Open Access

Role of microRNAs and long non‑coding RNAs in glucocorticoid signaling (Review)

  • Authors:
    • Katerina Pierouli
    • Louis Papageorgiou
    • Thanasis Mitsis
    • Eleni Papakonstantinou
    • Io Diakou
    • Stefanos Leptidis
    • Markezina Sigala
    • Konstantina Dragoumani
    • Demetrios A. Spandidos
    • Flora Bacopoulou
    • George P. Chrousos
    • George Ν. Goulielmos
    • Elias Eliopoulos
    • Dimitrios Vlachakis
  • View Affiliations

  • Published online on: November 11, 2022     https://doi.org/10.3892/ijmm.2022.5203
  • Article Number: 147
  • Copyright: © Pierouli et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The synthesis and release of glucocorticoids in living organisms are related to their response to unfavorable stressful conditions in order to maintain homeostatic functions and survive. One such hormone in humans is cortisol, which is produced by the hypothalamic‑pituitary‑adrenal cortex axis and binds with the glucocorticoid receptor (GR) following its secretion. GR controls a number of distinct gene networks. Non‑coding RNAs (ncRNAs), such as microRNAs (miRNAs) and long non‑coding RNAs (lncRNAs), regulate the expression and function of GR, having a considerable impact on various biological processes and treatment approaches for numerous disorders. In the present review, the GR pathways and signaling as part of the stress response system are discussed. A detailed report on the role of miRNAs and lncRNAs in glucocorticoid signaling is also presented.

1. Introduction

All living organisms must cope with a number of adversities during their lifetime and have to maintain a complex dynamic equilibrium called homeostasis (1). The state of threatened (or perceived as such) homeostasis is known as stress and the intrinsic or extrinsic forces that lead to such a state are called stressors (2). Severe or prolonged stress has been associated with several pathological conditions, such as the deregulation of the immune system, cardiovascular disease, neuropsychiatric disorders, metabolic disorders, endocrine disorders, and growth, as well as development impairments (3,4). Thus, in response to stressors, an organism activates multiple complex and dynamic processes in an effort to restore homeostasis, while the involved processes form the stress response system (5). The stress response system leads to several behavioral and physiological changes, such as increased awareness and enhanced analgesia, as well as an increased respiratory rate and the inhibition of general vegetative functions (6). The stress response system is comprised mainly of two components, the hypothalamic-pituitary-adrenal cortex (HPA) axis and the locus coeruleus/norepinephrine autonomic nervous system (7).

The HPA axis is the neuroendocrine link between stress and an organism's physiological response to such a state (8). A stressor induces a chain of events in the brain that activates the HPA axis. Specifically, neurons whose bodies are located in the paraventricular nucleus of the hypothalamus secrete corticotropin-releasing hormone (CRH) and arginine-vasopressin (AVP) into the hypophysial portal system. Then, they target the anterior lobe of the pituitary gland. There, CRH and AVP stimulate proopiomelanocortin cells, which in turn release adrenocorticotropic hormone (ACTH), also known as corticotropin. Finally, ACTH is released into the bloodstream and acts on the cortex of the adrenal glands, thus triggering glucocorticoid (GC) production (cortisol in humans and corticosterone in rats) (9,10). GCs subsequently self-regulate the HPA axis through a negative feedback loop (11). Although elevations in cortisol levels are the physiological response of the body to fear or threat and are beneficial to promoting survival, chronic exposure to stress results in long-term cortisol exposure. This prolonged exposition can lead to a broad range of issues, including the emergence of metabolic syndrome, obesity, cancer, mental health disorders, cardiovascular disease and increased susceptibility to infections (5).

GC signaling is related to the function of non-coding RNAs (ncRNAs) which have been extensively studied for their post-transcriptional involvement in the regulation of gene expression. It has been shown that microRNAs (miRNAs/miRs) can function as direct regulators of GC receptors (GRs) via the hybridization of various miRNAs with the 3′untranslated region (3′UTR) of the GR transcript, interrupting protein synthesis (12). Furthermore, miRNAs are directly or indirectly influenced by GCs, while it has been shown that the expression level of miRNAs is dependent on endogenous GC levels (13). In addition, long ncRNAs (lncRNAs), such as growth arrest-specific 5 (GAS5) (14) and EDN1-AS (15), regulate the expression levels of GRs, interfering with the GC signaling pathways. It is therefore paramount to increase our understanding of how nc-RNAs, mainly via the action of miRNAs and lncRNAs, regulate GR and are regulated by GCs, building an intricate regulatory network to fine-tune the body's response to various stimuli.

2. Glucocorticoid signaling

Endogenous GCs, as the final products of the HPA axis, are the main regulators of the stress response system (16). GCs are steroid hormones and as main regulators of the stress response system, are involved in numerous biological processes, including immune response, metabolism, as well as developmental, cognitive and behavioral functions (17). Endogenous GC secretion displays both ultradian and circadian rhythms. The circadian rhythm refers to physical, mental and behavioral changes that follow a 24-h cycle. Specifically, the circadian peak in GC release occurs in the early morning in diurnal animals and in the early night in nocturnal animals. On the other hand, the ultradian pulses, which are cycles having a duration shorter than a day, but longer than an hour, display a frequency of approximately one or two per hour (18,19), with nocturnal animals exhibiting an increase in amplitude towards the end of the day (20). GCs can bind either the GR or the mineralocorticoid receptor (MR) (21). The MR has a 10-fold higher affinity for naturally occurring GCs than the GR, leading to large receptor occupancy, whereas GR is only activated during circadian peak and stress response (22). Therefore, the actions of GCs are mainly exerted through the GR. GRs are expressed in almost every cell of the body and there are several GR isoforms generated from one gene, due to alternative splicing and various post-translational modifications (14). In particular, GCs function as ligands that bind GR transcripts and enable their action, with the full GRα transcript being the predominant isoform (23,24). It should also be mentioned that due to this wide range of functions GCs display, primarily via their capacity to influence immune response, synthetic GCs, such as dexamethasone and fludrocortisone have been developed and are widely prescribed for the treatment of several diseases, including rheumatic, pulmonary, gastroenterological, and cutaneous diseases (25).

Both GR and MR are members of a structurally-related protein superfamily known as nuclear receptors (NRs) (26). NRs function as transcription factors, regulating gene expression (27). NRs exhibit a characteristic structure, and so does GR. In particular, the GR is encoded by the NR subfamily 3 group C member 1 (NR3C1) gene and its functional and structural domains include an N-terminal domain (NTD), a DNA-binding domain (DBD), a hinge region (HR) and a C-terminal ligand-binding domain (LBD) (17). More specifically, the NTD contains an activation function (AF)-1 region that interacts with coregulators (28). Additionally, the highly conserved DNA-binding domain contains two zinc finger motifs and binds specific DNA sequences called GC response elements (GREs) (29). The highly flexible HR contains an amino terminus that is an essential part of the DBD and is involved in receptor dimerization, while the flexibility provided by HR to the GR allows a receptor dimmer to interact with multiple GREs (30). The somewhat conserved C-terminal LBD binds ligands, which in turn lead to conformational changes within the LBD that modulate a second AF region (AF-2) and enable interaction with specific coregulators (28).

In the absence of GCs, GR is located in the cytoplasm, where it is bound to a number of chaperone proteins that render it inactive (29). GR is first bound by heat shock protein (Hsp) 40 kDa (Hsp40), heat shock cognate 71 kDa protein and the Hsp70-Hsp90 organizing protein (Hop), while at later stages, it is bound by Hsp90, FK506-binding proteins (FKBPs) and prostaglandin E synthase 3 (PTGES3/p23) (31). Specifically, following receptor translation, Hsp70 binds the receptor in the cytosol, which leads to the unfolding of the GR and LBD inactivation. This process is accelerated by Hsp40 binding. The Hsp40/Hsp70-GR complex is then recruited to interact with Hsp90 via Hop. Hop, Hsp40 and Hsp70 are then dislodged from the Hsp90-GR complex upon Hsp90 binding, ATP and the subsequent association of FKBPs cochaperones, such as FKBP51 and FKBP52. Hsp90 interaction with GR folds the receptor and leads to a functional LBD. The heterocomplex is stabilized via p23 binding through the N-terminus and middle domains, promoting a conformation with a high affinity for corticosteroids (17,32). The binding of a ligand to the receptor's LBD leads to conformational alterations that change the proteins which comprise the heterocomplex, while also promoting GR dimerization and the subsequent translocation to the nucleus, where it can act as a transcriptional regulator (33,34). GR import is a rapid and active process that relies on GR association with Hsp90, FKBP52 and importin-a. The GR complex is transported into the nucleus by dynein along the cytoskeleton and through the nuclear pore complex (34). Once in the nucleus, the activated GR can either activate or repress gene transcription. As regards the GR signaling pathways, there are three primary ways the receptors interact with molecules and intervene in gene expression. For instance, GR can either bind directly to DNA in specific sequences, or can tether itself to other DNA-bound transcription factors. The third is via direct binding to DNA and interaction with neighboring DNA-bound transcription factors. In particular, transactivation can be achieved directly through GR homodimer binding to a GRE found in gene promoter regions, or indirectly, where GR functions as a monomer and co-operates with other transcription factors to induce transcription (35,36). Transrepression can also be either direct, via GR homodimer or, preferably, monomer binding to a negative GRE, or indirect via GR monomer binding to a pro-inflammatory transcription factor, such as NF-κB (35-37). GR remains bound to DNA for a specific time period which could be influenced by the bound ligand. This influence may be due to differences in ligand-induced conformational changes (38). Following ligand disengagement, GR dissociates from DNA and is either degraded by the proteasome or exported from the nucleus, which is an inactive process, most likely occurring through passive diffusion (34). This system enables the cell to rapidly respond to environmental changes and exercise its effects via the intricate networks established around GR activity.

3. GR pathways

The anti-inflammatory abilities of GCs highlight their critical role in the regulation of the immune system. Specifically, GR inhibits major regulators of pro-inflammatory pathways, such as transcription factors activator protein-1 (AP-1) and the aforementioned NF-κB (39). AP-1 is a protein complex composed of a Jun protein family member dimerized with another Jun protein or with a Fos protein, and enhances the expression of a number of cytokines, such as IL-1 and IL-2. In particular, GR binds the ubiquitously expressed c-Fos/c-Jun dimers via a sequence-specific to c-Fos and inhibits the DNA binding and transactivation abilities of this AP-1 heterocomplex (40). In mammals, the NF-κB protein family of transcription factors includes several proteins, such as p65 and p105/p50, which interact and form distinct homodimers or heterodimers with transcriptional regulation abilities (41). NF-κB binds to specific response elements and regulates the expression of genes encoding proteins, such as pro-inflammatory cytokines, chemokines, receptors and adhesion molecules (42). GR can inhibit NF-κB directly by interacting with the p65 and p50 subunits, or indirectly by inducing the expression of the TSC22D3 gene, which encodes GC-induced leucine zipper, a protein that binds NF-κB and suppresses its function (39). Under pathological conditions, stress-induced GCs may suppress cell-mediated immunity and may thus lead to viral infection susceptibility and tumor development (43). To illustrate the multifaceted interaction of NR3C1 with other proteins, a gene association network of genes related to the NR3C1 gene was developed using the Genemania algorithm (44) (Fig. 1). A total of 21 genes were detected in this network, including FKBPs, which interact with, or are influenced by GRα.

GCs also play a pivotal role in development. In utero GC levels influence embryonic development and affect adult physiology and pathophysiology, with a prime example being increased susceptibility to cardiovascular disease (45). In particular, GR in cardiomyocytes may induce the expression of genes, such as peroxisome proliferator-activated receptor γ coactivator 1α, which regulates cardiac mitochondrial capacity and genes encoding other regulators of cardiac metabolism, such as peroxisome proliferator-activated receptor α, Krüppel-like factor 15 and lipin 1 (lpn1). These regulators of cardiac metabolism play critical roles in the transition from the fetal to the neonatal heart (46). Additionally, GR expression in mesenchymal cells has been proven to be necessary for lung maturation. Mesenchymal GR regulates the development, morphological differentiation and remodeling of the lungs in neonates. Mesenchymal and GR-null mice have been shown to exhibit lung cell hyperplasia, suggesting a crucial function of GC signaling during lung development, via a decrease in proliferation and the thinning of interstitial cells for effective gas exchange (47). Furthermore, GC excess may lead to numerous developmental issues, culminating in congenital disabilities. Specifically, GR can induce the expression of Dickkopf-related protein 1, one of the main inhibitors of the Wnt/β-Catenin pathway which among others, participates in stem cell renewal, cell differentiation and cell proliferation during development (48).

Cardiovascular diseases are directly connected to stress (49). Exposure to a high concentration of GCs, either endogenous or exogenous, is associated with an increased risk of heart failure, ischemic heart disease and hypertension. The cardiovascular and pulmonary system needs to maintain a specific balance of GC levels, while either too high or too low concentrations have adverse effects on the system functions. It has been shown that mouse models with either GR deletion or an altered GR expression in cardiomyocytes exhibit cardiovascular pathologies, irregularly shaped and disorganized myofibrils at embryonic stages, larger hearts, as well as cardiac fibrosis. Males also develop cardiomyocyte hypertrophy, suggesting that the sexually dimorphic actions of GCs can occur in the heart as well (50). Furthermore, research on GR functions in rapid GC-induced hypertension has revealed that following vascular smooth muscle-specific GR deletion, there is a decrease in hypertension in mice (51). The mechanism used by GCs to regulate blood pressure is through the vascular GRs. Additionally, the endothelial GR appears to protect against lipopolysaccharide (LPS)-induced septic shock, via the repression of the release of NF-κB and IL-6, two inflammatory cytokines. On the other hand, the overexpression of GR is associated with atrio-ventricular block and bradycardia, with the first being easily reversed, once the levels of GR are restored (52).

GCs also participate in several metabolic processes. In the liver, GCs can stimulate gluconeogenesis by inducing phosphoenolpyruvate carboxykinase and glucose-6-phosphatase action, while in skeletal muscle and white adipose tissue, they lower glucose uptake and utilization by interfering with the insulin signaling pathway (53,54). In skeletal muscle specifically, GCs decrease insulin receptor substrate 1 transcription, a downstream signaling molecule of the insulin signaling pathway, and increase the transcription of protein tyrosine phosphatase type 1B and p38 mitogen-activated protein kinases (p38 MAPK), which counter insulin action (55). Of note, excessive GC action has been found to be associated with several metabolic diseases, such as type II diabetes and obesity (56).

GCs influence several central nervous system (CNS) processes. GCs' effect on the CNS is both cell-type and stress-type specific (57). GCs have been shown to regulate stress responses, apoptosis and long-term potentiation (58). These hormones can cross the blood-brain-barrier and alter synaptic physiology, stress responsiveness and behavior (59). Specifically, in the brain, following a GC peak, areas involved in emotional responses and simple behavioral strategies exhibit an enhanced activity, while in the aftermath of stress, areas involved in higher cognitive functions are activated and allow individuals to associate stressful events with a specific context and thus store information for future use (60). Maladaptive behavioral responses to stress have been shown to be associated with mood disorders (61). According to in vivo studies, transgenic mice where the GR was deleted in the CNS cells (brain-specific GR knockout), exhibited increased basal corticosterone levels and reduced anxiety. Considering that corticosterone is the rodent analog of human cortisol, the results can be attributed to the loss of central feedback inhibition in the HPA axis stress system. Moreover, mice which did not produce GR in the forebrain exhibited elevated MR expression in the hippocampus. Several variations of GR knockout mice in the hypothalamus resulted in sex specific differences in the HPA axis function, where females exhibited elevated corticosterone levels at the lowest point of the circadian cycle, while males did the same after intense stress. Notably, the depletion of GR in dopamine-receptive neurons resulted in mice with phenotypes of social aversion due to chronic stress (47).

The involvement of stress system dysregulation and more specifically, the dysregulation of the HPA axis in neuropsychiatric disorders is equally unsurprising, given the robust and dynamic nature of stress biology. Affective disorders, including major depressive disorder, bipolar disorder, anxiety and panic disorders, schizophrenia and post-traumatic stress disorder are considered anxiety disorders in which the key neural pathways that regulate the stress response do not function optimally (62). The increased reactivity to threatening stimuli, the reduced ability to finish the stress response, and/or the suboptimal coupling between internal emotional states and the external environment are part of stress system dysregulation. Of these, the latter dysfunction may contribute to mood shifting in an extreme and seemingly random manner in bipolar illness or to 'stick' in a negative manner in major depression. Although both disorders are primarily inherited (63,64), vulnerability to these is directly linked to how the individual responds to the environment (65). Indeed, according to the study by Wray et al (66), who examined ~460,000 individuals for the genetics of depression, it appeared that 'all people carry a greater or lesser number of genetic risk factors for major depression', where these are due to the dysregulation of the HPA axis in depression. Major depression is the result of the dysregulation of various genes related to cellular development, cell repair and growth factors. An example of these genes is the fibroblast growth factor (FGF) gene family, which is dysregulated in major depression (67). FGF2, which is a member of the FGF family, has been detected to function as an 'endogenous antidepressant'. Its levels in the hippocampus and frontal cortex are low in depression in humans, and they are reduced in rodents as a result of repeated social stress, and have been shown to modulate the HPA axis via GR (68). On the other hand, the levels of FGF9, another gene member of the FGF family, are increased in the depressed brain, as this gene functions as a vulnerability factor and is increased by chronic stress in animal models (67), and its selective inhibition in the hippocampus reduces anxiety behavior. Of note, treatment with FGF2 during early life leads to epigenetic changes to GR in the hippocampus, increasing its association with a repressive histone, H3K9Me3 (69), which leads to reduced levels of NR3C1 expression and with a lower number of GRs in the hippocampus (70). In addition to major depression, schizophrenia and bipolar disorder are also associated with the dysregulation of HPA axis activity, under basal conditions and during stress. The decreased mRNA expression of the GR has been observed in both illnesses using multiple post-mortem tissue cohorts and brain regions (71). Moreover, the increased expression of a truncated GR protein isoform has been reproducibly demonstrated in the prefrontal cortex of patients with schizophrenia and bipolar disorder (72,73). Notably, in conjunction with the decreased expression of multiple GR mRNA transcripts and the increased expression of the functional, truncated GRα-D1 protein isoform 35, the altered expression, and the dysregulation of FKBP5, PTGES3 and BAG1 mRNAs have been also detected. Taken together, widespread stress-signaling alterations are detected in both schizophrenia and bipolar disorder (71).

4. miRNAs in glucocorticoid signaling

miRNAs are a class of small ncRNAs ~22 nt in length, which exert their effects by binding to the 3′UTR region of target mRNAs. Compared to other ncRNA classes, they have been the focus of extensive research for the past two decades, due to both their post-transcriptional regulatory capacity, as well as their therapeutic potential (74). To date, >1,900 miRNAs have been discovered in the human genome and more specifically, according to miRbase (release 22.1), there are 1,917 precursors and 2,654 mature miRNAs (GRCh38) (75-81), with regulatory activity spanning across a multitude of biological processes, such as development, metabolism, inflammation, as well as GR protein expression. miRNAs have long been established as dynamic regulators during development, but most importantly during environmental adaptation. As such, their physiological function in the regulation of GR expression is also predominantly in CNS development, as well as the brain's response to both pre- and post-natal challenges (82). Nevertheless, prolonged exposure to environmental stressors can lead to extensive neuronal reprogramming accompanied by altered GC response (83,84). Additionally, due to the extensive use of GCs for cancer treatment (85), accompanied by the central roles of miRNAs in numerous types of cancer (86), investigating the association between miRNAs and GC signaling is pivotal for developing successful therapeutic approaches.

The present review mainly focuses on two aspects of the miRNA regulation of GC signaling. In the first part, focus is placed on miRNAs that have been shown to be direct regulators of GRs, investigating their effects in both brain disorders, as well as in cancer development and treatment. In the second part, miRNAs directly or indirectly influenced by GCs are reviewed (Table I).

Table I

MicroRNAs and glucocorticoid activity.

Table I

MicroRNAs and glucocorticoid activity.

GR regulatorsBrain-relatedCancer-related Immune/inflammatory-relatedOther
miR-124(90-92)(99)(91)
miR-18a(94)
miR-137(96)
miR-142-3p(100)(99)
miR-101a, -96, -433(101)
miR-130b, -181, -636(103)
miR-1-3p, -128, -370, -28(122)
miR-29a(122,123)
miR-340(128)

Regulated by GCsBrain-relatedCancer-related Immune/inflammatory-relatedOther

miR-155(104)(105-107)
miR-511(108-110)
miR-98(112)
miR-101(113)
miR-17-92 cluster(114-116)
miR-221/222(118)
miR-125b(120)

[i] The table presents relations and references (numbers in parentheses) to miRNAs that function as regulators of GR or are regulated by GCs, in brain-related, cancer-related, immune-related or other diseases. GC, glucocorticoid; GR, glucocorticoid receptor.

miRNAs and GRs

NR3C1, the mRNA transcript of GRα protein, has been shown to harbor a number of conserved miRNA target sites in its 3′UTR (12). Using the well-established miRNA target prediction algorithm TargetScan (release 7.2; March, 2018) (87), 135 conserved putative miRNA 8-mer binding sites (with a <-0.1 context score) were predicted for the human NR3C1 transcript (Table SI). Extending this search to the gene-network level with the use of the Genemania NR3C1 network, as shown in Fig. 1, which includes 21 proteins predicted to interact and form a network with GRα and miRDB (88,89), a miRNA target prediction on the 3′UTR region algorithm which allows for the concurrent investigation of multiple genes. Thus, this made it possible to identify 120 miRNAs predicted to regulate more than one of the proteins of the Genemania network, via targeting their respective mRNA 3′UTRs (Fig. 2). This simple analysis serves to show the extensive regulatory effects multiple miRNAs can have on different levels of the GC network, as well as the great fine-tuning potential they can apply on GRα levels. Thus far, a number of these miRNAs have already been shown to interact with GRα and thus exert their regulatory effects.

First and foremost is miR-124, a miRNA induced by GC treatment, which has been the subject of extensive studies in relation to GR expression. Using in vitro experiments in Jurkat T-cells and T-cells from healthy volunteers and patients with sepsis, miR-124 was previously shown to decrease the expression of GRα by direct binding of the 3′UTR of the NR3C1 transcript (90). This direct binding was also further validated in experiments using 293/293T cells (91). Notably, miR-124 levels have been found to be increased in patients with major depressive disorder (92,93), while depression-like behaviors in mice were able to be effectively treated by miR-124 antagomir administration, effectively blocking the inhibitory activity of miR-124 and establishing it as both a biomarker and an interesting therapeutic target for depression treatment (91). miR-124a, which is exclusively expressed in the brain and is the most abundant miRNA in the vertebrate central nervous system, was shown to regulate GR expression by direct binding of the GR mRNA in P19 cells, a well-established neuronal differentiation cell line (94). A similar inhibitory activity was also observed for miR-18, although the binding potential for the GR transcript was not observed in vitro (94). Nevertheless, additional studies have implicated miR-18a as a direct post-transcriptional regulator of GR. Direct binding was observed in cultured neuronal cells, while miR-18a overexpression was concomitant with suppressed GR protein levels in a model of stress-susceptible rats, further establishing the direct GR regulation potential of the miR-18 family (95). miR-137, a miRNA identified as a potential regulator in schizophrenia, as well as other brain disorders, has been implicated in neuronal plasticity through GR-dependent signaling, while in silico target prediction algorithms identified NR3C1 transcript as its putative target in both humans and mice (96). In a model of depression, following chronic unpredictable mild stress, miR-382-5p levels were found to be elevated with the concomitant suppression of GR levels in the hippocampus of rats. si-miR-382-5p treatment restored GR levels, as well as its downstream target levels of BDNF and p-TrkB62 (97).

A number of miRNAs that regulate GR expression have been implicated in various types of cancer. In the case of miR-124, it has also been shown that it can exert indirect regulatory effects on the GC response and sensitivity. The upregulation of GRα expression, accompanied by increased cAMP levels and the decreased expression of phosphodiesterase 4B, has been observed under stable miR-124 expression, in diffuse large B-cells lymphoma cell lines (98). Additionally, in patients with acute lymphoblastic leukemia (ALL), miR-124 levels have been shown to be increased, while this miRNA contributes to GC resistance, decreased apoptosis and decreased GR expression (99). Another miRNA that has been shown to be able to directly bind the NR3C1 transcript and regulate its expression in various instances is miR-142-3p. This miRNA, which is upregulated in patients with leukemia, has been shown to be able to regulate GC response in T-leukemic cells by directly binding with the 3′UTR of the GRα mRNA (100). miR-142-3p was also identified in the study by Riester et al (101), where following ACTH stimulation, miR-101a, miR-142-3p, miR-96 and miR-433 were identified as putative direct regulators of GRα. Following modification of the 3′UTR of the NR3C1 transcript in an in vitro experiment, it was shown that these four miRNAs could directly bind the 3′UTR and inhibit GRα expression by up to 40% (101). By investigating GC sensitivity in multiple myeloma cell lines, Tessel et al (102) identified miR-130b, miR-181a and miR-636 as miRNAs with direct binding potential to the 3′UTR of the NR3C1 transcript, while also exhibiting differential expression between GC-sensitive and resistant cell lines. miR-130b overexpression in vitro caused decreased GR expression, concomitant with GC resistance, and decreased GC-induced apoptotic effects. Due to the common usage of GCs as a treatment for various forms of cancer, including leukemia and multiple myeloma, miRNA expression patterns can be used to identify GC responsiveness and resistance development (12,103).

GC regulation of miRNAs

Several miRNAs are involved in the mechanisms of the immune and inflammatory response, which are modulated and regulated by GCs. One of the first oncogenic miRNAs with elevated levels in several types of cancer (104), is miR-155. Increased levels of miR-155 have been reported in fibroblasts and macrophages in rheumatoid arthritis, where this miRNA appears to contribute to increased expression of chemokines and pro-inflammatory cytokines (105,106). It was previously demonstrated that the administration of GC dexamethasone to primary macrophages and macrophage cell lines, spleen and liver cells of LPS-injected mice, inhibited the expression of miR-155, suggesting that a decrease in miR-155 expression is an anti-inflammatory result of GCs (107,108).

miR-511 is a miRNA produced by the fifth intron of mRNA encoding the C-type mannose receptor CD206. More specifically, miR-511-5p contributes to impaired sensitivity to LPS and reduced expression of the classically activated (M1) macrophage marker IL-12, targeting the Toll-like receptor 4 and IL-12p40 subunit (109). In addition, miR-511-5p targets the p55 TNF receptor mRNA, thus providing TNF resistance, in cases where there is increased expression of this miRNA due to high levels of endogenous GCs (110). In conclusion, changes in the expression level of miR-511 are dependent on levels of endogenous GCs and have an impact on the differentiation and activation of myeloid cells, as a byproduct of altered MRC1 gene expression (111).

A member of the let-7 miRNA family miR-98 is another GC-induced miRNA that reduces the expression of the p75 TNF receptor in T-lymphocytes by targeting Tnfrsf1b mRNA. Moreover, this miRNA is considered to target 3′UTR of IL-13, which plays a pathogenic role in asthma. In this way, miR-98 may contribute to the therapeutic effects of GCs on asthma (112). miR-101 is involved in inflammatory responses in myeloid and other cells and targets the 3′UTR of dual specificity phosphatase 1 (DUSP1) to regulate the activation of p38 MAPK in macrophages. Numerous studies have demonstrated that GCs in combination with pro-inflammatory stimuli, enhance DUSP1 expression, leading to the inactivation of p38 MAPK. In the study by Zhu et al (113), it was shown that GCs inhibit the expression of miR-101, thus leading to prolonged expression of DUSP1, which in turn led to reduced activation of p38 MARK and consequently to a decreased expression of p38 MARK-dependent inflammatory mediators.

The miR-17-92 cluster consists of six miRNAs, miR-17, -18a, -19a, -19b-1, -20a and 92a-1, which are derived from a precursor RNA transcribed from chromosome 13 (114). This complex targets the PTEN tumor suppressor, which regulates inositol phosphate signaling, and a pro-apoptotic member of the Bcl-2 protein family, Bim which regulates cytochrome c-induced apoptosis (115,116). According to several studies, GCs function as inhibitors of this cluster, thus exerting several therapeutic effects on lymphoma, while the failure of GCs to inhibit the miR-17-92 cluster resulted in resistance to apoptosis. In conclusion, the miR-17-92 cluster plays an important role in regulating responses to pro- and anti-apoptotic signals (116,117).

Finally, two miRNAs that have been linked to GC-induced apoptosis resistance in multiple myeloma (MM) are miR-221/222 (118) and miR-125b (119). The increased expression of these miRNAs has been shown to be associated with the attenuation of cell death pathways, such as apoptosis led by tumor protein 53. miR-125b expression levels are regulated by GCs as part of a possible self-limiting mechanism of GCs pro-apoptotic effects (119). On the other hand, a detailed analysis of miR-221 expression demonstrates the specificity of its expression in different types of cancer. More specifically, the increased expression of miR-221 in MM has been shown to induce resistance to GC-induced apoptosis (118), contrary to certain types of ALL, where resistance to GC-induced apoptosis is due to the reduced expression of miR-221 (120).

Performing RNA-sequencing and microarray expression analysis of male and female mouse hearts with a cardiomyocyte-specific knockout of the GR, Cruz-Topete et al (121) were able to identify 130 miRNAs whose expression was sex- and GR-dependent. Of these miRNAs, 25 were responsible for the vast majority of the differences observed between male and female hearts, including prominent heart failure biomarkers, such as miR-1-3p, miR-128, miR-370 and miR-28. miR-29a and miR-340 overexpression in L929 cells was previously found to be associated with a significant reduction in GR expression, suggesting a direct binding role for both miRNAs and the GR receptor transcript (122). Of note, in another study, the overexpression of miR-29a in vivo ameliorated GC-induced bone tissue destruction (123), while miR-340 overexpression decreased GR protein levels in the mouse placenta and affected sensitivity against activity-based anorexia (124).

Last but not least, a few recent studies have also identified GR as a potential miRNA regulator. In an in vitro study, miR-22 expression was shown to be elevated in AR42J cells following the induction of apoptosis. A GR binding site was identified in the promoter region of miR-22, while GR was shown to be able to repress miR-22 expression (125). In addition, a recent profiling study in triple-negative breast cancer suggested the ability of GR to influence multiple miRNA expression profiles (126), while in the study by Tejos-Bravo et al (127), neuron-specific GR knockout mice exhibited altered miRNA expression profiles in a sex-dependent manner. Such effects exhibit the great complexity inherent in miRNA regulation of GC activity, but also suggest an additional fine-tuning potential via the possible integration of feedback loops.

miRNA regulation of GR-chaperone proteins

As previously mentioned, until the moment of GC activation GR remains located in the cytoplasm, forming a complex with various chaperone proteins. As such, a number of secondary-degree regulatory potential exists via miRNA-mediated regulation of these chaperone proteins. While not thoroughly investigated, such effects have already been identified. miR-511, whose effects were previously discussed, was found to be able to directly bind the 3′UTR of FKBP5, suppressing GC-induced FKBP51 expression, while increasing neuronal development (128). In a similar manner, miR-124 was also shown to be able to target FKBP5 (129). Of note miR-142 and miR-340, miRNAs with the established binding of GRα transcript, were predicted also to target the 3′UTR of FKBP5 (128). Taking into account the numerous chaperone proteins that GRα interacts with, as well as the indicative gene regulatory network already established (Fig. 1), it is clear that the miRNA-mediated regulation of GC response is a very complex and multilayered process, which has only begun to be explored.

5. lncRNAs in glucocorticoid signaling

As already mentioned, GC signaling involves various signal transduction cascades in the cell. In recent years, studies focusing on the roles of ncRNAs have increased, including their roles in regulating the transcriptional activity of GR and other NRs (Table II). lncRNAs are a class of ncRNAs, which consist of >200 nucleotides and are derived from various regions in the genome, such as promoters, enhancers, introns, UTRs, overlapping or non-coding isoforms of coding genes, antisense to other transcripts and pseudogenes (130,131). According to previous studies, lncRNAs have been observed in the majority of organisms, such as animals (132), plants (133), fungi (134) and even viruses (135), without however evolving conservation among species. Following technological advancements and novel laboratory techniques, numerous data related to the roles of lncRNAs in a variety of vital biological processes have been gathered (136) and more specifically in transcription (137), alternative splicing (138), translation, cell cycle (139), apoptosis (140) and heat shock response (141). Several lncRNAs create RNA-protein, RNA-DNA and RNA-RNA complexes, while they have also been associated with chromatin modification and guiding transcription factors to specific genomic DNA targets. Last but not least, a number of lncRNAs have been found to be associated with various diseases, including cancer, myocardial infractions and Alzheimer's disease (142,143).

Table II

lncRNAs and glucocorticoid activity.

Table II

lncRNAs and glucocorticoid activity.

lncRNAsType of interaction Metabolism-related Immune/inflammatory-relatedCancer-relatedOther
GR-related
 GAS5GR regulator(14)(144-145,147)(148)(146)
 EDN1-ASRegulated by GR(15,149-150)
 PSORS1C3Regulated by GR(152-154)
Related to other receptors
 SRAAR, ER, GR and PR regulator(156)(155)
 PRNCR1, PCGEM1AR regulators(157)

[i] Type of interaction between the lncRNAs and corresponding receptors, and references (numbers in parentheses) to lncRNAs related to GR or to other NR in metabolism-related, immune/inflammatory-related, cancer-related or other diseases. lncRNA, long non-coding RNA; GC, glucocorticoid; GR, glucocorticoid receptor; growth arrest-specific 5; PSORS1C3, psoriasis susceptibility 1 candidate 3; SRA, steroid RNA coactivator; AR, androgen receptor; ER, estrogen receptor; PR, progesterone receptor.

GAS5 is a lncRNA that has been of immense interest to researchers in recent years and is involved in GR activity. As its name suggests, GAS5 inhibits cell growth caused by a lack of nutrients or growth factors. In the study by Kino et al (14), it was shown that GAS5 functions as an inhibitor of the transcriptional activity of GR and other steroid receptors (SRs). More specifically, the sequence of GAS5 contains two GC response elements-mimetic sequences at nucleotides 539-544 (GRE-1) and 553-559 (GRE-2). Thus, GAS5 acts as a competitor for GR binding through the GRE regions, decreasing GR-mediated gene activation and ultimately affecting cell survival and metabolic activity during nutrient deficiency. GAS5 overexpression significantly inhibits the transcription of GR target genes, including genes encoding the cellular inhibitor of apoptosis 2 and serum/GC-responsive kinase 1 (14).

However, an interesting observation was the fact that during growth inhibition or lack of nutrients in cells, the accumulation of GAS5 was reported in the organs of mice which are involved in metabolism, i.e., in the liver and adipose tissue, by modifying the mTOR signaling pathway, contrary to organs involved in the immune system, such as the thymus gland, spleen and the brain. Recent studies have demonstrated that the levels of GAS5 in cells can vary. In particular, it has been suggested that GAS5 exerts regulatory activity on GR in the immune system, independently of nutrient availability. This is evidenced by the different expression of GAS5 in leukocytes of patients with inflammatory or autoimmune diseases (144), as well as its role in the GC response of children with diseases, such as inflammatory bowel disease (145), multiple sclerosis (146), human beta cell dysfunction (147) and acute myeloid leukemia (148).

EDN1-AS is a lncRNA, which appears to interact with GR. This lncRNA is located antisense of the endothelin 1 gene, which is a peptide hormone that acts on the vascular system as a vasoconstrictor, while in the kidneys it affects blood pressure through diuretic and natriuretic effects (15). Its aberrant quantity has been shown to be associated with pathological conditions, such as hypertension (149) and chronic kidney disease (150). EDN1-AS is expressed in multiple human cell types, including the kidneys. In the study by Douma et al (15) in a human kidney proximal tubule cell line (HK-2), the promoter of lncRNA EDN1-AS appeared to have a GRE sequence, which could be recognized and bound by GR, as well as the MR, representing a new mechanism for regulating ET-1 expression. Using CRISPR/Cas9 for the deletion of the GRE element from the EDN1-AS promoter, abolished the binding of GR to EDN1-AS and resulted in increased EDN1-AS expression with a concomitant increase in endothelin 1 expression and cell proliferation. The inhibitory effect of GR binding to the EDN1-AS promoter is therefore inferred (15).

Psoriasis susceptibility 1 candidate 3 (PSORS1C3) is a lncRNA whose sequence is adjacent to the octamer-binding transcription factor 4 (OCT4) gene. As is well known, the transcription factor OCT4 plays regulatory roles in oncogenesis, stemness and in response to stress. PSORS1C3 has shown to be associated with diseases, such as psoriasis (151) and other immune-mediated diseases, such as acute anterior uveitis (152), as well as major depressive disorder (153). According to the study by Mirzadeh Azad et al (154), this lncRNA has two endogenously active promoters, promoters 0 and 1 and two sets of transcripts, small and large variants. A GRE sequence was identified upstream of promoter 0 of PSORS1C3, where GR binds and acts as either an enhancer or a repressor of the expression of target genes. More specifically, that study demonstrated the positive effect of GR on the expression level of OCT4 and small variants of PSORS1C3, which may reflect a new regulatory pathway of cell proliferation and the stress response through the regulation of expression levels of OCT4. On the other hand, the binding of GR to promoter 0 of large variants of lncRNA PSORS1C3 exerted an inhibitory effect, suggesting its function as a pro-inflammatory factor. Thus, GR functions as a regulator of the expression of the lncRNA PSORS1C3, which in turn regulates and moderates the expression of the OCT4 factor in non-multipotent cells, as the PSORS1C3 promoter 0 region acts as an enhancer for the neighboring OCT4 gene (154).

In general, several lncRNAs have been recorded to regulate the transcriptional activity of several SRs, including GR. A prototype lncRNA is the steroid RNA coactivator (SRA) which increases the transcriptional activity of the androgen receptor (AR), estrogen receptor, GR and progesterone receptor. The SRA regulates the transcriptional activity by binding to the SRA stem-loop-interacting RNA binding-protein and the RNA-induced silencing complex complex (155,156). In addition, according to Yang et al (157), lncRNAs PRNCR1 and PCGEM1, which are expressed primarily in the prostate gland, bind AR to its DNA binding domain, and suppress receptor activity, playing a significant role in the development of prostate cancer. In summary, lncRNAs appear to regulate the expression levels of various target genes of NRs, including GR. For this reason, future research is required to elucidate the function of lncRNAs in regulating gene expression through their interaction with transcription factors, such as GR in GC signaling pathways.

6. Discussion

The stress response system is related to the production of endogenous GCs. This system, particularly the interaction between cortisol and GR, which is described by the GR pathways, has attracted increasing attention in recent years due to its medical interest in developing therapeutic approaches. Although GR is derived from a single gene, its function differs based on the different isomers which are produced due to alternative splicing and alternative translation initiation mechanisms. It has been shown that GCs bind in a similar manner to all the isomers (158); however, GRs differ in their subcellular distribution and gene regulatory profiles, affecting the human organism via multiple mechanisms. It should be noted that some GR polymorphisms have been linked to GC resistance and a healthier metabolic profile, whereas others seem to be associated with GC hypersensitivity increasing cardiovascular risk (159). It has been proposed that the regulation of GC signaling is related to pathological conditions, such as cancer, heart diseases, diabetes and other metabolic disorders.

Recently, numerous studies have implicated GC signaling in cancer progression or prevention, depending on the cell type. In one case, animal models of human breast cancer revealed that GCs inhibit tumor cell apoptosis, while in other cases, synthetic GCs are used to induce apoptotic cell death in malignant lymphoid cells (e.g., lymphoma). Even though pharmacologic GC therapy is frequently administered to cancer patients to reduce the associated side effects of chemotherapy, further investigations on the results of the treatment on patients need to be conducted as GC application may contribute to tumor growth (160). Other studies have suggested that GR signaling in cardiomyocytes is critical for the normal development and function of the heart (161,162). In a previous study (161), cardiomyocyte-specific GR overexpression led to bradycardia, while GR inhibition resulted in cardiac hypertrophy, systolic dysfunction, and impaired maturation. Further research is required in order to determine the precise molecular pathways and genes through which cardiomyocyte GC signaling can either promote or protect against heart pathology (161). It should also be mentioned that, apart from the endogenous GCs, studies have suggested the use of dexamethasone for the treatment of post-operative nausea and vomiting, as it is a synthetic GC with anti-inflammatory and immunosuppressant properties, with 20- to 30-fold the binding affinity for GR of endogenous cortisol (163).

Due to the ability of miRNAs to regulate multiple targets, while every 3′UTR can harbor multiple binding sites for different microRNAs, it is easily apparent that post-transcriptional regulation via miRNAs forms a highly complex and sensitive network. Under physiological conditions, the miRNA-mediated modulation of GR expression is mainly involved in fine-tuning GC responses during CNS development. However, a number of pathological responses have implicated the interaction between the mRNA transcript of GR protein and different miRNAs, as observed in brain disorders, as well as in cancer development and treatment. Taking into account the negative role of miRNAs in steroidogenesis, the association between miRNAs and GR expression warrants further in-depth investigations. Currently, numerous miRNAs have already been identified that have the capacity to directly bind NR3C1 3′UTR and exercise their regulatory effects, such as miR-124-3p and miR-142-3p, among several others. Nevertheless, the concept of a single cause-single target, while beneficial for past drug development, is gradually being pushed aside, as the multi-factorial nature of regulation becomes more and more apparent. These types of interactions offer increased sensitivity and fine-tuning potential to environmental changes. More systematic approaches are therefore necessary that will offer a holistic regulatory view. Combining existing target prediction tools with network generation packages, their complexity instantly emerges (Fig. 2). Additional regulatory levels are also exercised through alternative splicing and the production of other GR isoforms. This is the case for miR-124-3p, where differential splicing produces the GRb isoform, no longer harboring its binding site, which in turn acts as a negative inhibitor of GRα (90,162). Tying this to recent studies, demonstrating that the GR can itself influence miRNA expression profiles (125-127), the intricacies of miRNA-GR regulation and response to GCs are becoming exponentially complex.

Similar to miRNAs, lncRNAs have a variety of regulatory roles in gene expression. Their ability to form RNA-protein, RNA-DNA and RNA-RNA complexes enable them to be involved in cellular processes, such as apoptosis, translation, cell cycle and heat shock response. They play key roles in various disease cases, such as numerous types of cancer, myocardial infraction and neurodegenerative diseases, such as Alzheimer's disease (142,143). lncRNAs are further involved in various biological processes, including GC signaling, which connects to extensive pathways related to the immune, nervous and metabolic systems. An example is lncRNA GAS5, which regulates the GR response via direct binding through its GRE sequences, thus having an impact on the gene expression of the GR target-genes (14). EDN1-AS and PSORS1C3 are two additional lncRNAs whose expression is regulated by the GR. In both cases, these lncRNAs are bound with the GR via GREs, which in turn acts as their inhibitor, affecting the expression of genes that interact with them, the EDN1 (15) and OCT4 (154) genes, respectively.

It thus becomes clear that GC signaling has extensive and important functions in the immune, nervous system and related metabolic responses in the context of homeostasis. Its activity has been associated with numerous pathological conditions, such as cancer, metabolic and neurodegenerative diseases. It is easily apparent that GC signaling is part of an intricate regulatory network, heavily involving post-transcriptional regulation via ncRNAs, in an effort to maintain the fine-tuning potential the body needs to respond to ever-shifting environmental conditions. Such a network is comprised of numerous miRNAs, lncRNAs and potentially yet unverified actors. Their complex associations are just beginning to be unraveled, but are already demanding new analysis paradigms to be adopted. To this aim, combinatory bioinformatic approaches need to be employed and new tools developed that can investigate such effects in a more systematic manner. Only then can a better understanding of GC activity be obtained and the utilization of its therapeutic potential can effectively be achieved.

7. Conclusions

ncRNAs are an intriguing field of study. Their unique properties, as well as their ability to be involved in vital cellular processes, render them suitable pharmaceutical targets and biomarkers. GC signaling participates in a regulatory network that includes post-transcriptional regulation via ncRNAs, including various miRNAs and lncRNAs. These molecules can either act as regulators of GR activity or be regulated by endogenous GCs, affecting the expression of GC-mediated genes. Generally, GC activity is associated with several pathological conditions, including cancer, neurodegenerative and metabolic diseases. Therefore, the development of more effective therapies for these diseases requires a better understanding of GC signaling that includes interacting regulatory ncRNAs.

Supplementary Data

Availability of data and materials

Not applicable.

Authors' contributions

All authors (KP, LP, TM, EP, ID, SL, MS, KD, DAS, FB, GPC, GNG, EE and DV) contributed to the conceptualization and design of the study, as well as in the writing, drafting, revising, editing and reviewing of the manuscript. Data authentication is not applicable. All authors have read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

DAS is the Editor-in-Chief for the journal, but had no personal involvement in the reviewing process, or any influence in terms of adjudicating on the final decision, for this article. The other authors declare that they have no competing interests.

Acknowledgments

Not applicable.

Funding

The authors would like to acknowledge funding from the following organizations: i) AdjustEBOVGP-Dx (RIA2018EF-2081): Biochemical Adjustments of native EBOV Glycoprotein in Patient Sample to Unmask target Epitopes for Rapid Diagnostic Testing. A European and Developing Countries Clinical Trials Partnership (EDCTP2) under the Horizon 2020 'Research and Innovation Actions' DESCA; and ii) 'MilkSafe: A novel pipeline to enrich formula milk using omics technologies', a research co-financed by the European Regional Development Fund of the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH-CREATE-INNOVATE (project code: T2EDK-02222).

Abbreviations:

ACTH

adrenocorticotropic hormone

AF

activation function

ALL

acute lymphoblastic leukemia

AR

androgen receptor

AVP

arginine-vasopressin

CNS

central nervous system

CRH

corticotropic-releasing hormone

DBD

DNA-binding domain

DUSP1

dual specificity phosphatase 1

FKBPs

FK506-binding proteins

GAS5

growth arrest-specific 5

GC

glucocorticoid

GR

glucocorticoid receptor

GRE

glucocorticoid response element

Hop

Hsp70-Hsp90 organizing protein

HPA

hypothalamic-pituitary-adrenal cortex

HR

hinge region

Hsp

heat shock protein

LBD

ligand-binding domain

lncRNA

long non-coding RNA

LPS

lipopolysaccharide

miRNAs/miRs

microRNAs

MM

multiple myeloma

MR

mineralocorticoid receptor

ncRNA

non-coding RNA

NTD

N-terminal domain

PTGES3/p23

prostaglandin E synthase 3

SRs

steroid receptors

References

1 

Chrousos GP: Stress and disorders of the stress system. Nat Rev Endocrinol. 5:374–381. 2009. View Article : Google Scholar : PubMed/NCBI

2 

Tsigos C, Kyrou I, Kassi E and Chrousos GP: Stress: Endocrine Physiology and Pathophysiology. Endotext. Feingold KR, Anawalt B, Boyce A, Chrousos G, de Herder WW, Dungan K, Grossman A, Hershman JM, Hofland J, Kaltsas G, et al: MDText.com, Inc. Copyright© 20002021, MDText.com, Inc. South Dartmouth, MA: 2000

3 

Charmandari E, Kino T, Souvatzoglou E and Chrousos GP: Pediatric stress: Hormonal mediators and human development. Horm Res. 59:161–179. 2003.PubMed/NCBI

4 

Yaribeygi H, Panahi Y, Sahraei H, Johnston TP and Sahebkar A: The impact of stress on body function: A review. EXCLI J. 16:1057–1072. 2017.PubMed/NCBI

5 

Russell G and Lightman S: The human stress response. Nat Re Endocrinol. 15:525–534. 2019. View Article : Google Scholar

6 

Smith SM and Vale WW: The role of the hypothalamicpituitary-adrenal axis in neuroendocrine responses to stress. Dialogues Clin Neurosci. 8:383–395. 2006. View Article : Google Scholar

7 

Nicolaides NC, Charmandari E, Kino T and Chrousos GP: Stress-related and circadian secretion and target tissue actions of glucocorticoids: Impact on Health. Front Endocrinol (Lausanne). 8:70. 2017. View Article : Google Scholar

8 

Dunlavey CJ: Introduction to the Hypothalamic-pituitary-adrenal axis: Healthy and dysregulated stress responses, developmental stress and neurodegeneration. J Undergrad Neurosci Educ. 16:R59–R60. 2018.PubMed/NCBI

9 

DeMorrow S: Role of the Hypothalamic-pituitary-adrenal axis in health and disease. Int J Mol Sci. 19:9862018. View Article : Google Scholar

10 

Chrousos GP: Stressors, stress, and neuroendocrine integration of the adaptive response. The 1997 hans selye memorial lecture. Ann N Y Acad Sci. 851:311–335. 1998. View Article : Google Scholar : PubMed/NCBI

11 

Evanson NK, Tasker JG, Hill MN, Hillard CJ and Herman JP: Fast feedback inhibition of the HPA axis by glucocorticoids is mediated by endocannabinoid signaling. Endocrinology. 151:4811–4819. 2010. View Article : Google Scholar : PubMed/NCBI

12 

Wang H, Gou X, Jiang T and Ouyang J: The effects of microRNAs on glucocorticoid responsiveness. J Cancer Res Clin Oncoly. 143:1005–1011. 2017. View Article : Google Scholar

13 

Kawa MP, Sobuś A, Litwińska Z, Osowicz-Korolonek L, Cymbaluk-Płoska A, Stecewicz I, Zagrodnik E, Romanowska H, Walczak M, Syrenicz A and Machaliński B: Expression of selected angiogenesis-related small microRNAs in patients with abnormally increased secretion of glucocorticoids. Endokryno Pol. 70:489–495. 2019. View Article : Google Scholar

14 

Kino T, Hurt DE, Ichijo T, Nader N and Chrousos GP: Noncoding RNA gas5 is a growth arrest- and starvation-associated repressor of the glucocorticoid receptor. Sci Signal. 3:ra82010. View Article : Google Scholar : PubMed/NCBI

15 

Douma LG, Solocinski K, Masten SH, Barral DH, Barilovits SJ, Jeffers LA, Alder KD, Patel R, Wingo CS, Brown KD, et al: EDN1-AS, a novel long non-coding rna regulating endothelin-1 in human proximal tubule cells. Front Physiol. 11:2092020. View Article : Google Scholar :

16 

Silverman MN and Sternberg EM: Glucocorticoid regulation of inflammation and its functional correlates: From HPA axis to glucocorticoid receptor dysfunction. Ann N Y Acad Sci. 1261:55–63. 2012. View Article : Google Scholar

17 

Timmermans S, Souffriau J and Libert C: A general introduction to glucocorticoid biology. Front Immunol. 10:1545. 2019. View Article : Google Scholar : PubMed/NCBI

18 

Flynn BP: Glucocorticoid ultradian rhythms. Curr Opin Endocrine Metabolic Res. 25:1003622022. View Article : Google Scholar

19 

Kalafatakis K, Russell GM, Ferguson SG, Grabski M, Harmer CJ, Munafò MR, Marchant N, Wilson A, Brooks JC, Thakrar J, et al: Glucocorticoid ultradian rhythmicity differentially regulates mood and resting state networks in the human brain: A randomised controlled clinical trial. Psychoneuroendocrinology. 124:1050962021. View Article : Google Scholar :

20 

Dickmeis T: Glucocorticoids and the circadian clock. J Ndocrinol. 200:3–22. 2009. View Article : Google Scholar

21 

Sevilla LM and Pérez P: Roles of the Glucocorticoid and mineralocorticoid receptors in skin pathophysiology. Int J Mol Sci. 19:19062018. View Article : Google Scholar :

22 

Sarabdjitsingh RA, Meijer OC and de Kloet ER: Specificity of glucocorticoid receptor primary antibodies for analysis of receptor localization patterns in cultured cells and rat hippocampus. Brain Res. 1331:1–11. 2010. View Article : Google Scholar : PubMed/NCBI

23 

Desmet SJ and De Bosscher K: Glucocorticoid receptors: Finding the middle ground. J Clin Invest. 127:1136–1145. 2017. View Article : Google Scholar : PubMed/NCBI

24 

Nicolaides NC, Skyrla E, Vlachakis D, Psarra AM, Moutsatsou P, Sertedaki A, Kossida S and Charmandari E: Functional characterization of the hGRαT556I causing Chrousos syndrome. Eur J Clin Invest. 46:42–49. 2016. View Article : Google Scholar

25 

Paragliola RM, Papi G, Pontecorvi A and Corsello SM: Treatment with synthetic glucocorticoids and the hypothalamus-pituitary-Adrenal Axis. Int J Mol Sci. 18:22012017. View Article : Google Scholar :

26 

Mazaira GI, Zgajnar NR, Lotufo CM, Daneri-Becerra C, Sivils JC, Soto OB, Cox MB and Galigniana MD: The nuclear receptor field: A historical overview and future challenges. Nucl Receptor Res. 5:1013202018. View Article : Google Scholar : PubMed/NCBI

27 

Porter BA, Ortiz MA, Bratslavsky G and Kotula L: Structure and function of the nuclear receptor superfamily and current targeted therapies of prostate cancer. Cancers (Basel). 11:18522019. View Article : Google Scholar

28 

Weikum ER, Okafor CD, D'Agostino EH, Colucci JK and Ortlund EA: Structural analysis of the glucocorticoid receptor ligand-binding domain in complex with triamcinolone acetonide and a fragment of the atypical coregulator, small heterodimer partner. Mol Pharmacol. 92:12–21. 2017. View Article : Google Scholar : PubMed/NCBI

29 

Tan CK and Wahli W: A trilogy of glucocorticoid receptor actions. Proc Natl Acad Sci USA. 113:1115–1117. 2016. View Article : Google Scholar : PubMed/NCBI

30 

Nicolaides NC, Galata Z, Kino T, Chrousos GP and Charmandari E: The human glucocorticoid receptor: Molecular basis of biologic function. Steroids. 75:1–12. 2010. View Article : Google Scholar :

31 

Kaziales A, Barkovits K, Marcus K and Richter K: Glucocorticoid receptor complexes form cooperatively with the Hsp90 co-chaperones Pp5 and FKBPs. Sci Rep. 10:10733. 2020. View Article : Google Scholar : PubMed/NCBI

32 

Baker JD, Ozsan I, Rodriguez Ospina S, Gulick D and Blair LJ: Hsp90 heterocomplexes regulate steroid hormone receptors: From stress response to psychiatric disease. Int J Mol Sci. 20:792018. View Article : Google Scholar

33 

Louw A: GR Dimerization and the Impact of GR Dimerization on GR protein stability and half-life. Front Immunol. 10:1693. 2019. View Article : Google Scholar : PubMed/NCBI

34 

Robertson S, Hapgood JP and Louw A: Glucocorticoid receptor concentration and the ability to dimerize influence nuclear translocation and distribution. Steroids. 78:182–194. 2013. View Article : Google Scholar

35 

Frego L and Davidson W: Conformational changes of the glucocorticoid receptor ligand binding domain induced by ligand and cofactor binding, and the location of cofactor binding sites determined by hydrogen/deuterium exchange mass spectrometry. Protein Sci. 15:722–730. 2006. View Article : Google Scholar : PubMed/NCBI

36 

Vandevyver S, Dejager L and Libert C: On the trail of the glucocorticoid receptor: Into the nucleus and back. Traffic. 13:364–374. 2012. View Article : Google Scholar

37 

Hudson WH, Youn C and Ortlund EA: The structural basis of direct glucocorticoid-mediated transrepression. Nat Struct Mol Biol. 20:53–58. 2013. View Article : Google Scholar :

38 

Groeneweg FL, van Royen ME, Fenz S, Keizer VI, Geverts B, Prins J, de Kloet ER, Houtsmuller AB, Schmidt TS and Schaaf MJ: Quantitation of glucocorticoid receptor DNA-binding dynamics by single-molecule microscopy and FRAP. PLoS One. 9:e905322014. View Article : Google Scholar : PubMed/NCBI

39 

Quatrini L and Ugolini S: New insights into the cell- and tissue-specificity of glucocorticoid actions. Cell Mol Immunol. 18:269–278. 2021. View Article : Google Scholar

40 

Petta I, Dejager L, Ballegeer M, Lievens S, Tavernier J, De Bosscher K and Libert C: The Interactome of the glucocorticoid receptor and its influence on the actions of glucocorticoids in combatting inflammatory and infectious diseases. Microbiol Mol Biol Rev. 80:495–522. 2016. View Article : Google Scholar : PubMed/NCBI

41 

Oeckinghaus A and Ghosh S: The NF-kappaB family of transcription factors and its regulation. Cold Spring Harb Perspect Biol. 1:a0000342009. View Article : Google Scholar

42 

Rao NA, McCalman MT, Moulos P, Francoijs KJ, Chatziioannou A, Kolisis FN, Alexis MN, Mitsiou DJ and Stunnenberg HG: Coactivation of GR and NFKB alters the repertoire of their binding sites and target genes. Genome Res. 21:1404–1416. 2011. View Article : Google Scholar : PubMed/NCBI

43 

Shimba A and Ikuta K: Control of immunity by glucocorticoids in health and disease. Semin Immunopathol. 42:669–680. 2020. View Article : Google Scholar : PubMed/NCBI

44 

Warde-Farley D, Donaldson SL, Comes O, Zuberi K, Badrawi R, Chao P, Franz M, Grouios C, Kazi F, Lopes CT, et al: The GeneMANIA prediction server: Biological network integration for gene prioritization and predicting gene function. Nucleic Acids Res. 38:W214–W220. 2010. View Article : Google Scholar : PubMed/NCBI

45 

Wilson KS, Tucker CS, Al-Dujaili EA, Holmes MC, Hadoke PW, Kenyon CJ and Denvir MA: Early-life glucocorticoids programme behaviour and metabolism in adulthood in zebrafish. J Endocrinol. 230:125–142. 2016. View Article : Google Scholar : PubMed/NCBI

46 

Rog-Zielinska EA, Craig MA, Manning JR, Richardson RV, Gowans GJ, Dunbar DR, Gharbi K, Kenyon CJ, Holmes MC, Hardie DG, et al: Glucocorticoids promote structural and functional maturation of foetal cardiomyocytes: A role for PGC-1α. Cell Death Differ. 22:1106–1116. 2015. View Article : Google Scholar

47 

Whirledge S and DeFranco DB: Glucocorticoid signaling in health and disease: Insights from tissue-specific GR knockout mice. Endocrinology. 159:46–64. 2018. View Article : Google Scholar :

48 

Meszaros K and Patocs A: Glucocorticoids influencing Wnt/β-catenin pathway; multiple sites, heterogeneous effects. Molecules. 25:14892020. View Article : Google Scholar

49 

Steptoe A and Kivimäki M: Stress and cardiovascular disease. Nat Rev Cardiol. 9:360–370. 2012. View Article : Google Scholar : PubMed/NCBI

50 

Duma D, Collins JB, Chou JW and Cidlowski JA: Sexually dimorphic actions of glucocorticoids provide a link to inflammatory diseases with gender differences in prevalence. Sci Signal. 3:ra742010. View Article : Google Scholar : PubMed/NCBI

51 

Goodwin JE, Zhang J and Geller DS: A critical role for vascular smooth muscle in acute glucocorticoid-induced hypertension. J Am Soc Nephrol. 19:1291–1299. 2008. View Article : Google Scholar : PubMed/NCBI

52 

Goodwin JE, Feng Y, Velazquez H and Sessa WC: Endothelial glucocorticoid receptor is required for protection against sepsis. Proc Natl Acad Sci USA. 110:306–311. 2013. View Article : Google Scholar :

53 

Akalestou E, Genser L and Rutter GA: Glucocorticoid metabolism in obesity and following weight loss. Front Endocrinol (Lausanne). 11:592020. View Article : Google Scholar

54 

Kuo T, McQueen A, Chen TC and Wang JC: Regulation of glucose homeostasis by glucocorticoids. Adv Exp Med Biol. 872:99–126. 2015. View Article : Google Scholar : PubMed/NCBI

55 

Ferris HA and Kahn CR: New mechanisms of glucocorticoid-induced insulin resistance: Make no bones about it. J Clin Invest. 122:3854–3857. 2012. View Article : Google Scholar : PubMed/NCBI

56 

Vegiopoulos A and Herzig S: Glucocorticoids, metabolism and metabolic diseases. Mol Cell Endocrinol. 275:43–61. 2007. View Article : Google Scholar : PubMed/NCBI

57 

Madalena KM and Lerch JK: The effect of glucocorticoid and glucocorticoid receptor interactions on brain, spinal cord, and glial cell plasticity. Neural Plast. 2017:86409702017. View Article : Google Scholar : PubMed/NCBI

58 

Chen H, Lombès M and Le Menuet D: Glucocorticoid receptor represses brain-derived neurotrophic factor expression in neuron-like cells. Mol Brain. 10:122017. View Article : Google Scholar : PubMed/NCBI

59 

Myers B, McKlveen JM and Herman JP: Glucocorticoid actions on synapses, circuits, and behavior: Implications for the energetics of stress. Front Neuroendocrinol. 35:180–196. 2014. View Article : Google Scholar

60 

Joëls M: Corticosteroids and the brain. J Endocrinol. 238:R121–R130. 2018. View Article : Google Scholar : PubMed/NCBI

61 

Fietta P and Fietta P: Glucocorticoids and brain functions. Riv Biol. 100:403–418. 2007.

62 

McEwen BS and Akil H: Revisiting the stress concept: Implications for affective disorders. J Neurosci. 40:12–21. 2020. View Article : Google Scholar : PubMed/NCBI

63 

Smoller JW and Finn CT: Family, twin, and adoption studies of bipolar disorder. Am J Med Genet C Semin Med Genet. 123C:48–58. 2003. View Article : Google Scholar : PubMed/NCBI

64 

Geschwind DH and Flint J: Genetics and genomics of psychiatric disease. Science. 349:1489–1494. 2015. View Article : Google Scholar : PubMed/NCBI

65 

Akil H, Gordon J, Hen R, Javitch J, Mayberg H, McEwen B, Meaney MJ and Nestler EJ: Treatment resistant depression: A multi-scale, systems biology approach. Neurosci Biobehav Rev. 84:272–288. 2018. View Article : Google Scholar

66 

Wray NR, Ripke S, Mattheisen M, Trzaskowski M, Byrne EM, Abdellaoui A, Adams MJ, Agerbo E, Air TM, Andlauer TMF, et al: Genome-wide association analyses identify 44 risk variants and refine the genetic architecture of major depression. Nat Genet. 50:668–681. 2018. View Article : Google Scholar : PubMed/NCBI

67 

Aurbach EL, Inui EG, Turner CA, Hagenauer MH, Prater KE, Li JZ, Absher D, Shah N, Blandino P Jr, Bunney WE, et al: Fibroblast growth factor 9 is a novel modulator of negative affect. Proc Natl Acad Sci USA. 112:11953–11958. 2015. View Article : Google Scholar : PubMed/NCBI

68 

Salmaso N, Stevens HE, McNeill J, ElSayed M, Ren Q, Maragnoli ME, Schwartz ML, Tomasi S, Sapolsky RM, Duman R and Vaccarino FM: Fibroblast growth factor 2 modulates hypothalamic pituitary axis activity and anxiety behavior through glucocorticoid receptors. Biol Psychiatry. 80:479–489. 2016. View Article : Google Scholar : PubMed/NCBI

69 

Chaudhury S, Aurbach EL, Sharma V, Blandino P Jr, Turner CA, Watson SJ and Akil H: FGF2 is a target and a trigger of epigenetic mechanisms associated with differences in emotionality: Partnership with H3K9me3. Proc Natl Acad Sci USA. 111:11834–11839. 2014. View Article : Google Scholar : PubMed/NCBI

70 

Tyrka AR, Parade SH, Eslinger NM, Marsit CJ, Lesseur C, Armstrong DA, Philip NS, Josefson B and Seifer R: Methylation of exons 1D, 1F, and 1H of the glucocorticoid receptor gene promoter and exposure to adversity in preschool-aged children. Dev Psychopathol. 27:577–585. 2015. View Article : Google Scholar : PubMed/NCBI

71 

Sinclair D, Fillman SG, Webster MJ and Weickert CS: Dysregulation of glucocorticoid receptor co-factors FKBP5, BAG1 and PTGES3 in prefrontal cortex in psychotic illness. Sci Rep. 3:35392013. View Article : Google Scholar : PubMed/NCBI

72 

Sinclair D, Tsai SY, Woon HG and Weickert CS: Abnormal glucocorticoid receptor mRNA and protein isoform expression in the prefrontal cortex in psychiatric illness. Neuropsychopharmacology. 36:2698–2709. 2011. View Article : Google Scholar : PubMed/NCBI

73 

Sinclair D, Webster MJ, Fullerton JM and Weickert CS: Glucocorticoid receptor mRNA and protein isoform alterations in the orbitofrontal cortex in schizophrenia and bipolar disorder. BMC Psychiatry. 12:842012. View Article : Google Scholar : PubMed/NCBI

74 

Kloosterman WP and Plasterk RH: The diverse functions of microRNAs in animal development and disease. Dev Cell. 11:441–450. 2006. View Article : Google Scholar : PubMed/NCBI

75 

Kozomara A and Griffiths-Jones S: miRBase: Integrating microRNA annotation and deep-sequencing data. Nucleic Acids Res. 39:D152–D157. 2011. View Article : Google Scholar :

76 

Griffiths-Jones S: miRBase: MicroRNA Sequences and Annotation. Curr Protoc Bioinformatics. Chapter 12: Unit 12.9.1-10. 2010. View Article : Google Scholar : PubMed/NCBI

77 

Griffiths-Jones S, Saini HK, van Dongen S and Enright AJ: miRBase: Tools for microRNA genomics. Nucleic Acids Res. 36:D154–D158. 2008. View Article : Google Scholar :

78 

Griffiths-Jones S: miRBase: The microRNA sequence database. Methods Mol Biol. 342:129–138. 2006.PubMed/NCBI

79 

Griffiths-Jones S, Grocock RJ, van Dongen S, Bateman A and Enright AJ: miRBase: MicroRNA sequences, targets and gene nomenclature. Nucleic Acids Res. 34:D140–D144. 2006. View Article : Google Scholar :

80 

Griffiths-Jones S: The microRNA Registry. Nucleic Acids Res. 32:D109–D111. 2004. View Article : Google Scholar :

81 

Kozomara A, Birgaoanu M and Griffiths-Jones S: miRBase: From microRNA sequences to function. Nucleic Acids Res. 47:D155–D162. 2018. View Article : Google Scholar : PubMed/NCBI

82 

Hollins SL and Cairns MJ: MicroRNA: Small RNA mediators of the brains genomic response to environmental stress. Prog Neurobiol. 143:61–81. 2016. View Article : Google Scholar : PubMed/NCBI

83 

de Kloet ER, Fitzsimons CP, Datson NA, Meijer OC and Vreugdenhil E: Glucocorticoid signaling and stress-related limbic susceptibility pathway: About receptors, transcription machinery and microRNA. Brain Res. 1293:129–141. 2009. View Article : Google Scholar : PubMed/NCBI

84 

Moisiadis VG and Matthews SG: Glucocorticoids and fetal programming part 2: Mechanisms. Nat Rev Endocrinol. 10:403–411. 2014. View Article : Google Scholar : PubMed/NCBI

85 

Pufall MA: Glucocorticoids and cancer. Adv Exp Med Biol. 872:315–333. 2015. View Article : Google Scholar : PubMed/NCBI

86 

Peng Y and Croce CM: The role of MicroRNAs in human cancer. Signal Transduct Target Ther. 1:150042016. View Article : Google Scholar : PubMed/NCBI

87 

Agarwal V, Bell GW, Nam JW and Bartel DP: Predicting effective microRNA target sites in mammalian mRNAs. Elife. 4:e050052015. View Article : Google Scholar :

88 

Chen Y and Wang X: miRDB: An online database for prediction of functional microRNA targets. Nucleic Acids Res. 48:D127–D131. 2020. View Article : Google Scholar :

89 

Liu W and Wang X: Prediction of functional microRNA targets by integrative modeling of microRNA binding and target expression data. Genome Biol. 20:182019. View Article : Google Scholar : PubMed/NCBI

90 

Ledderose C, Möhnle P, Limbeck E, Schütz S, Weis F, Rink J, Briegel J and Kreth S: Corticosteroid resistance in sepsis is influenced by microRNA-124-induced downregulation of glucocorticoid receptor-α. Crit Care Med. 40:2745–2753. 2012. View Article : Google Scholar : PubMed/NCBI

91 

Wang SS, Mu RH, Li CF, Dong SQ, Geng D, Liu Q and Yi LT: microRNA-124 targets glucocorticoid receptor and is involved in depression-like behaviors. Prog Neuropsychopharmacol Biol Psychiatry. 79:417–425. 2017. View Article : Google Scholar : PubMed/NCBI

92 

Roy B, Dunbar M, Shelton RC and Dwivedi Y: Identification of MicroRNA-124-3p as a putative epigenetic signature of major depressive disorder. Neuropsychopharmacology. 42:864–875. 2017. View Article : Google Scholar :

93 

Dwivedi Y: microRNA-124: A putative therapeutic target and biomarker for major depression. Expert Opin Ther Targets. 21:653–656. 2017. View Article : Google Scholar : PubMed/NCBI

94 

Vreugdenhil E, Verissimo CS, Mariman R, Kamphorst JT, Barbosa JS, Zweers T, Champagne DL, Schouten T, Meijer OC, de Kloet ER and Fitzsimons CP: MicroRNA 18 and 124a down-regulate the glucocorticoid receptor: Implications for glucocorticoid responsiveness in the brain. Endocrinology. 150:2220–2228. 2009. View Article : Google Scholar : PubMed/NCBI

95 

Uchida S, Nishida A, Hara K, Kamemoto T, Suetsugi M, Fujimoto M, Watanuki T, Wakabayashi Y, Otsuki K, McEwen BS and Watanabe Y: Characterization of the vulnerability to repeated stress in Fischer 344 rats: Possible involvement of microRNA-mediated down-regulation of the glucocorticoid receptor. Eur J Neurosci. 27:2250–2261. 2008. View Article : Google Scholar : PubMed/NCBI

96 

Vallès A, Martens GJ, De Weerd P, Poelmans G and Aschrafi A: MicroRNA-137 regulates a glucocorticoid receptor-dependent signalling network: Implications for the etiology of schizophrenia. J Psychiatry Neurosc. 39:312–320. 2014. View Article : Google Scholar

97 

Li S, Ma H, Yuan X, Zhou X, Wan Y and Chen S: MicroRNA-382-5p targets nuclear receptor subfamily 3 group C member 1 to regulate depressive-like behaviors induced by chronic unpredictable mild stress in rats. Neuropsychiatr Dis Treat. 16:2053–2061. 2020. View Article : Google Scholar : PubMed/NCBI

98 

Kim J, Jeong D, Nam J, Aung TN, Gim JA, Park KU and Kim SW: MicroRNA-124 regulates glucocorticoid sensitivity by targeting phosphodiesterase 4B in diffuse large B cell lymphoma. Gene. 558:173–180. 2015. View Article : Google Scholar : PubMed/NCBI

99 

Liang YN, Tang YL, Ke ZY, Chen YQ, Luo XQ, Zhang H and Huang LB: MiR-124 contributes to glucocorticoid resistance in acute lymphoblastic leukemia by promoting proliferation, inhibiting apoptosis and targeting the glucocorticoid receptor. J Steroid Biochem Mol Biol. 172:62–68. 2017. View Article : Google Scholar : PubMed/NCBI

100 

Lv M, Zhang X, Jia H, Li D, Zhang B, Zhang H, Hong M, Jiang T, Jiang Q, Lu J, et al: An oncogenic role of miR-142-3p in human T-cell acute lymphoblastic leukemia (T-ALL) by targeting glucocorticoid receptor-alpha and cAMP/PKA pathways. Leukemia. 26:769–777. 2012. View Article : Google Scholar

101 

Riester A, Issler O, Spyroglou A, Rodrig SH, Chen A and Beuschlein F: ACTH-dependent regulation of microRNA as endogenous modulators of glucocorticoid receptor expression in the adrenal gland. Endocrinology. 153:212–222. 2012. View Article : Google Scholar

102 

Tessel MA, Benham AL, Krett NL, Rosen ST and Gunaratne PH: Role for microRNAs in regulating glucocorticoid response and resistance in multiple myeloma. Horm Cancer. 2:182–189. 2011. View Article : Google Scholar : PubMed/NCBI

103 

Sionov RV: MicroRNAs and glucocorticoid-induced apoptosis in lymphoid malignancies. ISRN Hematol. 2013:3482122013. View Article : Google Scholar : PubMed/NCBI

104 

Vigorito E, Kohlhaas S, Lu D and Leyland R: miR-155: An ancient regulator of the immune system. Immunol Rev. 253:146–157. 2013. View Article : Google Scholar : PubMed/NCBI

105 

Elmesmari A, Fraser AR, Wood C, Gilchrist D, Vaughan D, Stewart L, McSharry C, McInnes IB and Kurowska-Stolarska M: MicroRNA-155 regulates monocyte chemokine and chemokine receptor expression in Rheumatoid Arthritis. Rheumatology (Oxford). 55:2056–2065. 2016. View Article : Google Scholar

106 

Kurowska-Stolarska M, Alivernini S, Ballantine LE, Asquith DL, Millar NL, Gilchrist DS, Reilly J, Ierna M, Fraser AR, Stolarski B, et al: MicroRNA-155 as a proinflammatory regulator in clinical and experimental arthritis. Proc Natl Acad Sci USA. 108:11193–11198. 2011. View Article : Google Scholar : PubMed/NCBI

107 

Wang ZH, Liang YB, Tang H, Chen ZB, Li ZY, Hu XC and Ma ZF: Dexamethasone down-regulates the expression of microRNA-155 in the livers of septic mice. PLoS One. 8:e805472013. View Article : Google Scholar : PubMed/NCBI

108 

Zheng Y, Xiong S, Jiang P, Liu R, Liu X, Qian J, Zheng X and Chu Y: Glucocorticoids inhibit lipopolysaccharide-mediated inflammatory response by downregulating microRNA-155: A novel anti-inflammation mechanism. Free Radic Biol Med. 52:1307–1317. 2012. View Article : Google Scholar : PubMed/NCBI

109 

Curtale G, Renzi TA, Drufuca L, Rubino M and Locati M: Glucocorticoids downregulate TLR4 signaling activity via its direct targeting by miR-511-5p. Eur J Immunol. 47:2080–2089. 2017. View Article : Google Scholar : PubMed/NCBI

110 

Puimège L, Van Hauwermeiren F, Steeland S, Van Ryckeghem S, Vandewalle J, Lodens S, Dejager L, Vandevyver S, Staelens J, Timmermans S, et al: Glucocorticoid-induced microRNA-511 protects against TNF by down-regulating TNFR1. EMBO Mol Med. 7:1004–1017. 2015. View Article : Google Scholar : PubMed/NCBI

111 

Clayton SA, Jones SW, Kurowska-Stolarska M and Clark AR: The role of microRNAs in glucocorticoid action. J Biol Chemistry. 293:1865–1874. 2018. View Article : Google Scholar

112 

Davis TE, Kis-Toth K, Szanto A and Tsokos GC: Glucocorticoids suppress T cell function by up-regulating microRNA-98. Arthritis Rheum. 65:1882–1890. 2013. View Article : Google Scholar : PubMed/NCBI

113 

Zhu QY, Liu Q, Chen JX, Lan K and Ge BX: MicroRNA-101 targets MAPK phosphatase-1 to regulate the activation of MAPKs in macrophages. J Immunol. 185:7435–7442. 2010. View Article : Google Scholar : PubMed/NCBI

114 

Mogilyansky E and Rigoutsos I: The miR-17/92 cluster: A comprehensive update on its genomics, genetics, functions and increasingly important and numerous roles in health and disease. Cell Death Differ. 20:1603–1614. 2013. View Article : Google Scholar : PubMed/NCBI

115 

Molitor is JK, McColl KS and Distelhorst CW: Glucocorticoid-mediated repression of the oncogenic microRNA cluster miR-17~92 contributes to the induction of Bim and initiation of apoptosis. Mol Endocrinol. 25:409–420. 2011. View Article : Google Scholar

116 

Harada M, Pokrovskaja-Tamm K, Söderhäll S, Heyman M, Grander D and Corcoran M: Involvement of miR17 pathway in glucocorticoid-induced cell death in pediatric acute lymphoblastic leukemia. Leuk Lymphoma. 53:2041–2050. 2012. View Article : Google Scholar : PubMed/NCBI

117 

Palagani A, Op de Beeck K, Naulaerts S, Diddens J, Sekhar Chirumamilla C, Van Camp G, Laukens K, Heyninck K, Gerlo S, Mestdagh P, et al: Ectopic microRNA-150-5p transcription sensitizes glucocorticoid therapy response in MM1S multiple myeloma cells but fails to overcome hormone therapy resistance in MM1R cells. PLoS One. 9:e1138422014. View Article : Google Scholar : PubMed/NCBI

118 

Zhao JJ, Chu ZB, Hu Y, Lin J, Wang Z, Jiang M, Chen M, Wang X, Kang Y, Zhou Y, et al: Targeting the miR-221-222/PUMA/BAK/BAX pathway abrogates dexamethasone resistance in multiple myeloma. Cancer Res. 75:4384–4397. 2015. View Article : Google Scholar : PubMed/NCBI

119 

Murray MY, Rushworth SA, Zaitseva L, Bowles KM and Macewan DJ: Attenuation of dexamethasone-induced cell death in multiple myeloma is mediated by miR-125b expression. Cell Cycle. 12:2144–2153. 2013. View Article : Google Scholar : PubMed/NCBI

120 

Kotani A, Ha D, Hsieh J, Rao PK, Schotte D, den Boer ML, Armstrong SA and Lodish HF: miR-128b is a potent glucocorticoid sensitizer in MLL-AF4 acute lymphocytic leukemia cells and exerts cooperative effects with miR-221. Blood. 114:4169–4178. 2009. View Article : Google Scholar : PubMed/NCBI

121 

Cruz-Topete D, Oakley RH, Xu X and Cidlowski JA: Glucocorticoid receptor signaling is critical for microRNA Gender-specific regulation of gene expression in the adult mouse heart. FASEB J. 31:687.4. 2017.

122 

Jung SH, Wang Y, Kim T, Tarr A, Reader B, Powell N and Sheridan JF: Molecular mechanisms of repeated social defeat-induced glucocorticoid resistance: Role of microRNA. Brain Behav Immun. 44:195–206. 2015. View Article : Google Scholar

123 

Ko JY, Chuang PC, Ke HJ, Chen YS, Sun YC and Wang FS: MicroRNA-29a mitigates glucocorticoid induction of bone loss and fatty marrow by rescuing Runx2 acetylation. Bone. 81:80–88. 2015. View Article : Google Scholar : PubMed/NCBI

124 

Schroeder M, Jakovcevski M, Polacheck T, Drori Y, Luoni A, Röh S, Zaugg J, Ben-Dor S, Albrecht C and Chen A: Placental miR-340 mediates vulnerability to activity based anorexia in mice. Nat Commun. 9:15962018. View Article : Google Scholar : PubMed/NCBI

125 

Fu Q, Liu CJ, Zhang X, Zhai ZS, Wang YZ, Hu MX, Xu XL, Zhang HW and Qin T: Glucocorticoid receptor regulates expression of microRNA-22 and downstream signaling pathway in apoptosis of pancreatic acinar cells. World J Gastroenterol. 24:5120–5130. 2018. View Article : Google Scholar : PubMed/NCBI

126 

Buschmann D, González R, Kirchner B, Mazzone C, Pfaffl MW, Schelling G, Steinlein O and Reithmair M: Glucocorticoid receptor overexpression slightly shifts microRNA expression patterns in triple-negative breast cancer. Int J Oncol. 52:1765–1776. 2018.PubMed/NCBI

127 

Tejos-Bravo M, Oakley RH, Whirledge SD, Corrales WA, Silva JP, García-Rojo G, Toledo J, Sanchez W, Román-Albasini L, Aliaga E, et al: Deletion of hippocampal Glucocorticoid receptors unveils sex-biased microRNA expression and neuronal morphology alterations in mice. Neurobiol Stress. 14:1003062021. View Article : Google Scholar : PubMed/NCBI

128 

Zheng D, Sabbagh JJ, Blair LJ, Darling AL, Wen X and Dickey CA: MicroRNA-511 binds to FKBP5 mRNA, which encodes a chaperone protein, and regulates neuronal differentiation. J Biol Chemistry. 291:17897–17906. 2016. View Article : Google Scholar

129 

Pelleymounter LL, Moon I, Johnson JA, Laederach A, Halvorsen M, Eckloff B, Abo R and Rossetti S: A novel application of pattern recognition for accurate SNP and indel discovery from high-throughput data: Targeted resequencing of the glucocorticoid receptor co-chaperone FKBP5 in a Caucasian population. Mol Genet Metab. 104:457–469. 2011. View Article : Google Scholar : PubMed/NCBI

130 

Mercer TR, Dinger ME and Mattick JS: Long non-coding RNAs: Insights into functions. Nat Rev Genet. 10:155–159. 2009. View Article : Google Scholar : PubMed/NCBI

131 

Wilusz JE, Sunwoo H and Spector DL: Long noncoding RNAs: Functional surprises from the RNA world. Genes Dev. 23:1494–1504. 2009. View Article : Google Scholar : PubMed/NCBI

132 

Clemson CM, McNeil JA, Willard HF and Lawrence JB: XIST RNA paints the inactive X chromosome at interphase: Evidence for a novel RNA involved in nuclear/chromosome structure. J Cell Biol. 132:259–275. 1996. View Article : Google Scholar : PubMed/NCBI

133 

Swiezewski S, Liu F, Magusin A and Dean C: Cold-induced silencing by long antisense transcripts of an arabidopsis polycomb target. Nature. 462:799–802. 2009. View Article : Google Scholar : PubMed/NCBI

134 

Houseley J, Rubbi L, Grunstein M, Tollervey D and Vogelauer M: A ncRNA modulates histone modification and mRNA induction in the yeast GAL gene cluster. Mol Cell. 32:685–695. 2008. View Article : Google Scholar : PubMed/NCBI

135 

Reeves MB, Davies AA, McSharry BP, Wilkinson GW and Sinclair JH: Complex I binding by a virally encoded RNA regulates mitochondria-induced cell death. Science. 316:1345–1348. 2007. View Article : Google Scholar : PubMed/NCBI

136 

Clark MB and Mattick JS: Long noncoding RNAs in cell biology. Semin Cell Dev Biol. 22:366–376. 2011. View Article : Google Scholar : PubMed/NCBI

137 

Martianov I, Ramadass A, Serra Barros A, Chow N and Akoulitchev A: Repression of the human dihydrofolate reductase gene by a non-coding interfering transcript. Nature. 445:666–670. 2007. View Article : Google Scholar : PubMed/NCBI

138 

Tripathi V, Ellis JD, Shen Z, Song DY, Pan Q, Watt AT, Freier SM, Bennett CF, Sharma A, Bubulya PA, et al: The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Mol Cell. 39:925–938. 2010. View Article : Google Scholar : PubMed/NCBI

139 

Mourtada-Maarabouni M, Hedge VL, Kirkham L, Farzaneh F and Williams GT: Growth arrest in human T-cells is controlled by the non-coding RNA growth-arrest-specific transcript 5 (GAS5). J Cell Sci. 121:939–946. 2008. View Article : Google Scholar : PubMed/NCBI

140 

Huarte M, Guttman M, Feldser D, Garber M, Koziol MJ, Kenzelmann-Broz D, Khalil AM, Zuk O, Amit I, Rabani M, et al: A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell. 142:409–419. 2010. View Article : Google Scholar : PubMed/NCBI

141 

Place RF and Noonan EJ: Non-coding RNAs turn up the heat: An emerging layer of novel regulators in the mammalian heat shock response. Cell Stress Chaperones. 19:159–172. 2014. View Article : Google Scholar :

142 

Jarroux J, Morillon A and Pinskaya M: History, discovery, and classification of lncRNAs. Adv Exp Med Biol. 1008:1–46. 2017. View Article : Google Scholar : PubMed/NCBI

143 

Ma L, Bajic VB and Zhang Z: On the classification of long non-coding RNAs. RNA Biol. 10:925–933. 2013. View Article : Google Scholar : PubMed/NCBI

144 

Mayama T, Marr AK and Kino T: Differential expression of glucocorticoid receptor noncoding RNA repressor Gas5 in autoimmune and inflammatory diseases. Horm Metab Res. 48:550–557. 2016. View Article : Google Scholar : PubMed/NCBI

145 

Lucafò M, Di Silvestre A, Romano M, Avian A, Antonelli R, Martelossi S, Naviglio S, Tommasini A, Stocco G, Ventura A, Decorti G and De Iudicibus S: Role of the long non-coding RNA growth Arrest-Specific 5 in glucocorticoid response in children with inflammatory bowel disease. Basic Clin Pharmacol Toxicol. 122:87–93. 2018. View Article : Google Scholar

146 

Gharesouran J, Taheri M, Sayad A, Ghafouri-Fard S, Mazdeh M and Omrani MD: The growth arrest-specific Transcript 5 (GAS5) and nuclear receptor Subfamily 3 Group C Member 1 (NR3C1): Novel Markers involved in multiple sclerosis. Int J Mol Cell Med. 7:102–110. 2018.PubMed/NCBI

147 

Esguerra JLS, Ofori JK, Nagao M, Shuto Y, Karagiannopoulos A, Fadista J, Sugihara H, Groop L and Eliasson L: Glucocorticoid induces human beta cell dysfunction by involving riborepressor GAS5. LincRNA Mol Metab. 32:160–167. 2020. View Article : Google Scholar

148 

Ketab FNG, Gharesouran J, Ghafouri-Fard S, Dastar S, Mazraeh SA, Hosseinzadeh H, Moradi M, Javadlar M, Hiradfar A, Rezamand A, et al: Dual biomarkers long non-coding RNA GAS5 and its target, NR3C1, contribute to acute myeloid leukemia. Exp Mol Pathol. 114:1043992020. View Article : Google Scholar : PubMed/NCBI

149 

Pulido T, Adzerikho I, Channick RN, Delcroix M, Galiè N, Ghofrani HA, Jansa P, Jing ZC, Le Brun FO, Mehta S, et al: Macitentan and morbidity and mortality in pulmonary arterial hypertension. N Engl J Med. 369:809–818. 2013. View Article : Google Scholar : PubMed/NCBI

150 

Speed JS and Pollock DM: Endothelin, kidney disease, and hypertension. Hypertension. 61:1142–1145. 2013. View Article : Google Scholar : PubMed/NCBI

151 

Holm SJ, Sánchez F, Carlén LM, Mallbris L, Ståhle M and O'Brien KP: HLA-Cw*0602 associates more strongly to psoriasis in the Swedish population than variants of the novel 6p213 gene PSORS1C3. Acta Derm Venereol. 85:2–8. 2005. View Article : Google Scholar

152 

Robinson PC, Leo PJ, Pointon JJ, Harris J, Cremin K, Bradbury LA; Wellcome Trust Case Control Consortium; Australasian Osteoporosis Genetics Consortium (AOGC); Stebbings S, Harrison AA, et al: The genetic associations of acute anterior uveitis and their overlap with the genetics of ankylosing spondylitis. Genes Immun. 17:46–51. 2016. View Article : Google Scholar

153 

Murphy TM, Crawford B, Dempster EL, Hannon E, Burrage J, Turecki G, Kaminsky Z and Mill J: Methylomic profiling of cortex samples from completed suicide cases implicates a role for PSORS1C3 in major depression and suicide. Transl Psychiatry. 7:e9892017. View Article : Google Scholar : PubMed/NCBI

154 

Mirzadeh Azad F, Malakootian M and Mowla SJ: lncRNA PSORS1C3 is regulated by glucocorticoids and fine-tunes OCT4 expression in non-pluripotent cells. Sci Rep. 9:83702019. View Article : Google Scholar : PubMed/NCBI

155 

Redfern AD, Colley SM, Beveridge DJ, Ikeda N, Epis MR, Li X, Foulds CE, Stuart LM, Barker A, Russell VJ, et al: RNA-induced silencing complex (RISC) Proteins PACT, TRBP, and Dicer are SRA binding nuclear receptor coregulators. Proc Natl Acad Sci USA. 110:6536–6541. 2013. View Article : Google Scholar : PubMed/NCBI

156 

Hatchell EC, Colley SM, Beveridge DJ, Epis MR, Stuart LM, Giles KM, Redfern AD, Miles LE, Barker A, MacDonald LM, et al: SLIRP, a small SRA binding protein, is a nuclear receptor corepressor. Mol Cell. 22:657–668. 2006. View Article : Google Scholar : PubMed/NCBI

157 

Yang L, Lin C, Jin C, Yang JC, Tanasa B, Li W, Merkurjev D, Ohgi KA, Meng D, Zhang J, et al: lncRNA-dependent mechanisms of androgen-receptor-regulated gene activation programs. Nature. 500:598–602. 2013. View Article : Google Scholar : PubMed/NCBI

158 

Vitellius G, Trabado S, Bouligand J, Delemer B and Lombès M: Pathophysiology of Glucocorticoid Signaling. Ann Endocrinol (Paris). 79:98–106. 2018. View Article : Google Scholar

159 

Nicolaides NC, Geer EB, Vlachakis D, Roberts ML, Psarra AM, Moutsatsou P, Sertedaki A, Kossida S and Charmandari E: A novel mutation of the hGR gene causing Chrousos syndrome. Eur J Clin Invest. 45:782–791. 2015. View Article : Google Scholar : PubMed/NCBI

160 

Volden PA and Conzen SD: The influence of glucocorticoid signaling on tumor progression. Brain Behav Immun. 30(Suppl): S26–S31. 2013. View Article : Google Scholar

161 

Oakley RH and Cidlowski JA: Glucocorticoid signaling in the heart: A cardiomyocyte perspective. J Steroid Biochem Mol Biol. 153:27–34. 2015. View Article : Google Scholar : PubMed/NCBI

162 

Oakley RH and Cidlowski JA: The biology of the glucocorticoid receptor: New signaling mechanisms in health and disease. J Allergy Clin Immunol. 132:1033–1044. 2013. View Article : Google Scholar : PubMed/NCBI

163 

Zabirowicz ES and Gan TJ: 34-Pharmacology of Postoperative Nausea and Vomiting. Pharmacology and Physiology for Anesthesia (Second Edition). Hemmings HC and Egan TD: Elsevier; Philadelphia: pp. 671–692. 2019, View Article : Google Scholar

164 

Flor M: R interface to D3 chord diagrams. Chorddiag. https://github.com/mattflor/chorddiag. Accessed February 18, 2022.

Related Articles

Journal Cover

December-2022
Volume 50 Issue 6

Print ISSN: 1107-3756
Online ISSN:1791-244X

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Pierouli K, Papageorgiou L, Mitsis T, Papakonstantinou E, Diakou I, Leptidis S, Sigala M, Dragoumani K, Spandidos DA, Bacopoulou F, Bacopoulou F, et al: Role of microRNAs and long non‑coding RNAs in glucocorticoid signaling (Review). Int J Mol Med 50: 147, 2022
APA
Pierouli, K., Papageorgiou, L., Mitsis, T., Papakonstantinou, E., Diakou, I., Leptidis, S. ... Vlachakis, D. (2022). Role of microRNAs and long non‑coding RNAs in glucocorticoid signaling (Review). International Journal of Molecular Medicine, 50, 147. https://doi.org/10.3892/ijmm.2022.5203
MLA
Pierouli, K., Papageorgiou, L., Mitsis, T., Papakonstantinou, E., Diakou, I., Leptidis, S., Sigala, M., Dragoumani, K., Spandidos, D. A., Bacopoulou, F., Chrousos, G. P., Goulielmos, G. Ν., Eliopoulos, E., Vlachakis, D."Role of microRNAs and long non‑coding RNAs in glucocorticoid signaling (Review)". International Journal of Molecular Medicine 50.6 (2022): 147.
Chicago
Pierouli, K., Papageorgiou, L., Mitsis, T., Papakonstantinou, E., Diakou, I., Leptidis, S., Sigala, M., Dragoumani, K., Spandidos, D. A., Bacopoulou, F., Chrousos, G. P., Goulielmos, G. Ν., Eliopoulos, E., Vlachakis, D."Role of microRNAs and long non‑coding RNAs in glucocorticoid signaling (Review)". International Journal of Molecular Medicine 50, no. 6 (2022): 147. https://doi.org/10.3892/ijmm.2022.5203