Next Article in Journal
GPNMB Extracellular Fragment Protects Melanocytes from Oxidative Stress by Inhibiting AKT Phosphorylation Independent of CD44
Next Article in Special Issue
Pituitary–Adrenal Responses and Glucocorticoid Receptor Expression in Critically Ill Patients with COVID-19
Previous Article in Journal
Skin Barrier Dysregulation in Psoriasis
Previous Article in Special Issue
Glucocorticoid Resistance: Interference between the Glucocorticoid Receptor and the MAPK Signalling Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Primary Generalized Glucocorticoid Resistance and Hypersensitivity Syndromes: A 2021 Update

1
First Department of Pediatrics, Division of Endocrinology, Metabolism and Diabetes, “Aghia Sophia” Children’s Hospital, National and Kapodistrian University of Athens Medical School, 11527 Athens, Greece
2
Center of Clinical, Experimental Surgery and Translational Research, Division of Endocrinology and Metabolism, Biomedical Research Foundation of the Academy of Athens, 11527 Athens, Greece
3
University Research Institute of Maternal and Child Health and Precision Medicine, University of Athens, 11527 Athens, Greece
4
Department of Molecular Genetics, Function and Therapy, The Cyprus Institute of Neurology and Genetics, Nicosia 2371, Cyprus
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(19), 10839; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms221910839
Submission received: 6 September 2021 / Revised: 25 September 2021 / Accepted: 27 September 2021 / Published: 7 October 2021
(This article belongs to the Special Issue Glucocorticoid Signaling Pathway: From Bench to Bedside)

Abstract

:
Glucocorticoids are the final products of the neuroendocrine hypothalamic–pituitary—adrenal axis, and play an important role in the stress response to re-establish homeostasis when it is threatened, or perceived as threatened. These steroid hormones have pleiotropic actions through binding to their cognate receptor, the human glucocorticoid receptor, which functions as a ligand-bound transcription factor inducing or repressing the expression of a large number of target genes. To achieve homeostasis, glucocorticoid signaling should have an optimal effect on all tissues. Indeed, any inappropriate glucocorticoid effect in terms of quantity or quality has been associated with pathologic conditions, which are characterized by short-term or long-lasting detrimental effects. Two such conditions, the primary generalized glucocorticoid resistance and hypersensitivity syndromes, are discussed in this review article. Undoubtedly, the tremendous progress of structural, molecular, and cellular biology, in association with the continued progress of biotechnology, has led to a better and more in-depth understanding of these rare endocrinologic conditions, as well as more effective therapeutic management.

1. Glucocorticoid Signaling

Glucocorticoids are steroid hormones, which are biosynthesized in the zona fasciculata of the adrenal cortex, as well as in many other extra-adrenal organs, including the intestine, thymus, and skin [1,2]. These lipophilic molecules are secreted inthe peripheral circulation in response to stressors, as well as in an ultradian and circadian fashion [3,4,5,6], and play a fundamental role in the establishment of resting and threatened homeostasis [7]. Indeed, glucocorticoids act as the end-products of a neuroendocrine axis, the hypothalamic–pituitary–adrenal (HPA) axis. This influences a broad spectrum of physiologic functions, such as carbohydrate, fat, and protein metabolism; cardiovascular tone, memory, behavior, and cognition; as well as inflammatory and immune responses [8,9,10]. In addition to their essential homeostatic functions, glucocorticoids can change the percentage of methylation in many cytosine–guanine dinucleotides (CpG) within a growing number of promoters by altering the expression and/or the activity of methyltransferases or demethylases, ultimately leading to epigenetic modifications of gene expression [11,12]. All these pleiotropic glucocorticoid effects are mediated by genomic, non-genomic, and mitochondrial actions of a ubiquitously expressed protein, the human glucocorticoid receptor (hGR), which is a member of the steroid receptor family of transcription factors [13].
The hGR is expressed by the NR3C1 gene, which is located on chromosome 5 (5q31.3), and contains 10 exons. Exon 1 is a non-expressing DNA region, and contains multiple regulatory sites that influence the expression of the NR3C1 gene positively or negatively. Exons 2-9α/9β contribute to the expression of hGR [14]. The alternative usage of exons 9α or 9β produces the two main protein isoforms: the hGRα and the hGRβ. The hGRα is expressed in almost every cell, and is predominantly located in the cytoplasmic region of the target cell. It is capable of binding natural glucocorticoids (cortisol in humans, corticosterone in rodents) and synthetic glucocorticoid analogues, and mediates all the above-mentioned glucocorticoid actions [15]. On the other hand, the hGRβ was initially reported as an inhibitor of the transcriptional activity of hGRα via several molecular mechanisms [16,17,18,19]. Accumulating evidence indicates that this isoform is substantially involved in insulin signal transduction, pathways of gluconeogenesis, and inflammatory processes in mouse liver [19,20,21]. Furthermore, the hGRβ was found to contribute to the formation of gliomas and migration of bladder malignant cells, suggesting an evolving role of this receptor isoform in crucial steps of carcinogenesis [19,22,23,24]. In addition to these two protein isoforms, Lu and Cidlowski proved the existence of eight additional GRα isoforms (hGRα-A, hGRα-B, hGRα-C1, hGRα-C2, hGRα-C3, hGRα-D1, hGRα-D2, and hGRα-D3), which are generated by alternative initiations of GRα mRNA translation [25]. These isoforms have progressively shorter N-terminal domains and different properties in terms of subcellular localization, expression, and function [26,27]. We speculated that the same molecular mechanisms may occur in the process of hGRβ mRNA translation into eight beta isoforms [15].
Within the target cell, glucocorticoids signal through their cognate receptor, which is located mainly in the cytoplasm, and forms a protein complex with immunophillins, such as FKBP51 and FKBP52, and heat shock proteins (e.g., HSP90 and HSP70) (Figure 1) [28,29]. Upon ligand binding, the structure of the receptor changes, thereby allowing the GR to dissociate from the interacting proteins and then move to the nucleus. In the nucleus, the ligand-activated hGRα forms dimers (homo-or heterodimers) that bind to specified DNA sequences, which are termed glucocorticoid response elements (GREs). GREs are located in the regulatory or promoter regions of glucocorticoid-responsive genes, thereby increasing or decreasing their expression [30]. In addition, the ligand-bound hGRα may influence the transcription rate of several non-target genes by interacting as a monomer with other transcription factors, including the nuclear factor-κB (NF-κB) [31], the activator protein-1 (AP-1) [32], and signal transducers and activators of transcription (STATs) [33]. In addition to genomic actions, glucocorticoids may induce some rapid effects. These “nongenomic” effects occur in cells that do not have a nucleus, and do not depend on transcription/translation processes [34,35]. The molecular mechanisms, which are responsible for the nongenomic glucocorticoid effects, remain under investigation; however, in vitro studies have uncovered the role of a membrane-anchored hGRα, which activates rapid kinase signaling pathways, including those of the mitogen-activated protein kinases (MAPK) or the phosphatidylinositol 3-kinase (PI3K) (Figure 1) [36]. Finally, the GR has also been detected in mitochondria, suggesting a functional role of the receptor in key mitochondrial functions, including apoptosis, production of energy, calcium homeostasis, thermogenesis, and response to stressors [37]. More than two decades ago, GREs were identified in the regulatory regions of the mitochondrial genome, termed as “D-loop”, indicating the GR as regulator of mitochondrial transcription [37,38]. In addition to direct binding of GR onto mitochondrial GREs, the expression of some mitochondrial genes is indirectly regulated by interactions between the nuclear GR and GRE, leading to the induction of mitochondrial RNA-processing enzymes and transcription factors, as well as important nuclear respiratory factors (Figure 1) [39,40].
In humans, the glucocorticoid signal transduction pathway exerts its functions in an inverted U—shaped activity-effect curve [8,40,41]. The optimal effect contributes to homeostasis or eustasis, and is achieved in the middle of glucocorticoid signaling activity. Suboptimal effects may be present on each side of the inverted U—shaped curve, and can cause insufficient adaptation. This state is termed allostasis or cacostasis. Indeed, the impaired activation of the glucocorticoid signal transduction may have short- or long-term detrimental effects for humans. These conditions may present with clinical manifestations of glucocorticoid resistance or glucocorticoid hypersensitivity [40,41].

2. Primary Generalized Glucocorticoid Resistance or Chrousos Syndrome

Primary generalized glucocorticoid resistance, or Chrousos syndrome, is an extremely rare endocrine disorder that affects all tissues expressing the hGR, and is characterized by a generalized, partial decreased sensitivity to glucocorticoids [40,41,42,43,44,45,46,47,48,49,50,51]. Chrousos syndrome can be inherited in an autosomal recessive or dominant fashion or may be sporadic due to de novo genetic defects (point mutations, deletions or insertions) in the NR3C1 gene [50,51]. In patients with Chrousos syndrome, the presence of defective hGRs in both hypothalamus and pituitary results in impaired glucocorticoid negative feedback loops, causing compensatory hypersecretion of corticotropin-releasing hormone (CRH), arginine vasopressin (AVP), and adrenocorticotropic hormone (ACTH) (Figure 2). The resultant high plasma ACTH concentrations lead to adrenal hyperplasia; stimulation of the adrenal cortex; activation of biosynthesis; and the release of cortisol, adrenal androgens (androstenedione, dehydroepiandrosterone (DHEA), and DHEA-sulfate (DHEAS)), and steroids with mineralocorticoid activity (corticosterone and deoxycorticosterone) (Figure 2) [50,51].
According to the pathophysiologic mechanisms discussed above, patients with Chrousos syndrome might present with biochemical hypercortisolism without manifestations of Cushing’s syndrome [51]. The clinical phenotype of the condition ranges from asymptomatic cases to clinical manifestations of mineralocorticoid excess (hypertension and hypokalemic alkalosis) and/or androgen excess, including ambiguous genitalia at birth with 46, XX, hirsutism, acne, male-pattern hair loss, precocious puberty, amenorrhea, hypofertility and menstrual irregularities in females, and oligospermia in males [40,41,42,43,44,45,46,47,48,49,50,51]. The increased secretion of CRH and AVP may cause anxiety and depression, while hypercortisolism may explain the chronic fatigue in these patients [51]. Glucocorticoid deficiency has been described in rare cases: in adult patients with chronic fatigue [47,52,53]; in a 2-year-old girl with generalized seizures, hypoglycemia, hypokalemia, increased arterial pressure and premature pubarche [54]; and in a newborn with hypoglycemia, severe hypertension, easy “fatigability” with feeding and growth hormone deficiency [55]. The molecular basis of Chrousos syndrome has been attributed to inherited or de novo genetic defects in the NR3C1 gene, which are depicted in Figure 3 and presented in Table 1 [52,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86].
The diagnostic workup of Chrousos syndrome includes a complete medical history with particular emphasis on symptoms and signs suggestive of impaired activity of the HPA axis or possible CNS dysfunction (e.g., visual impairment, headaches or seizures) [41,42,43,44,49]. In adolescent girls and women, any irregularity in menstrual cycles should be documented. The growth velocity and puberty progress in children and adolescents should be carefully assessed through clinical examination. In addition, specific attention should be given to any of the above-discussed clinical manifestations indicative of mineralocorticoid or/and androgen excess [41,42,43,44,49]. The initial endocrinologic evaluation includes the determination of the morning (08:00 h) concentrations of serum cortisol, plasma ACTH, serum aldosterone, plasma renin activity (recumbent and upright), androgens (androstenedione, testosterone, DHEA and DHEAS), and insulin. Triglycerides, total cholesterol, LDL, HDL, and fasting glucose concentrations should be simultaneously measured at 08:00 h. The confirmation of diagnosis of Chrousos syndrome depends on the determination of the 24 h urinary free cortisol (UFC) excretion on a minimum of two consecutive days. The 24 h UFC excretion might be increased by 50-fold in comparison with the upper normal values, while serum cortisol concentrations might reach 7-fold higher in comparison with the highest values of its normal range. Moreover, ACTH concentrations might be normal or elevated, whereas the circadian rhythms of the HPA axis remain within the normal range [41,42,43,44,49].
To evaluate how the HPA axis responds to exogenous synthetic glucocorticoids, patients with suspected Chrousos syndrome should undergo a dexamethasone suppression test. According to the standard protocol, progressively increasing doses of the synthetic glucocorticoid analogue dexamethasone (0.3, 0.6, 1.0, 1.5, 2.0, 2.5, and 3.0 mg) should be given per os to the patients at midnight every other day [41,42,43,44,49]. Serum cortisol concentrations are measured at 08.00 h the following morning. Crucially, serum dexamethasone concentrations should also be measured to eliminate the possibility of increased clearance or reduced absorption of this synthetic glucocorticoid analogue. Due to the impaired negative feedback loops, patients with Chrousos syndrome display resistance of the HPA axis following a dexamethasone suppression test. In addition to endocrinologic evaluation, a pituitary magnetic resonance imaging scan should be performed to exclude the presence of pituitary adenoma [41,42,43,44,49].
Thymidine incorporation assays and dexamethasone-binding assays performed on mononuclear cells of systemic circulation remain the appropriate ex vivo methods to confirm the diagnosis of Chrousos syndrome [41,42,43,44,49]. Compared to control participants, affected subjects show resistance to the suppression of phytohemaglutinin-stimulated thymidine incorporation following exposure to dexamethasone, and display lower binding of the administered tritiated dexamethasone to their mutant receptors. Finally, the identification of the genetic defects in patients with Chrousos syndrome is achieved by the Sanger sequencing of the NR3C1 gene, including the junctions of introns and exons [41,42,43,44,49].
Mineralocorticoid-sparing synthetic glucocorticoids, such as dexamethasone, should be administered per os at high doses only in symptomatic patients with Chrousos syndrome (1–3 mg given once daily at night) [41,42,43,44,49]. The main aim of treatment in these patients is to effectively suppress the increased secretion of ACTH, thereby decreasing the concentrations of adrenal-derived steroids with mineralocorticoid or androgenic activity. To avoid the development of an ACTH-secreting adenoma, the dose of dexamethasone in affected patients should be strictly titrated, based on the clinical manifestations and biochemical findings of the patients [41,42,43,44,49].

3. Recent Advances in Primary Generalized Glucocorticoid Resistance or Chrousos Syndrome

In 2018, Vitellius et al. provided, for the first time, an estimation of the prevalence of NR3C1 mutations in 100 patients who had bilateral adrenal hyperplasia, increased arterial pressure, and/or hypercortisolism, but without any stigmata of Cushing’s syndrome [79,87]. Notably, they found that five of these patients (5%) carried novel heterozygous NR3C1 mutations. The two mutations (R469X and R477S) were identified in the DNA-binding domain, and the rest of them (R491X, Q501H, and L672P) were identified in the ligand-binding domain of hGRα [79,87]. This high prevalence indicates that a number of cases with NR3C1 genetic defects might be present and not identified. Consequently, clinicians should recommend Sanger sequencing of the NR3C1 in patients with symptoms and signs suggestive of Chrousos syndrome [87].
Tatsi and coworkers recently described an interesting case of a female patient who initially presented at almost 3 years old with hypertension, hypoglycemic seizures, and hypokalemia [68]. The patient had endocrinologic findings suggestive of Chrousos syndrome, and was subsequently treated with antihypertensive medications, potassium supplementation, and dexamethasone at a dose of 1 mg daily. The case was first published by Nader et al., who identified a novel heterozygous point NR3C1 mutation (c.2141G > A, p.R714Q) by using mRNA/cDNA, and proceeded to the functional characterization of this point mutation [54]. At the age of 11 years 9 months, the patient was receiving a 12 mg dose of dexamethasone daily, and was found to have several neuroimaging findings consistent with severe hypertensive encephalopathy. The daily dose of dexamethasone was increased to 14 mg to achieve better control of arterial hypertension [68]. Using whole-exome sequencing, Tatsi et al. identified two nonsynonymous heterozygous mutations in the NR3C1 gene: the one previously published by Nader et al. [54], and a novel mutation in exon 2, c.592G > T, p.E198X. To the best of our knowledge, this is the first published case of two heterozygous point mutations being harbored in the NR3C1 gene, suggesting that it is better to use whole-exome sequencing when searching for NR3C1 genetic defects [68].
In addition to the significant contribution of next-generation sequencing to the identification of novel NR3C1 mutations in patients with Chrousos syndrome, the last decade has seen the tremendous progress of structural biology in an attempt to determine how a specific genetic defect of the NR3C1 gene leads to such conformational change(s) of the hGRα, ultimately leading to tissue resistance to glucocorticoids [72,74,75,76,88,89]. In addition to our previously published work on the structural biology of hGRαV423A [74], hGRαV575G [75], hGRαH726R [76], and hGRαT556I [72], two recent publications shed more light on the structural alterations caused by already known NR3C1 mutations that resulted in the clinical manifestations of Chrousos syndrome [88,89]. Monteiro et al. performed computational analysis for eight mutations located in the ligand-binding domain of the hGRα to provide a deeper understanding on the atom scale of how they cause glucocorticoid resistance [88]. They found that these mutations caused decreased binding to cortisol, as well as alterations in a number of loop conformations due to changes in the motion of the mutant protein [88]. In a second recently published study, Kaziales et al. used in vitro and in silico methods to describe the structural biology of the mutant hGRαL773P receptor [89]. The authors demonstrated significant conformational alterations regarding the binding of the mutant receptor to dexamethasone, to an NCOA-2 co-activator peptide, to DNA, and to HSP90 [89].

4. Primary Generalized Glucocorticoid Hypersensitivity Syndrome

Primary generalized glucocorticoid hypersensitivity is also extremely rare, and represents the inverse condition of primary generalized glucocorticoid resistance syndrome. This syndrome is characterized by increased tissue glucocorticoid sensitivity; therefore, affected patients usually present with the cardinal clinical manifestations of metabolic syndrome, such as obesity, dyslipidemia, insulin resistance, and hypertension. However, the concentrations of serum cortisol are low or even undetectable due to compensatory hypoactivation of the HPA axis (Figure 2) [40,42,43,44]. The molecular basis of this condition remains under investigation. A small number of cases with primary generalized glucocorticoid hypersensitivity syndrome have been associated with NR3C1 polymorphisms, including the N363S and Bcl1, or mutations [40,42,43,44,90].
The first case of primary generalized glucocorticoid hypersensitivity syndrome was a woman who harbored an activating NR3C1 mutation and presented with visceral obesity, type 2 diabetes, increased arterial pressure, and dyslipidemia [90]. Sequencing analysis of the NR3C1 gene showed a heterozygous point substitution G→C at nucleotide position 1201, which caused a substitution of aspartic acid (D) for histidine (H) at amino acid position 401, which is located in the N-terminal domain of hGRα. Charmandari et al. performed in vitro studies to investigate the function of hGRαD401H [90]. Transactivation studies showed that hGRαD401H induced the expression of glucocorticoid target genes in response to increasing concentrations of dexamethasone. However, the mutation D401H did not alter the affinity of the mutant receptor for the ligand, had no effect on the time required to complete its cytoplasmic to nucleus translocation, did not affect its ability to bind to GREs of target genes, and did not change its ability to interact with the GRIP1 coactivator [90]. Notably, a number of cases with primary generalized glucocorticoid hypersensitivity syndrome and no mutated NR3C1 gene have also been reported [91,92,93,94,95].
Our research team has also described a novel case of transient primary generalized glucocorticoid hypersensitivity syndrome, and performed transcriptomics during the acute and the resolution phase of the disease [96]. The case index was a 9-year-old girl with a long-standing history of weight gain, moon facies, buffalo hump, acanthosis nigricans, purple skin striae, myopathy, hirsutism, and gradual deceleration of the growth velocity. Remarkably, the 08:00 h plasma ACTH and serum cortisol concentrations were undetectable (<1 pg/mL and 0.025 mcg/dL, respectively), and both remained at low concentrations during the 24 h period. A CRH stimulation test was performed and revealed no increase in the concentrations of plasma ACTH and serum cortisol. A pituitary MRI scan was normal [96]. Both the clinical manifestations and endocrinologic findings were suggestive of generalized glucocorticoid hypersensitivity. Sequencing analysis of the NR3C1 gene did not identify any activating mutations or polymorphisms associated with glucocorticoid hypersensitivity. Notably, the clinical features of the patient were gradually eliminated over the next 3 months, indicating the diagnosis of transient generalized glucocorticoid hypersensitivity [96]. RNA was isolated from the peripheral blood mononuclear cells at the acute phase and the resolution phase of the disease, and RNA samples were subjected to transcriptomic analysis using RNA sequencing. Of note, 106 genes were induced, and 797 were repressed during the acute phase, as compared with the resolution phase. Most of them were identified as NF-κB responsive genes, suggesting that a viral or bacterial protein might function as a hGRα co-activator, leading to increased transcriptional activity of the receptor [96].

5. Novel Insights in Primary Generalized Glucocorticoid Hypersensitivity Syndrome

In addition to the D401H mutation and the well-studied polymorphisms of the NR3C1 gene, a novel hyperactive G1376T single-nucleotide polymorphism (SNP) isoform was recently identified in hospitalized patients with burn injuries on at least 20% of their total body surface area [97]. This SNP resulted in a single substitution of glutamine to valine at amino acid position 459, which is located in the DNA-binding domain of the receptor. The hGRαG459V-mediated transcriptional activity was investigated in reporter assays following exposure to hydrocortisone, the hGR antagonist RU486 (mifepristone) alone, or RU486 followed by hydrocortisone [97]. Upon exposure to hydrocortisone, the hGRαG459V displayed 30-fold greater transcriptional activity compared to the wild-type hGRα. Surprisingly, the hGRαG459V showed elevated transcriptional activity when exposed to RU486. When transfected cells were treated with RU486 followed by hydrocortisone, the transcriptional activity of the mutant receptor was repressed, but greater than that noted upon exposure only to RU486 [97]. These unique transcriptional properties of the mutant receptor hGRαG459V might be attributed to the localization of this substitution. Further studies are needed to investigate the function of the mutant receptor [97].
From a therapeutic point of view, Liu et al. reported a case of a 27-year-old male with clinical manifestations and endocrinologic findings suggestive of primary generalized glucocorticoid hypersensitivity [98]. A pituitary MRI scan was normal, and an adrenal computed tomography (CT) scan revealed bilateral adrenal atrophy and an increased content of subcutaneous fat in the abdomen. The patient was treated using mifepristone with beneficial effects on electrolyte balance, lipid metabolism, and bone turnover [98]. Larger studies should confirm the therapeutic effects of mifepristone in patients with this rare condition.

6. Beyond NR3C1 Gene Mutations and Polymorphisms: Tissue Glucocorticoid Sensitivity in the –OMICs Era

Primary generalized glucocorticoid resistance and hypersensitivity syndromes represent the two sides of the inverted U-shaped curve of the homeostatic function of glucocorticoid signaling. Undoubtedly, these allostatic conditions may have detrimental adverse effects for humans [8]. However, significant variability in tissue glucocorticoid sensitivity has been documented among healthy adults, as evidenced by differences in therapeutic response and in the frequency of glucocorticoid adverse effects. In an attempt to identify a motif that could distinguish healthy individuals with differences in tissue glucocorticoid sensitivity, we recruited 101 adults with unremarkable medical history [99,100]. The participants were given 0.25 mg of dexamethasone at midnight, and serum cortisol concentrations were measured the following morning at 08:00 h. According to their serum cortisol concentrations, participants were rank-ordered into the 10% most glucocorticoid sensitive and the 10% most glucocorticoid resistant. One month later, DNA, RNA, and plasma samples were collected from the 22 subjects. A sequencing analysis of the NR3C1 gene did not identify any mutations or polymorphisms in the 22 selected subjects [99,100]. In addition, no statistically significant difference was found in telomere length between the glucocorticoid-resistant and the glucocorticoid sensitive group [99]. A transcriptomic analysis through RNA sequencing revealed a large percentage of differentially expressed genes, which play an important role in the maintenance of telomere length, systemic lupus erythematosus, Alzheimer’s disease, and Parkinson’s disease [99]. Untargeted metabolomics analysis in plasma samples from the selected participants showed a detrimental profile of metabolites in the most glucocorticoid-sensitive compared to the most glucocorticoid-resistant of the healthy subjects, and offered a metabolic motif that could distinguish between the two groups of participants in subsequent larger studies [100]. We speculate that the observed differences among healthy individuals at the –OMICS level might be attributed to epigenetic modifications that occurred during acute or chronic stress at early life stages. Nevertheless, tissue glucocorticoid sensitivity, in the context of primary generalized glucocorticoid resistance, hypersensitivity syndromes, or healthy individuals, remains at the forefront of translational research successfully connecting clinical observations with laboratory findings.

Author Contributions

Conceptualization: N.C.N. and E.C.; writing—original draft preparation: N.C.N.; writing—review and editing: N.C.N. and E.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by internal funds provided by the Division of Endocrinology and Metabolism, Biomedical Research Foundation of the Academy of Athens, Athens, Greece.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nicolaides, N.C.; Charmandari, E.; Chrousos, G.P. Overview of Glucocorticoids. In Encyclopedia of Endocrine Diseases, 2nd ed.; Huhtaniemi, I., Martini, L., Eds.; Elsevier BV: New York, NY, USA, 2018; Volume 3, pp. 64–71. [Google Scholar]
  2. Slominski, R.M.; Tuckey, R.; Manna, P.R.; Jetten, A.M.; Postlethwaite, A.; Raman, C.; Slominski, A.T. Extra-adrenal glucocorticoid biosynthesis: Implications for autoimmune and inflammatory disorders. Genes Immun. 2020, 21, 150–168. [Google Scholar] [CrossRef]
  3. Focke, C.M.; Iremonger, K.J. Rhythmicity matters: Circadian and ultradian patterns of HPA axis activity. Mol. Cell. Endocrinol. 2020, 501, 110652. [Google Scholar] [CrossRef]
  4. Nicolaides, N.C.; Charmandari, E.; Kino, T.; Chrousos, G.P. Stress-Related and Circadian Secretion and Target Tissue Actions of Glucocorticoids: Impact on Health. Front. Endocrinol. 2017, 8, 70. [Google Scholar] [CrossRef]
  5. Agorastos, A.; Nicolaides, N.C.; Bozikas, V.P.; Chrousos, G.P.; Pervanidou, P. Multilevel Interactions of Stress and Circadian System: Implications for Traumatic Stress. Front. Psychiatry 2020, 10. [Google Scholar] [CrossRef] [PubMed]
  6. Nicolaides, N.C.; Kyratzi, E.; Lamprokostopoulou, A.; Chrousos, G.P.; Charmandari, E. Stress, the Stress System and the Role of Glucocorticoids. Neuroimmunomodulation 2015, 22, 6–19. [Google Scholar] [CrossRef] [PubMed]
  7. Stavrou, S.; Critselis, E.; Darviri, C.; Charmandari, E.; Nicolaides, N.C.; Chrousos, G.P. Paediatric stress: From neuroendocrinology to contemporary disorders. Eur. J. Clin. Investig. 2017, 47, 262–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Chrousos, G.P. Stress and disorders of the stress system. Nat. Rev. Endocrinol. 2009, 5, 374–381. [Google Scholar] [CrossRef] [PubMed]
  9. Hardy, R.S.; Raza, K.; Cooper, M.S. Therapeutic glucocorticoids: Mechanisms of actions in rheumatic diseases. Nat. Rev. Rheumatol. 2020, 16, 133–144. [Google Scholar] [CrossRef] [Green Version]
  10. Syed, A.P.; Greulich, F.; Ansari, S.A.; Uhlenhaut, N.H. Anti-inflammatory glucocorticoid action: Genomic insights and emerging concepts. Curr. Opin. Pharmacol. 2020, 53, 35–44. [Google Scholar] [CrossRef] [PubMed]
  11. Zannas, A.S.; Chrousos, G.P. Epigenetic programming by stress and glucocorticoids along the human lifespan. Mol. Psychiatry 2017, 22, 640–646. [Google Scholar] [CrossRef]
  12. Bartlett, A.A.; Lapp, H.E.; Hunter, R.G. Epigenetic Mechanisms of the Glucocorticoid Receptor. Trends Endocrinol. Metab. 2019, 30, 807–818. [Google Scholar] [CrossRef]
  13. Nicolaides, N.C.; Chrousos, G.; Kino, T. Glucocorticoid Receptor; Feingold, K.R., Anawalt, B., Boyce, A., Chrousos, G., de Herder, W.W., Dhatariya, K., Dungan, K., Grossman, A., Hershman, J.M., Hofland, J., et al., Eds.; MDText.com, Inc.: South Dartmouth, MA, USA, 2000. [Google Scholar]
  14. Nicolaides, N.C.; Galata, Z.; Kino, T.; Chrousos, G.P.; Charmandari, E. The human glucocorticoid receptor: Molecular basis of biologic function. Steroids 2010, 75, 1–12. [Google Scholar] [CrossRef] [Green Version]
  15. Chrousos, G.P.; Kino, T. Intracellular Glucocorticoid Signaling: A Formerly Simple System Turns Stochastic. Sci. Signal. 2005, 2005, pe48. [Google Scholar] [CrossRef]
  16. Bamberger, C.M.; de Castro, M.; Chrousos, G.P. Glucocorticoid receptor beta, a potential endogenous inhibitor of glucocorticoid action in humans. J. Clin. Investig. 1995, 95, 2435–2441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Charmandari, E.; Chrousos, G.P.; Ichijo, T.; Bhattacharyya, N.; Vottero, A.; Souvatzoglou, E.; Kino, T. The Human Glucocorticoid Receptor (hGR) β Isoform Suppresses the Transcriptional Activity of hGRα by Interfering with Formation of Active Coactivator Complexes. Mol. Endocrinol. 2005, 19, 52–64. [Google Scholar] [CrossRef] [Green Version]
  18. Yudt, M.R.; Jewell, C.M.; Bienstock, R.; Cidlowski, J.A. Molecular Origins for the Dominant Negative Function of Human Glucocorticoid Receptor Beta. Mol. Cell. Biol. 2003, 23, 4319–4330. [Google Scholar] [CrossRef] [Green Version]
  19. Ramos-Ramírez, P.; Tliba, O. Glucocorticoid Receptor β (GRβ): Beyond Its Dominant-Negative Function. Int. J. Mol. Sci. 2021, 22, 3649. [Google Scholar] [CrossRef]
  20. Stechschulte, L.; Wuescher, L.; Marino, J.; Hill, J.; Eng, C.; Hinds, T., Jr. Glucocorticoid Receptor β Stimulates Akt1 Growth Pathway by Attenuation of PTEN. J. Biol. Chem. 2014, 289, 17885–17894. [Google Scholar] [CrossRef] [Green Version]
  21. He, B.; Cruz-Topete, D.; Oakley, R.H.; Xiao, X.; Cidlowski, J.A. Human Glucocorticoid Receptor β Regulates Gluconeogenesis and Inflammation in Mouse Liver. Mol. Cell. Biol. 2016, 36, 714–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yin, Y.; Zhang, X.; Li, Z.; Deng, L.; Jiao, G.; Zhang, B.; Xie, P.; Mu, H.; Qiao, W.; Zou, J. Glucocorticoid receptor β regulates injury-mediated astrocyte activation and contributes to glioma pathogenesis via modulation of β-catenin/TCF transcriptional activity. Neurobiol. Dis. 2013, 59, 165–176. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Q.; Lu, P.; Shi, Z.-F.; Xu, Y.-J.; Xiang, J.; Wang, Y.-X.; Deng, L.-X.; Xie, P.; Yin, Y.; Zhang, B.; et al. Glucocorticoid Receptor β Acts as a Co-activator of T-Cell Factor 4 and Enhances Glioma Cell Proliferation. Mol. Neurobiol. 2014, 52, 1106–1118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. McBeth, L.; Nwaneri, A.C.; Grabnar, M.; Demeter, J.; Nestor-Kalinoski, A.; Hinds, J.T.D. Glucocorticoid receptor beta increases migration of human bladder cancer cells. Oncotarget 2016, 7, 27313–27324. [Google Scholar] [CrossRef] [Green Version]
  25. Lu, N.Z.; Cidlowski, J.A. Translational Regulatory Mechanisms Generate N-Terminal Glucocorticoid Receptor Isoforms with Unique Transcriptional Target Genes. Mol. Cell 2005, 18, 331–342. [Google Scholar] [CrossRef] [PubMed]
  26. Lu, N.Z.; Collins, J.B.; Grissom, S.F.; Cidlowski, J. Selective Regulation of Bone Cell Apoptosis by Translational Isoforms of the Glucocorticoid Receptor. Mol. Cell. Biol. 2007, 27, 7143–7160. [Google Scholar] [CrossRef] [Green Version]
  27. Wu, I.; Shin, S.C.; Cao, Y.; Bender, I.K.; Jafari, N.; Feng, G.; Lin, S.; Cidlowski, J.; Schleimer, R.P.; Lu, N.Z. Selective glucocorticoid receptor translational isoforms reveal glucocorticoid-induced apoptotic transcriptomes. Cell Death Dis. 2013, 4, e453. [Google Scholar] [CrossRef] [PubMed]
  28. Grad, I.; Picard, D. The glucocorticoid responses are shaped by molecular chaperones. Mol. Cell. Endocrinol. 2007, 275, 2–12. [Google Scholar] [CrossRef]
  29. Pratt, W.B.; Toft, D.O. Steroid Receptor Interactions with Heat Shock Protein and Immunophilin Chaperones. Endocr. Rev. 1997, 18, 306–360. [Google Scholar] [CrossRef]
  30. Nicolaides, N.C.; Chrousos, G.P. Adrenal Cortex Hormones. In Hormonal Signaling in Biology and Medicine: Comprehensive Modern Endocrinology; Litwack, G., Ed.; Academic Press: Cambridge, MA, USA, 2020; pp. 619–633. [Google Scholar]
  31. Barnes, P.J.; Karin, M. Nuclear Factor-κB—A Pivotal Transcription Factor in Chronic Inflammatory Diseases. N. Engl. J. Med. 1997, 336, 1066–1071. [Google Scholar] [CrossRef]
  32. Karin, M.; Chang, L. AP-1—glucocorticoid receptor crosstalk taken to a higher level. J. Endocrinol. 2001, 169, 447–451. [Google Scholar] [CrossRef] [Green Version]
  33. DiDonato, J.A.; Saatcioglu, F.; Karin, M. Molecular Mechanisms of Immunosuppression and Anti-inflammatory Activities by Glucocorticoids. Am. J. Respir. Crit. Care Med. 1996, 154, S11–S15. [Google Scholar] [CrossRef]
  34. Johnstone, W.M.; Honeycutt, J.L.; Deck, C.A.; Borski, R.J. Nongenomic glucocorticoid effects and their mechanisms of action in vertebrates. Int. Rev. Cell Mol. Biol. 2019, 346, 51–96. [Google Scholar] [CrossRef] [PubMed]
  35. Panettieri, R.A.; Schaafsma, D.; Amrani, Y.; Koziol-White, C.; Ostrom, R.; Tliba, O. Non-genomic Effects of Glucocorticoids: An Updated View. Trends Pharmacol. Sci. 2019, 40, 38–49. [Google Scholar] [CrossRef] [PubMed]
  36. Nicolaides, N.C.; Kino, T.; Roberts, M.L.; Katsantoni, E.; Sertedaki, A.; Moutsatsou, P.; Psarra, A.-M.; Chrousos, G.P.; Charmandari, E. The Role of S-Palmitoylation of the Human Glucocorticoid Receptor (hGR) in Mediating the Nongenomic Glucocorticoid Actions. J. Mol. Biochem. 2017, 6, 3–12. [Google Scholar] [PubMed]
  37. Lapp, H.E.; Bartlett, A.A.; Hunter, R.G. Stress and glucocorticoid receptor regulation of mitochondrial gene expression. J. Mol. Endocrinol. 2019, 62, R121–R128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Demonacos, C.; Djordjevic-Markovic, R.; Tsawdaroglou, N.; Sekeris, C. The mitochondrion as a primary site of action of glucocorticoids: The interaction of the glucocorticoid receptor with mitochondrial DNA sequences showing partial similarity to the nuclear glucocorticoid responsive elements. J. Steroid Biochem. Mol. Biol. 1995, 55, 43–55. [Google Scholar] [CrossRef]
  39. Lee, S.-R.; Kim, H.-K.; Song, I.-S.; Youm, J.; Dizon, L.A.; Jeong, S.-H.; Ko, T.-H.; Heo, H.-J.; Ko, K.S.; Rhee, B.D.; et al. Glucocorticoids and their receptors: Insights into specific roles in mitochondria. Prog. Biophys. Mol. Biol. 2013, 112, 44–54. [Google Scholar] [CrossRef]
  40. Nicolaides, N.C.; Charmandari, E. Novel insights into the molecular mechanisms underlying generalized glucocorticoid resistance and hypersensitivity syndromes. Hormones 2017, 16, 124–138. [Google Scholar] [CrossRef] [Green Version]
  41. Nicolaides, N.; Lamprokostopoulou, A.; Sertedaki, A.; Charmandari, E. Recent advances in the molecular mechanisms causing primary generalized glucocorticoid resistance. Hormones 2016, 15. [Google Scholar] [CrossRef]
  42. Charmandari, E. Primary Generalized Glucocorticoid Resistance and Hypersensitivity. Horm. Res. Paediatr. 2011, 76, 145–155. [Google Scholar] [CrossRef]
  43. Charmandari, E. Primary Generalized Glucocorticoid Resistance and Hypersensitivity: The End-Organ Involvement in the Stress Response. Sci. Signal. 2012, 5, pt5. [Google Scholar] [CrossRef]
  44. Charmandari, E.; Kino, T.; Chrousos, G.P. Primary Generalized Familial and Sporadic Glucocorticoid Resistance (Chrousos Syndrome) and Hypersensitivity. Ghrelin Syst. 2013, 24, 67–85. [Google Scholar] [CrossRef] [Green Version]
  45. Nicolaides, N.C.; Charmandari, E.; Chrousos, G.P.; Kino, T. Recent advances in the molecular mechanisms determining tissue sensitivity to glucocorticoids: Novel mutations, circadian rhythm and ligand-induced repression of the human glucocorticoid receptor. BMC Endocr. Disord. 2014, 14, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Charmandari, E.; Kino, T. Chrousos syndrome: A seminal report, a phylogenetic enigma and the clinical implications of glucocorticoid signalling changes. Eur. J. Clin. Investig. 2010, 40, 932–942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Charmandari, E.; Kino, T.; Ichijo, T.; Chrousos, G.P. Generalized Glucocorticoid Resistance: Clinical Aspects, Molecular Mechanisms, and Implications of a Rare Genetic Disorder. J. Clin. Endocrinol. Metab. 2008, 93, 1563–1572. [Google Scholar] [CrossRef] [Green Version]
  48. Chrousos, G. Q&A: Primary generalized glucocorticoid resistance. BMC Med. 2011, 9, 27. [Google Scholar] [CrossRef] [Green Version]
  49. Nicolaides, N.C.; Charmandari, E. Chrousos syndrome: From molecular pathogenesis to therapeutic management. Eur. J. Clin. Investig. 2015, 45, 504–514. [Google Scholar] [CrossRef]
  50. Nicolaides, N.C.; Charmandari, E. Glucocorticoid Resistance. Galanin 2019, 111, 85–102. [Google Scholar] [CrossRef]
  51. Martins, C.S.; de Castro, M. Generalized and tissue specific glucocorticoid resistance. Mol. Cell. Endocrinol. 2021, 530, 111277. [Google Scholar] [CrossRef]
  52. Chrousos, G.P.; Vingerhoeds, A.; Brandon, D.; Eil, C.; Pugeat, M.; Devroede, M.; Loriaux, D.L.; Lipsett, M.B. Primary cortisol resistance in man. A glucocorticoid receptor-mediated disease. J. Clin. Investig. 1982, 69, 1261–1269. [Google Scholar] [CrossRef]
  53. Chrousos, G.P.; Detera-Wadleigh, S.D.; Karl, M. Syndromes of Glucocorticoid Resistance. Ann. Intern. Med. 1993, 119, 1113. [Google Scholar] [CrossRef] [Green Version]
  54. Nader, N.; Bachrach, B.E.; Hurt, D.E.; Gajula, S.; Pittman, A.; Lescher, R.; Kino, T. A Novel Point Mutation in Helix 10 of the Human Glucocorticoid Receptor Causes Generalized Glucocorticoid Resistance by Disrupting the Structure of the Ligand-Binding Domain. J. Clin. Endocrinol. Metab. 2010, 95, 2281–2285. [Google Scholar] [CrossRef] [Green Version]
  55. McMahon, S.K.; Pretorius, C.J.; Ungerer, J.P.J.; Salmon, N.J.; Conwell, L.S.; Pearen, M.A.; Batch, J.A. Neonatal Complete Generalized Glucocorticoid Resistance and Growth Hormone Deficiency Caused by a Novel Homozygous Mutation in Helix 12 of the Ligand Binding Domain of the Glucocorticoid Receptor Gene (NR3C1). J. Clin. Endocrinol. Metab. 2010, 95, 297–302. [Google Scholar] [CrossRef] [Green Version]
  56. Hurley, D.M.; Accili, D.; Stratakis, C.A.; Karl, M.; Vamvakopoulos, N.; Rorer, E.; Constantine, K.; Taylor, S.I.; Chrousos, G.P. Point mutation causing a single amino acid substitution in the hormone binding domain of the glucocorticoid receptor in familial glucocorticoid resistance. J. Clin. Investig. 1991, 87, 680–686. [Google Scholar] [CrossRef] [PubMed]
  57. Charmandari, E.; Kino, T.; Souvatzoglou, E.; Vottero, A.; Bhattacharyya, N.; Chrousos, G.P. Natural Glucocorticoid Receptor Mutants Causing Generalized Glucocorticoid Resistance: Molecular Genotype, Genetic Transmission, and Clinical Phenotype. J. Clin. Endocrinol. Metab. 2004, 89, 1939–1949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Karl, M.; Lamberts, S.W.; Detera-Wadleigh, S.D.; Encío, I.J.; Stratakis, C.A.; Hurley, D.M.; Accili, D.; Chrousos, G.P. Familial glucocorticoid resistance caused by a splice site deletion in the human glucocorticoid receptor gene. J. Clin. Endocrinol. Metab. 1993, 76, 683–689. [Google Scholar] [CrossRef] [PubMed]
  59. Malchoff, D.M.; Brufsky, A.; Reardon, G.; McDermott, P.; Javier, E.C.; Bergh, C.H.; Rowe, D.; Malchoff, C.D. A mutation of the glucocorticoid receptor in primary cortisol resistance. J. Clin. Investig. 1993, 91, 1918–1925. [Google Scholar] [CrossRef] [Green Version]
  60. Karl, M.; Lamberts, S.W.; Koper, J.W.; Katz, D.A.; Huizenga, N.E.; Kino, T.; Haddad, B.R.; Hughes, M.R.; Chrousos, G.P. Cushing’s disease preceded by generalized glucocorticoid resistance: Clinical consequences of a novel, dominant-negative glucocorticoid receptor mutation. Proc. Assoc. Am. Physicians 1996, 108, 8863343. [Google Scholar]
  61. Kino, T.; Stauber, R.H.; Resau, J.H.; Pavlakis, G.N.; Chrousos, G.P. Pathologic Human GR Mutant Has a Transdominant Negative Effect on the Wild-Type GR by Inhibiting Its Translocation into the Nucleus: Importance of the Ligand-Binding Domain for Intracellular GR Trafficking. J. Clin. Endocrinol. Metab. 2001, 86, 5600–5608. [Google Scholar] [CrossRef]
  62. Ruiz, M.; Lind, U.; Gåfvels, M.; Eggertsen, G.; Carlstedt-Duke, J.; Nilsson, L.; Holtmann, M.; Stierna, P.; Wikström, A.-C.; Werner, S. Characterization of two novel mutations in the glucocorticoid receptor gene in patients with primary cortisol resistance. Clin. Endocrinol. 2001, 55, 363–371. [Google Scholar] [CrossRef]
  63. Charmandari, E.; Kino, T.; Ichijo, T.; Zachman, K.; Alatsatianos, A.; Chrousos, G.P. Functional Characterization of the Natural Human Glucocorticoid Receptor (hGR) Mutants hGRαR477H and hGRαG679S Associated with Generalized Glucocorticoid Resistance. J. Clin. Endocrinol. Metab. 2006, 91, 1535–1543. [Google Scholar] [CrossRef] [Green Version]
  64. Mendonca, B.B.; Leite, M.V.; De Castro, M.; Kino, T.; Elias, L.L.K.; Bachega, T.A.S.; Arnhold, I.J.P.; Chrousos, G.P.; Latronico, A.C. Female Pseudohermaphroditism Caused by a Novel Homozygous Missense Mutation of the GR Gene. J. Clin. Endocrinol. Metab. 2002, 87, 1805–1809. [Google Scholar] [CrossRef]
  65. Vottero, A.; Kino, T.; Combe, H.; LeComte, P.; Chrousos, G.P. A Novel, C-Terminal Dominant Negative Mutation of the GR Causes Familial Glucocorticoid Resistance through Abnormal Interactions with p160 Steroid Receptor Coactivators. J. Clin. Endocrinol. Metab. 2002, 87, 2658–2667. [Google Scholar] [CrossRef] [PubMed]
  66. Charmandari, E.; Raji, A.; Kino, T.; Ichijo, T.; Tiulpakov, A.; Zachman, K.; Chrousos, G.P. A Novel Point Mutation in the Ligand-Binding Domain (LBD) of the Human Glucocorticoid Receptor (hGR) Causing Generalized Glucocorticoid Resistance: The Importance of the C Terminus of hGR LBD in Conferring Transactivational Activity. J. Clin. Endocrinol. Metab. 2005, 90, 3696–3705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Charmandari, E.; Kino, T.; Ichijo, T.; Jubiz, W.; Mejia, L.; Zachman, K.; Chrousos, G.P. A Novel Point Mutation in Helix 11 of the Ligand-Binding Domain of the Human Glucocorticoid Receptor Gene Causing Generalized Glucocorticoid Resistance. J. Clin. Endocrinol. Metab. 2007, 92, 3986–3990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Tatsi, C.; Xekouki, P.; Nioti, O.; Bachrach, B.; Belyavskaya, E.; Lyssikatos, C.; Stratakis, C.A. A novel mutation in the glucocorticoid receptor gene as a cause of severe glucocorticoid resistance complicated by hypertensive encephalopathy. J. Hypertens. 2019, 37, 1475–1481. [Google Scholar] [CrossRef] [PubMed]
  69. Bouligand, J.; Delemer, B.; Hecart, A.-C.; Meduri, G.; Viengchareun, S.; Amazit, L.; Trabado, S.; Fève, B.; Guiochon-Mantel, A.; Young, J.; et al. Familial Glucocorticoid Receptor Haploinsufficiency by Non-Sense Mediated mRNA Decay, Adrenal Hyperplasia and Apparent Mineralocorticoid Excess. PLoS ONE 2010, 5, e13563. [Google Scholar] [CrossRef]
  70. Trebble, P.; Matthews, L.; Blaikley, J.; Wayte, A.W.O.; Black, G.C.M.; Wilton, A.; Ray, D.W. Familial glucocorticoid resistance caused by a novel frameshift glucocorticoid receptor mutation. J. Clin. Endocrinol. Metab. 2010, 95, E490–E499. [Google Scholar] [CrossRef] [Green Version]
  71. Zhu, H.-J.; Dai, Y.-F.; Wang, O.; Li, M.; Lu, L.; Zhao, W.-G.; Xing, X.-P.; Pan, H.; Li, N.-S.; Gong, F.-Y. Generalized glucocorticoid resistance accompanied with an adrenocortical adenoma and caused by a novel point mutation of human glucocorticoid receptor gene. Chin. Med. J. 2011, 124, 551–555. [Google Scholar] [PubMed]
  72. Nicolaides, N.C.; Skyrla, E.; Vlachakis, D.; Psarra, A.-M.G.; Moutsatsou, P.; Sertedaki, A.; Kossida, S.; Charmandari, E. Functional Characterization of the hGR? T556I Causing Chrousos Syndrome. Eur. J. Clin. Investig. 2015, 46, 42–49. [Google Scholar] [CrossRef]
  73. Donner, K.M.; Hiltunen, T.P.; Jänne, O.A.; Sane, T.; Kontula, K. Generalized glucocorticoid resistance caused by a novel two-nucleotide deletion in the hormone-binding domain of the glucocorticoid receptor gene NR3C1. Eur. J. Endocrinol. 2013, 168, K9–K18. [Google Scholar] [CrossRef] [Green Version]
  74. Roberts, M.L.; Kino, T.; Nicolaides, N.C.; Hurt, D.E.; Katsantoni, E.; Sertedaki, A.; Komianou, F.; Kassiou, K.; Chrousos, G.P.; Charmandari, E. A novel point mutation in the DNA-binding domain (DBD) of the human glucocorticoid receptor causes primary generalized glucocorticoid resistance by disrupting the hydrophobic structure of its DBD. J. Clin. Endocrinol. Metab. 2013, 98, E790–E795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Nicolaides, N.C.; Roberts, M.; Kino, T.; Braatvedt, G.; Hurt, D.E.; Katsantoni, E.; Sertedaki, A.; Chrousos, G.P.; Charmandari, E. A Novel Point Mutation of the Human Glucocorticoid Receptor Gene Causes Primary Generalized Glucocorticoid Resistance Through Impaired Interaction with the LXXLL Motif of the p160 Coactivators: Dissociation of the Transactivating and Transreppressive Activities. J. Clin. Endocrinol. Metab. 2014, 99, E902–E907. [Google Scholar] [CrossRef] [Green Version]
  76. Nicolaides, N.C.; Geer, E.B.; Vlachakis, D.; Roberts, M.L.; Psarra, A.-M.; Moutsatsou, P.; Sertedaki, A.; Kossida, S.; Charmandari, E. A novel mutation of thehGRgene causing Chrousos syndrome. Eur. J. Clin. Investig. 2015, 45, 782–791. [Google Scholar] [CrossRef] [PubMed]
  77. Velayos, T.; Grau, G.; Rica, I.; de Nanclares, G.P.; Gaztambide, S. Glucocorticoid resistance syndrome caused by two novel mutations in the NR3C1 gene. Endocrinol. Nutr. 2016, 63, 369–371. [Google Scholar] [CrossRef] [PubMed]
  78. Vitellius, G.; Fagart, J.; Delemer, B.; Amazit, L.; Ramos, N.; Bouligand, J.; Billan, F.L.; Castinetti, F.; Guiochon-Mantel, A.; Trabado, S.; et al. Three Novel Heterozygous Point Mutations ofNR3C1Causing Glucocorticoid Resistance. Hum. Mutat. 2016, 37, 794–803. [Google Scholar] [CrossRef]
  79. Vitellius, G.; Trabado, S.; Hoeffel-Fornes, C.; Bouligand, J.; Bennet, A.; Castinetti, F.; Decoudier, B.; Guiochon-Mantel, A.; Lombès, M.; Delemer, B.; et al. Significant prevalence of NR3C1 mutations in incidentally discovered bilateral adrenal hyperplasia: Results of the French MUTA-GR Study. Eur. J. Endocrinol. 2018, 178, 411–423. [Google Scholar] [CrossRef] [Green Version]
  80. Al Argan, R.; Saskin, A.; Yang, J.W.; D’Agostino, M.D.; Rivera, J. Glucocorticoid resistance syndrome caused by a novel NR3C1 point mutation. Endocr. J. 2018, 65, 1139–1146. [Google Scholar] [CrossRef] [Green Version]
  81. Cannavò, S.; Benvenga, S.; Messina, E.; Moleti, M.; Ferraù, F. Comment to ‘Glucocorticoid resistance syndrome caused by a novel NR3C1 point mutation’ by Al Argan et al. Endocr. J. 2019, 66, 657. [Google Scholar] [CrossRef] [Green Version]
  82. Molnár, Á.; Patócs, A.; Likó, I.; Nyírő, G.; Rácz, K.; Tóth, M.; Sármán, B. An unexpected, mild phenotype of glucocorticoid resistance associated with glucocorticoid receptor gene mutation case report and review of the literature. BMC Med. Genet. 2018, 19, 37. [Google Scholar] [CrossRef] [Green Version]
  83. Vitellius, G.; Delemer, B.; Caron, P.; Chabre, O.; Bouligand, J.; Pussard, E.; Trabado, S.; Lombes, M. Impaired 11β-Hydroxysteroid Dehydrogenase Type 2 in Glucocorticoid-Resistant Patients. J. Clin. Endocrinol. Metab. 2019, 104, 5205–5216. [Google Scholar] [CrossRef]
  84. Lin, L.; Wu, X.; Hou, Y.; Zheng, F.; Xu, R. A Novel Mutation in the Glucocorticoid Receptor Gene Causing Resistant Hypertension: A Case Report. Am. J. Hypertens. 2019, 32, 1126–1128. [Google Scholar] [CrossRef] [PubMed]
  85. Ma, L.; Tan, X.; Li, J.; Long, Y.; Xiao, Z.; De, J.; Ren, Y.; Tian, H.; Chen, T. A Novel Glucocorticoid Receptor Mutation in Primary Generalized Glucocorticoid Resistance Disease. Endocr. Pr. 2020, 26, 651–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Paragliola, R.M.; Costella, A.; Corsello, A.; Urbani, A.; Concolino, P. A Novel Pathogenic Variant in the N-Terminal Domain of the Glucocorticoid Receptor, Causing Glucocorticoid Resistance. Mol. Diagn. Ther. 2020, 24, 473–485. [Google Scholar] [CrossRef] [PubMed]
  87. Nicolaides, N.C.; Chrousos, G.P. Bilateral adrenal incidentalomas and NR3C1 mutations causing glucocorticoid resistance: Is there an association? Eur. J. Endocrinol. 2018, 179, C1–C4. [Google Scholar] [CrossRef] [PubMed]
  88. Monteiro, L.; Franco, O.; Alencar, S.; Porto, W. Deciphering the structural basis for glucocorticoid resistance caused by missense mutations in the ligand binding domain of glucocorticoid receptor. J. Mol. Graph. Model. 2019, 92, 216–226. [Google Scholar] [CrossRef]
  89. Kaziales, A.; Rührnößl, F.; Richter, K. Glucocorticoid resistance conferring mutation in the C-terminus of GR alters the receptor conformational dynamics. Sci. Rep. 2021, 11, 1–14. [Google Scholar] [CrossRef]
  90. Charmandari, E.; Ichijo, T.; Jubiz, W.; Baid, S.; Zachman, K.; Chrousos, G.P.; Kino, T. A Novel Point Mutation in the Amino Terminal Domain of the Human Glucocorticoid Receptor (hGR) Gene Enhancing hGR-Mediated Gene Expression. J. Clin. Endocrinol. Metab. 2008, 93, 4963–4968. [Google Scholar] [CrossRef] [Green Version]
  91. Iida, S.; Nakamura, Y.; Fujii, H.; Nishimura, J.-I.; Tsugawa, M.; Gomi, M.; Fukata, J.; Tarui, S.; Moriwaki, K.; Kitani, T. A Patient with Hypocortisolism and Cushing’s Syndrome-Like Manifestations: Cortisol Hyperreactive Syndrome. J. Clin. Endocrinol. Metab. 1990, 70, 729–737. [Google Scholar] [CrossRef] [PubMed]
  92. Fujii, H.; Iida, S.; Gomi, M.; Tsugawa, M.; Kitani, T.; Moriwaki, K. Augmented induction by dexamethasone of metallothionein IIa messenger ribonucleic acid in fibroblasts from a patient with cortisol hyperreactive syndrome. J. Clin. Endocrinol. Metab. 1993, 76, 445–449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Newfield, R.S.; Kalaitzoglou, G.; Licholai, T.; Chilton, D.; Ashraf, J.; Thompson, E.B.; New, M.I. Normocortisolemic Cushing’s Syndrome Initially Presenting with Increased Glucocorticoid Receptor Numbers1. J. Clin. Endocrinol. Metab. 2000, 85, 14–21. [Google Scholar] [CrossRef] [Green Version]
  94. Russcher, H.; Smit, P.; van Rossum, E.F.C.; van den Akker, E.L.T.; Brinkmann, A.O.d.; Heide, L.J.M.; de Jong, F.H.; Koper, J.W.; Lamberts, S.W.J. Strategies for the Characterization of Disorders in Cortisol Sensitivity. J. Clin. Endocrinol. Metab. 2006, 91, 694–701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Krysiak, R.; Okopień, B. Glucocorticoid hypersensitivity syndrome—A case report. West Indian Med. J. 2012, 61, 844–846. [Google Scholar] [PubMed]
  96. Nicolaides, N.C.; Lamprokostopoulou, A.; Polyzos, A.; Kino, T.; Katsantoni, E.; Triantafyllou, P.; Christophoridis, A.; Katzos, G.; Dracopoulou, M.; Sertedaki, A.; et al. Transient generalized glucocorticoid hypersensitivity. Eur. J. Clin. Investig. 2015, 45, 1306–1315. [Google Scholar] [CrossRef] [PubMed]
  97. Grigsby, M.J.; Green, T.L.; Lim, D.; Cho, K.; Greenhalgh, D.G. A Novel Human Glucocorticoid Receptor Variant, G459V, is Hyperactive in Response to Steroids. Shock 2021, 56, 318–324. [Google Scholar] [CrossRef]
  98. Liu, Y.; Han, M.; Yang, J.; Xu, Q.; Xu, L.; Ren, Y.; Xiang, C.; Liu, Z.; Zhang, Y. Primary Generalized Glucocorticoid Hypersensitivity Treated with Mifepristone: A Case Report. Int. J. Gen. Med. 2020, 13, 825–831. [Google Scholar] [CrossRef]
  99. Nicolaides, N.C.; Polyzos, A.; Koniari, E.; Lamprokostopoulou, A.; Papageorgiou, I.; Golfinopoulou, E.; Papathanasiou, C.; Sertedaki, A.; Thanos, D.; Chrousos, G.P.; et al. Transcriptomics in tissue glucocorticoid sensitivity. Eur. J. Clin. Investig. 2019, 49, e13129. [Google Scholar] [CrossRef]
  100. Nicolaides, N.; Ioannidi, M.-K.; Koniari, E.; Papageorgiou, I.; Bartzeliotou, A.; Sertedaki, A.; Klapa, M.; Charmandari, E. Untargeted Plasma Metabolomics Unravels a Metabolic Signature for Tissue Sensitivity to Glucocorticoids in Healthy Subjects: Its Implications in Dietary Planning for a Healthy Lifestyle. Nutrients 2021, 13, 2120. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Signaling pathways of genomic, nongenomic and mitochondrial glucocorticoid actions. cPLA2α: cytosolic phospholipase A2 alpha; DRIP/TRAP: vitamin D receptor-interacting protein/thyroid hormone receptor-associated protein complex; eNOS: endothelial nitric oxide synthetase; FKBP: immunophilins; GR: glucocorticoid receptor; HSP: heat shock proteins; MAPK: mitogen-activated protein kinases; NO: nitric oxide; p160: nuclear receptor coactivators p160; PI3K: phosphatidylinositol 3-kinase; SWI/SNF: switching/sucrose non-fermenting complex (modified from ref. [6]).
Figure 1. Signaling pathways of genomic, nongenomic and mitochondrial glucocorticoid actions. cPLA2α: cytosolic phospholipase A2 alpha; DRIP/TRAP: vitamin D receptor-interacting protein/thyroid hormone receptor-associated protein complex; eNOS: endothelial nitric oxide synthetase; FKBP: immunophilins; GR: glucocorticoid receptor; HSP: heat shock proteins; MAPK: mitogen-activated protein kinases; NO: nitric oxide; p160: nuclear receptor coactivators p160; PI3K: phosphatidylinositol 3-kinase; SWI/SNF: switching/sucrose non-fermenting complex (modified from ref. [6]).
Ijms 22 10839 g001
Figure 2. Pathophysiologic mechanisms of primary generalized glucocorticoid resistance and hypersensitivity syndromes. ACTH: adrenocorticotropic hormone; AVP: arginine-vasopressin; CRH: corticotropin-releasing hormone; FKBP: immunophilins; GR: glucocorticoid receptor; HSP: heat shock proteins.
Figure 2. Pathophysiologic mechanisms of primary generalized glucocorticoid resistance and hypersensitivity syndromes. ACTH: adrenocorticotropic hormone; AVP: arginine-vasopressin; CRH: corticotropin-releasing hormone; FKBP: immunophilins; GR: glucocorticoid receptor; HSP: heat shock proteins.
Ijms 22 10839 g002
Figure 3. Schematic representation of the published genetic defects causing Chrousos syndrome in the NR3C1 gene (A) and in the linearized hGR protein (B) (modified with permission from ref. [13], 2020, Nicolaides, N.C.; Chrousos, G.; Kino, T.).
Figure 3. Schematic representation of the published genetic defects causing Chrousos syndrome in the NR3C1 gene (A) and in the linearized hGR protein (B) (modified with permission from ref. [13], 2020, Nicolaides, N.C.; Chrousos, G.; Kino, T.).
Ijms 22 10839 g003
Table 1. Mutations of the NR3C1 Gene Causing Chrousos Syndrome: Molecular Mechanisms and Clinical Manifestations.
Table 1. Mutations of the NR3C1 Gene Causing Chrousos Syndrome: Molecular Mechanisms and Clinical Manifestations.
ReferencecDNAAmino AcidMolecular MechanismsGenotypePhenotype
Chrousos et al. 1982 [52]
Hurley et al. 1991 [56]
Charmandari et al. 2004 [57]
1922 (A→T)641 (D→V)Transactivation ↓
Affinity for ligand ↓ (x3)
Nuclear translocation: 22 min
Abnormal interaction with GRIP1
HomozygousHypertension
Hypokalemic alkalosis
Karl et al. 1993 [58]4 bp deletion
in exon-intron 6
hGRα number: 50% of control
Inactivation of the affected allele
HeterozygousHirsutism
Male-pattern hair-loss
Menstrual irregularities
Malchoff et al. 1993 [59]
Charmandari et al. 2004 [57]
2185 (G→A)729 (V→I)Transactivation ↓
Affinity for ligand ↓ (x2)
Nuclear translocation: 120 min
Abnormal interaction with GRIP1
HomozygousPrecocious puberty
Hyperandrogenism
Karl et al. 1996 [60]
Kino et al. 2001 [61]
Charmandari et al. 2004 [57]
1676 (T→A)559 (I→N)Transactivation ↓
Decrease in hGR binding sites
Transdominance (+)
Nuclear translocation: 180 min
Abnormal interaction with GRIP1
HeterozygousHypertension
Oligospermia
Infertility
Ruiz et al. 2001 [62]
Charmandari et al. 2006 [63]
1430 (G→A)477 (R→H)Transactivation ↓
No DNA binding
Nuclear translocation: 20 min
HeterozygousHirsutism
Fatigue
Hypertension
Ruiz et al. 2001 [62]
Charmandari et al. 2006 [63]
2035 (G→A)679 (G→S)Transactivation ↓
Affinity for ligand ↓ (x2)
Nuclear translocation: 30 min
Abnormal interaction with GRIP1
HeterozygousHirsutism
Fatigue
Hypertension
Mendonca et al. 2002 [64]
Charmandari et al. 2004 [57]
1712 (T→C)571 (V→A)Transactivation ↓
Affinity for ligand ↓ (x6)
Nuclear translocation: 25 min
Abnormal interaction with GRIP1
HomozygousAmbiguous genitalia
Hypertension
Hypokalemia
Hyperandrogenism
Vottero et al. 2002 [65]
Charmandari et al. 2004 [57]
2241 (T→G)747 (I→M)Transactivation ↓
Transdominance (+)
Affinity for ligand ↓ (x2)
Nuclear translocation ↓
Abnormal interaction with GRIP1
HeterozygousCystic acne
Hirsutism
Oligo-amenorrhea
Charmandari et al. 2005 [66] 2318 (T→C)773 (L→P)Transactivation ↓
Transdominance (+)
Affinity for ligand ↓ (x2.6)
Nuclear translocation: 30 min
Abnormal interaction with GRIP1
HeterozygousFatigue
Anxiety
Acne
Hirsutism
Hypertension
Charmandari et al. 2007 [67] 2209 (T→C)737 (F→L)Transactivation ↓
Transdominance (+)
Affinity for ligand ↓ (x 1.5)
Nuclear translocation: 180 min
HeterozygousHypertension
Hypokalemia
Nader et al. 2010 [54]

Tatsi et al. 2019 [68]
2141 (G→A)714 (R→Q)Transactivation ↓
Transdominance (+)
Affinity for ligand ↓ (x2)
Nuclear translocation ↓
Abnormal interaction with GRIP1
HeterozygousHypoglycemia
Hypokalemia
Hypertension
Mild clitoromegaly
Advanced bone age
Precocious pubarche
McMahon et al. 2010 [55]2 bp deletion
at nt 2318-9
773Transactivation ↓
Affinity for ligand: absent
No suppression of IL-6
HomozygousHypoglycemia
Fatigability with feeding
Hypertension
Bouligand et al. 2010 [69]1405 (C→T)469 (R→X)Transactivation ↓
Ligand-binding sites ↓
No DNA binding
No nuclear translocation
HeterozygousAdrenal hyperplasia
Hypertension
Hypokalemia
Trebble et al. 2010 [70]1835delCSer612Tyrfs*Protein expression: (ࢤ)
Ligand binding: (ࢤ)
Cytoplasm to nuclear translocation: (ࢤ)
Dominant negative effect: Yes
Transactivation: (ࢤ)
HeterozygousFatigue
Zhu Hui-juan et al. 2011 [71]
Nicolaides et al. 2016 [72]
1667 (G→T)556 (T→I)Transactivation ↓
Transrepression ↑
Affinity for ligand ↓
Nuclear translocation ↓
Abnormal interaction with GRIP1
HeterozygousAdrenal incidentaloma
Donner et al. 2012 [73]2317_2318delCTLeu773Valfs*Protein expression: slightly reduced
Transactivation: Decreased
Dominant negative effect: No
Ligand binding: (ࢤ)
HeterozygousHypertension
Roberts et al. 2013 [74]1268 (T→C) 423 (V→A)Transactivation ↓
Affinity for ligand: Normal
No DNA binding
Nuclear translocation: 35 min
Interaction with GRIP1: Normal
HeterozygousFatigue
Anxiety
Hypertension
Nicolaides et al. 2014 [75]1724 (T→G)575 (V→G)Transactivation ↓
Transrepression ↑
Affinity for ligand ↓ (x2)
Nuclear translocation ↓
Abnormal interaction with GRIP1
HeterozygousMelanoma
Asymptomatic daughters
Nicolaides et al. 2015 [76]2177 (A→G)726 (H→R)Transactivation ↓
Transrepression ↓
Affinity for ligand ↓ (x2)
Nuclear translocation ↓
Abnormal interaction with GRIP1
HeterozygousHirsutism
Acne
Alopecia
Anxiety
Fatigue
Irregular menstrual cycles
Velayos et al. 2016 [77]1429 (C→T)477 (R→C)Not studied yetHeterozygousMild hirsutism
Asymptomatic mother
Velayos et al. 2016 [77] 1762_1763insTTAC588 (H→L*5)Not studied yetHeterozygousHirsutism
Anxiety
Chronic fatigue
Vitellius et al. 2016 [78]1429 (C→A)477 (R→S)No Transactivation
Affinity for ligand: Normal
No DNA binding
Nuclear translocation ↓
HeterozygousAdrenal incidentaloma
Vitellius et al. 2016 [78]1433 (A→G)478 (Y→C)Transactivation ↓
Affinity for ligand: Normal
DNA binding ↓
Nuclear translocation ↓
HeterozygousAdrenal incidentaloma
Vitellius et al. 2016 [78]2015 (T→C)672 (L→P)No Transactivation
No Affinity for ligand
No DNA binding
No Nuclear translocation
HeterozygousAdrenal incidentaloma
Vitellius et al. 2018 [79]1471C>TArg491TerTransactivation: (ࢤ)HeterozygousBilateral adrenal hyperplasia
Vitellius et al. 2018 [79]1503G>TGln501HisTransactivation: DecreasedHeterozygousBilateral adrenal hyperplasia
Al Argan et al. 2018 [80]1392del465 (I→S)Not studied yetHeterozygousChronic fatigue
Acne
Hirsutism
Hypomenorrhea
Cannavò et al. 2019 [81]1915C>GLeu595ValNot performedNot availableHirsutism
Amenorrhea
Hypertension
Molnár et al. 2018 [82]2141G→A714 (R→Q)Studied in Nader et al. 2010HeterozygousInfertility
Asymptomatic sister
Vitellius et al. 2019 [83]1980T>GTyr660TerTransactivation: (ࢤ)HeterozygousHypertension
Tatsi et al. 2019 [68]592G>TGlu198TerNot performedCompound heterozygousHypertensive encephalopathy
Lin et al. 2019 [84]26C>GPro9ArgNot performedHeterozygousHypertension
Ma et al. 2020 [85]1652C>ASer551TyrLigand binding: Decreased
Cytoplasm to nuclear translocation: Decreased
Transactivation: Decreased
HomozygousFatigue
Hypokalemia
Hypertension
Polyuria
Paragliola et al. 2020 [86]367G>TGlu123TerNot performedHeterozygousChronic fatigue
Anxiety
Hirsutism
Irregular menstrual cycles
Infertility
↓: decreased; +: present; (x2): two-fold increased; ↑: increased; (ࢤ): absent.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nicolaides, N.C.; Charmandari, E. Primary Generalized Glucocorticoid Resistance and Hypersensitivity Syndromes: A 2021 Update. Int. J. Mol. Sci. 2021, 22, 10839. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms221910839

AMA Style

Nicolaides NC, Charmandari E. Primary Generalized Glucocorticoid Resistance and Hypersensitivity Syndromes: A 2021 Update. International Journal of Molecular Sciences. 2021; 22(19):10839. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms221910839

Chicago/Turabian Style

Nicolaides, Nicolas C., and Evangelia Charmandari. 2021. "Primary Generalized Glucocorticoid Resistance and Hypersensitivity Syndromes: A 2021 Update" International Journal of Molecular Sciences 22, no. 19: 10839. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms221910839

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop