson L, Ljunghall S, Akerstrom G. et al. Hyperparathyroidism presenting as the first lesion in multiple endocrine neoplasia type 1. Am J Med 1987; 82: 731-737
- 75 Trump D, Farren B, Wooding C. et al. Clinical studies of multiple endocrine neoplasia type 1 (MEN1). QJM 1996; 89: 653-669
- 76 Carty SE, Helm AK, Amico JA. et al The variable penetrance and spectrum of manifestations of multiple endocrine neoplasia type 1. Surgery 1998; 124: 1106-1113 discussion 1113-1104
- 77 Machens A, Schaaf L, Karges W. et al. Age-related penetrance of endocrine tumours in multiple endocrine neoplasia type 1 (MEN1): A multicentre study of 258 gene carriers. Clin Endocrinol (Oxf) 2007; 67: 613-622
- 78 Sakurai A, Suzuki S, Kosugi S. et al. Multiple endocrine neoplasia type 1 in Japan: Establishment and analysis of a multicentre database. Clin Endocrinol (Oxf) 2012; 76: 533-539
- 79 Giusti F, Cianferotti L, Boaretto F. et al. Multiple endocrine neoplasia syndrome type 1: Institution, management, and data analysis of a nationwide multicenter patient database. Endocrine 2017; 58: 349-359
- 80 Romanet P, Mohamed A, Giraud S. et al. UMD-MEN1 database: An overview of the 370 MEN1 variants present in 1676 patients from the French population. J Clin Endocrinol Metab 2019; 104: 753-764
- 81 Waguespack SG. Beyond the "3 Ps": A critical appraisal of the non-endocrine manifestations of multiple endocrine neoplasia type 1. Front Endocrinol (Lausanne) 2022; 13: 1029041
- 82 Chandrasekharappa SC, Guru SC, Manickam P. et al. Positional cloning of the gene for multiple endocrine neoplasia-type 1. Science 1997; 276: 404-407
- 83 Lemmens I, Van de Ven WJ, Kas K. et al. Identification of the multiple endocrine neoplasia type 1 (MEN1) gene. The European Consortium on MEN1. Hum Mol Genet 1997; 6: 1177-1183
- 84 Cebrian A, Ruiz-Llorente S, Cascon A. et al. Mutational and gross deletion study of the MEN1 gene and correlation with clinical features in Spanish patients. J Med Genet 2003; 40: e72
- 85 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 2008; 29: 22-32
- 86 de Laat JM, van der Luijt RB, Pieterman CR. et al. MEN1 redefined, a clinical comparison of mutation-positive and mutation-negative patients. BMC Med 2016; 14: 182
- 87 Beijers H, Stikkelbroeck NML, Mensenkamp AR. et al. Germline and somatic mosaicism in a family with multiple endocrine neoplasia type 1 (MEN1) syndrome. Eur J Endocrinol 2019; 180: K15-K19
- 88 Coppin L, Ferriere A, Crepin M. et al. Diagnosis of mosaic mutations in the MEN1 gene by next generation sequencing. Eur J Endocrinol 2019; 180: L1-L3
- 89 Schaaf L, Pickel J, Zinner K. et al. Developing effective screening strategies in multiple endocrine neoplasia type 1 (MEN 1) on the basis of clinical and sequencing data of German patients with MEN 1. Exp Clin Endocrinol Diabetes 2007; 115: 509-517
- 90 La P, Desmond A, Hou Z. et al. Tumor suppressor menin: The essential role of nuclear localization signal domains in coordinating gene expression. Oncogene 2006; 25: 3537-3546
- 91 Matkar S, Thiel A, Hua X. Menin: A scaffold protein that controls gene expression and cell signaling. Trends Biochem Sci 2013; 38: 394-402
- 92 Verges B, Boureille F, Goudet P. et al. Pituitary disease in MEN type 1 (MEN1): Data from the France-Belgium MEN1 multicenter study. J Clin Endocrinol Metab 2002; 87: 457-465
- 93 de Laat JM, Dekkers OM, Pieterman CR. et al. Long-term natural course of pituitary tumors in patients with MEN1: Results from the DutchMEN1 study group (DMSG). J Clin Endocrinol Metab 2015; 100: 3288-3296
- 94 Wu Y, Gao L, Guo X. et al. Pituitary adenomas in patients with multiple endocrine neoplasia type 1: A single-center experience in China. Pituitary 2019; 22: 113-123
- 95 Le Bras M, Leclerc H, Rousseau O. et al. Pituitary adenoma in patients with multiple endocrine neoplasia type 1: A cohort study. Eur J Endocrinol 2021; 185: 863-873
- 96 Trouillas J, Labat-Moleur F, Sturm N. et al. Pituitary tumors and hyperplasia in multiple endocrine neoplasia type 1 syndrome (MEN1): A case-control study in a series of 77 patients versus 2509 non-MEN1 patients. Am J Surg Pathol 2008; 32: 534-543
- 97 Farrell WE, Azevedo MF, Batista DL. et al. Unique gene expression profile associated with an early-onset multiple endocrine neoplasia (MEN1)-associated pituitary adenoma. J Clin Endocrinol Metab 2011; 96: E1905-E1914
- 98 Rix M, Hertel NT, Nielsen FC. et al. Cushing's disease in childhood as the first manifestation of multiple endocrine neoplasia syndrome type 1. Eur J Endocrinol 2004; 151: 709-715
- 99 Al Brahim NY, Rambaldini G, Ezzat S. et al. Complex endocrinopathies in MEN-1: Diagnostic dilemmas in endocrine oncology. Endocr Pathol 2007; 18: 37-41
- 100 Stratakis CA, Tichomirowa MA, Boikos S. et al. The role of germline AIP, MEN1, PRKAR1A, CDKN1B and CDKN2C mutations in causing pituitary adenomas in a large cohort of children, adolescents, and patients with genetic syndromes. Clin Genet 2010; 78: 457-463
- 101 Simonds WF, Varghese S, Marx SJ. et al. Cushing's syndrome in multiple endocrine neoplasia type 1. Clin Endocrinol (Oxf) 2012; 76: 379-386
- 102 Uraki S, Ariyasu H, Doi A. et al. Hypersecretion of ACTH and PRL from pituitary adenoma in MEN1, adequately managed by medical therapy. Endocrinol Diabetes Metab Case Rep 2017; 2017: 17-0027
- 103 Makri A, Bonella MB, Keil MF. et al. Children with MEN1 gene mutations may present first (and at a young age) with Cushing disease. Clin Endocrinol (Oxf) 2018; 89: 437-443
- 104 Herath M, Parameswaran V, Thompson M. et al. Paediatric and young adult manifestations and outcomes of multiple endocrine neoplasia type 1. Clin Endocrinol (Oxf) 2019; 91: 633-638
- 105 Agarwal SK. Exploring the tumors of multiple endocrine neoplasia type 1 in mouse models for basic and preclinical studies. Int J Endocr Oncol 2014; 1: 153-161
- 106 Harding B, Lemos MC, Reed AA. et al. Multiple endocrine neoplasia type 1 knockout mice develop parathyroid, pancreatic, pituitary and adrenal tumours with hypercalcaemia, hypophosphataemia and hypercorticosteronaemia. Endocr Relat Cancer 2009; 16: 1313-1327
- 107 Pellegata NS, Quintanilla-Martinez L, Siggelkow H. et al. Germ-line mutations in p27Kip1 cause a multiple endocrine neoplasia syndrome in rats and humans. Proc Natl Acad Sci U S A 2006; 103: 15558-15563
- 108 Georgitsi M, Raitila A, Karhu A. et al. Germline CDKN1B/p27Kip1 mutation in multiple endocrine neoplasia. J Clin Endocrinol Metab 2007; 92: 3321-3325
- 109 Agarwal SK, Mateo CM, Marx SJ. Rare germline mutations in cyclin-dependent kinase inhibitor genes in multiple endocrine neoplasia type 1 and related states. J Clin Endocrinol Metab 2009; 94: 1826-1834
- 110 Frederiksen A, Rossing M, Hermann P. et al. Clinical features of multiple endocrine neoplasia type 4: Novel pathogenic variant and review of published cases. J Clin Endocrinol Metab 2019; 104: 3637-3646
- 111 Molatore S, Marinoni I, Lee M. et al. A novel germline CDKN1B mutation causing multiple endocrine tumors: Clinical, genetic and functional characterization. Hum Mutat 2010; 31: E1825-E1835
- 112 Belar O, De la Hoz C, Pérez-Nanclares G. et al. Novel mutations in MEN1, CDKN1B and AIP genes in patients with multiple endocrine neoplasia type 1 syndrome in Spain. Clin Endocrinol (Oxf) 2012; 76: 719-724
- 113 Malanga D, De GS, Riccardi M. et al. Functional characterization of a rare germline mutation in the gene encoding the cyclin-dependent kinase inhibitor p27Kip1 (CDKN1B) in a Spanish patient with multiple endocrine neoplasia (MEN)-like phenotype. Eur J Endocrinol 2012; 166: 551-560
- 114 Occhi G, Regazzo D, Trivellin G. et al. A novel mutation in the upstream open reading frame of the CDKN1B gene causes a MEN4 phenotype. PLoS Genet 2013; 9: e1003350
- 115 Pardi E, Mariotti S, Pellegata NS. et al. Functional characterization of a CDKN1B mutation in a Sardinian kindred with multiple endocrine neoplasia type 4 (MEN4). Endocr Connect 2015; 4: 1-8
- 116 Sambugaro S, Di Ruvo M, Ambrosio MR. et al. Early onset acromegaly associated with a novel deletion in CDKN1B 5'UTR region. Endocrine 2015; 49: 58-64
- 117 Polyak K, Kato JY, Solomon MJ. et al. p27Kip1, a cyclin-Cdk inhibitor, links transforming growth factor-beta and contact inhibition to cell cycle arrest. Genes Dev 1994; 8: 9-22
- 118 Chu IM, Hengst L, Slingerland JM. The Cdk inhibitor p27 in human cancer: Prognostic potential and relevance to anticancer therapy. Nat Rev Cancer 2008; 8: 253-267
- 119 Fero ML, Rivkin M, Tasch M. et al. A syndrome of multiorgan hyperplasia with features of gigantism, tumorigenesis, and female sterility in p27(Kip1)-deficient mice. Cell 1996; 85: 733-744 S0092-8674(00)81239-8
- 120 Kiyokawa H, Kineman RD, Manova-Todorova KO. et al. Enhanced growth of mice lacking the cyclin-dependent kinase inhibitor function of p27(Kip1). Cell 1996; 85: 721-732 S0092-8674(00)81238-6
- 121 Nakayama K, Ishida N, Shirane M. et al. Mice lacking p27(Kip1) display increased body size, multiple organ hyperplasia, retinal dysplasia, and pituitary tumors. Cell 1996; 85: 707-720 S0092-8674(00)81237-4
- 122 Dahia PL, Aguiar RC, Honegger J. et al. Mutation and expression analysis of the p27/kip1 gene in corticotrophin-secreting tumours. Oncogene 1998; 16: 69-76
- 123 Lidhar K, Korbonits M, Jordan S. et al. Low expression of the cell cycle inhibitor p27Kip1 in normal corticotroph cells, corticotroph tumors, and malignant pituitary tumors. J Clin Endocrinol Metab 1999; 84: 3823-3830
- 124 Korbonits M, Chahal HS, Kaltsas G. et al. Expression of phosphorylated p27(Kip1) protein and Jun activation domain-binding protein 1 in human pituitary tumors. J Clin Endocrinol Metab 2002; 87: 2635-2643
- 125 Singeisen H, Renzulli MM, Pavlicek V. et al. Multiple endocrine neoplasia type 4: A new member of the MEN family. Endocr Connect 2023; 12: e220411
- 126 Chasseloup F, Pankratz N, Lane J. et al. Germline CDKN1B loss-of-function variants cause pediatric Cushing's disease with or without an MEN4 phenotype. J Clin Endocrinol Metab 2020; 105: 1983-2005
- 127 Pappa V, Papageorgiou S, Papageorgiou E. et al. A novel p27 gene mutation in a case of unclassified myeloproliferative disorder. Leuk Res 2005; 29: 229-231
- 128 Tichomirowa MA, Lee M, Barlier A. et al. Cyclin dependent kinase inhibitor 1B (CDKN1B) gene variants in AIP mutation-negative familial isolated pituitary adenomas (FIPA) kindreds. Endocr Relat Cancer 2012; 19: 233-241
- 129 Ruiz-Heredia Y, Sánchez-Vega B, Onecha E. et al. Mutational screening of newly diagnosed multiple myeloma patients by deep targeted sequencing. Haematologica 2018; 103: e544-e548
- 130 Denes J, Swords F, Rattenberry E. et al. Heterogeneous genetic background of the association of pheochromocytoma/paraganglioma and pituitary adenoma: Results from a large patient cohort. J Clin Endocrinol Metab 2015; 100: E531-E541
- 131 Xekouki P, Szarek E, Bullova P. et al. Pituitary adenoma with paraganglioma/pheochromocytoma (3PAs) and succinate dehydrogenase defects in humans and mice. J Clin Endocrinol Metabolism 2015; 100: E710-E719
- 132 O'Toole SM, Denes J, Robledo M. et al. 15 years of paraganglioma: The association of pituitary adenomas and phaeochromocytomas or paragangliomas. Endocr Relat Cancer 2015; 22: T105-T122
- 133 Loughrey PB, Baker G, Herron B. et al. Invasive ACTH-producing pituitary gland neoplasm secondary to MSH2 mutation. Cancer Genet 2021; 256-257: 36-39
- 134 Bezawork-Geleta A, Rohlena J, Dong L. et al. Mitochondrial complex II: At the crossroads. Trends Biochem Sci 2017; 42: 312-325
- 135 Dahia PL, Ross KN, Wright ME. et al. A HIF1alpha regulatory loop links hypoxia and mitochondrial signals in pheochromocytomas. PLoS Genet 2005; 1: 72-80
- 136 Guzy RD, Sharma B, Bell E. et al. Loss of the SdhB, but Not the SdhA, subunit of complex II triggers reactive oxygen species-dependent hypoxia-inducible factor activation and tumorigenesis. Mol Cell Biol 2008; 28: 718-731
- 137 Lopez-Jimenez E, Gomez-Lopez G, Leandro-Garcia LJ. et al. Research resource: Transcriptional profiling reveals different pseudohypoxic signatures in SDHB and VHL-related pheochromocytomas. Mol Endocrinol 2010; 24: 2382-2391
- 138 Loughrey PB, Roncaroli F, Healy E. et al. Succinate dehydrogenase and MYC-associated factor X mutations in pituitary neuroendocrine tumours. Endocr Relat Cancer 2022; 29: R157-R172
- 139 Tufton N, Roncaroli F, Hadjidemetriou I. et al. Pituitary carcinoma in a patient with an SDHB mutation. Endocr Pathol 2017; 28: 320-325
- 140 Nakajima A, Kurihara H, Yagita H. et al. Mitochondrial extrusion through the cytoplasmic vacuoles during cell death. J Biol Chem 2008; 283: 24128-24135
- 141 Steiner AL, Goodman AD, Powers SR. Study of a kindred with pheochromocytoma, medullary thyroid carcinoma, hyperparathyroidism and Cushing's disease: Multiple endocrine neoplasia, type 2. Medicine (Baltimore) 1968; 47: 371-409
- 142 Naziat A, Karavitaki N, Thakker R. et al. Confusing genes: A patient with MEN2A and Cushing's disease. Clin Endocrinol (Oxf) 2013; 78: 966-968
- 143 Johnston PC, Kennedy L, Recinos PF. et al. Cushing's disease and co-existing phaeochromocytoma. Pituitary 2016; 19: 654-656
- 144 Carney JA, Gordon H, Carpenter PC. et al. The complex of myxomas, spotty pigmentation, and endocrine overactivity. Medicine (Baltimore) 1985; 64: 270-283
- 145 Kirschner LS, Carney JA, Pack SD. et al. Mutations of the gene encoding the protein kinase A type I-alpha regulatory subunit in patients with the Carney complex. Nat Genet 2000; 26: 89-92
- 146 Bertherat J, Horvath A, Groussin L. et al. Mutations in regulatory subunit type 1A of cyclic adenosine 5'-monophosphate-dependent protein kinase (PRKAR1A): Phenotype analysis in 353 patients and 80 different genotypes. J Clin Endocrinol Metab 2009; 94: 2085-2091
- 147 Stratakis CA, Kirschner LS, Carney JA. Clinical and molecular features of the Carney complex: Diagnostic criteria and recommendations for patient evaluation. J Clin Endocrinol Metab 2001; 86: 4041-4046
- 148 Rothenbuhler A, Stratakis CA. Clinical and molecular genetics of Carney complex. Best Pract Res Clin Endocrinol Metab 2010; 24: 389-399
- 149 Hernández-Ramírez LC, Trivellin G, Stratakis CA. Cyclic 3',5'-adenosine monophosphate (cAMP) signaling in the anterior pituitary gland in health and disease. Mol Cell Endocrinol 2018; 463: 72-86
- 150 Pack SD, Kirschner LS, Pak E. et al. Genetic and histologic studies of somatomammotropic pituitary tumors in patients with the "complex of spotty skin pigmentation, myxomas, endocrine overactivity and schwannomas" (Carney complex). J Clin Endocr Metab 2000; 85: 3860-3865
- 151 Stergiopoulos SG, Abu-Asab MS, Tsokos M. et al. Pituitary pathology in Carney complex patients. Pituitary 2004; 7: 73-82
- 152 Lonser RR, Mehta GU, Kindzelski BA. et al. Surgical management of Carney complex-associated pituitary pathology. Neurosurgery 2017; 80: 780-786
- 153 Kaltsas GA, Kola B, Borboli N. et al. Sequence analysis of the PRKAR1A gene in sporadic somatotroph and other pituitary tumours. Clin Endocrinol (Oxf) 2002; 57: 443-448
- 154 Sandrini F, Kirschner LS, Bei T. et al. PRKAR1A, one of the Carney complex genes, and its locus (17q22-24) are rarely altered in pituitary tumours outside the Carney complex. J Med Genet 2002; 39: e78
- 155 Yamasaki H, Mizusawa N, Nagahiro S. et al. GH-secreting pituitary adenomas infrequently contain inactivating mutations of PRKAR1A and LOH of 17q23-24. Clin Endocrinol (Oxf) 2003; 58: 464-470
- 156 Basson CT, Aretz HT. Case records of the Massachusetts General Hospital. Weekly clinicopathological exercises. Case 11-2002. A 27-year-old woman with two intracardiac masses and a history of endocrinopathy. N Engl J Med 2002; 346: 1152-1158
- 157 Hernández-Ramírez LC, Tatsi C, Lodish MB. et al. Corticotropinoma as a component of Carney complex. J Endocr Soc 2017; 1: 918-925
- 158 Kiefer FW, Winhofer Y, Iacovazzo D. et al. PRKAR1A mutation causing pituitary-dependent Cushing disease in a patient with Carney complex. Eur J Endocrinol 2017; 177: K7-K12
- 159 Stratakis CA. Clinical genetics of multiple endocrine neoplasias, Carney complex and related syndromes. J Endocrinol Invest 2001; 24: 370-383
- 160 Donis-Keller H, Dou S, Chi D. et al. Mutations in the RET proto-oncogene are associated with MEN 2A and FMTC. Hum Mol Genet 1993; 2: 851-856
- 161 Mulligan LM, Kwok JB, Healey CS. et al. Germ-line mutations of the RET proto-oncogene in multiple endocrine neoplasia type 2A. Nature 1993; 363: 458-460
- 162 Hofstra RM, Landsvater RM, Ceccherini I. et al. A mutation in the RET proto-oncogene associated with multiple endocrine neoplasia type 2B and sporadic medullary thyroid carcinoma. Nature 1994; 367: 375-376
- 163 Eng C. Multiple Endocrine Neoplasia Type 2. In: Adam MP, Mirzaa GM, Pagon RA, eds. GeneReviews. Seattle (WA): University of Washington, Seattle; 1999
- 164 Wang X. Structural studies of GDNF family ligands with their receptors-Insights into ligand recognition and activation of receptor tyrosine kinase RET. Biochim Biophys Acta 2013; 1834: 2205-2212
- 165 Canibano C, Rodriguez NL, Saez C. et al. The dependence receptor Ret induces apoptosis in somatotrophs through a Pit-1/p53 pathway, preventing tumor growth. EMBO J 2007; 26: 2015-2028
- 166 Saito T, Miura D, Taguchi M. et al. Coincidence of multiple endocrine neoplasia type 2A with acromegaly. Am J Med Sci 2010; 340: 329-331
- 167 Heinlen JE, Buethe DD, Culkin DJ. et al. Multiple endocrine neoplasia 2a presenting with pheochromocytoma and pituitary macroadenoma. ISRN Oncol 2011; 2011: 732452
- 168 Kasturi K, Fernandes L, Quezado M. et al. Cushing disease in a patient with multiple endocrine neoplasia type 2B. J Clin Transl Endocrinol Case Rep 2017; 4: 1-4
- 169 Wells SA, Pacini F, Robinson BG. et al. Multiple endocrine neoplasia type 2 and familial medullary thyroid carcinoma: An update. J Clin Endocrinol Metab 2013; 98: 3149-3164
- 170 Schultz KAP, Stewart DR, Kamihara J. et al. DICER1 Tumor Predisposition. In: Adam MP, Mirzaa GM, Pagon RA et al, eds. GeneReviews(R); Seattle (WA): 2020
- 171 Slade I, Bacchelli C, Davies H. et al. DICER1 syndrome: Clarifying the diagnosis, clinical features and management implications of a pleiotropic tumour predisposition syndrome. J Med Genet 2011; 48: 273-278
- 172 Schultz KAP, Williams GM, Kamihara J. et al. DICER1 and associated conditions: Identification of at-risk individuals and recommended surveillance strategies. Clin Cancer Res 2018; 24: 2251-2261
- 173 Hill DA, Ivanovich J, Priest JR. et al. DICER1 mutations in familial pleuropulmonary blastoma. Science 2009; 325: 965
- 174 de Kock L, Wang YC, Revil T. et al. High-sensitivity sequencing reveals multi-organ somatic mosaicism causing DICER1 syndrome. J Med Genet 2016; 53: 43-52
- 175 Klein S, Lee H, Ghahremani S. et al. Expanding the phenotype of mutations in DICER1: Mosaic missense mutations in the RNase IIIb domain of DICER1 cause GLOW syndrome. J Med Genet 2014; 51: 294-302
- 176 Brenneman M, Field A, Yang J. et al. Temporal order of RNase IIIb and loss-of-function mutations during development determines phenotype in pleuropulmonary blastoma / DICER1 syndrome: A unique variant of the two-hit tumor suppression model. F1000Res 2015; 4: 214
- 177 de Boer CM, Eini R, Gillis AM. et al. DICER1 RNase IIIb domain mutations are infrequent in testicular germ cell tumours. BMC Res Notes 2012; 5: 569
- 178 Heravi-Moussavi A, Anglesio MS, Cheng SW. et al. Recurrent somatic DICER1 mutations in nonepithelial ovarian cancers. N Engl J Med 2012; 366: 234-242
- 179 de Kock L, Plourde F, Carter MT. et al. Germ-line and somatic DICER1 mutations in a pleuropulmonary blastoma. Pediatr Blood Cancer 2013; 60: 2091-2092
- 180 Wu MK, Sabbaghian N, Xu B. et al. Biallelic DICER1 mutations occur in Wilms tumours. J Pathol 2013; 230: 154-164
- 181 Tomiak E, de KL, Grynspan D. et al. DICER1 mutations in an adolescent with cervical embryonal rhabdomyosarcoma (cERMS). Pediatr Blood Cancer 2014; 61: 568-569
- 182 de Kock L, Sabbaghian N, Soglio DBD. et al. Exploring the association between DICER1 mutations and differentiated thyroid carcinoma. J Clin Endocr Metab 2014; 99: E1072-E1077
- 183 de Kock L, Sabbaghian N, Druker H. et al. Germ-line and somatic DICER1 mutations in pineoblastoma. Acta Neuropathol 2014; 128: 583-595
- 184 de Kock L, Sabbaghian N, Plourde F. et al. Pituitary blastoma: A pathognomonic feature of germ-line DICER1 mutations. Acta Neuropathol 2014; 128: 111-122
- 185 Doros LA, Rossi CT, Yang J. et al. DICER1 mutations in childhood cystic nephroma and its relationship to DICER1-renal sarcoma. Mod Pathol 2014; 27: 1267-1280
- 186 Murray MJ, Bailey S, Raby KL. et al. Serum levels of mature microRNAs in DICER1-mutated pleuropulmonary blastoma. Oncogenesis 2014; 3: e87
- 187 Pugh TJ, Yu W, Yang J. et al. Exome sequencing of pleuropulmonary blastoma reveals frequent biallelic loss of TP53 and two hits in DICER1 resulting in retention of 5p-derived miRNA hairpin loop sequences. Oncogene 2014; 33: 5295-5302
- 188 Anglesio MS, Wang Y, Yang W. et al. Cancer-associated somatic DICER1 hotspot mutations cause defective miRNA processing and reverse-strand expression bias to predominantly mature 3p strands through loss of 5p strand cleavage. J Pathol 2013; 229: 400-409
- 189 Foulkes WD, Priest JR, Duchaine TF. DICER1: Mutations, microRNAs and mechanisms. Nat Rev Cancer 2014; 14: 662-672
- 190 Bernstein E, Caudy AA, Hammond SM. et al. Role for a bidentate ribonuclease in the initiation step of RNA interference. Nature 2001; 409: 363-366
- 191 Chendrimada TP, Gregory RI, Kumaraswamy E. et al. TRBP recruits the Dicer complex to Ago2 for microRNA processing and gene silencing. Nature 2005; 436: 740-744
- 192 Scheithauer BW, Kovacs K, Horvath E. et al. Pituitary blastoma. Acta Neuropathol 2008; 116: 657-666
- 193 Nadaf J, de Kock L, Chong AS. et al. Molecular characterization of DICER1-mutated pituitary blastoma. Acta Neuropathol 2021; 141: 929-944
- 194 Chong AS, Han H, Albrecht S. et al. DICER1 syndrome in a young adult with pituitary blastoma. Acta Neuropathol 2021; 142: 1071-1076
- 195 Liu APY, Kelsey MM, Sabbaghian N. et al. Clinical outcomes and complications of pituitary blastoma. J Clin Endocrinol Metab 2021; 106: 351-363
- 196 Liu AP, Li KK, Chow C. et al. Expanding the clinical and molecular spectrum of pituitary blastoma. Acta Neuropathol 2022; 143: 415-417
- 197 Xu Y, Zou R, Wang J. et al. The role of the cancer testis antigen PRAME in tumorigenesis and immunotherapy in human cancer. Cell Prolif 2020; 53: e12770
- 198 Martínez de LaPiscina I, Hernández-Ramírez LC, Portillo N. et al. Rare germline DICER1 variants in pediatric patients with Cushing's disease: What is their role?. Front Endocrinol (Lausanne) 2020; 11: 433
- 199 Idos G, Valle L. Lynch Syndrome. In: Adam MP, Feldman J, Mirzaa GM et al, eds. GeneReviews. Seattle (WA): University of Washington; 2004
- 200 Fishel R, Lescoe MK, Rao MR. et al. The human mutator gene homolog MSH2 and its association with hereditary nonpolyposis colon cancer. Cell 1993; 75: 1027-1038
- 201 Leach FS, Nicolaides NC, Papadopoulos N. et al. Mutations of a mutS homolog in hereditary nonpolyposis colorectal cancer. Cell 1993; 75: 1215-1225
- 202 Bronner CE, Baker SM, Morrison PT. et al. Mutation in the DNA mismatch repair gene homologue hMLH1 is associated with hereditary non-polyposis colon cancer. Nature 1994; 368: 258-261
- 203 Nicolaides NC, Papadopoulos N, Liu B. et al. Mutations of two PMS homologues in hereditary nonpolyposis colon cancer. Nature 1994; 371: 75-80
- 204 Palombo F, Hughes M, Jiricny J. et al. Mismatch repair and cancer. Nature 1994; 367: 417
- 205 Papadopoulos N, Nicolaides NC, Wei YF. et al. Mutation of a mutL homolog in hereditary colon cancer. Science 1994; 263: 1625-1629
- 206 Miyaki M, Konishi M, Tanaka K. et al. Germline mutation of MSH6 as the cause of hereditary nonpolyposis colorectal cancer. Nat Genet 1997; 17: 271-272
- 207 Kovacs ME, Papp J, Szentirmay Z. et al. Deletions removing the last exon of TACSTD1 constitute a distinct class of mutations predisposing to Lynch syndrome. Hum Mutat 2009; 30: 197-203
- 208 Ligtenberg MJ, Kuiper RP, Chan TL. et al. Heritable somatic methylation and inactivation of MSH2 in families with Lynch syndrome due to deletion of the 3' exons of TACSTD1. Nat Genet 2009; 41: 112-117
- 209 Giardiello FM, Allen JI, Axilbund JE. et al. Guidelines on genetic evaluation and management of Lynch syndrome: A consensus statement by the US Multi-Society Task Force on colorectal cancer. Gastroenterology 2014; 147: 502-526
- 210 Ijsselsteijn R, Jansen JG, de Wind N. DNA mismatch repair-dependent DNA damage responses and cancer. DNA Repair 2020; 93: 102923
- 211 Therkildsen C, Ladelund S, Rambech E. et al. Glioblastomas, astrocytomas and oligodendrogliomas linked to Lynch syndrome. Eur J Neurol 2015; 22: 717-724
- 212 Bengtsson D, Joost P, Aravidis C. et al. Corticotroph pituitary carcinoma in a patient with Lynch syndrome (LS) and pituitary tumors in a nationwide LS cohort. J Clin Endocrinol Metab 2017; 102: 3928-3932
- 213 Uraki S, Ariyasu H, Doi A. et al. Atypical pituitary adenoma with MEN1 somatic mutation associated with abnormalities of DNA mismatch repair genes; MLH1 germline mutation and MSH6 somatic mutation. Endocr J 2017; 64: 895-906
- 214 Voisin MR, Almeida JP, Perez-Ordonez B. et al. Recurrent undifferentiated carcinoma of the sella in a patient with lynch syndrome. World Neurosurg 2019; 132: 219-222
- 215 Teuber J, Reinhardt A, Reuss D. et al. Aggressive pituitary adenoma in the context of Lynch syndrome: A case report and literature review on this rare coincidence. Br J Neurosurg 2021; 1-6
- 216 Curatolo P, Bombardieri R, Jozwiak S. Tuberous sclerosis. Lancet 2008; 372: 657-668
- 217 Northrup H, Aronow ME, Bebin EM. et al. Updated international tuberous sclerosis complex diagnostic criteria and surveillance and management recommendations. Pediatr Neurol 2021; 123: 50-66
- 218 The European Chromosome 16 Tuberous Sclerosis Consortium. Identification and characterization of the tuberous sclerosis gene on chromosome 16. Cell 1993; 75: 1305-1315
- 219 van Slegtenhorst M, de Hoogt R, Hermans C. et al. Identification of the tuberous sclerosis gene TSC1 on chromosome 9q34. Science 1997; 277: 805-808
- 220 Verhoef S, Bakker L, Tempelaars AM. et al. High rate of mosaicism in tuberous sclerosis complex. Am J Hum Genet 1999; 64: 1632-1637
- 221 Qin W, Kozlowski P, Taillon BE. et al. Ultra deep sequencing detects a low rate of mosaic mutations in tuberous sclerosis complex. Hum Genet 2010; 127: 573-582
- 222 Crino PB. Evolving neurobiology of tuberous sclerosis complex. Acta Neuropathol 2013; 125: 317-332
- 223 Castro AF, Rebhun JF, Clark GJ. et al. Rheb binds tuberous sclerosis complex 2 (TSC2) and promotes S6 kinase activation in a rapamycin- and farnesylation-dependent manner. J Biol Chem 2003; 278: 32493-32496
- 224 Tee AR, Manning BD, Roux PP. et al. Tuberous sclerosis complex gene products, Tuberin and Hamartin, control mTOR signaling by acting as a GTPase-activating protein complex toward Rheb. Curr Biol 2003; 13: 1259-1268
- 225 Dibble CC, Elis W, Menon S. et al. TBC1D7 is a third subunit of the TSC1-TSC2 complex upstream of mTORC1. Mol Cell 2012; 47: 535-546
- 226 Saxton RA, Sabatini DM. mTOR Signaling in Growth, Metabolism, and Disease. Cell 2017; 169: 361-371
- 227 Panwar V, Singh A, Bhatt M. et al. Multifaceted role of mTOR (mammalian target of rapamycin) signaling pathway in human health and disease. Signal Transduct Target Ther 2023; 8: 375
- 228 Yeung RS, Katsetos CD, Klein-Szanto A. Subependymal astrocytic hamartomas in the Eker rat model of tuberous sclerosis. Am J Pathol 1997; 151: 1477-1486
- 229 Galaction-Nitelea O, Dociu I, Murgu V. A case of tuberous sclerosis with acromegaly. Rev Med Interna Neurol Psihiatr Neurochir Dermatovenerol Neurol Psihiatr Neurochir 1978; 23: 253-262
- 230 Hoffman WH, Perrin JC, Halac E. et al. Acromegalic gigantism and tuberous sclerosis. J Pediatr 1978; 93: 478-480
- 231 Bloomgarden ZT, McLean GW, Rabin D. Autonomous hyperprolactinemia in tuberous sclerosis. Arch Intern Med 1981; 141: 1513-1515
- 232 Tigas S, Carroll PV, Jones R. et al. Simultaneous Cushing's disease and tuberous sclerosis; a potential role for TSC in pituitary ontogeny. Clin Endocrinol (Oxf) 2005; 63: 694-695
- 233 Nandagopal R, Vortmeyer A, Oldfield EH. et al. Cushing's syndrome due to a pituitary corticotropinoma in a child with tuberous sclerosis: An association or a coincidence. Clin Endocrinol (Oxf) 2007; 67: 639-641
- 234 Regazzo D, Gardiman MP, Theodoropoulou M. et al. Silent gonadotroph pituitary neuroendocrine tumor in a patient with tuberous sclerosis complex: Evaluation of a possible molecular link. Endocrinol Diabetes Metab Case Rep 2018; 2018
- 235 Yehia L, Eng C. PTEN Hamartoma Tumor Syndrome. In: Adam MP, Feldman J, Mirzaa GM et al, eds. GeneReviews(R); Seattle (WA): 1993
- 236 Lloyd KM, Dennis M. Cowden's disease. A possible new symptom complex with multiple system involvement. Ann Intern Med 1963; 58: 136-142
- 237 Efstathiadou ZA, Sapranidis M, Anagnostis P. et al. Unusual case of Cowden-like syndrome, neck paraganglioma, and pituitary adenoma. Head Neck 2014; 36: E12-E16
- 238 Amatya N, Piziak V. Abstract #821: Recurrent pituitary apoplexy in Cowden syndrome: A case report. Endocr Pract 2017; 23: 173
- 239 Srichomkwun P, Houngngam N, Boonchaya-Anant P. et al. Cowden syndrome and pituitary tumours. QJM 2018; 111: 735-736
- 240 Zhang H, Li J, Lee M. et al. Pituitary carcinoma in a patient with Cowden syndrome. Am J Case Rep 2022; 23: e934846
- 241 Valdes-Socin H, Poncin J, Stevens V. et al. Familial isolated pituitary adenomas unrelated to MEN1 mutations: A follow-up of 27 patients. 10th Meeting of the Belgian Endocrine Society. 2000
- 242 Daly AF, Rixhon M, Adam C. et al. High prevalence of pituitary adenomas: A cross-sectional study in the province of Liege, Belgium. J Clin Endocrinol Metab 2006; 91: 4769-4775
- 243 Daly AF, Tichomirowa MA, Petrossians P. et al. Clinical characteristics and therapeutic responses in patients with germ-line AIP mutations and pituitary adenomas: An international collaborative study. J Clin Endocrinol Metab 2010; 95: E373-E383
- 244 Igreja S, Chahal HS, King P. et al. Characterization of aryl hydrocarbon receptor interacting protein (AIP) mutations in familial isolated pituitary adenoma families. Hum Mutat 2010; 31: 950-960
- 245 Hernández-Ramírez LC, Gabrovska P, Dénes J. et al. Landscape of familial isolated and young-onset pituitary adenomas: Prospective diagnosis in AIP mutation carriers. J Clin Endocrinol Metab 2015; 100: E1242-E1254
- 246 Vierimaa O, Georgitsi M, Lehtonen R. et al. Pituitary adenoma predisposition caused by germline mutations in the AIP gene. Science 2006; 312: 1228-1230
- 247 Daly AF, Vanbellinghen JF, Khoo SK. et al. Aryl hydrocarbon receptor-interacting protein gene mutations in familial isolated pituitary adenomas: Analysis in 73 families. J Clin Endocrinol Metab 2007; 92: 1891-1896
- 248 Iacovazzo D, Hernández-Ramírez LC, Korbonits M. Sporadic pituitary adenomas: The role of germline mutations and recommendations for genetic screening. Expert Rev Endocrinol Metab 2017; 12: 143-153
- 249 Trivellin G, Korbonits M. AIP and its interacting partners. J Endocrinol 2011; 210: 137-155
- 250 Chahal HS, Trivellin G, Leontiou CA. et al. Somatostatin analogs modulate AIP in somatotroph adenomas: The role of the ZAC1 pathway. J Clin Endocrinol Metab 2012; 97: E1411-E1420
- 251 Tuominen I, Heliovaara E, Raitila A. et al. AIP inactivation leads to pituitary tumorigenesis through defective Galpha-cAMP signaling. Oncogene 2015; 34: 1174-1184
- 252 Hernández-Ramírez LC, Morgan RML, Barry S. et al. Multi-chaperone function modulation and association with cytoskeletal proteins are key features of the function of AIP in the pituitary gland. Oncotarget 2018; 9: 9177-9198
- 253 Garcia-Rendueles AR, Chenlo M, Oroz-Gonjar F. et al. RET signalling provides tumorigenic mechanism and tissue specificity for AIP-related somatotrophinomas. Oncogene 2021; 40: 6354-6368
- 254 Sun D, Stopka-Farooqui U, Barry S. et al. Aryl hydrocarbon receptor interacting protein maintains germinal venter B cells through suppression of BCL6 degradation. Cell Rep 2019; 27: 1461-1471 e1464
- 255 Solis-Fernandez G, Montero-Calle A, Sanchez-Martinez M. et al. Aryl-hydrocarbon receptor-interacting protein regulates tumorigenic and metastatic properties of colorectal cancer cells driving liver metastasis. Br J Cancer 2022; 126: 1604-1615
- 256 Barry S, Carlsen E, Marques P. et al. Tumor microenvironment defines the invasive phenotype of AIP-mutation-positive pituitary tumors. Oncogene 2019; 38: 5381-5395
- 257 Georgitsi M, Raitila A, Karhu A. et al. Molecular diagnosis of pituitary adenoma predisposition caused by aryl hydrocarbon receptor-interacting protein gene mutations. Proc Natl Acad Sci U S A 2007; 104: 4101-4105
- 258 Cazabat L, Bouligand J, Salenave S. et al. Germline AIP mutations in apparently sporadic pituitary adenomas: Prevalence in a prospective single-center cohort of 443 patients. J Clin Endocrinol Metab 2012; 97: E663-E670
- 259 Beckers A, Aaltonen LA, Daly AF. et al. Familial isolated pituitary adenomas (FIPA) and the pituitary adenoma predisposition due to mutations in the aryl hydrocarbon receptor interacting protein (AIP) gene. Endocrine Reviews 2013; 34: 239-277
- 260 Hernández-Ramírez LC, Trivellin G, Stratakis CA. Role of phosphodiesterases on the function of aryl hydrocarbon receptor-interacting protein (AIP) in the pituitary gland and on the evaluation of AIP gene variants. Horm Metab Res 2017; 49: 286-295
- 261 Nguyen JT, Ferriere A, Tabarin A. Case report: Complete restoration of the HPA axis function in Cushing's disease with drug treatment. Front Endocrinol (Lausanne) 2024; 15: 1337741
- 262 Trofimiuk-Muldner M, Domagala B, Sokolowski G. et al. AIP gene germline variants in adult Polish patients with apparently sporadic pituitary macroadenomas. Front Endocrinol (Lausanne) 2023; 14: 1098367
- 263 Mistry A, Solomou A, Vignola ML. et al Investigating the role of AIP in pituitary tumourigenesis. Endocrine Abstracts 2019; 65 OC2.2
- 264 Trivellin G, Daly AF, Faucz FR. et al. Gigantism and acromegaly due to Xq26 microduplications and GPR101 mutation. N Engl J Med 2014; 371: 2363-2374
- 265 Trivellin G, Hernández-Ramírez LC, Swan J. et al. An orphan G-protein-coupled receptor causes human gigantism and/or acromegaly: Molecular biology and clinical correlations. Best Pract Res Clin Endocrinol Metab 2018; 32: 125-140
- 266 Trivellin G, Correa RR, Batsis M. et al. Screening for GPR101 defects in pediatric pituitary corticotropinomas. Endocr Relat Cancer 2016; 23: 357-365
- 267 Zhang Q, Peng C, Song J. et al. Germline mutations in CDH23, encoding cadherin-related 23, are associated with both familial and sporadic pituitary adenomas. Am J Hum Genet 2017; 100: 817-823
- 268 Back N, Mains RE, Eipper BA. PAM: Diverse roles in neuroendocrine cells, cardiomyocytes, and green algae. FEBS J 2022; 289: 4470-4496
- 269 Trivellin G, Daly AF, Hernández-Ramírez LC. et al. Germline loss-of-function PAM variants are enriched in subjects with pituitary hypersecretion. Front Endocrinol (Lausanne) 2023; 14: 1166076
- 270 De Sousa SMC, Shen A, Yates CJ. et al. PAM variants in patients with thyrotrophinomas, cyclical Cushing's disease and prolactinomas. Front Endocrinol (Lausanne) 2023; 14: 1305606
- 271 Zukerberg LR, Patrick GN, Nikolic M. et al. Cables links Cdk5 and c-Abl and facilitates Cdk5 tyrosine phosphorylation, kinase upregulation, and neurite outgrowth. Neuron 2000; 26: 633-646
- 272 Wu CL, Kirley SD, Xiao H. et al. Cables enhances cdk2 tyrosine 15 phosphorylation by Wee1, inhibits cell growth, and is lost in many human colon and squamous cancers. Cancer Res 2001; 61: 7325-7332
- 273 Huang JR, Tan GM, Li Y. et al. The emerging role of Cables1 in cancer and other diseases. Mol Pharmacol 2017; 92: 240-245
- 274 Shi Z, Park HR, Du Y. et al. Cables1 complex couples survival signaling to the cell death machinery. Cancer Res 2015; 75: 147-158
- 275 Kirley SD, Rueda BR, Chung DC. et al. Increased growth rate, delayed senescense and decreased serum dependence characterize cables-deficient cells. Cancer Biol Ther 2005; 4: 654-658
- 276 Zukerberg LR, DeBernardo RL, Kirley SD. et al. Loss of cables, a cyclin-dependent kinase regulatory protein, is associated with the development of endometrial hyperplasia and endometrial cancer. Cancer Res 2004; 64: 202-208
- 277 Kirley SD, D'Apuzzo M, Lauwers GY. et al. The Cables gene on chromosome 18Q regulates colon cancer progression in vivo. Cancer Biol Ther 2005; 4: 861-863
- 278 Roussel-Gervais A, Couture C, Langlais D. et al. The Cables1 gene in glucocorticoid regulation of pituitary corticotrope growth and Cushing disease. J Clin Endocrinol Metab 2016; 101: 513-522
- 279 Hernández-Ramírez LC, Gam R, Valdés N. et al. Loss-of-function mutations in the CABLES1 gene are a novel cause of Cushing's disease. Endocr Relat Cancer 2017; 24: 379-392
- 280 Franco-Álvarez AL, Torres-Morán M, Rebollar-Vega R. et al. OR2802 A novel CABLES1 missense variant associated with Cushing's disease disrupts protein structure and stability. J Endocr Soc 2023; 7 bvad114.1364
- 281 Clayton R, Burden AC, Schrieber V. et al. Secondary pituitary hyperplasia in Addison's disease. Lancet 1977; 2: 954-956
- 282 Dluhy RG, Moore TJ, Williams GH. Sella turcica enlargement and primary adrenal nsufficiency. An Intern Med 1978; 89: 513-514
- 283 Himsworth RL, Lewis JG, Rees LH. A possible ACTH secreting tumour of the pituitary developing in a conventionally treated case of Addison's disease. Clin Endocrinol (Oxf) 1978; 9: 131-139
- 284 Jara-Albarran A, Bayort J, Caballero A. et al. Probable pituitary adenoma with adrenocorticotropin hypersecretion (corticotropinoma) secondary to Addison's disease. J Clin Endocrinol Metab 1979; 49: 236-241
- 285 Aanderud S, Bassoe HH. A pituitary tumour with possible ACTH and TSH hypersecretion in a patient with Addison's disease and primary hypothyroidism. Acta Endocrinol (Copenh) 1980; 95: 181-184
- 286 Krautli B, Muller J, Landolt AM. et al. ACTH-producing pituitary adenomas in Addison's disease: Two cases treated by transsphenoidal microsurgery. Acta Endocrinol (Copenh) 1982; 99: 357-363
- 287 Scheithauer BW, Kovacs K, Randall RV. The pituitary gland in untreated Addison's disease. A histologic and immunocytologic study of 18 adenohypophyses. Arch Pathol Lab Med 1983; 107: 484-487
- 288 Yanase T, Sekiya K, Ando M. et al. Probable ACTH-secreting pituitary tumour in association with Addison's disease. Acta Endocrinol (Copenh) 1985; 110: 36-41
- 289 Sugiyama K, Kimura M, Abe T. et al. Hyper-adrenocorticotropinemia in a patient with Addison's disease after treatment with corticosteroids. Intern Med 1996; 35: 555-559
- 290 Fan S, Jiang Y, Yao Y. et al. Pituitary ACTH-secreting adenoma in Addison's disease: A case report. Clin Neurol Neurosurg 2013; 115: 2543-2546
- 291 Royrvik EC, Husebye ES. The genetics of autoimmune Addison disease: Past, present and future. Nat Rev Endocrinol 2022; 18: 399-412
- 292 Auer MK, Nordenstrom A, Lajic S. et al. Congenital adrenal hyperplasia. Lancet 2023; 401: 227-244
- 293 Horrocks PM, Franks S, Hockley AD. et al. An ACTH-secreting pituitary tumour arising in a patient with congenital adrenal hyperplasia. Clin Endocrinol (Oxf) 1982; 17: 457-468
- 294 Boronat M, Carrillo A, Ojeda A. et al. Clinical manifestations and hormonal profile of two women with Cushing's disease and mild deficiency of 21-hydroxylase. J Endocrinol Invest 2004; 27: 583-590
- 295 Haase M, Schott M, Kaminsky E. et al. Cushing's disease in a patient with steroid 21-hydroxylase deficiency. Endocr J 2011; 58: 699-706
- 296 McCabe ER. DAX1: Increasing complexity in the roles of this novel nuclear receptor. Mol Cell Endocrinol 2007; 265-266: 179-182
- 297 Peter M, Viemann M, Partsch CJ. et al. Congenital adrenal hypoplasia: Clinical spectrum, experience with hormonal diagnosis, and report on new point mutations of the DAX-1 gene. J Clin Endocrinol Metab 1998; 83: 2666-2674
- 298 Reutens AT, Achermann JC, Ito M. et al. Clinical and functional effects of mutations in the DAX-1 gene in patients with adrenal hypoplasia congenita. J Clin Endocrinol Metab 1999; 84: 504-511
- 299 Guran T, Buonocore F, Saka N. et al. Rare causes of primary adrenal insufficiency: Genetic and clinical characterization of a large nationwide cohort. J Clin Endocrinol Metab 2016; 101: 284-292
- 300 Tabarin A, Achermann JC, Recan D. et al. A novel mutation in DAX1 causes delayed-onset adrenal insufficiency and incomplete hypogonadotropic hypogonadism. J Clin Invest 2000; 105: 321-328
- 301 Mantovani G, Ozisik G, Achermann JC. et al. Hypogonadotropic hypogonadism as a presenting feature of late-onset X-linked adrenal hypoplasia congenita. J Clin Endocrinol Metab 2002; 87: 44-48
- 302 Achermann JC, Vilain EJ. NR0B1-Related Adrenal Hypoplasia Congenita. In: Adam MP, Mirzaa GM, Pagon RA et al, eds. GeneReviews(R); Seattle (WA): 2018
- 303 Weiss L, Mellinger RC. Congenital adrenal hypoplasia--an X-linked disease. J Med Genet 1970; 7: 27-32
- 304 Muscatelli F, Strom TM, Walker AP. et al. Mutations in the DAX-1 gene give rise to both X-linked adrenal hypoplasia congenita and hypogonadotropic hypogonadism. Nature 1994; 372: 672-676
- 305 Zanaria E, Muscatelli F, Bardoni B. et al. An unusual member of the nuclear hormone receptor superfamily responsible for X-linked adrenal hypoplasia congenita. Nature 1994; 372: 635-641
- 306 Ikeda Y, Swain A, Weber TJ. et al. Steroidogenic factor 1 and Dax-1 colocalize in multiple cell lineages: Potential links in endocrine development. Mol Endocrinol 1996; 10: 1261-1272
- 307 De Menis E, Roncaroli F, Calvari V. et al. Corticotroph adenoma of the pituitary in a patient with X-linked adrenal hypoplasia congenita due to a novel mutation of the DAX-1 gene. Eur J Endocrinol 2005; 153: 211-215
- 308 Martins CS, Camargo RC, Coeli-Lacchini FB. et al. USP8 mutations and cell cycle regulation in corticotroph adenomas. Horm Metab Res 2020; 52: 117-123
- 309 Lin AL, Rudneva VA, Richards AL. et al. Genome-wide loss of heterozygosity predicts aggressive, treatment-refractory behavior in pituitary neuroendocrine tumors. Acta Neuropathol 2024; 147: 85
Correspondence
Publication History
Received: 22 March 2024
Received: 16 May 2024
Accepted: 03 June 2024
Accepted Manuscript online:
03 June 2024
Article published online:
03 July 2024
© 2024. Thieme. All rights reserved.
Georg Thieme Verlag KG
Rüdigerstraße 14, 70469 Stuttgart, Germany
-
References
- 1 Mete O, Grossman A, Yamada S. et al. Pituitary Tumours: Anterior Pituitary Neuroendocrine Tumours (PitNETs)/ PitNETs of TPIT Lineage/ Corticotroph PitNET/Adenoma. In: Osamura RY, Asa SL, eds. WHO Classification of Tumours Editorial Board Endocrine and neuroendocrine tumours. 5 edn. Lyon (France): International Agency for Research on Cancer; 2022
- 2 Etxabe J, Vázquez JA. Morbidity and mortality in Cushing's disease: An epidemiological approach. Clin Endocrinol (Oxf) 1994; 40: 479-484
- 3 Lindholm J, Juul S, Jorgensen JO. et al. Incidence and late prognosis of Cushing's syndrome: A population-based study. J Clin Endocrinol Metab 2001; 86: 117-123
- 4 Clayton RN, Jones PW, Reulen RC. et al. Mortality in patients with Cushing's disease more than 10 years after remission: A multicentre, multinational, retrospective cohort study. Lancet Diabetes Endocrinol 2016; 4: 569-576
- 5 Ragnarsson O, Olsson DS, Papakokkinou E. et al. Overall and disease-specific mortality in patients with Cushing disease: A Swedish nationwide study. J Clin Endocrinol Metab 2019; 104: 2375-2384
- 6 Rubinstein G, Osswald A, Hoster E. et al. Time to diagnosis in Cushing's syndrome: A meta-analysis based on 5367 patients. J Clin Endocrinol Metab 2020; 105: e12-e22
- 7 Lindsay JR, Nansel T, Baid S. et al. Long-term impaired quality of life in Cushing's syndrome despite initial improvement after surgical remission. J Clin Endocrinol Metab 2006; 91: 447-453
- 8 Espinosa-de-Los-Monteros AL, Sosa E, Martinez N. et al. Persistence of Cushing's disease symptoms and comorbidities after surgical cure: A long-term, integral evaluation. Endocr Pract 2013; 19: 252-258
- 9 Fleseriu M, Auchus R, Bancos I. et al. Consensus on diagnosis and management of Cushing's disease: A guideline update. Lancet Diabetes Endocrinol 2021; 9: 847-875
- 10 Feelders RA, Newell-Price J, Pivonello R. et al. Advances in the medical treatment of Cushing's syndrome. Lancet Diabetes Endocrinol 2019; 7: 300-312
- 11 Fleseriu M, Varlamov EV, Hinojosa-Amaya JM. et al. An individualized approach to the management of Cushing disease. Nat Rev Endocrinol 2023; 19: 581-599
- 12 Ma ZY, Song ZJ, Chen JH. et al. Recurrent gain-of-function USP8 mutations in Cushing's disease. Cell Res 2015; 25: 306-317
- 13 Reincke M, Sbiera S, Hayakawa A. et al. Mutations in the deubiquitinase gene USP8 cause Cushing's disease. Nat Genet 2015; 47: 31-38
- 14 Song ZJ, Reitman ZJ, Ma ZY. et al. The genome-wide mutational landscape of pituitary adenomas. Cell Res 2016; 26: 1255-1259
- 15 Chen J, Jian X, Deng S. et al. Identification of recurrent USP48 and BRAF mutations in Cushing's disease. Nat Commun 2018; 9: 3171
- 16 Sbiera S, Pérez-Rivas LG, Taranets L. et al. Driver mutations in USP8 wild-type Cushing's disease. Neuro Oncol 2019; 21: 1273-1283
- 17 Casar-Borota O, Boldt HB, Engstrom BE. et al. Corticotroph aggressive pituitary tumors and carcinomas frequently harbor ATRX mutations. J Clin Endocrinol Metab 2021; 106: 1183-1194
- 18 Pérez-Rivas LG, Simon J, Albani A. et al. TP53 mutations in functional corticotroph tumors are linked to invasion and worse clinical outcome. Acta Neuropathol Commun 2022; 10: 139
- 19 Williamson EA, Ince PG, Harrison D. et al. G-protein mutations in human pituitary adrenocorticotrophic hormone-secreting adenomas. Eur J Clin Invest 1995; 25: 128-131
- 20 Riminucci M, Collins MT, Lala R. et al. An R201H activating mutation of the GNAS1 (Gsalpha) gene in a corticotroph pituitary adenoma. Mol Pathol 2002; 55: 58-60
- 21 Sbiera S, Kunz M, Weigand I. et al. The new genetic landscape of Cushing's disease: Deubiquitinases in the spotlight. Cancers (Basel) 2019; 11
- 22 Hayashi K, Inoshita N, Kawaguchi K. et al. The USP8 mutational status may predict drug susceptibility in corticotroph adenomas of Cushing's disease. Eur J Endocrinol 2016; 174: 213-226
- 23 Faucz FR, Tirosh A, Tatsi C. et al. Somatic USP8 gene mutations are a common cause of pediatric Cushing disease. J Clin Endocrinol Metab 2017; 102: 2836-2843
- 24 Albani A, Pérez-Rivas LG, Dimopoulou C. et al. The USP8 mutational status may predict long-term remission in patients with Cushing's disease. Clin Endocrinol (Oxf) 2018; 89: 454-458
- 25 Ballmann C, Thiel A, Korah HE. et al. USP8 mutations in pituitary Cushing adenomas-targeted analysis by next-generation sequencing. J Endocr Soc 2018; 2: 266-278
- 26 Bujko M, Kober P, Boresowicz J. et al. USP8 mutations in corticotroph adenomas determine a distinct gene expression profile irrespective of functional tumour status. Eur J Endocrinol 2019; 181: 615-627
- 27 Losa M, Mortini P, Pagnano A. et al. Clinical characteristics and surgical outcome in USP8-mutated human adrenocorticotropic hormone-secreting pituitary adenomas. Endocrine 2019; 63: 240-246
- 28 Weigand I, Knobloch L, Flitsch J. et al. Impact of USP8 gene mutations on protein deregulation in Cushing disease. J Clin Endocrinol Metab 2019; 104: 2535-2546
- 29 Sesta A, Cassarino MF, Terreni M. et al. Ubiquitin-specific protease =8 mutant corticotrope adenomas present unique secretory and molecular features and shed light on the role of ubiquitylation on ACTH processing. Neuroendocrinology 2020; 110: 119-129
- 30 Pasternak-Pietrzak K, Faucz FR, Stratakis CA. et al. Is there a common cause for paediatric Cushing's disease?. Endokrynol Pol 2021; 72: 104-107
- 31 Treppiedi D, Barbieri AM, Di Muro G. et al. Genetic profiling of a cohort of Italian patients with ACTH-secreting pituitary tumors and characterization of a novel USP8 gene variant. Cancers (Basel) 2021; 13: 4022
- 32 Andonegui-Elguera S, Silva-Roman G, Pena-Martinez E. et al. The genomic landscape of corticotroph tumors: From silent adenomas to ACTH-secreting carcinomas. Int J Mol Sci 2022; 23: 4861
- 33 Hernández-Ramírez LC, Pankratz N, Lane J. et al. Genetic drivers of Cushing's disease: Frequency and associated phenotypes. Genet Med 2022; 24: 2516-2525
- 34 Shichi H, Fukuoka H, Kanzawa M. et al. Responsiveness to DDAVP in Cushing's disease is associated with USP8 mutations through enhancing AVPR1B promoter activity. Pituitary 2022; 25: 496-507
- 35 Neou M, Villa C, Armignacco R. et al. Pangenomic classification of pituitary neuroendocrine tumors. Cancer Cell 2020; 37: 123-134.e125
- 36 Mizuno E, Kitamura N, Komada M. 14-3-3-dependent inhibition of the deubiquitinating activity of UBPY and its cancellation in the M phase. Exp Cell Res 2007; 313: 3624-3634
- 37 Albani A, Pérez-Rivas LG, Tang S. et al. Improved pasireotide response in USP8 mutant corticotroph tumours in vitro. Endocr Relat Cancer 2022; 29: 503-511
- 38 Cohen M, Persky R, Stegemann R. et al. Germline USP8 mutation associated with pediatric Cushing disease and other clinical features: A new syndrome. J Clin Endocrinol Metab 2019; 104: 4676-4682
- 39 Mizuno E, Iura T, Mukai A. et al. Regulation of epidermal growth factor receptor down-regulation by UBPY-mediated deubiquitination at endosomes. Mol Biol Cell 2005; 16: 5163-5174
- 40 Ronchi CL, Peverelli E, Herterich S. et al. Landscape of somatic mutations in sporadic GH-secreting pituitary adenomas. Eur J Endocrinol 2016; 174: 363-372
- 41 Bi WL, Horowitz P, Greenwald NF. et al. Landscape of genomic alterations in pituitary adenomas. Clin Cancer Res 2017; 23: 1841-1851
- 42 Pérez-Rivas LG, Osswald A, Knosel T. et al. Expression and mutational status of USP8 in tumors causing ectopic ACTH secretion syndrome. Endocr Relat Cancer 2017; 24: L73-L77
- 43 Pérez-Rivas LG, Theodoropoulou M, Ferrau F. et al. The gene of the ubiquitin-specific protease 8 is frequently mutated in adenomas causing Cushing's disease. J Clin Endocrinol Metab 2015; 100: E997-E1004
- 44 Castellnou S, Vasiljevic A, Lapras V. et al. SST5 expression and USP8 mutation in functioning and silent corticotroph pituitary tumors. Endocr Connect 2020; 9: 243-253
- 45 Mossakowska BJ, Rusetska N, Konopinski R. et al. The expression of cell cycle-related genes in USP8-mutated corticotroph neuroendocrine pituitary tumors and their possible role in cell cycle-targeting treatment. Cancers (Basel) 2022; 14: 5594
- 46 Kober P, Rusetska N, Mossakowska BJ. et al. The expression of glucocorticoid and mineralocorticoid receptors in pituitary tumors causing Cushing's disease and silent corticotroph tumors. Front Endocrinol (Lausanne) 2023; 14: 1124646
- 47 Treppiedi D, Marra G, Di Muro G. et al. P720R USP8 mutation is associated with a better responsiveness to pasireotide in ACTH-secreting PitNETs. Cancers (Basel) 2022; 14: 2455
- 48 Pérez-Rivas LG, Theodoropoulou M, Puar TH. et al. Somatic USP8 mutations are frequent events in corticotroph tumor progression causing Nelson's tumor. Eur J Endocrinol 2018; 178: 59-65
- 49 Wanichi IQ, de Paula Mariani BM, Frassetto FP. et al. Cushing's disease due to somatic USP8 mutations: A systematic review and meta-analysis. Pituitary 2019; 22: 435-442
- 50 Abraham AP, Pai R, Beno DL. et al. USP8, USP48, and BRAF mutations differ in their genotype-phenotype correlation in Asian Indian patients with Cushing's disease. Endocrine 2022; 75: 549-559
- 51 Tatsi C, Pankratz N, Lane J. et al. Large genomic aberrations in corticotropinomas are associated with greater aggressiveness. J Cli Endocrinol Metab 2019; 104: 1792-1801
- 52 Zhou A, Lin K, Zhang S. et al. Gli1-induced deubiquitinase USP48 aids glioblastoma tumorigenesis by stabilizing Gli1. EMBO Rep 2017; 18: 1318-1330
- 53 Vila G, Papazoglou M, Stalla J. et al. Sonic hedgehog regulates CRH signal transduction in the adult pituitary. FASEB J 2005; 19: 281-283
- 54 Vila G, Theodoropoulou M, Stalla J. et al. Expression and function of sonic hedgehog pathway components in pituitary adenomas: Evidence for a direct role in hormone secretion and cell proliferation. J Clin Endocrinol Metab 2005; 90: 6687-6694
- 55 Karl M, Von Wichert G, Kempter E. et al. Nelson's syndrome associated with a somatic frame shift mutation in the glucocorticoid receptor gene. J Clin Endocrinol Metab 1996; 81: 124-129
- 56 Miao H, Liu Y, Lu L. et al. Effect of 3 NR3C1 mutations in the pathogenesis of pituitary ACTH adenoma. Endocrinology 2021; 162: bqab167
- 57 Theodoropoulou M. Glucocorticoid receptors are making a comeback in corticotroph tumorigenesis. Endocrinology 2022; 163: bqab257
- 58 Hurley DM, Accili D, Stratakis CA. et al. Point mutation causing a single amino acid substitution in the hormone binding domain of the glucocorticoid receptor in familial glucocorticoid resistance. J Clin Invest 1991; 87: 680-686
- 59 Karl M, Lamberts SW, Koper JW. et al. Cushing's disease preceded by generalized glucocorticoid resistance: Clinical consequences of a novel, dominant-negative glucocorticoid receptor mutation. Proc Assoc Am Physicians 1996; 108: 296-307
- 60 Davies H, Bignell GR, Cox C. et al. Mutations of the BRAF gene in human cancer. Nature 2002; 417: 949-954
- 61 Ikenoue T, Hikiba Y, Kanai F. et al. Functional analysis of mutations within the kinase activation segment of B-Raf in human colorectal tumors. Cancer Res 2003; 63: 8132-8137
- 62 Wellbrock C, Ogilvie L, Hedley D. et al. V599EB-RAF is an oncogene in melanocytes. Cancer Res 2004; 64: 2338-2342
- 63 Proietti I, Skroza N, Michelini S. et al. BRAF inhibitors: Molecular targeting and immunomodulatory actions. Cancers (Basel) 2020; 12: 1823
- 64 Levy A, Hall L, Yeudall WA. et al. p53 gene mutations in pituitary adenomas: Rare events. Clin Endocrinol (Oxf) 1994; 41: 809-814
- 65 Tanizaki Y, Jin L, Scheithauer BW. et al P53 gene mutations in pituitary carcinomas. Endocr Pathol 2007; 18: 217-222 [doi]
- 66 Kawashima ST, Usui T, Sano T. et al P53 gene mutation in an atypical corticotroph adenoma with Cushing's disease. Clin Endocrinol (Oxf) 2009; 70: 656-657 CEN3404 [pii] [doi]
- 67 Pinto EM, Siqueira SA, Cukier P. et al. Possible role of a radiation-induced p53 mutation in a Nelson's syndrome patient with a fatal outcome. Pituitary 2011; 14: 400-404
- 68 Saeger W, Mawrin C, Meinhardt M. et al. Two pituitary neuroendocrine tumors (PitNETs) with very high proliferation and TP53 mutation – High-grade PitNET or PitNEC?. Endocr Pathol 2022; 33: 257-262
- 69 Sumislawski P, Rotermund R, Klose S. et al. ACTH-secreting pituitary carcinoma with TP53, NF1, ATRX and PTEN mutations Case report and review of the literature. Endocrine 2022; 76: 228-236
- 70 Dyer MA, Qadeer ZA, Valle-Garcia D. et al. ATRX and DAXX: Mechanisms and mutations. Cold Spring Harb Perspect Med 2017; 7: a026567
- 71 Jiao Y, Shi C, Edil BH. et al. DAXX/ATRX, MEN1, and mTOR pathway genes are frequently altered in pancreatic neuroendocrine tumors. Science 2011; 331: 1199-1203
- 72 Casar-Borota O, Botling J, Granberg D. et al. Serotonin, ATRX, and DAXX expression in pituitary adenomas: Markers in the differential diagnosis of neuroendocrine tumors of the sellar region. Am J Surg Pathol 2017; 41: 1238-1246
- 73 Alzoubi H, Minasi S, Gianno F. et al. Alternative lengthening of telomeres (ALT) and telomerase reverse transcriptase promoter methylation in recurrent adult and primary pediatric pituitary neuroendocrine tumors. Endocr Pathol 2022; 33: 494-505
- 74 Benson L, Ljunghall S, Akerstrom G. et al. Hyperparathyroidism presenting as the first lesion in multiple endocrine neoplasia type 1. Am J Med 1987; 82: 731-737
- 75 Trump D, Farren B, Wooding C. et al. Clinical studies of multiple endocrine neoplasia type 1 (MEN1). QJM 1996; 89: 653-669
- 76 Carty SE, Helm AK, Amico JA. et al The variable penetrance and spectrum of manifestations of multiple endocrine neoplasia type 1. Surgery 1998; 124: 1106-1113 discussion 1113-1104
- 77 Machens A, Schaaf L, Karges W. et al. Age-related penetrance of endocrine tumours in multiple endocrine neoplasia type 1 (MEN1): A multicentre study of 258 gene carriers. Clin Endocrinol (Oxf) 2007; 67: 613-622
- 78 Sakurai A, Suzuki S, Kosugi S. et al. Multiple endocrine neoplasia type 1 in Japan: Establishment and analysis of a multicentre database. Clin Endocrinol (Oxf) 2012; 76: 533-539
- 79 Giusti F, Cianferotti L, Boaretto F. et al. Multiple endocrine neoplasia syndrome type 1: Institution, management, and data analysis of a nationwide multicenter patient database. Endocrine 2017; 58: 349-359
- 80 Romanet P, Mohamed A, Giraud S. et al. UMD-MEN1 database: An overview of the 370 MEN1 variants present in 1676 patients from the French population. J Clin Endocrinol Metab 2019; 104: 753-764
- 81 Waguespack SG. Beyond the "3 Ps": A critical appraisal of the non-endocrine manifestations of multiple endocrine neoplasia type 1. Front Endocrinol (Lausanne) 2022; 13: 1029041
- 82 Chandrasekharappa SC, Guru SC, Manickam P. et al. Positional cloning of the gene for multiple endocrine neoplasia-type 1. Science 1997; 276: 404-407
- 83 Lemmens I, Van de Ven WJ, Kas K. et al. Identification of the multiple endocrine neoplasia type 1 (MEN1) gene. The European Consortium on MEN1. Hum Mol Genet 1997; 6: 1177-1183
- 84 Cebrian A, Ruiz-Llorente S, Cascon A. et al. Mutational and gross deletion study of the MEN1 gene and correlation with clinical features in Spanish patients. J Med Genet 2003; 40: e72
- 85 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 2008; 29: 22-32
- 86 de Laat JM, van der Luijt RB, Pieterman CR. et al. MEN1 redefined, a clinical comparison of mutation-positive and mutation-negative patients. BMC Med 2016; 14: 182
- 87 Beijers H, Stikkelbroeck NML, Mensenkamp AR. et al. Germline and somatic mosaicism in a family with multiple endocrine neoplasia type 1 (MEN1) syndrome. Eur J Endocrinol 2019; 180: K15-K19
- 88 Coppin L, Ferriere A, Crepin M. et al. Diagnosis of mosaic mutations in the MEN1 gene by next generation sequencing. Eur J Endocrinol 2019; 180: L1-L3
- 89 Schaaf L, Pickel J, Zinner K. et al. Developing effective screening strategies in multiple endocrine neoplasia type 1 (MEN 1) on the basis of clinical and sequencing data of German patients with MEN 1. Exp Clin Endocrinol Diabetes 2007; 115: 509-517
- 90 La P, Desmond A, Hou Z. et al. Tumor suppressor menin: The essential role of nuclear localization signal domains in coordinating gene expression. Oncogene 2006; 25: 3537-3546
- 91 Matkar S, Thiel A, Hua X. Menin: A scaffold protein that controls gene expression and cell signaling. Trends Biochem Sci 2013; 38: 394-402
- 92 Verges B, Boureille F, Goudet P. et al. Pituitary disease in MEN type 1 (MEN1): Data from the France-Belgium MEN1 multicenter study. J Clin Endocrinol Metab 2002; 87: 457-465
- 93 de Laat JM, Dekkers OM, Pieterman CR. et al. Long-term natural course of pituitary tumors in patients with MEN1: Results from the DutchMEN1 study group (DMSG). J Clin Endocrinol Metab 2015; 100: 3288-3296
- 94 Wu Y, Gao L, Guo X. et al. Pituitary adenomas in patients with multiple endocrine neoplasia type 1: A single-center experience in China. Pituitary 2019; 22: 113-123
- 95 Le Bras M, Leclerc H, Rousseau O. et al. Pituitary adenoma in patients with multiple endocrine neoplasia type 1: A cohort study. Eur J Endocrinol 2021; 185: 863-873
- 96 Trouillas J, Labat-Moleur F, Sturm N. et al. Pituitary tumors and hyperplasia in multiple endocrine neoplasia type 1 syndrome (MEN1): A case-control study in a series of 77 patients versus 2509 non-MEN1 patients. Am J Surg Pathol 2008; 32: 534-543
- 97 Farrell WE, Azevedo MF, Batista DL. et al. Unique gene expression profile associated with an early-onset multiple endocrine neoplasia (MEN1)-associated pituitary adenoma. J Clin Endocrinol Metab 2011; 96: E1905-E1914
- 98 Rix M, Hertel NT, Nielsen FC. et al. Cushing's disease in childhood as the first manifestation of multiple endocrine neoplasia syndrome type 1. Eur J Endocrinol 2004; 151: 709-715
- 99 Al Brahim NY, Rambaldini G, Ezzat S. et al. Complex endocrinopathies in MEN-1: Diagnostic dilemmas in endocrine oncology. Endocr Pathol 2007; 18: 37-41
- 100 Stratakis CA, Tichomirowa MA, Boikos S. et al. The role of germline AIP, MEN1, PRKAR1A, CDKN1B and CDKN2C mutations in causing pituitary adenomas in a large cohort of children, adolescents, and patients with genetic syndromes. Clin Genet 2010; 78: 457-463
- 101 Simonds WF, Varghese S, Marx SJ. et al. Cushing's syndrome in multiple endocrine neoplasia type 1. Clin Endocrinol (Oxf) 2012; 76: 379-386
- 102 Uraki S, Ariyasu H, Doi A. et al. Hypersecretion of ACTH and PRL from pituitary adenoma in MEN1, adequately managed by medical therapy. Endocrinol Diabetes Metab Case Rep 2017; 2017: 17-0027
- 103 Makri A, Bonella MB, Keil MF. et al. Children with MEN1 gene mutations may present first (and at a young age) with Cushing disease. Clin Endocrinol (Oxf) 2018; 89: 437-443
- 104 Herath M, Parameswaran V, Thompson M. et al. Paediatric and young adult manifestations and outcomes of multiple endocrine neoplasia type 1. Clin Endocrinol (Oxf) 2019; 91: 633-638
- 105 Agarwal SK. Exploring the tumors of multiple endocrine neoplasia type 1 in mouse models for basic and preclinical studies. Int J Endocr Oncol 2014; 1: 153-161
- 106 Harding B, Lemos MC, Reed AA. et al. Multiple endocrine neoplasia type 1 knockout mice develop parathyroid, pancreatic, pituitary and adrenal tumours with hypercalcaemia, hypophosphataemia and hypercorticosteronaemia. Endocr Relat Cancer 2009; 16: 1313-1327
- 107 Pellegata NS, Quintanilla-Martinez L, Siggelkow H. et al. Germ-line mutations in p27Kip1 cause a multiple endocrine neoplasia syndrome in rats and humans. Proc Natl Acad Sci U S A 2006; 103: 15558-15563
- 108 Georgitsi M, Raitila A, Karhu A. et al. Germline CDKN1B/p27Kip1 mutation in multiple endocrine neoplasia. J Clin Endocrinol Metab 2007; 92: 3321-3325
- 109 Agarwal SK, Mateo CM, Marx SJ. Rare germline mutations in cyclin-dependent kinase inhibitor genes in multiple endocrine neoplasia type 1 and related states. J Clin Endocrinol Metab 2009; 94: 1826-1834
- 110 Frederiksen A, Rossing M, Hermann P. et al. Clinical features of multiple endocrine neoplasia type 4: Novel pathogenic variant and review of published cases. J Clin Endocrinol Metab 2019; 104: 3637-3646
- 111 Molatore S, Marinoni I, Lee M. et al. A novel germline CDKN1B mutation causing multiple endocrine tumors: Clinical, genetic and functional characterization. Hum Mutat 2010; 31: E1825-E1835
- 112 Belar O, De la Hoz C, Pérez-Nanclares G. et al. Novel mutations in MEN1, CDKN1B and AIP genes in patients with multiple endocrine neoplasia type 1 syndrome in Spain. Clin Endocrinol (Oxf) 2012; 76: 719-724
- 113 Malanga D, De GS, Riccardi M. et al. Functional characterization of a rare germline mutation in the gene encoding the cyclin-dependent kinase inhibitor p27Kip1 (CDKN1B) in a Spanish patient with multiple endocrine neoplasia (MEN)-like phenotype. Eur J Endocrinol 2012; 166: 551-560
- 114 Occhi G, Regazzo D, Trivellin G. et al. A novel mutation in the upstream open reading frame of the CDKN1B gene causes a MEN4 phenotype. PLoS Genet 2013; 9: e1003350
- 115 Pardi E, Mariotti S, Pellegata NS. et al. Functional characterization of a CDKN1B mutation in a Sardinian kindred with multiple endocrine neoplasia type 4 (MEN4). Endocr Connect 2015; 4: 1-8
- 116 Sambugaro S, Di Ruvo M, Ambrosio MR. et al. Early onset acromegaly associated with a novel deletion in CDKN1B 5'UTR region. Endocrine 2015; 49: 58-64
- 117 Polyak K, Kato JY, Solomon MJ. et al. p27Kip1, a cyclin-Cdk inhibitor, links transforming growth factor-beta and contact inhibition to cell cycle arrest. Genes Dev 1994; 8: 9-22
- 118 Chu IM, Hengst L, Slingerland JM. The Cdk inhibitor p27 in human cancer: Prognostic potential and relevance to anticancer therapy. Nat Rev Cancer 2008; 8: 253-267
- 119 Fero ML, Rivkin M, Tasch M. et al. A syndrome of multiorgan hyperplasia with features of gigantism, tumorigenesis, and female sterility in p27(Kip1)-deficient mice. Cell 1996; 85: 733-744 S0092-8674(00)81239-8
- 120 Kiyokawa H, Kineman RD, Manova-Todorova KO. et al. Enhanced growth of mice lacking the cyclin-dependent kinase inhibitor function of p27(Kip1). Cell 1996; 85: 721-732 S0092-8674(00)81238-6
- 121 Nakayama K, Ishida N, Shirane M. et al. Mice lacking p27(Kip1) display increased body size, multiple organ hyperplasia, retinal dysplasia, and pituitary tumors. Cell 1996; 85: 707-720 S0092-8674(00)81237-4
- 122 Dahia PL, Aguiar RC, Honegger J. et al. Mutation and expression analysis of the p27/kip1 gene in corticotrophin-secreting tumours. Oncogene 1998; 16: 69-76
- 123 Lidhar K, Korbonits M, Jordan S. et al. Low expression of the cell cycle inhibitor p27Kip1 in normal corticotroph cells, corticotroph tumors, and malignant pituitary tumors. J Clin Endocrinol Metab 1999; 84: 3823-3830
- 124 Korbonits M, Chahal HS, Kaltsas G. et al. Expression of phosphorylated p27(Kip1) protein and Jun activation domain-binding protein 1 in human pituitary tumors. J Clin Endocrinol Metab 2002; 87: 2635-2643
- 125 Singeisen H, Renzulli MM, Pavlicek V. et al. Multiple endocrine neoplasia type 4: A new member of the MEN family. Endocr Connect 2023; 12: e220411
- 126 Chasseloup F, Pankratz N, Lane J. et al. Germline CDKN1B loss-of-function variants cause pediatric Cushing's disease with or without an MEN4 phenotype. J Clin Endocrinol Metab 2020; 105: 1983-2005
- 127 Pappa V, Papageorgiou S, Papageorgiou E. et al. A novel p27 gene mutation in a case of unclassified myeloproliferative disorder. Leuk Res 2005; 29: 229-231
- 128 Tichomirowa MA, Lee M, Barlier A. et al. Cyclin dependent kinase inhibitor 1B (CDKN1B) gene variants in AIP mutation-negative familial isolated pituitary adenomas (FIPA) kindreds. Endocr Relat Cancer 2012; 19: 233-241
- 129 Ruiz-Heredia Y, Sánchez-Vega B, Onecha E. et al. Mutational screening of newly diagnosed multiple myeloma patients by deep targeted sequencing. Haematologica 2018; 103: e544-e548
- 130 Denes J, Swords F, Rattenberry E. et al. Heterogeneous genetic background of the association of pheochromocytoma/paraganglioma and pituitary adenoma: Results from a large patient cohort. J Clin Endocrinol Metab 2015; 100: E531-E541
- 131 Xekouki P, Szarek E, Bullova P. et al. Pituitary adenoma with paraganglioma/pheochromocytoma (3PAs) and succinate dehydrogenase defects in humans and mice. J Clin Endocrinol Metabolism 2015; 100: E710-E719
- 132 O'Toole SM, Denes J, Robledo M. et al. 15 years of paraganglioma: The association of pituitary adenomas and phaeochromocytomas or paragangliomas. Endocr Relat Cancer 2015; 22: T105-T122
- 133 Loughrey PB, Baker G, Herron B. et al. Invasive ACTH-producing pituitary gland neoplasm secondary to MSH2 mutation. Cancer Genet 2021; 256-257: 36-39
- 134 Bezawork-Geleta A, Rohlena J, Dong L. et al. Mitochondrial complex II: At the crossroads. Trends Biochem Sci 2017; 42: 312-325
- 135 Dahia PL, Ross KN, Wright ME. et al. A HIF1alpha regulatory loop links hypoxia and mitochondrial signals in pheochromocytomas. PLoS Genet 2005; 1: 72-80
- 136 Guzy RD, Sharma B, Bell E. et al. Loss of the SdhB, but Not the SdhA, subunit of complex II triggers reactive oxygen species-dependent hypoxia-inducible factor activation and tumorigenesis. Mol Cell Biol 2008; 28: 718-731
- 137 Lopez-Jimenez E, Gomez-Lopez G, Leandro-Garcia LJ. et al. Research resource: Transcriptional profiling reveals different pseudohypoxic signatures in SDHB and VHL-related pheochromocytomas. Mol Endocrinol 2010; 24: 2382-2391
- 138 Loughrey PB, Roncaroli F, Healy E. et al. Succinate dehydrogenase and MYC-associated factor X mutations in pituitary neuroendocrine tumours. Endocr Relat Cancer 2022; 29: R157-R172
- 139 Tufton N, Roncaroli F, Hadjidemetriou I. et al. Pituitary carcinoma in a patient with an SDHB mutation. Endocr Pathol 2017; 28: 320-325
- 140 Nakajima A, Kurihara H, Yagita H. et al. Mitochondrial extrusion through the cytoplasmic vacuoles during cell death. J Biol Chem 2008; 283: 24128-24135
- 141 Steiner AL, Goodman AD, Powers SR. Study of a kindred with pheochromocytoma, medullary thyroid carcinoma, hyperparathyroidism and Cushing's disease: Multiple endocrine neoplasia, type 2. Medicine (Baltimore) 1968; 47: 371-409
- 142 Naziat A, Karavitaki N, Thakker R. et al. Confusing genes: A patient with MEN2A and Cushing's disease. Clin Endocrinol (Oxf) 2013; 78: 966-968
- 143 Johnston PC, Kennedy L, Recinos PF. et al. Cushing's disease and co-existing phaeochromocytoma. Pituitary 2016; 19: 654-656
- 144 Carney JA, Gordon H, Carpenter PC. et al. The complex of myxomas, spotty pigmentation, and endocrine overactivity. Medicine (Baltimore) 1985; 64: 270-283
- 145 Kirschner LS, Carney JA, Pack SD. et al. Mutations of the gene encoding the protein kinase A type I-alpha regulatory subunit in patients with the Carney complex. Nat Genet 2000; 26: 89-92
- 146 Bertherat J, Horvath A, Groussin L. et al. Mutations in regulatory subunit type 1A of cyclic adenosine 5'-monophosphate-dependent protein kinase (PRKAR1A): Phenotype analysis in 353 patients and 80 different genotypes. J Clin Endocrinol Metab 2009; 94: 2085-2091
- 147 Stratakis CA, Kirschner LS, Carney JA. Clinical and molecular features of the Carney complex: Diagnostic criteria and recommendations for patient evaluation. J Clin Endocrinol Metab 2001; 86: 4041-4046
- 148 Rothenbuhler A, Stratakis CA. Clinical and molecular genetics of Carney complex. Best Pract Res Clin Endocrinol Metab 2010; 24: 389-399
- 149 Hernández-Ramírez LC, Trivellin G, Stratakis CA. Cyclic 3',5'-adenosine monophosphate (cAMP) signaling in the anterior pituitary gland in health and disease. Mol Cell Endocrinol 2018; 463: 72-86
- 150 Pack SD, Kirschner LS, Pak E. et al. Genetic and histologic studies of somatomammotropic pituitary tumors in patients with the "complex of spotty skin pigmentation, myxomas, endocrine overactivity and schwannomas" (Carney complex). J Clin Endocr Metab 2000; 85: 3860-3865
- 151 Stergiopoulos SG, Abu-Asab MS, Tsokos M. et al. Pituitary pathology in Carney complex patients. Pituitary 2004; 7: 73-82
- 152 Lonser RR, Mehta GU, Kindzelski BA. et al. Surgical management of Carney complex-associated pituitary pathology. Neurosurgery 2017; 80: 780-786
- 153 Kaltsas GA, Kola B, Borboli N. et al. Sequence analysis of the PRKAR1A gene in sporadic somatotroph and other pituitary tumours. Clin Endocrinol (Oxf) 2002; 57: 443-448
- 154 Sandrini F, Kirschner LS, Bei T. et al. PRKAR1A, one of the Carney complex genes, and its locus (17q22-24) are rarely altered in pituitary tumours outside the Carney complex. J Med Genet 2002; 39: e78
- 155 Yamasaki H, Mizusawa N, Nagahiro S. et al. GH-secreting pituitary adenomas infrequently contain inactivating mutations of PRKAR1A and LOH of 17q23-24. Clin Endocrinol (Oxf) 2003; 58: 464-470
- 156 Basson CT, Aretz HT. Case records of the Massachusetts General Hospital. Weekly clinicopathological exercises. Case 11-2002. A 27-year-old woman with two intracardiac masses and a history of endocrinopathy. N Engl J Med 2002; 346: 1152-1158
- 157 Hernández-Ramírez LC, Tatsi C, Lodish MB. et al. Corticotropinoma as a component of Carney complex. J Endocr Soc 2017; 1: 918-925
- 158 Kiefer FW, Winhofer Y, Iacovazzo D. et al. PRKAR1A mutation causing pituitary-dependent Cushing disease in a patient with Carney complex. Eur J Endocrinol 2017; 177: K7-K12
- 159 Stratakis CA. Clinical genetics of multiple endocrine neoplasias, Carney complex and related syndromes. J Endocrinol Invest 2001; 24: 370-383
- 160 Donis-Keller H, Dou S, Chi D. et al. Mutations in the RET proto-oncogene are associated with MEN 2A and FMTC. Hum Mol Genet 1993; 2: 851-856
- 161 Mulligan LM, Kwok JB, Healey CS. et al. Germ-line mutations of the RET proto-oncogene in multiple endocrine neoplasia type 2A. Nature 1993; 363: 458-460
- 162 Hofstra RM, Landsvater RM, Ceccherini I. et al. A mutation in the RET proto-oncogene associated with multiple endocrine neoplasia type 2B and sporadic medullary thyroid carcinoma. Nature 1994; 367: 375-376
- 163 Eng C. Multiple Endocrine Neoplasia Type 2. In: Adam MP, Mirzaa GM, Pagon RA, eds. GeneReviews. Seattle (WA): University of Washington, Seattle; 1999
- 164 Wang X. Structural studies of GDNF family ligands with their receptors-Insights into ligand recognition and activation of receptor tyrosine kinase RET. Biochim Biophys Acta 2013; 1834: 2205-2212
- 165 Canibano C, Rodriguez NL, Saez C. et al. The dependence receptor Ret induces apoptosis in somatotrophs through a Pit-1/p53 pathway, preventing tumor growth. EMBO J 2007; 26: 2015-2028
- 166 Saito T, Miura D, Taguchi M. et al. Coincidence of multiple endocrine neoplasia type 2A with acromegaly. Am J Med Sci 2010; 340: 329-331
- 167 Heinlen JE, Buethe DD, Culkin DJ. et al. Multiple endocrine neoplasia 2a presenting with pheochromocytoma and pituitary macroadenoma. ISRN Oncol 2011; 2011: 732452
- 168 Kasturi K, Fernandes L, Quezado M. et al. Cushing disease in a patient with multiple endocrine neoplasia type 2B. J Clin Transl Endocrinol Case Rep 2017; 4: 1-4
- 169 Wells SA, Pacini F, Robinson BG. et al. Multiple endocrine neoplasia type 2 and familial medullary thyroid carcinoma: An update. J Clin Endocrinol Metab 2013; 98: 3149-3164
- 170 Schultz KAP, Stewart DR, Kamihara J. et al. DICER1 Tumor Predisposition. In: Adam MP, Mirzaa GM, Pagon RA et al, eds. GeneReviews(R); Seattle (WA): 2020
- 171 Slade I, Bacchelli C, Davies H. et al. DICER1 syndrome: Clarifying the diagnosis, clinical features and management implications of a pleiotropic tumour predisposition syndrome. J Med Genet 2011; 48: 273-278
- 172 Schultz KAP, Williams GM, Kamihara J. et al. DICER1 and associated conditions: Identification of at-risk individuals and recommended surveillance strategies. Clin Cancer Res 2018; 24: 2251-2261
- 173 Hill DA, Ivanovich J, Priest JR. et al. DICER1 mutations in familial pleuropulmonary blastoma. Science 2009; 325: 965
- 174 de Kock L, Wang YC, Revil T. et al. High-sensitivity sequencing reveals multi-organ somatic mosaicism causing DICER1 syndrome. J Med Genet 2016; 53: 43-52
- 175 Klein S, Lee H, Ghahremani S. et al. Expanding the phenotype of mutations in DICER1: Mosaic missense mutations in the RNase IIIb domain of DICER1 cause GLOW syndrome. J Med Genet 2014; 51: 294-302
- 176 Brenneman M, Field A, Yang J. et al. Temporal order of RNase IIIb and loss-of-function mutations during development determines phenotype in pleuropulmonary blastoma / DICER1 syndrome: A unique variant of the two-hit tumor suppression model. F1000Res 2015; 4: 214
- 177 de Boer CM, Eini R, Gillis AM. et al. DICER1 RNase IIIb domain mutations are infrequent in testicular germ cell tumours. BMC Res Notes 2012; 5: 569
- 178 Heravi-Moussavi A, Anglesio MS, Cheng SW. et al. Recurrent somatic DICER1 mutations in nonepithelial ovarian cancers. N Engl J Med 2012; 366: 234-242
- 179 de Kock L, Plourde F, Carter MT. et al. Germ-line and somatic DICER1 mutations in a pleuropulmonary blastoma. Pediatr Blood Cancer 2013; 60: 2091-2092
- 180 Wu MK, Sabbaghian N, Xu B. et al. Biallelic DICER1 mutations occur in Wilms tumours. J Pathol 2013; 230: 154-164
- 181 Tomiak E, de KL, Grynspan D. et al. DICER1 mutations in an adolescent with cervical embryonal rhabdomyosarcoma (cERMS). Pediatr Blood Cancer 2014; 61: 568-569
- 182 de Kock L, Sabbaghian N, Soglio DBD. et al. Exploring the association between DICER1 mutations and differentiated thyroid carcinoma. J Clin Endocr Metab 2014; 99: E1072-E1077
- 183 de Kock L, Sabbaghian N, Druker H. et al. Germ-line and somatic DICER1 mutations in pineoblastoma. Acta Neuropathol 2014; 128: 583-595
- 184 de Kock L, Sabbaghian N, Plourde F. et al. Pituitary blastoma: A pathognomonic feature of germ-line DICER1 mutations. Acta Neuropathol 2014; 128: 111-122
- 185 Doros LA, Rossi CT, Yang J. et al. DICER1 mutations in childhood cystic nephroma and its relationship to DICER1-renal sarcoma. Mod Pathol 2014; 27: 1267-1280
- 186 Murray MJ, Bailey S, Raby KL. et al. Serum levels of mature microRNAs in DICER1-mutated pleuropulmonary blastoma. Oncogenesis 2014; 3: e87
- 187 Pugh TJ, Yu W, Yang J. et al. Exome sequencing of pleuropulmonary blastoma reveals frequent biallelic loss of TP53 and two hits in DICER1 resulting in retention of 5p-derived miRNA hairpin loop sequences. Oncogene 2014; 33: 5295-5302
- 188 Anglesio MS, Wang Y, Yang W. et al. Cancer-associated somatic DICER1 hotspot mutations cause defective miRNA processing and reverse-strand expression bias to predominantly mature 3p strands through loss of 5p strand cleavage. J Pathol 2013; 229: 400-409
- 189 Foulkes WD, Priest JR, Duchaine TF. DICER1: Mutations, microRNAs and mechanisms. Nat Rev Cancer 2014; 14: 662-672
- 190 Bernstein E, Caudy AA, Hammond SM. et al. Role for a bidentate ribonuclease in the initiation step of RNA interference. Nature 2001; 409: 363-366
- 191 Chendrimada TP, Gregory RI, Kumaraswamy E. et al. TRBP recruits the Dicer complex to Ago2 for microRNA processing and gene silencing. Nature 2005; 436: 740-744
- 192 Scheithauer BW, Kovacs K, Horvath E. et al. Pituitary blastoma. Acta Neuropathol 2008; 116: 657-666
- 193 Nadaf J, de Kock L, Chong AS. et al. Molecular characterization of DICER1-mutated pituitary blastoma. Acta Neuropathol 2021; 141: 929-944
- 194 Chong AS, Han H, Albrecht S. et al. DICER1 syndrome in a young adult with pituitary blastoma. Acta Neuropathol 2021; 142: 1071-1076
- 195 Liu APY, Kelsey MM, Sabbaghian N. et al. Clinical outcomes and complications of pituitary blastoma. J Clin Endocrinol Metab 2021; 106: 351-363
- 196 Liu AP, Li KK, Chow C. et al. Expanding the clinical and molecular spectrum of pituitary blastoma. Acta Neuropathol 2022; 143: 415-417
- 197 Xu Y, Zou R, Wang J. et al. The role of the cancer testis antigen PRAME in tumorigenesis and immunotherapy in human cancer. Cell Prolif 2020; 53: e12770
- 198 Martínez de LaPiscina I, Hernández-Ramírez LC, Portillo N. et al. Rare germline DICER1 variants in pediatric patients with Cushing's disease: What is their role?. Front Endocrinol (Lausanne) 2020; 11: 433
- 199 Idos G, Valle L. Lynch Syndrome. In: Adam MP, Feldman J, Mirzaa GM et al, eds. GeneReviews. Seattle (WA): University of Washington; 2004
- 200 Fishel R, Lescoe MK, Rao MR. et al. The human mutator gene homolog MSH2 and its association with hereditary nonpolyposis colon cancer. Cell 1993; 75: 1027-1038
- 201 Leach FS, Nicolaides NC, Papadopoulos N. et al. Mutations of a mutS homolog in hereditary nonpolyposis colorectal cancer. Cell 1993; 75: 1215-1225
- 202 Bronner CE, Baker SM, Morrison PT. et al. Mutation in the DNA mismatch repair gene homologue hMLH1 is associated with hereditary non-polyposis colon cancer. Nature 1994; 368: 258-261
- 203 Nicolaides NC, Papadopoulos N, Liu B. et al. Mutations of two PMS homologues in hereditary nonpolyposis colon cancer. Nature 1994; 371: 75-80
- 204 Palombo F, Hughes M, Jiricny J. et al. Mismatch repair and cancer. Nature 1994; 367: 417
- 205 Papadopoulos N, Nicolaides NC, Wei YF. et al. Mutation of a mutL homolog in hereditary colon cancer. Science 1994; 263: 1625-1629
- 206 Miyaki M, Konishi M, Tanaka K. et al. Germline mutation of MSH6 as the cause of hereditary nonpolyposis colorectal cancer. Nat Genet 1997; 17: 271-272
- 207 Kovacs ME, Papp J, Szentirmay Z. et al. Deletions removing the last exon of TACSTD1 constitute a distinct class of mutations predisposing to Lynch syndrome. Hum Mutat 2009; 30: 197-203
- 208 Ligtenberg MJ, Kuiper RP, Chan TL. et al. Heritable somatic methylation and inactivation of MSH2 in families with Lynch syndrome due to deletion of the 3' exons of TACSTD1. Nat Genet 2009; 41: 112-117
- 209 Giardiello FM, Allen JI, Axilbund JE. et al. Guidelines on genetic evaluation and management of Lynch syndrome: A consensus statement by the US Multi-Society Task Force on colorectal cancer. Gastroenterology 2014; 147: 502-526
- 210 Ijsselsteijn R, Jansen JG, de Wind N. DNA mismatch repair-dependent DNA damage responses and cancer. DNA Repair 2020; 93: 102923
- 211 Therkildsen C, Ladelund S, Rambech E. et al. Glioblastomas, astrocytomas and oligodendrogliomas linked to Lynch syndrome. Eur J Neurol 2015; 22: 717-724
- 212 Bengtsson D, Joost P, Aravidis C. et al. Corticotroph pituitary carcinoma in a patient with Lynch syndrome (LS) and pituitary tumors in a nationwide LS cohort. J Clin Endocrinol Metab 2017; 102: 3928-3932
- 213 Uraki S, Ariyasu H, Doi A. et al. Atypical pituitary adenoma with MEN1 somatic mutation associated with abnormalities of DNA mismatch repair genes; MLH1 germline mutation and MSH6 somatic mutation. Endocr J 2017; 64: 895-906
- 214 Voisin MR, Almeida JP, Perez-Ordonez B. et al. Recurrent undifferentiated carcinoma of the sella in a patient with lynch syndrome. World Neurosurg 2019; 132: 219-222
- 215 Teuber J, Reinhardt A, Reuss D. et al. Aggressive pituitary adenoma in the context of Lynch syndrome: A case report and literature review on this rare coincidence. Br J Neurosurg 2021; 1-6
- 216 Curatolo P, Bombardieri R, Jozwiak S. Tuberous sclerosis. Lancet 2008; 372: 657-668
- 217 Northrup H, Aronow ME, Bebin EM. et al. Updated international tuberous sclerosis complex diagnostic criteria and surveillance and management recommendations. Pediatr Neurol 2021; 123: 50-66
- 218 The European Chromosome 16 Tuberous Sclerosis Consortium. Identification and characterization of the tuberous sclerosis gene on chromosome 16. Cell 1993; 75: 1305-1315
- 219 van Slegtenhorst M, de Hoogt R, Hermans C. et al. Identification of the tuberous sclerosis gene TSC1 on chromosome 9q34. Science 1997; 277: 805-808
- 220 Verhoef S, Bakker L, Tempelaars AM. et al. High rate of mosaicism in tuberous sclerosis complex. Am J Hum Genet 1999; 64: 1632-1637
- 221 Qin W, Kozlowski P, Taillon BE. et al. Ultra deep sequencing detects a low rate of mosaic mutations in tuberous sclerosis complex. Hum Genet 2010; 127: 573-582
- 222 Crino PB. Evolving neurobiology of tuberous sclerosis complex. Acta Neuropathol 2013; 125: 317-332
- 223 Castro AF, Rebhun JF, Clark GJ. et al. Rheb binds tuberous sclerosis complex 2 (TSC2) and promotes S6 kinase activation in a rapamycin- and farnesylation-dependent manner. J Biol Chem 2003; 278: 32493-32496
- 224 Tee AR, Manning BD, Roux PP. et al. Tuberous sclerosis complex gene products, Tuberin and Hamartin, control mTOR signaling by acting as a GTPase-activating protein complex toward Rheb. Curr Biol 2003; 13: 1259-1268
- 225 Dibble CC, Elis W, Menon S. et al. TBC1D7 is a third subunit of the TSC1-TSC2 complex upstream of mTORC1. Mol Cell 2012; 47: 535-546
- 226 Saxton RA, Sabatini DM. mTOR Signaling in Growth, Metabolism, and Disease. Cell 2017; 169: 361-371
- 227 Panwar V, Singh A, Bhatt M. et al. Multifaceted role of mTOR (mammalian target of rapamycin) signaling pathway in human health and disease. Signal Transduct Target Ther 2023; 8: 375
- 228 Yeung RS, Katsetos CD, Klein-Szanto A. Subependymal astrocytic hamartomas in the Eker rat model of tuberous sclerosis. Am J Pathol 1997; 151: 1477-1486
- 229 Galaction-Nitelea O, Dociu I, Murgu V. A case of tuberous sclerosis with acromegaly. Rev Med Interna Neurol Psihiatr Neurochir Dermatovenerol Neurol Psihiatr Neurochir 1978; 23: 253-262
- 230 Hoffman WH, Perrin JC, Halac E. et al. Acromegalic gigantism and tuberous sclerosis. J Pediatr 1978; 93: 478-480
- 231 Bloomgarden ZT, McLean GW, Rabin D. Autonomous hyperprolactinemia in tuberous sclerosis. Arch Intern Med 1981; 141: 1513-1515
- 232 Tigas S, Carroll PV, Jones R. et al. Simultaneous Cushing's disease and tuberous sclerosis; a potential role for TSC in pituitary ontogeny. Clin Endocrinol (Oxf) 2005; 63: 694-695
- 233 Nandagopal R, Vortmeyer A, Oldfield EH. et al. Cushing's syndrome due to a pituitary corticotropinoma in a child with tuberous sclerosis: An association or a coincidence. Clin Endocrinol (Oxf) 2007; 67: 639-641
- 234 Regazzo D, Gardiman MP, Theodoropoulou M. et al. Silent gonadotroph pituitary neuroendocrine tumor in a patient with tuberous sclerosis complex: Evaluation of a possible molecular link. Endocrinol Diabetes Metab Case Rep 2018; 2018
- 235 Yehia L, Eng C. PTEN Hamartoma Tumor Syndrome. In: Adam MP, Feldman J, Mirzaa GM et al, eds. GeneReviews(R); Seattle (WA): 1993
- 236 Lloyd KM, Dennis M. Cowden's disease. A possible new symptom complex with multiple system involvement. Ann Intern Med 1963; 58: 136-142
- 237 Efstathiadou ZA, Sapranidis M, Anagnostis P. et al. Unusual case of Cowden-like syndrome, neck paraganglioma, and pituitary adenoma. Head Neck 2014; 36: E12-E16
- 238 Amatya N, Piziak V. Abstract #821: Recurrent pituitary apoplexy in Cowden syndrome: A case report. Endocr Pract 2017; 23: 173
- 239 Srichomkwun P, Houngngam N, Boonchaya-Anant P. et al. Cowden syndrome and pituitary tumours. QJM 2018; 111: 735-736
- 240 Zhang H, Li J, Lee M. et al. Pituitary carcinoma in a patient with Cowden syndrome. Am J Case Rep 2022; 23: e934846
- 241 Valdes-Socin H, Poncin J, Stevens V. et al. Familial isolated pituitary adenomas unrelated to MEN1 mutations: A follow-up of 27 patients. 10th Meeting of the Belgian Endocrine Society. 2000
- 242 Daly AF, Rixhon M, Adam C. et al. High prevalence of pituitary adenomas: A cross-sectional study in the province of Liege, Belgium. J Clin Endocrinol Metab 2006; 91: 4769-4775
- 243 Daly AF, Tichomirowa MA, Petrossians P. et al. Clinical characteristics and therapeutic responses in patients with germ-line AIP mutations and pituitary adenomas: An international collaborative study. J Clin Endocrinol Metab 2010; 95: E373-E383
- 244 Igreja S, Chahal HS, King P. et al. Characterization of aryl hydrocarbon receptor interacting protein (AIP) mutations in familial isolated pituitary adenoma families. Hum Mutat 2010; 31: 950-960
- 245 Hernández-Ramírez LC, Gabrovska P, Dénes J. et al. Landscape of familial isolated and young-onset pituitary adenomas: Prospective diagnosis in AIP mutation carriers. J Clin Endocrinol Metab 2015; 100: E1242-E1254
- 246 Vierimaa O, Georgitsi M, Lehtonen R. et al. Pituitary adenoma predisposition caused by germline mutations in the AIP gene. Science 2006; 312: 1228-1230
- 247 Daly AF, Vanbellinghen JF, Khoo SK. et al. Aryl hydrocarbon receptor-interacting protein gene mutations in familial isolated pituitary adenomas: Analysis in 73 families. J Clin Endocrinol Metab 2007; 92: 1891-1896
- 248 Iacovazzo D, Hernández-Ramírez LC, Korbonits M. Sporadic pituitary adenomas: The role of germline mutations and recommendations for genetic screening. Expert Rev Endocrinol Metab 2017; 12: 143-153
- 249 Trivellin G, Korbonits M. AIP and its interacting partners. J Endocrinol 2011; 210: 137-155
- 250 Chahal HS, Trivellin G, Leontiou CA. et al. Somatostatin analogs modulate AIP in somatotroph adenomas: The role of the ZAC1 pathway. J Clin Endocrinol Metab 2012; 97: E1411-E1420
- 251 Tuominen I, Heliovaara E, Raitila A. et al. AIP inactivation leads to pituitary tumorigenesis through defective Galpha-cAMP signaling. Oncogene 2015; 34: 1174-1184
- 252 Hernández-Ramírez LC, Morgan RML, Barry S. et al. Multi-chaperone function modulation and association with cytoskeletal proteins are key features of the function of AIP in the pituitary gland. Oncotarget 2018; 9: 9177-9198
- 253 Garcia-Rendueles AR, Chenlo M, Oroz-Gonjar F. et al. RET signalling provides tumorigenic mechanism and tissue specificity for AIP-related somatotrophinomas. Oncogene 2021; 40: 6354-6368
- 254 Sun D, Stopka-Farooqui U, Barry S. et al. Aryl hydrocarbon receptor interacting protein maintains germinal venter B cells through suppression of BCL6 degradation. Cell Rep 2019; 27: 1461-1471 e1464
- 255 Solis-Fernandez G, Montero-Calle A, Sanchez-Martinez M. et al. Aryl-hydrocarbon receptor-interacting protein regulates tumorigenic and metastatic properties of colorectal cancer cells driving liver metastasis. Br J Cancer 2022; 126: 1604-1615
- 256 Barry S, Carlsen E, Marques P. et al. Tumor microenvironment defines the invasive phenotype of AIP-mutation-positive pituitary tumors. Oncogene 2019; 38: 5381-5395
- 257 Georgitsi M, Raitila A, Karhu A. et al. Molecular diagnosis of pituitary adenoma predisposition caused by aryl hydrocarbon receptor-interacting protein gene mutations. Proc Natl Acad Sci U S A 2007; 104: 4101-4105
- 258 Cazabat L, Bouligand J, Salenave S. et al. Germline AIP mutations in apparently sporadic pituitary adenomas: Prevalence in a prospective single-center cohort of 443 patients. J Clin Endocrinol Metab 2012; 97: E663-E670
- 259 Beckers A, Aaltonen LA, Daly AF. et al. Familial isolated pituitary adenomas (FIPA) and the pituitary adenoma predisposition due to mutations in the aryl hydrocarbon receptor interacting protein (AIP) gene. Endocrine Reviews 2013; 34: 239-277
- 260 Hernández-Ramírez LC, Trivellin G, Stratakis CA. Role of phosphodiesterases on the function of aryl hydrocarbon receptor-interacting protein (AIP) in the pituitary gland and on the evaluation of AIP gene variants. Horm Metab Res 2017; 49: 286-295
- 261 Nguyen JT, Ferriere A, Tabarin A. Case report: Complete restoration of the HPA axis function in Cushing's disease with drug treatment. Front Endocrinol (Lausanne) 2024; 15: 1337741
- 262 Trofimiuk-Muldner M, Domagala B, Sokolowski G. et al. AIP gene germline variants in adult Polish patients with apparently sporadic pituitary macroadenomas. Front Endocrinol (Lausanne) 2023; 14: 1098367
- 263 Mistry A, Solomou A, Vignola ML. et al Investigating the role of AIP in pituitary tumourigenesis. Endocrine Abstracts 2019; 65 OC2.2
- 264 Trivellin G, Daly AF, Faucz FR. et al. Gigantism and acromegaly due to Xq26 microduplications and GPR101 mutation. N Engl J Med 2014; 371: 2363-2374
- 265 Trivellin G, Hernández-Ramírez LC, Swan J. et al. An orphan G-protein-coupled receptor causes human gigantism and/or acromegaly: Molecular biology and clinical correlations. Best Pract Res Clin Endocrinol Metab 2018; 32: 125-140
- 266 Trivellin G, Correa RR, Batsis M. et al. Screening for GPR101 defects in pediatric pituitary corticotropinomas. Endocr Relat Cancer 2016; 23: 357-365
- 267 Zhang Q, Peng C, Song J. et al. Germline mutations in CDH23, encoding cadherin-related 23, are associated with both familial and sporadic pituitary adenomas. Am J Hum Genet 2017; 100: 817-823
- 268 Back N, Mains RE, Eipper BA. PAM: Diverse roles in neuroendocrine cells, cardiomyocytes, and green algae. FEBS J 2022; 289: 4470-4496
- 269 Trivellin G, Daly AF, Hernández-Ramírez LC. et al. Germline loss-of-function PAM variants are enriched in subjects with pituitary hypersecretion. Front Endocrinol (Lausanne) 2023; 14: 1166076
- 270 De Sousa SMC, Shen A, Yates CJ. et al. PAM variants in patients with thyrotrophinomas, cyclical Cushing's disease and prolactinomas. Front Endocrinol (Lausanne) 2023; 14: 1305606
- 271 Zukerberg LR, Patrick GN, Nikolic M. et al. Cables links Cdk5 and c-Abl and facilitates Cdk5 tyrosine phosphorylation, kinase upregulation, and neurite outgrowth. Neuron 2000; 26: 633-646
- 272 Wu CL, Kirley SD, Xiao H. et al. Cables enhances cdk2 tyrosine 15 phosphorylation by Wee1, inhibits cell growth, and is lost in many human colon and squamous cancers. Cancer Res 2001; 61: 7325-7332
- 273 Huang JR, Tan GM, Li Y. et al. The emerging role of Cables1 in cancer and other diseases. Mol Pharmacol 2017; 92: 240-245
- 274 Shi Z, Park HR, Du Y. et al. Cables1 complex couples survival signaling to the cell death machinery. Cancer Res 2015; 75: 147-158
- 275 Kirley SD, Rueda BR, Chung DC. et al. Increased growth rate, delayed senescense and decreased serum dependence characterize cables-deficient cells. Cancer Biol Ther 2005; 4: 654-658
- 276 Zukerberg LR, DeBernardo RL, Kirley SD. et al. Loss of cables, a cyclin-dependent kinase regulatory protein, is associated with the development of endometrial hyperplasia and endometrial cancer. Cancer Res 2004; 64: 202-208
- 277 Kirley SD, D'Apuzzo M, Lauwers GY. et al. The Cables gene on chromosome 18Q regulates colon cancer progression in vivo. Cancer Biol Ther 2005; 4: 861-863
- 278 Roussel-Gervais A, Couture C, Langlais D. et al. The Cables1 gene in glucocorticoid regulation of pituitary corticotrope growth and Cushing disease. J Clin Endocrinol Metab 2016; 101: 513-522
- 279 Hernández-Ramírez LC, Gam R, Valdés N. et al. Loss-of-function mutations in the CABLES1 gene are a novel cause of Cushing's disease. Endocr Relat Cancer 2017; 24: 379-392
- 280 Franco-Álvarez AL, Torres-Morán M, Rebollar-Vega R. et al. OR2802 A novel CABLES1 missense variant associated with Cushing's disease disrupts protein structure and stability. J Endocr Soc 2023; 7 bvad114.1364
- 281 Clayton R, Burden AC, Schrieber V. et al. Secondary pituitary hyperplasia in Addison's disease. Lancet 1977; 2: 954-956
- 282 Dluhy RG, Moore TJ, Williams GH. Sella turcica enlargement and primary adrenal nsufficiency. An Intern Med 1978; 89: 513-514
- 283 Himsworth RL, Lewis JG, Rees LH. A possible ACTH secreting tumour of the pituitary developing in a conventionally treated case of Addison's disease. Clin Endocrinol (Oxf) 1978; 9: 131-139
- 284 Jara-Albarran A, Bayort J, Caballero A. et al. Probable pituitary adenoma with adrenocorticotropin hypersecretion (corticotropinoma) secondary to Addison's disease. J Clin Endocrinol Metab 1979; 49: 236-241
- 285 Aanderud S, Bassoe HH. A pituitary tumour with possible ACTH and TSH hypersecretion in a patient with Addison's disease and primary hypothyroidism. Acta Endocrinol (Copenh) 1980; 95: 181-184
- 286 Krautli B, Muller J, Landolt AM. et al. ACTH-producing pituitary adenomas in Addison's disease: Two cases treated by transsphenoidal microsurgery. Acta Endocrinol (Copenh) 1982; 99: 357-363
- 287 Scheithauer BW, Kovacs K, Randall RV. The pituitary gland in untreated Addison's disease. A histologic and immunocytologic study of 18 adenohypophyses. Arch Pathol Lab Med 1983; 107: 484-487
- 288 Yanase T, Sekiya K, Ando M. et al. Probable ACTH-secreting pituitary tumour in association with Addison's disease. Acta Endocrinol (Copenh) 1985; 110: 36-41
- 289 Sugiyama K, Kimura M, Abe T. et al. Hyper-adrenocorticotropinemia in a patient with Addison's disease after treatment with corticosteroids. Intern Med 1996; 35: 555-559
- 290 Fan S, Jiang Y, Yao Y. et al. Pituitary ACTH-secreting adenoma in Addison's disease: A case report. Clin Neurol Neurosurg 2013; 115: 2543-2546
- 291 Royrvik EC, Husebye ES. The genetics of autoimmune Addison disease: Past, present and future. Nat Rev Endocrinol 2022; 18: 399-412
- 292 Auer MK, Nordenstrom A, Lajic S. et al. Congenital adrenal hyperplasia. Lancet 2023; 401: 227-244
- 293 Horrocks PM, Franks S, Hockley AD. et al. An ACTH-secreting pituitary tumour arising in a patient with congenital adrenal hyperplasia. Clin Endocrinol (Oxf) 1982; 17: 457-468
- 294 Boronat M, Carrillo A, Ojeda A. et al. Clinical manifestations and hormonal profile of two women with Cushing's disease and mild deficiency of 21-hydroxylase. J Endocrinol Invest 2004; 27: 583-590
- 295 Haase M, Schott M, Kaminsky E. et al. Cushing's disease in a patient with steroid 21-hydroxylase deficiency. Endocr J 2011; 58: 699-706
- 296 McCabe ER. DAX1: Increasing complexity in the roles of this novel nuclear receptor. Mol Cell Endocrinol 2007; 265-266: 179-182
- 297 Peter M, Viemann M, Partsch CJ. et al. Congenital adrenal hypoplasia: Clinical spectrum, experience with hormonal diagnosis, and report on new point mutations of the DAX-1 gene. J Clin Endocrinol Metab 1998; 83: 2666-2674
- 298 Reutens AT, Achermann JC, Ito M. et al. Clinical and functional effects of mutations in the DAX-1 gene in patients with adrenal hypoplasia congenita. J Clin Endocrinol Metab 1999; 84: 504-511
- 299 Guran T, Buonocore F, Saka N. et al. Rare causes of primary adrenal insufficiency: Genetic and clinical characterization of a large nationwide cohort. J Clin Endocrinol Metab 2016; 101: 284-292
- 300 Tabarin A, Achermann JC, Recan D. et al. A novel mutation in DAX1 causes delayed-onset adrenal insufficiency and incomplete hypogonadotropic hypogonadism. J Clin Invest 2000; 105: 321-328
- 301 Mantovani G, Ozisik G, Achermann JC. et al. Hypogonadotropic hypogonadism as a presenting feature of late-onset X-linked adrenal hypoplasia congenita. J Clin Endocrinol Metab 2002; 87: 44-48
- 302 Achermann JC, Vilain EJ. NR0B1-Related Adrenal Hypoplasia Congenita. In: Adam MP, Mirzaa GM, Pagon RA et al, eds. GeneReviews(R); Seattle (WA): 2018
- 303 Weiss L, Mellinger RC. Congenital adrenal hypoplasia--an X-linked disease. J Med Genet 1970; 7: 27-32
- 304 Muscatelli F, Strom TM, Walker AP. et al. Mutations in the DAX-1 gene give rise to both X-linked adrenal hypoplasia congenita and hypogonadotropic hypogonadism. Nature 1994; 372: 672-676
- 305 Zanaria E, Muscatelli F, Bardoni B. et al. An unusual member of the nuclear hormone receptor superfamily responsible for X-linked adrenal hypoplasia congenita. Nature 1994; 372: 635-641
- 306 Ikeda Y, Swain A, Weber TJ. et al. Steroidogenic factor 1 and Dax-1 colocalize in multiple cell lineages: Potential links in endocrine development. Mol Endocrinol 1996; 10: 1261-1272
- 307 De Menis E, Roncaroli F, Calvari V. et al. Corticotroph adenoma of the pituitary in a patient with X-linked adrenal hypoplasia congenita due to a novel mutation of the DAX-1 gene. Eur J Endocrinol 2005; 153: 211-215
- 308 Martins CS, Camargo RC, Coeli-Lacchini FB. et al. USP8 mutations and cell cycle regulation in corticotroph adenomas. Horm Metab Res 2020; 52: 117-123
- 309 Lin AL, Rudneva VA, Richards AL. et al. Genome-wide loss of heterozygosity predicts aggressive, treatment-refractory behavior in pituitary neuroendocrine tumors. Acta Neuropathol 2024; 147: 85