Next Article in Journal
Immunomodulatory Properties and Osteogenic Activity of Polyetheretherketone Coated with Titanate Nanonetwork Structures
Next Article in Special Issue
CircRNAs as Potential Blood Biomarkers and Key Elements in Regulatory Networks in Gastric Cancer
Previous Article in Journal
Impacts of Commonly Used Edible Plants on the Modulation of Platelet Function
Previous Article in Special Issue
Gender-Dependent Deregulation of Linear and Circular RNA Variants of HOMER1 in the Entorhinal Cortex of Alzheimer’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The lncRNAs at X Chromosome Inactivation Center: Not Just a Matter of Sex Dosage Compensation

Department of Environmental, Biological, Pharmaceutical Sciences and Technologies, University of Campania “Luigi Vanvitelli”, 81100 Caserta, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(2), 611; https://doi.org/10.3390/ijms23020611
Submission received: 3 December 2021 / Revised: 30 December 2021 / Accepted: 5 January 2022 / Published: 6 January 2022
(This article belongs to the Special Issue RNA Regulatory Networks at the Crossroad of Human Diseases 2.0)

Abstract

:
Non-coding RNAs (ncRNAs) constitute the majority of the transcriptome, as the result of pervasive transcription of the mammalian genome. Different RNA species, such as lncRNAs, miRNAs, circRNA, mRNAs, engage in regulatory networks based on their reciprocal interactions, often in a competitive manner, in a way denominated “competing endogenous RNA (ceRNA) networks” (“ceRNET”): miRNAs and other ncRNAs modulate each other, since miRNAs can regulate the expression of lncRNAs, which in turn regulate miRNAs, titrating their availability and thus competing with the binding to other RNA targets. The unbalancing of any network component can derail the entire regulatory circuit acting as a driving force for human diseases, thus assigning “new” functions to “old” molecules. This is the case of XIST, the lncRNA characterized in the early 1990s and well known as the essential molecule for X chromosome inactivation in mammalian females, thus preventing an imbalance of X-linked gene expression between females and males. Currently, literature concerning XIST biology is becoming dominated by miRNA associations and they are also gaining prominence for other lncRNAs produced by the X-inactivation center. This review discusses the available literature to explore possible novel functions related to ceRNA activity of lncRNAs produced by the X-inactivation center, beyond their role in dosage compensation, with prospective implications for emerging gender-biased functions and pathological mechanisms.

1. Introduction

Non-coding RNAs (ncRNAs) constitute the largest class of the transcriptome, as the result of pervasive transcription of the mammalian genome, of which less than 2% encodes proteins [1]. Although the coding-independent functions of some ncRNAs, such as X-inactive transcript (XIST), were characterized in the early 1990s, the existence and biological relevance of the vast majority of ncRNAs were only gradually recognized a decade later, when advances in high-throughput sequencing technologies shed light on a plethora of RNA species, thus attracting tremendous interest in the field. It has become increasingly clear that ncRNAs are far from being “evolutionary junk”, and comprehension of their precious contribution to higher eukaryotes’ complexity is still at the very beginning.
Besides ribosomal RNA, ncRNA molecules can be broadly classified according to their size threshold: small or short ncRNAs, from a few to 200 nt, and long ncRNA (lncRNAs), longer than 200 nt, with a size up to several kilobases (up to 100 kb) [2].
Short ncRNAs comprise: transfer RNA (tRNA), involved in translation of mRNA; small nuclear RNA (snRNA) and small nucleolar RNA (snoRNA), involved in splicing and in ribosomal RNA modification, respectively; Piwi-interacting RNA (piRNA), involved in transposon repression; microRNA (miRNA), the most studied group of small ncRNAs, with a size of approximately 20 nt. miRNAs are post-transcriptional regulators of gene expression by binding to target transcripts, thus affecting their translation and/or stability [3]. Mammalian miRNAs are predicted to regulate up to 50% of all protein-coding genes, wherein each miRNA can bind different mRNAs and each mRNA can be targeted by various miRNAs, giving rise to complex regulatory networks that play key roles in almost all physiological pathways, but also in the pathogenesis of several diseases, especially cancer, where they can act as oncogenes (oncomiRs) or tumor suppressors [3,4,5].
lncRNAs constitute the largest class of ncRNAs in the mammalian genome and can be further classified into subclasses: long intergenic ncRNA (lincRNA), transcribed from intergenic regions, endowed with their own regulatory elements; sense lncRNAs, transcribed in the same direction as a coding gene, overlapping one or more exons or embedded in one of the introns (intronic lncRNA); antisense RNA (asRNA), transcribed as antisense strands compared to an overlapping known gene; pseudogenes, a version of coding genes that lost their protein-coding capacity due to mutations; circular RNA (circRNA), arising from backsplicing events of protein-coding transcripts that form covalently closed continuous loops; enhancer RNA (eRNA), deriving from enhancers endowed with enhancer-like functions [6].
lncRNAs and coding transcript biogenesis share various features: transcription, generally by RNA polymerase II, 5′-capping and 3′-polyadenylation modifications and also splicing. Additionally, miRNAs are transcribed by RNA polymerase II as long primary transcripts, and then subjected to a maturation compartmentalized between the nucleus and cytoplasm [3]. In contrast to mRNAs and miRNAs, largely operative in the cytoplasm, lncRNAs can be retained in the nucleus [7]. Unlike protein-coding transcripts and miRNAs sequences, lncRNAs are evolutionarily less conserved, however, conservation may be found in secondary structures that enable them to bind to proteins, DNA and other RNA molecules [8,9]. In the nucleus, lncRNAs can function as gene expression regulators by interacting with DNA, chromatin modifying complexes and/or various transcriptional regulators; in the cytoplasm, by binding proteins and RNA molecules, lncRNAs can still regulate gene expression at the post-transcriptional level by sponging miRNAs, regulating mRNA degradation and translation, and short open reading frames can even serve as templates for the synthesis of “micropeptides” [2,10]. Like miRNAs, lncRNA can also be secreted and found in the extracellular space [11]. Finally, lncRNA expression is finely regulated in physiological conditions, and their dysregulation contributes to pathogenesis of several diseases [7,12,13].
It is becoming increasingly clear that ncRNAs engage in regulatory networks based on interactions among the different RNA species, often in a competitive manner, in a way denominated “competing endogenous RNA (ceRNA) networks”, abbreviated as “ceRNET”; miRNAs and other ncRNAs modulate each other, since miRNAs can regulate the expression of lncRNAs, which in turn regulate miRNAs, titrating their availability and thus competing with the binding to other RNA targets [14,15,16]. In this scenario, coding transcripts themselves have a regulatory potential beyond their coding ability, when they compete for binding to shared miRNAs [16,17]. Indeed, transcripts from pseudogenes should also not be considered “evolutionary relics”, but can regulate expression of their related gene by competitively binding shared miRNAs [18]. In a physiological state, an optimal crosstalk occurs, since the shared pool of miRNAs is sufficient to target repression, whereas the unbalancing of any network component can derail the entire regulatory circuit acting as a driving force for human diseases. In this new perspective of RNA crosstalk, behind the conventional miRNA unidirectional regulation of target transcripts, different studies have highlighted additional roles of well-known lncRNAs as miRNA decoys with the effect of preventing their inhibitory binding to specific mRNA targets. Furthermore, it is also possible for lncRNAs and miRNAs with different preferential subcellular localizations to participate in the same ceRNA networks, even by shuttling between different compartments upon specific physiological or pathological conditions [16,19,20]. This is the case of XIST, the lncRNA well known for its role as a major and indispensable effector of X chromosome inactivation in mammalian females that prevents an imbalance of X-linked gene expression between females and males (see next paragraph); despite XIST’s preferential nuclear localization, and canonical miRNA function in the cytosol, many works revealed that XIST could also function as a ceRNA by sponging different miRNAs and thus de-repressing different protein-coding RNAs. Actually, literature concerning XIST biology is becoming dominated by miRNA associations and sponging activity [19], and this is also becoming true for other lncRNAs, also produced by the X-inactivation center (Xic) and involved in fine-tuning XIST expression. However, implications of those ceRNETs involving sex-specific lncRNAs in emerging gender-biased functions and pathological mechanisms are still poorly explored. With this perspective in mind, this review wants to explore possible new functions related to ceRNA activity of lncRNAs produced by the X-inactivation center, beyond their well-established role in the inactivation of the X chromosome of mammalian females.

2. Dosage Compensation and X-Inactivation Center

The presence of two X chromosomes in females and only one X chromosome in males requires mechanisms to equalize gene dosage between sexes and relative to the other chromosomes (autosomes), thus avoiding a potentially lethal double dose of genes localized on the X chromosome [21]. In mammals, dosage compensation between females and males is achieved through the process named X chromosome inactivation (XCI), i.e., the silencing of one of the two X chromosomes in females, a specialized form of heterochromatin formation established during early development and maintained through cell division and the adult body. This process, proposed for the first time by Lyon, ensures that the number of active X chromosomes in male and female cells is equalized with only one X chromosome being active (Xa) [21]. Different species employ different strategies for chromosome-wide inactivation with regard to the parental origin of the inactive X chromosome (Xi), being imprinted XCI in marsupials and monotremes, and random XCI in eutherians. The developmental timing and mode of chromosome selection for inactivation also varies. As an example, in mice, from the 8-cell to the morula stage, the paternal X is progressively inactivated, and then reactivated in the inner cell mass at the blastocyst stage, followed by random XCI, whereby both X chromosomes have the same probability of being inactivated and then epigenetically maintained; in humans, the two X chromosomes are active in the early cleavage stage, random XCI occurs at the morula stage and it persists in blastocysts in embryonic and extra-embryonic lineages [22]. However, a common feature of gene dosage compensation systems is their dependence on lncRNAs. In particular, in eutherians, the lncRNA XIST is the master regulator of the XCI process, whose genomic organization and function is widely conserved. XIST is transcribed by the X-inactivation center (XIC), then it coats and propagates along the future Xi, packaging it into transcriptional inactive heterochromatin through interaction with several protein partners [23].
Beyond the intricate molecular mechanisms underlying the XCI, it has also been shown that some genes escape X inactivation and are expressed from both the active and inactive X chromosome. Such genes are potential contributors to sexually dimorphic traits, to phenotypic variability among females heterozygous for X-linked conditions and to clinical abnormalities in patients with abnormal X chromosomes. Up to 15% of X-linked genes escape inactivation with large variability in their number and tissue distribution within a given individual and between individuals; the escape from silencing or skewed XCI allows the expression of some genes by both X chromosomes in females; in addition, skewed XCI may be also a consequence of early embryonic cell death [24,25,26].
The human XIC is the X-linked minimal genetic region that is necessary and sufficient to initiate XCI; it spans approximately a 1 Mb region in Xq13 [27]; it produces few coding transcripts, various uncharacterized inferred pseudogene transcripts and it is particularly enriched in genes producing lncRNAs, with some of them better characterized (see next paragraphs) and acting as activators or repressors engaged in a complex interplay regulating the monoallelic expression of XIST (Figure 1). Tridimensionally, the XIC is physically and spatially organized in two topologically associated domains (TADs): the TSIX TAD that includes a repressor of XCI, and the XIST TAD encompassing XCI activators. Despite the overall XIC gene synteny, order and orientation conservation, the human XIC locus is considerably expanded compared to the mouse ortholog region, and comprises different “actors’’ [28]; some differences can be observed in the organization and structure of the X-inactivation center, with a lesser degree of conservation for the TSIX TAD and dynamics of XIST expression during early embryogenesis. The generally poor conservation of the sequence of various lncRNAs implicated in a conserved process such as XCI suggests that rapidly evolving lncRNAs might favor adaptation to species-specific developmental/environmental constraints and acquire diversified functions. It is paradigmatic of the versatility of lncRNAs’ structure and function that an essential phenomenon may follow different routes, just sustained by ncRNAs.
Below there is a summary of the most well-characterized lncRNAs and their contribution to the mechanism of XCI.

2.1. XIST

XIST was one of the first long non-coding RNAs to be discovered in the early 1990s, first in humans and soon after in mice [23], being 17 kbp long in mice (Xist) and 19 kbp long in humans (XIST), transcribed by polymerase II, polyadenylated and retained in the nucleus [29]. Xist/XIST is the master regulator of XCI, transcribed exclusively from the (Xi), and then spreading along the X chromosome from which it was transcribed [30].
XIST is necessary and sufficient for starting the chromosome-wide gene silencing by cis-coating the future inactive X chromosome (Xi), mediating a cascade of epigenetic modifications and structural reshaping of the heterochromatin, culminating in the Xi becoming condensed in the so-called Barr body, which is maintained through multiple rounds of cell division [27,31,32]. Imaging studies revealed that X-linked genes, initially located at the periphery of the Xist RNA cloud, adopt a more internal position when silenced [33,34]. Exact mechanisms and factors whereby Xist initiates X inactivation are still under investigation; complex interactions occur between epigenetic and genomic features, such as genomic distance from the Xist locus, gene density and proximity to long interspersed nuclear elements (LINEs) that act as waystations to enhance Xist RNA coating throughout the Xi 3D space [35].
Despite the generally weak sequence conservation of Xist among eutherian mammalian, conservation is observed for the global gene structure and especially for the blocks of tandem repeats named A–F-repeats along its eight exons, cooperatively enabling Xist interaction with transcriptional factors, scaffold proteins and chromatin-modifying proteins, and thus mediating silencing and Xist localization on Xi [22,36,37,38,39]. In particular, the A-repeat, consisting of 7.5 copies of a 26 nt core sequence at the 5′ end of Xist RNA, is not involved in the coating mechanism, but in triggering gene silencing by recruiting the transcriptional repressor SPEN [40,41,42]; the A-repeat directly binds Polycomb repressive complex 2 (PRC2) components, thus mediating PRC2 recruitment to the Xi. The B-repeat, together with a short part of the C-repeat, is crucial for Xist spreading and for the recruitment of the Polycomb repressive complex 1 (PRC1), that could in turn recruit PRC2 [43,44,45]. The C-repeat, along with a ‘’nucleation center’’ for spreading across the Xi at the F-repeat, is involved in the interaction with the transcription factor Ying Yang 1 (YY1), that acts as an anchoring point to bridge between Xist RNA and the Xist gene [46] at an initial phase of X inactivation. Another interactor involved in Xist localization and X-linked gene silencing is the heterogeneous nuclear ribonucleoprotein U (hnRNP U), which interacts with Xist RNA by binding to exon 1 and 7 of both human and mouse XIST/Xist RNA [23] and it acts as a bridge between the matrix/scaffold-attached region in the Xi and Xist to facilitate spreading of the silencing machinery such as PRC2.
Of fundamental importance in the complex scenario of XCI is the recruitment of the Polycomb repressive complexes PRC1 and PRC2 by Xist RNA, indeed representing an important paradigm for chromatin regulation by long non-coding RNAs [47,48]. Then, other mechanisms are believed to act synergistically to maintain the inactive state, such as histone modifications, including loss of active histone marks (H3 and H4 acetylation, H3K4 methylation) and gain of repressive histone modifications (H3K9, H3K27 and H4K20 methylation), enrichment in the histone variant macroH2A and DNA methylation of the CpG island at the promoters of the X-linked genes [49].
A complex picture comes out from the abovementioned studies, whereby XIST acts as a trigger for Xi, and its structural versatility works as a scaffold for the recruitment of so many complexes to establish and maintain the inactive state.

2.2. JPX

Just proximal to XIST (JPX) is a long non-coding RNA transcribed from a gene within the X-inactivation center located ~10 kb upstream of XIST and expressed in the antisense direction. It lacks open reading frames but is relatively conserved in its 5′ exons. JPX was hinted as a key regulator in the positive arm of Xist regulation although the exact mechanisms of action need to be explored further [50]. The Jpx gene, also known as Enox, is less conserved than Ftx, the other positive regulator of Xist, showing a high identity with human JPX only at the level of one exon [28].
Experiments conducted on mouse embryonic stem cells (ESCs) revealed that Jpx escapes X chromosome inactivation, is upregulated during X-inactivation and it is required for the proper expression of Xist. Indeed, time-course measurements of both Jpx and Xist displayed increased Jpx RNA levels; this upregulation occurred in both XX and XY cells, however, whereas Xist induction followed Jpx regulation in female cells, Xist remained suppressed in male cells. Of note is the fact that the deletion of a single Jpx allele in female cells is sufficient to prevent X chromosome inactivation, thus reinforcing its role as an indispensable element in orchestrating the process. Moreover, through expressing Jpx from an autosomal transgene into ΔJpx/+ cells it was shown that Jpx levels were restored, implying that Jpx must therefore be able to act predominantly in trans, showing only a mild cis preference [50]. Based on overexpression experiments conducted on mouse ESCs, Sun and colleagues demonstrated the existence of a correlation between Jpx levels and Xist induction, indeed, during the different stages of differentiation, Xist showed a dose-dependent response to Jpx expression. One of the most interesting features of the regulation exerted by Jpx regards a sort of mechanism of molecular titration between Jpx and the RNA-binding protein CTCF that dictate Xist induction in opposite manners. Specifically, CTCF binds the Xist promoter P2 in mice and correlates with Xist repression. These findings were supported by the demonstration that CTCF overexpression blunted Xist upregulation and implied that CTCF is a blocking factor for Xist, acting in an opposite manner to Jpx. In addition, the fact that CTCF persisted at Xist P2 when Jpx was deficient hinted that Jpx and CTCF might be antagonistically linked. In the same study, in fact, new important evidence emerged: Jpx activates the Xist promoter by removing CTCF protein from the Xist promoter on the Xi [51]. A recent study developed a novel transgenic mouse system to demonstrate the regulatory mechanisms of lncRNA Jpx: the authors observed a dose-dependent relationship between Jpx copy number and Xist expression in mice, and that transgenic Jpx can activate the endogenous Xist in trans, suggesting that Jpx is sufficient to activate Xist expression in vivo; then, the authors proposed that Jpx is also able to act in cis [52], as previously found by others [53,54]. However, whether Jpx effectively acts in cis or in trans is still a controversy and needs to be further clarified.

2.3. FTX

Five prime to Xist (FTX) is another gene located upstream of XIST, within the XIC. It produces a long non-coding RNA that positively regulates the expression of XIST. Comparative analyses of ESTs revealed that the murine Ftx gene is composed of 15 exons spanning about 63 kbp, while human FTX comprises 12 exons spanning about 330 kbp and both mouse and human FTX genes give rise to various transcripts that originate from different combinations of alternative promoter usage and splicing. Strikingly, the Ftx gene harbors the conserved miRNA cluster miR-374b and 421 (miR-374 and miR-421 in mice) and only in humans is there a second cluster composed of miR-374a and miR-545, localized in intron b. Sequence analyses in humans revealed that miR-374a and miR-545 show similarities to miR-374b and miR-421, suggesting that they arise from duplication.
By means of mouse ES cells, it has been possible to demonstrate that Ftx is specifically upregulated during female cell differentiation at the onset of XCI, reminiscent of Xist expression, and the deletion of the promoter region of Ftx leads to local alterations in the chromatin structure. Strikingly, Xist is significantly downregulated in the absence of Ftx, suggesting that Ftx acts as an activator [55].
Additional evidence, regarding the involvement of Ftx for proper Xist expression dynamics and for efficient XCI progression, arose from a recent study in which several experiments at different levels were carried out in mutant mouse ESCs that were engineered by CRISPR-Cas/9, in order to delete a 9 kb region including the three putative promoters of Ftx. The data revealed that Ftx influences the efficiency of Xist activation and indicate that transcription from Ftx promoters is required for proper Xist expression in cis, independently of Ftx lncRNA transcripts produced from the locus or of Ftx-embedded miRs [56]. Moreover, previous observations that Ftx escapes XCI [55] and that it is required in cis for Xist accumulation were confirmed [56].
In this context, Ftx represents one important component of the delicate balance of activators and repressors in the multiplicity of layers operating in the organization of XCI and driving robust Xist upregulation.

2.4. TSIX

The Tsix gene expresses a non-coding transcript across the 3’ end and antisense to XIST, named “Tsix”, that is, Xist spelled in reverse order; conversely to Jpx and Ftx, it plays an antagonistic role in Xist expression. Tsix RNA is dynamically regulated during X inactivation: before X inactivation, Tsix is biallelically expressed but becomes monoallelically expressed at the onset of X inactivation, marking only the future active X and therefore raising the possibility that Tsix blocks Xist accumulation. By carrying out targeted deletion in female and male mouse cells of the 5′ CpG-rich domain of Tsix, it was demonstrated that Tsix encodes a single 40 kbp antisense transcript rather than multiple smaller RNAs and it is required for the random nature of X chromosome choice. Moreover, Tsix acts exclusively in cis and marks the future active X chromosome by blocking Xist accumulation [57]. Tsix was initially described as a 40 kb RNA encoded by a single exon, but later studies showed that it is, at least in part, subjected to splicing [58]. Even though Tsix plays an important role in mouse X inactivation, it is less conserved in humans at the sequence level [28].
Human TSIX produces a >30 kb transcript that is expressed only in cells of fetal origin; it is expressed from human XIC transgenes in mouse embryonic stem cells and from human embryoid-body-derived cells, but not from human adult somatic cells. Differences in the structure of human and murine genes suggest that human TSIX was truncated during evolution, and these differences could explain the fact that X inactivation is not imprinted in human placenta, thus raising questions about the role of TSIX in random X inactivation [59]. A later study from the same authors shed light on other interesting features regarding the differences between the two species; specifically using RNA FISH for cellular localization of transcripts in human fetal cells, they showed that human TSIX antisense transcripts were unable to repress XIST. Indeed, TSIX was transcribed only from the inactive X chromosome and was coexpressed with XIST, implying that the repression of Xist by mouse Tsix has no counterpart in humans, and TSIX is not the gene that protects the active X chromosome from random inactivation [60]. Regarding the role of Tsix in the chromatin modifications at the Xist locus, of note is the “dual effect” exerted by “opening” the chromatin structure along the Xist gene and “closing” it at the Xist promoter itself [61,62].
Overall, the interplay of different lncRNAs from the XIC secures the monoallelic Xist expression by acting as positive (FTX, JPX) and negative (TSIX) regulators and bringing about Xi in only one of the two X chromosomes in females.

3. ceRNA Activity of the lncRNAs from the XIC

3.1. ceRNETs Involving XIST

In recent years, literature concerning XIST biology has come to be dominated by miRNA associations. In particular, by searching “(XIST) AND (miRNA)” throughout PubMed, 257 results were retrieved that were published before 30 September 2021. A deep inspection of these articles led to the results presented in Table 1, selecting only those reporting miRNAs whose binding to XIST and the competitor transcripts was experimentally validated by RNA immunoprecipitation (RIP) and/or luciferase and/or RNA pull-down assays; then, the list of miRNAs and related mRNA targets was further processed by grouping similar “biological contexts” or similar “effects” or common components of the networks to envisage the potential overall framework. In addition, studies performed on mice and rats were also included, based on the consideration that despite some differences in nucleotide sequence, human lncRNA homologs may functionally compensate for the loss of mouse homologs [63].
The different studies mainly related the miRNA sponging activity of XIST to an oncogenic role in different types of cancer. In the case of colon cancer, the studies consistently point to an oncogenic role of XIST: it was upregulated in colon cancer tissues compared to the non-tumoral tissues and correlated to poor prognosis; it sponges miR-34a, thus preventing miRNA binding to WNT1, thus triggering the Wnt/beta-catenin pathway activation and then promoting proliferation and invasion of colon cancer cells [64]. A functionally similar axis involving XIST, miR-486-5p and neuropilin-2 has also been found to be involved in the epithelial–mesenchymal transition of colorectal cancer cells, thus participating in cancer progression [65]. Intriguingly, the XIST ceRNA activity is also involved in chemoresistance: as an example, XIST competes with serum and glucocorticoid-inducible kinase 1 (SGK1) by sponging the shared miR-124, and thus enhancing doxorubicin resistance [66]. Recently, complex lncRNA/miRNA/mRNA networks based on XIST/miR-500a-3p, miR-370-3p, miR-2467-3p, miR-512-3p/XBP have been demonstrated to promote cell proliferation and aggravated tumor growth in vivo by regulating endoplasmic reticulum stress response and cell apoptosis [67].
In the context of gastrointestinal tumors, XIST was also found to be overexpressed in gastric cancer tissues and associated with an aggressive tumor phenotype and adverse prognosis; XIST’s oncogenic power was demonstrated both in vitro and in vivo and at least partly attributed to the sponging activity of miR-101, resulting in the de-repression of the Polycomb group protein enhancer of zeste homolog 2 (EZH2) and promotion of cancer progression and metastasis [68]. Similarly, XIST upregulated Paxillin (PXN) expression by competitively binding to miR-132, and thus promoting carcinogenesis [69].
An oncogenic role of XIST mediated by its ceRNA activity has also been consistently indicated by different studies of non-small-cell lung cancer (NSCLC). In one pathway, XIST upregulation in NSLC tissues led to sequestering of miR-367 and miR-141, resulting in the de-repression of their target zinc finger E-box binding homeobox 1 (ZEB2), a transcriptional repressor of E-cadherin, and promoting cell invasion and metastasis; both in vitro and in vivo experiments link XIST to TGF-beta-induced epithelial–mesenchymal transition via miRNA crosstalk [70]. Different axes involving XIST/shared miRNA/mRNA target, and reported in Table 1, point to a role of XIST in cell proliferation, migration, invasion, EMT and chemoresistance of lung cancer, prospectively suggesting new therapeutic targets.
Similarly to ZEB2, ZEB1 from the same protein family has also been linked to an oncogenic role of XIST in another kind of tumor: in pancreatic cancer, XIST was frequently upregulated, especially in metastatic tissues, and sponges the tumor suppressor miR-429, thus acting as a ceRNA for ZEB1 that was upregulated; in particular, the axis XIST/miR-429/ZEB1 is critical for cell migration, invasion and EMT [71]. In the same pathological pathway, a role for the XIST/miR-141-5p/TGF-β2 axis has been reported [72].
Both ZEB1 and 2 have also been found to be involved in retinoblastoma, the most frequent eye malignancy in childhood, where XIST’s oncogenic role has been demonstrated in terms of promotion of proliferation, migration, invasion and EMT; again, XIST acts as a sponge for miR-101, thus resulting in the upregulation of its targets ZEB1 and ZEB2 [73]. Other works reported in Table 1 consistently indicate the ceRNA activity of XIST as a mechanism promoting retinoblastoma progression. In the context of pediatric tumors, in neuroblastoma an oncogenic role for XIST has also been determined and related to its ceRNA activity. In particular, XIST depletion repressed tumor growth in vivo and increased radiosensitivity, arrested cell cycle progression and impeded proliferation of neuroblastoma cells in vitro; mechanistically, XIST modulated L1 cell adhesion molecular (L1CAM) expression by competitively binding to miR-375 [74].
The oncogenic role of XIST has also been coherently reported by different papers studying another tumors of the nervous system: in glioma tissues and cell lines, XIST is significantly upregulated and related to poorer clinical and pathologic features and shorter survival time; its knockdown inhibits cell proliferation, migration and invasion in vitro and reduces tumor growth in vivo [75,76,77,78]. Again, the mechanism of action seems to be based on regulatory RNA networks involving XIST, miRNAs and their main targets. In one pathway, XIST participates in glioblastoma progression by enhancing glucose metabolism via sponging miR-126 and then preventing its binding to insulin receptor substrate 1 (IRS1), thus modulating the IRS1/PI3K/Akt pathway [75]. Functionally similar axes based on XIST/miR-133a/SOX4 and XIST/miR-329-3p/CREB1 networks have been found to be involved in promoting cell proliferation and metastasis [78,79].
The ceRNA activity of XIST versus miR-34a has an oncogenic effect in two different tumors, thyroid cancer and nasopharyngeal carcinoma, due to de-repression of different targets in different cell contexts, i.e., MET and E2F3, respectively [80,81].
With regard to female cancers, an oncogenic role has been attributed to XIST and associated with ceRNA activity. XIST was upregulated in cervical cancer tissues and cell lines and predicted unfavorable prognosis of patients. In vitro and in vivo experiments showed that it contributes to progression of cancer through different axes, i.e., miR-200a/Fus, miR-140-5p/ORC1, miR-889-3p/SIX1, that could be collectively explored for development of new therapeutic methods [82,83,84]. In ovarian cancer, XIST, miR-140-5p and the transcriptional factor FOXP3 engage a ceRNET, driving ovarian cancer cell progression: XIST was found upregulated in cancer tissues and correlated to poor prognosis of patients; conversely, miR-149-3p was found downregulated with the consequent upregulation of its target, FOXP3; XIST knock-down elevated miR-149-3p to suppress migration and invasion of ovarian cancer cells, and restricted tumor growth in vivo [85]. A similar regulatory network has been identified, whereby XIST regulated proliferation, invasion and migration via the miR-335/BCL2L2 axis [86].
An oncogenic role of XIST has been coherently indicated for other tumors, such as laryngeal squamous cell carcinoma, esophageal cancer and osteosarcoma (Table 1), whereas contrasting results have been published for hepatocellular carcinoma (HCC), where a tumor suppressor role has mainly been attributed. In particular, XIST could inhibit HCC cell proliferation and metastasis by sponging miR-92b, whose oncogenic role was demonstrated in vitro and in vivo and is at least partly due to the regulation of its target Smad7 [87]. Then, functionally similar networks were described involving XIST/miR-155-5p/SOX6, PTEN and XIST/miR-497-59/PDCD4; accordingly, XIST overexpression inhibited tumor growth in vivo [88,89]. These findings, together with the notion of a prevailing expression of XIST in females, may suggest an additional interpretation of the gender disparity of HCC occurrence (incidence rate 2–3 times higher in males than in females), to date related to sex hormones and cytokines [90,91].
Other phenomena linked to ceRNA activity of XIST are listed in Table 1. Among them, some papers accordingly indicate a proinflammatory role in different pathological pathways (Table 1) [92,93,94], including osteoarthritis [95,96]. XIST is upregulated in human osteoarthritis specimens and contributes to the pathogenesis by engaging different regulatory networks: it competitively binds to miR-376c-5p to increase the expression of osteopontin (OPN), exacerbating the inflammatory microenvironment and modulating the influence of proinflammatory M1 macrophages on chondrocyte apoptosis [95]; it binds to miR-211, thus increasing CXCR4, a major contributor to chondrocyte apoptosis [97]; it promotes extracellular matrix degradation by sponging miR-1277-5p, thus de-repressing matrix metalloproteinase 13 (MMP-13) and ADAM metallopeptidase with thrombospondin type 1 motif 5 (ADAMTS5) [98]; finally, two additional regulatory axes have been recently reported, miR-130/STAT3 and miR-149-5p/DNMT3A, where XIST was also confirmed to promote pathogenesis [96,99].
In the context of cardiovascular diseases, XIST was found upregulated in coronary heart disease tissues and post-myocardial infarction cells, where it suppressed cell proliferation and promoted apoptosis, at least partly by XIST/miR-130-3p/PDE4D [100]. Similar effects were observed in myocardial cell apoptosis in acute myocardial infarction model rats through XIST/miR-449/Notch1 [101]. Other works also point to a role of XIST in cardiovascular homeostasis, although in different model systems, and with some contrasting results, probably due to different disease models, and thus requiring further investigation [102,103].
The complex picture emerging from the data reported in Table 1 indicates that XIST is implicated in many pathways other than dosage compensation, especially in cancer, where it is often upregulated and acts as an oncogene by derailing the RNA networks involved in the control of cell proliferation, invasion, EMT and chemoresistance. However, much work is still required to envisage a common scenario, also complicated by the fact that a lncRNA may carry various binding sites for interaction with different miRNAs, and combination/coexpression of miRNAs and potential targets in different cell contexts may drive different, sometimes contrasting, effects. As an example, the interaction between XIST and miR-34a has been validated in different cell contexts, and implicated in nasopharyngeal carcinoma and colon and thyroid cancer, by networking with different mRNA targets, i.e., E2F3, WNT1 and MET, respectively [65,81,82], as shown in Table 1. Vice versa, the same mRNAs in combination with different XIST/miRNA axes can have relevance in different pathological contexts; as an example, the ZEB1/ZEB2 family has been implicated in NSCLC, pancreatic cancer, retinoblastoma and HCC in combination with different XIST/miRNA axes [71,72,74,104], as shown in Table 1.
Table 1. ceRNETs involving lncRNAs from XIC.
Table 1. ceRNETs involving lncRNAs from XIC.
Competing Endogenous RNAs
lncRNAsmRNAsShared miRNAsContextEffectRef.
XISTWNT1miR-34aColon cancerOncogenic role[64]
neuropilin-2miR-486-5p Oncogenic role[65]
XBP-1miR-500a-3p,miR-370-3p, miR-2467-3p, miR-512-3p Oncogenic role[67]
PAX5miR-338-3p Oncogenic role[105]
HIF-1AmiR-93-5p [106]
SGK1miR-124 Doxorubicin resistance[66]
ROR1miR-30a-5p Chemoresistance[107]
WEE1miR-125b-2-3p Oncogenic role and chemoresistance[108]
EZH2miR-101Gastric cancerOncogenic role[68]
PXNmiR-132 Oncogenic role[69]
JAK2miR-337 Oncogenic role[109]
ZEB2miR-367, miR-141 Oncogenic role[70]
Notch-1miR-137NSCLCOncogenic role[110]
SOD2miR-335 Oncogenic role[111]
RING1 miR-744 Oncogenic role[112]
PAX6miR-142-5p Oncogenic role[113]
ATG7miR-17 Cisplatin resistance[114]
MDM2miR-363-3pLUADOncogenic role[115]
ZEB1miR-429Pancreatic cancerOncogenic role[71]
TGF-β2 miR-141-3p Oncogenic role[72]
Notch1miR-137 Oncogenic role[116]
ZEB1, ZEB2miR-101RetinoblastomaOncogenic role[73]
STAT3 miR-124 Oncogenic role[117]
SOX4miR-140-5p Oncogenic role[118]
BDNFmiR-191-5p Oncogenic role[119]
STX17miR-361-3p Oncogenic role[120]
L1CAMmiR-375NeuroblastomaOncogenic role[74]
bFGF (FGF2)miR-424-5pPituitary neuroendocrine tumorOncogenic role[121]
IRS1miR-126GliomaOncogenic role[75]
SOX4miR-133a Oncogenic role[79]
CREB1miR-329-3p Oncogenic role[78]
ROCK1miR-448 Oncogenic role[122]
SP1, MGMTmiR-29c Temozolomide chemoresistance[77]
METmiR-34aThyroid cancerOncogenic role[80]
CLDN1miR-101-3p Oncogenic role[123]
E2F3miR-34a-5pNasopharyngeal carcinomaOncogenic role[81]
ADAM17miR-148a-3p Oncogenic role[124]
NEK5miR-381-3p Oncogenic role[125]
RECKmiR-30b Oncogenic role[126]
EZH2miR-124Laryngeal squamous cell carcinomaOncogenic role[127]
IRS1miR-144 Oncogenic role[128]
TRIB2miR-125b-5p Oncogenic role[129]
CDK6miR-494Esophageal cancerOncogenic role[130]
mTORmiR-375-3pOsteosarcomaOncogenic role[131]
RAB16miR-758 Oncogenic role[132]
ORC1miR-140-5pCervical cancerOncogenic role[83]
FusmiR-200a Oncogenic role[82]
SIX1miR-889-3p Oncogenic role[84]
FOXP3miR-149-3pOvarian cancerOncogenic role[85]
BCL2L2miR-335 Oncogenic role[86]
ANLNmiR-200c-3pBreast cancerDoxorubicin resistance[133]
GINS2miR-23a-3pMelanomaOncogenic role[134]
MYCmiR-29aAcute myeloid leukemiaOncogenic role[135]
Bcl-wmiR-497Extranodal natural killer/T-cell lymphomaOncogenic role[136]
Smad7miR-92bHCCTumor suppressor[87]
SOX6, PTENmiR-155-5p Tumor suppressor[88]
PDCD4miR-497-5p Tumor suppressor[89]
PDK1miR-139-5p Oncogenic role[137]
ZEB1/2miR-200b-3p Oncogenic role[104]
O-GlcNAc transferasemiR-424-5p Oncogenic role[138]
PIK3CAmiR-320a Oncogenic role[139]
P21miR-106b-5pRenal cell carcinomaTumor suppressor[140]
CUL3miR-15a-5pAcute kidney injuryPathogenesis promotion[141]
PDCD4miR-142-5pAcute kidney injuryPathogenesis promotion[142]
YAPmiR-194-5pWilms tumorOncogenic role[143]
CDKN1AmiR-93-5pDiabetic nephropathyPromotion of renal interstitial fibrosis[144]
TLR4miR-217Membranous nephropathyPromotion of podocyte apoptosis and disease development[145]
SOX-6miR-19bRenal fibrosisApoptosis and inflammation promotion[146]
NOD2miR-320AtherosclerosisPromotion of oxidative-LDL-induced cell injury[147]
TLR4miR-370-3pPneumoniaProinflammatory role in LPS-induced injury[92]
CCL16miR-30b-5pPneumoniaProinflammatory role in LPS-induced injury[148]
IRF2miR-204Respiratory distress syndrome (mice)Promotion of LPS-induced acute respiratory distress syndrome[149]
EGR3miR-200c-3pChronic obstructive pulmonary diseaseApoptosis and inflammation promotion[150]
IL-12AmiR-21Primary graft dysfunction in lung injuryInduction of neutrophil extracellular trap formation and dysfunction progression[151]
HMGB3miR-101-3pBronchopulmonar dysplasiaBP dysplasia promotion[152]
TLR5miR-154-5pNeuropathic pain development (rats)Neuropathic pain progression[153]
SIRT1miR-30d-5pDiabetesDiabetic peripheral neuropathy attenuation[154]
Nav1.7miR-146aSatellite glial cell activation and inflammatory pain (rats)Proinflammatory role[93]
STAT3let-7c-5pRheumatoid arthritis [155]
Smurf1miR-27aMicroglial cells (spinal cord injury, rats) [94]
NFAT5miR-29c-3pEpilepsy (rat model) [156]
NLRP3miRNA-223-3pRenal calculus (mouse model) [157]
OPNmiR-376c-5pOsteoarthritisPromotion of inflammatory microenvironment and chondrocyte apoptosis[95]
CXCR4miR-211 Promotion of chondrocyte apoptosis[96]
MMP-13, ADAMTS5 miR-1277-5p Promotion of extracellular matrix degradation[97]
DNMT3AmiR-149-5p Promotion of osteoarthritis[99]
STAT3miR-130a Promotion of inflammation and extracellular matrix degradation[112]
SIRT1miR-653-5p Protective role[158]
ZFPM2miR-203-3pFracture healingInterferes with proliferation and differentiation of osteoblasts[159]
AHNAKmiR-17-5pCervical ossification of the Posterior longitudinal ligamentPromotion of osteogenic differentiation[160]
PTENmiR-19Intervertebral disc degenerationAutophagy induction[161]
BACE1miR-124Alzheimer’s diseaseContribution to disease progression[162]
Sp1miR-199a-3pParkinson’s diseaseContribution to disease progression[163]
BACH1miR-98Cerebral injuryPromotion of neuronal injury[164]
TIPARPmiR-455-3p Promotion of neuronal injury[165]
FOXO3miR-27a-3p Promotion of neuronal injury[166]
IKKβmiR-96-5p Aggravation of neuronal apoptosis[167]
Itga5 or KLF4miR-92a Alleviation of cerebral vascular injury [168]
COL1A1 miR-29b-3pSkin fibroblasts (thermal injury)Promotion of extracellular matrix synthesis, proliferation and migration for wound healing[169]
HMGB1miR-29bHepatic stellate cells (alcoholic liver fibrogenesis)Enhancement ethanol-induced hepatic stellate cell autophagy and activation[170]
PDE4DmiR-130a-3pMyocardial infarctionPromotion of myocardial cell apoptosis and inhibition of cell proliferation[100]
Notch1miR-449Myocardial infarctionPromotes myocardial cell apoptosis[101]
S100BmiR-330-3pCardiomyocyte hypertrophyAntihypertrophy effect[102]
TLR2miR-101Cardiac hypertrophyPromotes the progression of cardiac hypertrophy[103]
FOXP2miR-122-5pHypoxia-induced H9c2 cardiomyocyte injuryAttenuates hypoxia-induced H9c2 cardiomyocyte injury[171]
c-FosmiR-150-5pSeptic myocardial injuryInduces pyroptosis[172]
PTENmiR-17Stanford type A aortic dissection (TAAD)Contribution to disease progression[173]
ELNmiR-29b-3p Thoracic aortic aneurysmAggravation of aortic smooth muscle cell apoptosis[174]
Arl2miR-214-3pAtrial fibrillationSuppression of myocardial pyroptosis[175]
FTXZEB2,HOXB9, NOB1, YY1miR-215Colorectal cancerOncogenic role[176]
RBPJmiR-590-5p Oncogenic role[177]
ZFXmiR-144Gastric cancerOncogenic role[178]
SIVA1miR-215 [179]
AEG-1miR-342-3pGliomaOncogenic role[180]
ALG3miR-342-3pDrug resistance in acute myeloid leukemiaDrug resistance[181]
c-MetmiR-186Bone marrow mesenchymal stem cellsOncogenic role[182]
WIF1,PTEN, WNT5Amir-374aHCCTumor suppressor[183]
FOXA2miR-200a-3pLung cancerTumor suppressor[100]
SOX7miR-21-5pEpileptiform hippocampal neurons (rat)Apoptosis inhibition[184]
Bcl2l2miR-29b-1-5pCardiomyocytes (mouse) Apoptosis inhibition[185]
Fmr1miR-410-3pMyocardial ischemia/reperfusion injuryAlleviation of hypoxia/reoxygenation-induced cardiomyocyte injury[186]
JPXNotch1miR-137Osteoclasts (osteoporosis)Osteogenic differentiation inhibition[187]
CCND2miR-145-5pLung cancerOncogenic role[188]
Twist1miR-33a-5p Oncogenic role[189]
CXCR6miR-197Gastric cancerOncogenic role[190]
CDH2miR-944Oral squamous cell carcinomaOncogenic role[191]
HIF-1alfamiR-18a-5pIntervertebral disc degeneration (human pulposus cells)Apoptosis inhibition[192]

3.2. ceRNETs Involving FTX

At the time of the preparation of this review, 27 results were retrieved by searching “(FTX) AND (miRNA)” throughout PubMed, and studies matching the criteria reported above were included in Table 1.
The FTX locus produces the lncRNA FTX and multiple intronic miRNAs (Figure 1). It was demonstrated that they are upregulated in colon cancer and cooperatively promote tumor progression. In particular, FTX interacted with DHX9 and Dicer and regulated A-to-I RNA editing and miRNA expression; miR-374b and miR-545 repressed tumor suppressors PTEN and RIG-I to increase proto-oncogenic PI3K-AKT signaling; miR-421 may have an autoregulatory effect on miR-374b and miR-545 [193]. The FTX oncogenic activity in colorectal cancer progression is also mediated by ceRNA activity: FTX binds to miR-215, thus preventing the inhibition of the miRNA targets ZEB2, HOXB9, NOB1 and YY1 related to cell proliferation, migration and invasion; consistently, FTX knock-down suppresses cell proliferation, migration and invasion, and inhibits growth and metastasis in vivo [176]. These results were confirmed by another study performed on colorectal cancer cells, showing the relevance of an additional functional axis, FTX/miR-590-5p/Recombination signal binding protein for the immunoglobulin kappa J region (RBPJ) [177].
Additionally, in gastric cancer FTX exerts a similar functional role via the miR-144/ZFX and miR-215-3p/SIVA1 regulatory axis in promoting tumorigenesis [178,179]; in addition, FTX was upregulated in tumor tissues in comparison to adjacent non-tumor tissues and correlated to poor prognosis [179].
FTX is also endowed with an oncogenic role in glioma, where it is upregulated and promotes cell proliferation and invasion by binding miR-342-3p, resulting in an increased expression of its target AEG-1 [180]. The interaction between FTX and miR-342 was also validated in acute myeloid leukemia cells, resulting in the upregulation of another miRNA target, the ALG3 mannosyltransferase, thus contributing to drug resistance in acute myeloid leukemia.
In contrast with reported role of FTX as a promoter of oncogenesis in different tumors (Table 1), in lung and HCC FTX can work as a tumor suppressor, again via ceRNA activity [183,194]. In particular, in HCC FTX inhibits proliferation and metastasis both in vitro and in vivo by repressing the silencing activity of miR-374a on its targets, i.e., WIF1, PTEN and WNT5A, as negative regulators of the WNT/beta-catenin signaling cascade, that indeed were inhibited, thus promoting cell epithelial–mesenchymal transition and invasion. FTX is expressed at a higher level in female livers compared with male livers; it was found downregulated in HCC tissues, but still maintained at higher levels in females compared with males [183]. It has been suggested that the tumor suppressor role and the reported expression pattern of FTX may contribute to HCC gender disparity, to date attributed solely to sex hormones: males are more susceptible than females to HCC, with average ratios between 2:1 and 4:1 [195,196]; in addition, male HCC patients suffer even worse prognoses than females patients and have shorter survival. Although the notion that estrogen and the estrogen receptor protect women from HCC, thus explaining the rising morbidity in post-menopausal women, and conversely that androgen and androgen receptors confer risk, the role of FTX in HCC, also related to HCC prognosis, and its higher expression in the female liver may contribute to cancer gender difference. Of note, a similar contribution to the HCC gender disparity has been hypothesized for XIST (see above), that is positively regulated by FTX, with an emerging picture that may warrant further investigating, and it may also have implications in gender medicine. Similar considerations could also be applied for glioma and colon and gastric cancer where both XIST and its positive regulator FTX seem to have an oncogenic role (Table 1).

3.3. ceRNETs Involving JPX

Ten results were retrieved by searching “(JPX) AND (miRNA)” throughout PubMed; by applying the criteria indicated above, ceRNA activity of JPX was validated for those cases reported in Table 1. Although few works have been published in the field to date, making difficult to speculate/draw conclusions, they demonstrate JPX ceRNET activity, mainly related to cancer.
As XIST, its positive regulator JPX has an oncogenic role in lung cancer, at least partly due to ceRNA mechanisms. First, it is significantly upregulated in NSCLC tissues compared with adjacent normal tissue and in lung cancer metastatic tissues, also correlating to poor survival and malignant phenotypes (tumor size, lymph node metastasis and TNM stage); it promoted cell proliferation in vitro and facilitated tumor growth in a xenograft mouse model [188,189]. Then, it was demonstrated that JPX increased cyclin D2 (CCND2) expression by competitively interacting with miR-145-5p, thus promoting cell proliferation and migration and contributing to cancer progression [188]. Furthermore, JPX also upregulated the transcription factor Twist1 by competitively sponging miR-33a-5p, thus inducing epithelial–mesenchymal transition (EMT) and cancer cell invasion [189]. Both ceRNA networks ultimately contribute to regulating lung tumorigenesis and metastasis, indicating potential therapeutic targets and novel biomarkers.
An oncogenic potential has also been verified in oral squamous cell carcinoma, where JPX contributes to cell proliferation, migration and invasion via the miR-944/CDH2 axis [191].
Finally, similarly to that reported for XIST and the other positive regulator FTX, JPX has an oncogenic role in gastric cancer and again the underlying molecular mechanism is based on a ceRNA mechanism: JPX increased the chemokine receptor CXCR6 by sponging its inhibitory miR-197 and thus promoting gastric cancer progression [190].

4. Implications for X Chromosome Aneuploidy Syndromes

In the scenario of ceRNA activity, miRNA binding sites can be envisioned as “the letters of an RNA code” by which transcripts communicate and regulate their relative expression levels [14]. Indeed, a transcript from a pseudogene can also regulate expression of its related gene by competitively binding shared miRNAs [18]. Genetic events such as copy number variation (amplification or loss) could induce “coding-independent” effects by altering the levels of microRNAs available for silencing particular transcripts. In this view, it should be noted that the X chromosome has the highest density of miRNA sequences compared to the other chromosomes, an evolutionarily conserved mammalian feature that equips females with a larger miRNA regulatory machinery than males [197,198]. Different syndromes related to X chromosome aneuploidies have been reported, characterized by either gain or loss of the entire chromosome (aneuploidy), or parts of the chromosome (structural abnormalities, e.g., isochromosomes) [199]. In particular, the partial or complete lack of a second X chromosome causes Turner syndrome (TS), estimated 1 in 2000 females. Forty to fifty percent of TS women have the 45,X karyotype, whereas the remaining cases are represented by mosaicism (mainly 45,X/46,XX mosaicism), where 20% of cases have alterations of the X chromosome (isochromosome Xq or ring X chromosome) and 10–12% of cases have differing amounts of Y chromosome material. Phenotypic traits include short stature, ovarian insufficiency and infertility, cardiac malformations, autoimmune diseases, metabolic disorders, neurocognitive problems [200]. The presence of an additional X chromosome in males causes Klinefelter syndrome (KS), with an incidence of 1 in every 660 male births. KS is characterized by a 47,XXY karyotype in about 80–90% of affected men, whereas the remaining cases are represented by mosaicism (usually 47,XXY/46,XY), higher grade sexual chromosome aneuploidies (e.g., 48,XXYY; 48,XXXY; 49,XXXXY) or X chromosome structural abnormalities. Phenotypic traits include tall stature, hypergonadotropic hypogonadism, infertility, type 2 diabetes, autoimmune disorders, neurocognitive problems, some cardiovascular abnormalities, obesity [201]. An additional X chromosome in females causes triple X syndrome (47,XXX), with an incidence of 1 in 1000 live-born girls. Phenotypic traits include a lower IQ, an increased incidence of psychiatric, language and autism-like disorders and premature ovarian failure [202]. In those syndromes, no obvious genotype–phenotype relationship has been established to date, besides the importance of the SHOX gene situated in the pseudo-autosomal region and therefore expressed on all sex chromosomes because it does not undergo X inactivation, and which has been shown to impact height and other skeletal changes, especially seen in patients with Turner syndrome [203,204], and explains part of the short stature in Turner syndrome and increased height in conditions with supernumerary sex chromosomes [205]. Patients with sex chromosome aneuploidies carrying the same karyotype can exhibit quite different traits and a range of comorbidities, suggesting a role of epigenetic mechanisms behind sex chromosome aneuploidy [206,207]. The absence or excess of escape X-linked miRNAs and other RNA species could greatly contribute to the observed variability in clinical traits, and it may be speculated that some of the ceRNETs discussed above could also be involved. Intriguingly, an increased expression of the lncRNAs XIST, JPX and TSIX was observed by comparing X chromosome RNA expression of blood cells in TS, female controls and triple X syndrome [208], as also found by others [209]. Moreover, the differential exon usage observed in the same study for several genes may have a functional significance in terms of ceRNA activity, beyond the coding potential, since they carry different binding sites for miRNAs.
Exploring ceRNET hypotheses in the context of X aneuploidy syndromes could contribute to filling the gap between puzzling clinical data and underlying molecular mechanisms, beyond a simple karyotype, also disclosing new perspectives in terms of innovative diagnostics tools and development of more effective therapeutic/patient-management strategies. As an example, higher expression of XIST has been linked to breast cancer pathogenesis and chemoresistance (Table 1) [133,210] and here the very low expression of XIST could at least partly underlie the very low risk of breast cancer among TS women and, vice versa, the higher XIST expression in KS men involved in the increased breast cancer incidence [211,212], which to date has not been explained convincingly by other data.

5. Conclusions

It has become increasingly clear that RNA molecules, either coding or non-coding, perform a variety of functions previously unexplored and their crosstalk provides regulatory networks governing numerous biological processes and whose derailing has pathological consequences, such as cancer.
Many years after its discovery, XIST, and other lncRNAs from the XIC, unveil their regulatory potential beyond their role in dosage compensation (Figure 2). While intriguing and somehow pioneering, the reported studies clearly demonstrate the ceRNET activity of some lncRNAs from the XIC, often focusing on the interaction of a single lncRNA with one miRNA shared with one competing target transcript; however, the possible high number of miRNA binding sites on the >19 kbp mature RNA such as XIST suggests a more complex potential landscape of shared miRNAs and their targets. Probably, the definition of such large-scale regulatory networks across the transcriptome will represent the next challenge. Moreover, no data have yet been published for many pseudogenes from the XIC (Figure 1), often overlapping with the main transcripts; their investigation and possible competing interactions with shared miRNAs may provide interesting findings, not only for “basic research”, but also for clinical perspectives. In fact, growing evidence linking pathological conditions to lncRNAs and their regulatory network is emerging, especially in cancer development and progression, with important implications in diagnostics and therapeutics. Specifically, different studies report the correlation of XIST expression with prognosis such as in the case of colon and gastric cancer, glioma and cervical and ovarian cancer [65,66,67,68,69,76,77,78,79,83,84,85,86] or FTX expression and gastric cancer prognosis [179] or JPX expression and poor survival of lung cancer patients [188,189]; probably, combination of the expression values of more than one RNA molecule involved in the same regulatory axis could enhance the diagnostic power, e.g., making it possible to stratify the patients according tumor stage, favorable/unfavorable outcome and even therapy effectiveness. Finally, understanding the versatility of RNA biology has become a stepping stone for RNA therapeutics that has definitely demonstrated its power [13]. Various therapeutic strategies can be based on boosting the expression level of a “beneficial” miRNA or inhibiting a harmful one; of note, different studies demonstrated that knock-down of a lncRNA such as XIST [68,69,75,76,77,78,79,83,84,85,86], or FTX [176] and JPX [188,189], exerts tumor-suppressive functions in vitro and in vivo, validating new therapeutic targets. In the near future, some of the predicted RNA interactions among lncRNAs/miRNAs/mRNAs that have not been included here could be experimentally validated, especially by in vivo studies, allowing a comprehensive view and envisaging new therapeutic opportunities by manipulating the network. Finally, the ceRNET activity of lncRNAs from the XIC highlighted in this review may suggest the importance of also considering the gender context in non-coding RNA studies, to fill the gap between clinical data and the understanding of molecular mechanisms underlying some sex-biased diseases, such as different tumor susceptibility (e.g., HCC).
The research field involving large-scale RNA regulatory networks is still in its infancy, and requires studies able to integrate computational analyses and new experimental platforms to fulfill the promise of significant advancement in the molecular knowledge of biological systems, and implications in sex-specific genome regulation and gender medicine.

Author Contributions

Conceptualization, N.P., C.S., A.D.P., A.R., investigation and data curation; N.P., C.S., writing—original draft preparation; all authors, review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

Research of A.R. and N.P. was funded by Valere grants from University of Campania “L. Vanvitelli”. Research of N.P. is also funded by “Associazione Famiglie di Soggetti con Deficit dell’Ormone della Crescita ed altre Patologie Rare” (A.Fa.D.O.C.).

Acknowledgments

The authors thank Claus Højbjerg Gravholt (Aarhus University Hospital, Denmark) for critically reading the manuscript and useful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef] [PubMed]
  2. Beermann, J.; Piccoli, M.T.; Viereck, J.; Thum, T. Non-coding RNAs in Development and Disease: Background, Mechanisms, and Therapeutic Approaches. Physiol. Rev. 2016, 96, 1297–1325. [Google Scholar] [CrossRef] [Green Version]
  3. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  4. Di Leva, G.; Garofalo, M.; Croce, C.M. MicroRNAs in cancer. Annu. Rev. Pathol. 2014, 9, 287–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Dragomir, M.P.; Knutsen, E.; Calin, G.A. Classical and noncanonical functions of miRNAs in cancers. Trends Genet. 2021. [Google Scholar] [CrossRef]
  6. Alessio, E.; Bonadio, R.S.; Buson, L.; Chemello, F.; Cagnin, S. A Single Cell but Many Different Transcripts: A Journey into the World of Long Non-Coding RNAs. Int. J. Mol. Sci. 2020, 21, 302. [Google Scholar] [CrossRef] [Green Version]
  7. Derrien, T.; Johnson, R.; Bussotti, G.; Tanzer, A.; Djebali, S.; Tilgner, H.; Guernec, G.; Martin, D.; Merkel, A.; Knowles, D.G.; et al. The GENCODE v7 catalog of human long noncoding RNAs: Analysis of their gene structure, evolution, and expression. Genome Res. 2012, 22, 1775–1789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Cruz, J.A.; Westhof, E. The dynamic landscapes of RNA architecture. Cell 2009, 136, 604–609. [Google Scholar] [CrossRef]
  9. Novikova, I.V.; Hennelly, S.P.; Tung, C.S.; Sanbonmatsu, K.Y. Rise of the RNA machines: Exploring the structure of long non-coding RNAs. J. Mol. Biol. 2013, 425, 3731–3746. [Google Scholar] [CrossRef]
  10. Diederichs, S. The four dimensions of noncoding RNA conservation. Trends Genet. 2014, 30, 121–123. [Google Scholar] [CrossRef] [PubMed]
  11. Pathania, A.S.; Challagundla, K.B. Exosomal Long Non-coding RNAs: Emerging Players in the Tumor Microenvironment. Mol. Ther. Nucleic Acids. 2020, 23, 1371–1383. [Google Scholar] [CrossRef] [PubMed]
  12. Cabili, M.; Trapnell, C.; Goff, L.; Koziol, M.; Tazon-Vega, B.; Regev, A.; Rinn, J.L. Integrative annotation of human large intergenic noncoding RNAs reveals global properties and specific subclasses. Genes Dev. 2011, 25, 1915–1927. [Google Scholar] [CrossRef] [Green Version]
  13. Winkle, M.; El-Daly, S.M.; Fabbri, M.; Calin, G.A. Noncoding RNA therapeutics-challenges and potential solutions. Nat. Rev. Drug Discov. 2021, 20, 629–651. [Google Scholar] [CrossRef] [PubMed]
  14. Salmena, L.; Poliseno, L.; Tay, Y.; Kats, L.; Pandolfi, P.P. A ceRNA hypothesis: The Rosetta Stone of a hidden RNA language? Cell 2011, 146, 353–358. [Google Scholar] [CrossRef] [Green Version]
  15. Tay, Y.; Rinn, J.; Pandolfi, P.P. The multilayered complexity of ceRNA crosstalk and competition. Nature 2014, 505, 344–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Chan, J.J.; Tay, Y. Noncoding RNA:RNA Regulatory Networks in Cancer. Int. J. Mol. Sci. 2018, 9, 1310. [Google Scholar] [CrossRef] [Green Version]
  17. Di Palo, A.; Siniscalchi, C.; Mosca, N.; Russo, A.; Potenza, N. A Novel ceRNA Regulatory Network Involving the Long Non-Coding Antisense RNA SPACA6P-AS, miR-125a and its mRNA Targets in Hepatocarcinoma Cells. Int. J. Mol. Sci. 2020, 21, 5068. [Google Scholar] [CrossRef]
  18. Poliseno, L.; Salmena, L.; Zhang, J.; Carver, B.; Haveman, W.J.; Pandolfi, P.P. A coding-independent function of gene and pseudogene mRNAs regulates tumour biology. Nature 2010, 465, 1033–1038. [Google Scholar] [CrossRef] [Green Version]
  19. Marshall, E.A.; Stewart, G.L.; Sage, A.P.; Lam, W.L.; Brown, C.J. Beyond sequence homology: Cellular biology limits the potential of XIST to act as a miRNA sponge. PLoS ONE 2019, 14, e0221371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Leung, A.K.L. The Whereabouts of microRNA Actions: Cytoplasm and Beyond. Trends Cell Biol. 2015, 25, 601–610. [Google Scholar] [CrossRef] [Green Version]
  21. Lyon, M.F. Gene action in the X-chromosome of the mouse (Mus musculus L.). Nature 1961, 190, 372–373. [Google Scholar] [CrossRef]
  22. Furlan, G.; Rougeulle, C. Function and evolution of the long noncoding RNA circuitry orchestrating X-chromosome inactivation in mammals. Wiley Interdiscip. Rev. RNA 2016, 7, 702–722. [Google Scholar] [CrossRef] [PubMed]
  23. Loda, A.; Heard, E. Xist RNA in action: Past, present, and future. PLoS Genet. 2019, 15, e1008333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Carrel, L.; Willard, H.F. X-inactivation profile reveals extensive variability in X-linked gene expression in females. Nature 2005, 434, 400–404. [Google Scholar] [CrossRef]
  25. Brooks, W.H. X chromosome inactivation and autoimmunity. Clin. Rev. Allergy Immunol. 2010, 39, 20–29. [Google Scholar] [CrossRef] [PubMed]
  26. Deng, X.; Berletch, J.B.; Nguyen, D.K.; Disteche, C.M. X chromosome regulation: Diverse patterns in development, tissues and disease. Nat. Rev. Genet. 2014, 15, 367–378. [Google Scholar] [CrossRef]
  27. Brown, C.J.; Lafreniere, R.G.; Powers, V.E.; Sebastio, G.; Ballabio, A.; Pettigrew, A.L.; Ledbetter, D.H.; Levy, E.; Craig, I.W.; Willard, H.F. Localization of the X inactivation centre on the human X chromosome in Xq13. Nature 1991, 349, 82–84. [Google Scholar] [CrossRef]
  28. Chureau, C.; Prissette, M.; Bourdet, A.; Barbe, V.; Cattolico, L.; Jones, L.; Eggen, A.; Avner, P.; Duret, L. Comparative sequence analysis of the X-inactivation center region in mouse, human, and bovine. Genome Res. 2002, 12, 894–908. [Google Scholar] [CrossRef] [PubMed]
  29. Richard, J.L.; Ogawa, Y. Understanding the Complex Circuitry of lncRNAs at the X-inactivation Center and Its Implications in Disease Conditions. Curr. Top. Microbiol. Immunol. 2016, 394, 1–27. [Google Scholar] [CrossRef] [PubMed]
  30. Weakley, S.M.; Wang, H.; Yao, Q.; Chen, C. Expression and function of a large non-coding RNA gene XIST in human cancer. World J. Surg. 2011, 35, 1751–1756. [Google Scholar] [CrossRef] [Green Version]
  31. Borsani, G.; Tonlorenzi, R.; Simmler, M.C.; Dandolo, L.; Arnaud, D.; Capra, V.; Grompe, M.; Pizzuti, A.; Muzny, D.; Lawrence, C.; et al. Characterization of a murine gene expressed from the inactive X chromosome. Nature 1991, 351, 325–329. [Google Scholar] [CrossRef]
  32. Brockdorff, N.; Ashworth, A.; Kay, G.F.; Cooper, P.; Smith, S.; McCabe, V.M.; Norris, D.P.; Penny, G.D.; Patel, D.; Rastan, S. Conservation of position and exclusive expression of mouse Xist from the inactive X chromosome. Nature 1991, 351, 329–331. [Google Scholar] [CrossRef]
  33. Chaumeil, J.; Le Baccon, P.; Wutz, A.; Heard, E. A novel role for Xist RNA in the formation of a repressive nuclear compartment into which genes are recruited when silenced. Genes Dev. 2006, 20, 2223–2237. [Google Scholar] [CrossRef] [Green Version]
  34. Clemson, C.M.; Hall, L.L.; Byron, M.; McNeil, J.; Lawrence, J.B. The X chromosome is organized into a gene-rich outer rim and an internal core containing silenced nongenic sequences. Proc. Natl. Acad. Sci. USA 2006, 103, 7688–7693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. E Sousa, L.B.A.; Jonkers, I.; Syx, L.; Dunkel, I.; Chaumeil, J.; Picard, C.; Foret, B.; Chen, C.J.; Lis, J.T.; Heard, E.; et al. Kinetics of Xist-induced gene silencing can be predicted from combinations of epigenetic and genomic features. Genome Res. 2019, 29, 1087–1099. [Google Scholar] [CrossRef] [Green Version]
  36. Sado, T.; Brockdorff, N. Advances in understanding chromosome silencing by the long non-coding RNA Xist. Philos. Trans. R Soc. Lond. B Biol. Sci. 2013, 368, 20110325. [Google Scholar] [CrossRef] [Green Version]
  37. Brown, C.J.; Hendrich, B.D.; Rupert, J.L.; Lafrenière, R.G.; Xing, Y.; Lawrence, J.; Willard, H.F. The human XIST gene: Analysis of a 17 kb inactive X-specific RNA that contains conserved repeats and is highly localized within the nucleus. Cell 1992, 71, 527–542. [Google Scholar] [CrossRef]
  38. Brockdorff, N.; Ashworth, A.; Kay, G.F.; McCabe, V.M.; Norris, D.P.; Cooper, P.J.; Swift, S.; Rastan, S. The product of the mouse Xist gene is a 15 kb inactive X-specific transcript containing no conserved ORF and located in the nucleus. Cell 1992, 71, 515–526. [Google Scholar] [CrossRef]
  39. Nesterova, T.B.; Slobodyanyuk, S.Y.; Elisaphenko, E.A.; Shevchenko, A.I.; Johnston, C.; Pavlova, M.E.; Rogozin, I.B.; Kolesnikov, N.N.; Brockdorff, N.; Zakian, S.M. Characterization of the genomic Xist locus in rodents reveals conservation of overall gene structure and tandem repeats but rapid evolution of unique sequence. Genome Res. 2001, 11, 833–849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Wutz, A.; Rasmussen, T.P.; Jaenisch, R. Chromosomal silencing and localization are mediated by different domains of Xist RNA. Nat. Genet. 2002, 30, 167–174. [Google Scholar] [CrossRef] [PubMed]
  41. Zhao, J.; Sun, B.K.; Erwin, J.A.; Song, J.J.; Lee, J.T. Polycomb proteins targeted by a short repeat RNA to the mouse X chromosome. Science 2008, 322, 750–756. [Google Scholar] [CrossRef] [Green Version]
  42. Nesterova, T.B.; Wei, G.; Coker, H.; Pintacuda, G.; Bowness, J.S.; Zhang, T.; Almeida, M.; Bloechl, B.; Moindrot, B.; Carter, E.J.; et al. Systematic allelic analysis defines the interplay of key pathways in X chromosome inactivation. Nat. Commun. 2019, 10, 3129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Colognori, D.; Sunwoo, H.; Kriz, A.J.; Wang, C.-Y.; Lee, J.T. Xist Deletional Analysis Reveals an Interdependency between Xist RNA and Polycomb Complexes for Spreading along the Inactive X. Mol. Cell 2019, 74, 101–117.e10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Tavares, L.; Dimitrova, E.; Oxley, D.; Webster, J.; Poot, R.; Demmers, J.; Bezstarosti, K.; Taylor, S.; Ura, H.; Koide, H.; et al. RYBP-PRC1 complexes mediate H2A ubiquitylation at polycomb target sites independently of PRC2 and H3K27me3. Cell 2012, 148, 664–678, Erratum in Cell 2012, 149, 1647–1648. [Google Scholar] [CrossRef] [Green Version]
  45. Pintacuda, G.; Wei, G.; Roustan, C.; Kirmizitas, B.A.; Solcan, N.; Cerase, A.; Castello, A.; Mohammed, S.; Moindrot, B.; Nesterova, T.B.; et al. hnRNPK Recruits PCGF3/5-PRC1 to the Xist RNA B-Repeat to Establish Polycomb-Mediated Chromosomal Silencing. Mol. Cell 2017, 68, 955–969. [Google Scholar] [CrossRef] [Green Version]
  46. Jeon, Y.; Lee, J.T. YY1 tethers Xist RNA to the inactive X nucleation center. Cell 2011, 146, 119–133. [Google Scholar] [CrossRef] [Green Version]
  47. Almeida, M.; Pintacuda, G.; Masui, O.; Koseki, Y.; Gdula, M.; Cerase, A.; Brown, D.; Mould, A.; Innocent, C.; Nakayama, M.; et al. PCGF3/5-PRC1 initiates Polycomb recruitment in X chromosome inactivation. Science 2017, 356, 1081–1084. [Google Scholar] [CrossRef]
  48. Żylicz, J.J.; Bousard, A.; Žumer, K.; Dossin, F.; Mohammad, E.; da Rocha, S.T.; Schwalb, B.; Syx, L.; Dingli, F.; Loew, D.; et al. The Implication of Early Chromatin Changes in X Chromosome Inactivation. Cell 2019, 176, 182–197.e23. [Google Scholar] [CrossRef] [Green Version]
  49. Chow, J.; Heard, E. X inactivation and the complexities of silencing a sex chromosome. Curr. Opin. Cell Biol. 2009, 21, 359–366. [Google Scholar] [CrossRef] [PubMed]
  50. Tian, D.; Sun, S.; Lee, J.T. The long noncoding RNA, Jpx, is a molecular switch for X chromosome inactivation. Cell 2010, 143, 390–403. [Google Scholar] [CrossRef] [Green Version]
  51. Sun, S.; Del Rosario, B.C.; Szanto, A.; Ogawa, Y.; Jeon, Y.; Lee, J.T. Jpx RNA activates Xist by evicting CTCF. Cell 2013, 153, 1537–1551. [Google Scholar] [CrossRef] [Green Version]
  52. Carmona, S.; Lin, B.; Chou, T.; Arroyo, K.; Sun, S. LncRNA Jpx induces Xist expression in mice using both trans and cis mechanisms. PLoS Genet. 2018, 14, e1007378. [Google Scholar] [CrossRef] [PubMed]
  53. Jonkers, I.; Barakat, T.S.; Achame, E.M.; Monkhorst, K.; Kenter, A.; Rentmeester, E.; Grosveld, F.; Grootegoed, J.A.; Gribnau, J. RNF12 is an X-Encoded dose-dependent activator of X chromosome inactivation. Cell 2009, 139, 999–1011. [Google Scholar] [CrossRef] [Green Version]
  54. Barakat, T.S.; Loos, F.; van Staveren, S.; Myronova, E.; Ghazvini, M.; Grootegoed, J.A.; Gribnau, J. The trans-activator RNF12 and cis-acting elements effectuate X chromosome inactivation independent of X-pairing. Mol. Cell 2014, 53, 965–978. [Google Scholar] [CrossRef] [Green Version]
  55. Chureau, C.; Chantalat, S.; Romito, A.; Galvani, A.; Duret, L.; Avner, P.; Rougeulle, C. Ftx is a non-coding RNA which affects Xist expression and chromatin structure within the X-inactivation center region. Hum. Mol. Genet. 2011, 20, 705–718. [Google Scholar] [CrossRef] [Green Version]
  56. Furlan, G.; Gutierrez Hernandez, N.; Huret, C.; Galupa, R.; van Bemmel, J.G.; Romito, A.; Heard, E.; Morey, C.; Rougeulle, C. The Ftx Noncoding Locus Controls X Chromosome Inactivation Independently of Its RNA Products. Mol. Cell 2018, 70, 462–472.e8. [Google Scholar] [CrossRef] [Green Version]
  57. Lee, J.T.; Lu, N. Targeted mutagenesis of Tsix leads to nonrandom X inactivation. Cell 1999, 99, 47–57. [Google Scholar] [CrossRef] [Green Version]
  58. Sado, T.; Wang, Z.; Sasaki, H.; Li, E. Regulation of imprinted X-chromosome inactivation in mice by Tsix. Development 2001, 128, 1275–1286. [Google Scholar] [CrossRef]
  59. Migeon, B.R.; Chowdhury, A.K.; Dunston, J.A.; McIntosh, I. Identification of TSIX, encoding an RNA antisense to human XIST, reveals differences from its murine counterpart: Implications for X inactivation. Am. J. Hum. Genet. 2001, 69, 951–960. [Google Scholar] [CrossRef] [Green Version]
  60. Migeon, B.R.; Lee, C.H.; Chowdhury, A.K.; Carpenter, H. Species differences in TSIX/Tsix reveal the roles of these genes in X-chromosome inactivation. Am. J. Hum. Genet. 2002, 71, 286–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Navarro, P.; Page, D.R.; Avner, P.; Rougeulle, C. Tsix-mediated epigenetic switch of a CTCF-flanked region of the Xist promoter determines the Xist transcription program. Genes Dev. 2006, 20, 2787–2792. [Google Scholar] [CrossRef] [Green Version]
  62. Navarro, P.; Chantalat, S.; Foglio, M.; Chureau, C.; Vigneau, S.; Clerc, P.; Avner, P.; Rougeulle, C. A role for non-coding Tsix transcription in partitioning chromatin domains within the mouse X-inactivation centre. Epigenetics Chromatin. 2009, 2, 8. [Google Scholar] [CrossRef] [Green Version]
  63. Karner, H.; Webb, C.H.; Carmona, S.; Liu, Y.; Lin, B.; Erhard, M.; Chan, D.; Baldi, P.; Spitale, R.C.; Sun, S. Functional Conservation of LncRNA JPX Despite Sequence and Structural Divergence. J. Mol. Biol. 2020, 432, 283–300. [Google Scholar] [CrossRef]
  64. Sun, N.; Zhang, G.; Liu, Y. Long non-coding RNA XIST sponges miR-34a to promotes colon cancer progression via Wnt/β-catenin signaling pathway. Gene 2018, 665, 141–148. [Google Scholar] [CrossRef]
  65. Liu, A.; Liu, L.; Lu, H. LncRNA XIST facilitates proliferation and epithelial-mesenchymal transition of colorectal cancer cells through targeting miR-486-5p and promoting neuropilin-2. J. Cell Physiol. 2019, 234, 13747–13761. [Google Scholar] [CrossRef]
  66. Zhu, J.; Zhang, R.; Yang, D.; Li, J.; Yan, X.; Jin, K.; Li, W.; Liu, X.; Zhao, J.; Shang, W.; et al. Knockdown of Long Non-Coding RNA XIST Inhibited Doxorubicin Resistance in Colorectal Cancer by Upregulation of miR-124 and Downregulation of SGK1. Cell Physiol. Biochem. 2018, 51, 113–128. [Google Scholar] [CrossRef]
  67. Wang, Y.; Zhang, J.; Zheng, S. The role of XBP-1-mediated unfolded protein response in colorectal cancer progression-a regulatory mechanism associated with lncRNA-miRNA-mRNA network. Cancer Cell Int. 2021, 21, 488. [Google Scholar] [CrossRef] [PubMed]
  68. Chen, D.L.; Ju, H.Q.; Lu, Y.X.; Chen, L.Z.; Zeng, Z.L.; Zhang, D.S.; Luo, H.Y.; Wang, F.; Qiu, M.Z.; Wang, D.S.; et al. Long non-coding RNA XIST regulates gastric cancer progression by acting as a molecular sponge of miR-101 to modulate EZH2 expression. J. Exp. Clin. Cancer Res. 2016, 35, 142. [Google Scholar] [CrossRef] [Green Version]
  69. Li, P.; Wang, L.; Li, P.; Hu, F.; Cao, Y.; Tang, D.; Ye, G.; Li, H.; Wang, D. Silencing lncRNA XIST exhibits antiproliferative and proapoptotic effects on gastric cancer cells by up-regulating microRNA-132 and down-regulating PXN. Aging 2020, 12, 14469–14481. [Google Scholar] [CrossRef] [PubMed]
  70. Li, C.; Wan, L.; Liu, Z.; Xu, G.; Wang, S.; Su, Z.; Zhang, Y.; Zhang, C.; Liu, X.; Lei, Z.; et al. Long non-coding RNA XIST promotes TGF-β-induced epithelial-mesenchymal transition by regulating miR-367/141-ZEB2 axis in non-small-cell lung cancer. Cancer Lett. 2018, 418, 185–195. [Google Scholar] [CrossRef] [PubMed]
  71. Shen, J.; Hong, L.; Yu, D.; Cao, T.; Zhou, Z.; He, S. LncRNA XIST promotes pancreatic cancer migration, invasion and EMT by sponging miR-429 to modulate ZEB1 expression. Int. J. Biochem. Cell Biol. 2019, 113, 17–26. [Google Scholar] [CrossRef]
  72. Sun, J.; Zhang, Y. LncRNA XIST enhanced TGF-β2 expression by targeting miR-141-3p to promote pancreatic cancer cells invasion. Biosci. Rep. 2019, 39, BSR20190332. [Google Scholar] [CrossRef] [Green Version]
  73. Cheng, Y.; Chang, Q.; Zheng, B.; Xu, J.; Li, H.; Wang, R. LncRNA XIST promotes the epithelial to mesenchymal transition of retinoblastoma via sponging miR-101. Eur. J. Pharmacol. 2019, 843, 210–216. [Google Scholar] [CrossRef] [PubMed]
  74. Yang, H.; Zhang, X.; Zhao, Y.; Sun, G.; Zhang, J.; Gao, Y.; Liu, Q.; Zhang, W.; Zhu, H. Downregulation of lncRNA XIST Represses Tumor Growth and Boosts Radiosensitivity of Neuroblastoma via Modulation of the miR-375/L1CAM Axis. Neurochem. Res. 2020, 45, 2679–2690. [Google Scholar] [CrossRef] [PubMed]
  75. Cheng, Z.; Luo, C.; Guo, Z. LncRNA-XIST/microRNA-126 sponge mediates cell proliferation and glucose metabolism through the IRS1/PI3K/Akt pathway in glioma. J. Cell Biochem. 2020, 121, 2170–2183. [Google Scholar] [CrossRef]
  76. Yao, Y.; Ma, J.; Xue, Y.; Wang, P.; Li, Z.; Liu, J.; Chen, L.; Xi, Z.; Teng, H.; Wang, Z.; et al. Knockdown of long non-coding RNA XIST exerts tumor-suppressive functions in human glioblastoma stem cells by up-regulating miR-152. Cancer Lett. 2015, 359, 75–86. [Google Scholar] [CrossRef]
  77. Du, P.; Zhao, H.; Peng, R.; Liu, Q.; Yuan, J.; Peng, G.; Liao, Y. LncRNA-XIST interacts with miR-29c to modulate the chemoresistance of glioma cell to TMZ through DNA mismatch repair pathway. Biosci. Rep. 2017, 37, BSR20170696. [Google Scholar] [CrossRef] [Green Version]
  78. Wang, Y.P.; Li, H.Q.; Chen, J.X.; Kong, F.G.; Mo, Z.H.; Wang, J.Z.; Huang, K.M.; Li, X.N.; Yan, Y. Overexpression of XIST facilitates cell proliferation, invasion and suppresses cell apoptosis by reducing radio-sensitivity of glioma cells via miR-329-3p/CREB1 axis. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 3190–3203. [Google Scholar] [CrossRef]
  79. Luo, C.; Quan, Z.; Zhong, B.; Zhang, M.; Zhou, B.; Wang, S.; Luo, X.; Tang, C. lncRNA XIST promotes glioma proliferation and metastasis through miR-133a/SOX4. Exp. Ther. Med. 2020, 19, 1641–1648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Liu, H.; Deng, H.; Zhao, Y.; Li, C.; Liang, Y. LncRNA XIST/miR-34a axis modulates the cell proliferation and tumor growth of thyroid cancer through MET-PI3K-AKT signaling. J. Exp. Clin. Cancer Res. 2018, 37, 279. [Google Scholar] [CrossRef]
  81. Song, P.; Ye, L.F.; Zhang, C.; Peng, T.; Zhou, X.H. Long non-coding RNA XIST exerts oncogenic functions in human nasopharyngeal carcinoma by targeting miR-34a-5p. Gene 2016, 592, 8–14. [Google Scholar] [CrossRef]
  82. Zhu, H.; Zheng, T.; Yu, J.; Zhou, L.; Wang, L. LncRNA XIST accelerates cervical cancer progression via upregulating Fus through competitively binding with miR-200a. Biomed. Pharmacother. 2018, 105, 789–797. [Google Scholar] [CrossRef]
  83. Chen, X.; Xiong, D.; Ye, L.; Wang, K.; Huang, L.; Mei, S.; Wu, J.; Chen, S.; Lai, X.; Zheng, L.; et al. Up-regulated lncRNA XIST contributes to progression of cervical cancer via regulating miR-140-5p and ORC1. Cancer Cell Int. 2019, 19, 45. [Google Scholar] [CrossRef]
  84. Liu, X.; Xie, S.; Zhang, J.; Kang, Y. Long Noncoding RNA XIST Contributes to Cervical Cancer Development Through Targeting miR-889-3p/SIX1 Axis. Cancer Biother. Radiopharm. 2020, 35, 640–649. [Google Scholar] [CrossRef]
  85. Jiang, R.; Zhang, H.; Zhou, J.; Wang, J.; Xu, Y.; Zhang, H.; Gu, Y.; Fu, F.; Shen, Y.; Zhang, G.; et al. Inhibition of long non-coding RNA XIST upregulates microRNA-149-3p to repress ovarian cancer cell progression. Cell Death Dis. 2021, 12, 145. [Google Scholar] [CrossRef] [PubMed]
  86. Meng, Q.; Wang, N.; Duan, G. Long non-coding RNA XIST regulates ovarian cancer progression via modulating miR-335/BCL2L2 axis. World J. Surg. Oncol. 2021, 19, 165. [Google Scholar] [CrossRef] [PubMed]
  87. Zhuang, L.K.; Yang, Y.T.; Ma, X.; Han, B.; Wang, Z.S.; Zhao, Q.Y.; Wu, L.Q.; Qu, Z.Q. MicroRNA-92b promotes hepatocellular carcinoma progression by targeting Smad7 and is mediated by long non-coding RNA XIST. Cell Death Dis. 2016, 7, e2203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Lin, X.Q.; Huang, Z.M.; Chen, X.; Wu, F.; Wu, W. XIST Induced by JPX Suppresses Hepatocellular Carcinoma by Sponging miR-155-5p. Yonsei Med. J. 2018, 59, 816–826. [Google Scholar] [CrossRef] [PubMed]
  89. Zhang, Y.; Zhu, Z.; Huang, S.; Zhao, Q.; Huang, C.; Tang, Y.; Sun, C.; Zhang, Z.; Wang, L.; Chen, H.; et al. lncRNA XIST regulates proliferation and migration of hepatocellular carcinoma cells by acting as miR-497-5p molecular sponge and targeting PDCD4. Cancer Cell Int. 2019, 19, 198. [Google Scholar] [CrossRef]
  90. Guy, J.; Peters, M.G. Liver disease in women: The influence of gender on epidemiology, natural history, and patient outcomes. Gastroenterol. Hepatol. 2013, 9, 633–639. [Google Scholar]
  91. Ruggieri, A.; Barbati, C.; Malorni, W. Cellular and molecular mechanisms involved in hepatocellular carcinoma gender disparity. Int. J. Cancer 2010, 127, 499–504. [Google Scholar] [CrossRef]
  92. Zhang, Y.; Zhu, Y.; Gao, G.; Zhou, Z. Knockdown XIST alleviates LPS-induced WI-38 cell apoptosis and inflammation injury via targeting miR-370-3p/TLR4 in acute pneumonia. Cell Biochem. Funct. 2019, 37, 348–358. [Google Scholar] [CrossRef] [PubMed]
  93. Sun, W.; Ma, M.; Yu, H.; Yu, H. Inhibition of lncRNA X inactivate-specific transcript ameliorates inflammatory pain by suppressing satellite glial cell activation and inflammation by acting as a sponge of miR-146a to inhibit Nav 1.7. J. Cell Biochem. 2018, 119, 9888–9898. [Google Scholar] [CrossRef] [PubMed]
  94. Zhao, Q.; Lu, F.; Su, Q.; Liu, Z.; Xia, X.; Yan, Z.; Zhou, F.; Qin, R. Knockdown of long noncoding RNA XIST mitigates the apoptosis and inflammatory injury of microglia cells after spinal cord injury through miR-27a/Smurf1 axis. Neurosci. Lett. 2020, 715, 134649. [Google Scholar] [CrossRef]
  95. Li, L.; Lv, G.; Wang, B.; Kuang, L. XIST/miR-376c-5p/OPN axis modulates the influence of proinflammatory M1 macrophages on osteoarthritis chondrocyte apoptosis. J. Cell Physiol. 2020, 235, 281–293. [Google Scholar] [CrossRef]
  96. Xiao, Y.; Liu, L.; Zheng, Y.; Liu, W.; Xu., Y. Kaempferol attenuates the effects of XIST/miR-130a/STAT3 on inflammation and extracellular matrix degradation in osteoarthritis. Future Med. Chem. 2021, 13, 1451–1464. [Google Scholar] [CrossRef] [PubMed]
  97. Li, L.; Lv, G.; Wang, B.; Kuang, L. The role of lncRNA XIST/miR-211 axis in modulating the proliferation and apoptosis of osteoarthritis chondrocytes through CXCR4 and MAPK signaling. Biochem. Biophys. Res. Commun. 2018, 503, 2555–2562. [Google Scholar] [CrossRef]
  98. Wang, T.; Liu, Y.; Wang, Y.; Huang, X.; Zhao, W.; Zhao, Z. Long non-coding RNA XIST promotes extracellular matrix degradation by functioning as a competing endogenous RNA of miR-1277-5p in osteoarthritis. Int. J. Mol. Med. 2019, 44, 630–642. [Google Scholar] [CrossRef] [Green Version]
  99. Liu, Y.; Liu, K.; Tang, C.; Shi, Z.; Jing, K.; Zheng, J. Long non-coding RNA XIST contributes to osteoarthritis progression via miR-149-5p/DNMT3A axis. Biomed. Pharmacother. 2020, 128, 110349. [Google Scholar] [CrossRef]
  100. Zhou, T.; Qin, G.; Yang, L.; Xiang, D.; Li, S. LncRNA XIST regulates myocardial infarction by targeting miR-130a-3p. J. Cell Physiol. 2019, 234, 8659–8667. [Google Scholar] [CrossRef]
  101. Zhang, M.; Liu, H.Y.; Han, Y.L.; Wang, L.; Zhai, D.D.; Ma, T.; Zhang, M.J.; Liang, C.Z.; Shen, Y. Silence of lncRNA XIST represses myocardial cell apoptosis in rats with acute myocardial infarction through regulating miR-449. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 8566–8572. [Google Scholar] [CrossRef]
  102. Chen, Y.; Liu, X.; Chen, L.; Chen, W.; Zhang, Y.; Chen, J.; Wu, X.; Zhao, Y.; Wu, X.; Sun, G. The long noncoding RNA XIST protects cardiomyocyte hypertrophy by targeting miR-330-3p. Biochem. Biophys. Res. Commun. 2018, 505, 807–815. [Google Scholar] [CrossRef]
  103. Xiao, L.; Gu, Y.; Sun, Y.; Chen, J.; Wang, X.; Zhang, Y.; Gao, L.; Li, L. The long noncoding RNA XIST regulates cardiac hypertrophy by targeting miR-101. J. Cell Physiol. 2019, 234, 13680–13692. [Google Scholar] [CrossRef]
  104. Liu, L.; Jiang, H.; Pan, H.; Zhu, X. LncRNA XIST promotes liver cancer progression by acting as a molecular sponge of miR-200b-3p to regulate ZEB1/2 expression. J. Int. Med. Res. 2021, 49, 3000605211016211. [Google Scholar] [CrossRef]
  105. Li, W.; He, Y.; Cheng, Z. Long noncoding RNA XIST knockdown suppresses the growth of colorectal cancer cells via regulating microRNA-338-3p/PAX5 axis. Eur. J. Cancer Prev. 2021, 30, 132–142. [Google Scholar] [CrossRef]
  106. Yang, L.G.; Cao, M.Z.; Zhang, J.; Li, X.Y.; Sun, Q.L. LncRNA XIST modulates HIF-1A/AXL signaling pathway by inhibiting miR-93-5p in colorectal cancer. Mol. Genet. Genom. Med. 2020, 8, e1112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Zhang, R.; Wang, Z.; Yu, Q.; Shen, J.; He, W.; Zhou, D.; Yu, Q.; Fan, J.; Gao, S.; Duan, L. Atractylenolide II reverses the influence of lncRNA XIST/miR-30a-5p/ROR1 axis on chemo-resistance of colorectal cancer cells. J. Cell Mol. Med. 2019, 23, 3151–3165. [Google Scholar] [CrossRef]
  108. Zeng, Z.L.; Lu, J.H.; Wang, Y.; Sheng, H.; Wang, Y.N.; Chen, Z.H.; Wu, Q.N.; Zheng, J.B.; Chen, Y.X.; Yang, D.D.; et al. The lncRNA XIST/miR-125b-2-3p axis modulates cell proliferation and chemotherapeutic sensitivity via targeting Wee1 in colorectal cancer. Cancer Med. 2021, 10, 2423–2441. [Google Scholar] [CrossRef] [PubMed]
  109. Zheng, W.; Li, J.; Zhou, X.; Cui, L.; Wang, Y. The lncRNA XIST promotes proliferation, migration and invasion of gastric cancer cells by targeting miR-337. Arab J. Gastroenterol. 2020, 21, 199–206. [Google Scholar] [CrossRef]
  110. Wang, X.; Zhang, G.; Cheng, Z.; Dai, L.; Jia, L.; Jing, X.; Wang, H.; Zhang, R.; Liu, M.; Jiang, T.; et al. Knockdown of LncRNA-XIST Suppresses Proliferation and TGF-β1-Induced EMT in NSCLC Through the Notch-1 Pathway by Regulation of miR-137. Genet. Test Mol. Biomark. 2018, 22, 333–342. [Google Scholar] [CrossRef] [PubMed]
  111. Liu, J.; Yao, L.; Zhang, M.; Jiang, J.; Yang, M.; Wang, Y. Downregulation of LncRNA-XIST inhibited development of non-small cell lung cancer by activating miR-335/SOD2/ROS signal pathway mediated pyroptotic cell death. Aging 2019, 11, 7830–7846. [Google Scholar] [CrossRef]
  112. Wang, J.; Cai, H.; Dai, Z.; Wang, G. Down-regulation of lncRNA XIST inhibits cell proliferation via regulating miR-744/RING1 axis in non-small cell lung cancer. Clin. Sci. 2019, 133, 1567–1579, Erratum in Clin. Sci. 2019, 133, 1825. [Google Scholar] [CrossRef] [PubMed]
  113. Jiang, Q.; Xing, W.; Cheng, J.; Yu, Y. Knockdown of lncRNA XIST Suppresses Cell Tumorigenicity in Human Non-Small Cell Lung Cancer by Regulating miR-142-5p/PAX6 Axis. Onco Targets Ther. 2020, 13, 4919–4929. [Google Scholar] [CrossRef] [PubMed]
  114. Sun, W.; Zu, Y.; Fu, X.; Deng, Y. Knockdown of lncRNA-XIST enhances the chemosensitivity of NSCLC cells via suppression of autophagy. Oncol. Rep. 2017, 38, 3347–3354. [Google Scholar] [CrossRef] [PubMed]
  115. Rong, H.; Chen, B.; Wei, X.; Peng, J.; Ma, K.; Duan, S.; He, J. Long non-coding RNA XIST expedites lung adenocarcinoma progression through upregulating MDM2 expression via binding to miR-363-3p. Thorac. Cancer 2020, 11, 659–671. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, P.J.; Pan, Y.H.; Wang, D.W.; You, D. Long non-coding RNA XIST promotes cell proliferation of pancreatic cancer through miR-137 and Notch1 pathway. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 12161–12170. [Google Scholar] [CrossRef]
  117. Hu, C.; Liu, S.; Han, M.; Wang, Y.; Xu, C. Knockdown of lncRNA XIST inhibits retinoblastoma progression by modulating the miR-124/STAT3 axis. Biomed. Pharmacother. 2018, 107, 547–554. [Google Scholar] [CrossRef] [PubMed]
  118. Wang, Y.; Sun, D.; Sheng, Y.; Guo, H.; Meng, F.; Song, T. XIST promotes cell proliferation and invasion by regulating miR-140-5p and SOX4 in retinoblastoma. World J. Surg. Oncol. 2020, 18, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Xu, Y.; Fu, Z.; Gao, X.; Wang, R.; Li, Q. Long non-coding RNA XIST promotes retinoblastoma cell proliferation, migration, and invasion by modulating microRNA-191-5p/brain derived neurotrophic factor. Bioengineered 2021, 12, 1587–1598. [Google Scholar] [CrossRef]
  120. Yang, L.L.; Li, Q.; Zhang, X.; Cao, T. Long non-coding RNA XIST confers aggressive progression via miR-361-3p/STX17 in retinoblastoma cells. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 10433–10444. [Google Scholar] [CrossRef]
  121. Zhou, K.; Li, S.; Du, G.; Fan, Y.; Wu, P.; Sun, H.; Zhang, T. LncRNA XIST depletion prevents cancer progression in invasive pituitary neuroendocrine tumor by inhibiting bFGF via upregulation of microRNA-424-5p. OncoTargets Ther. 2019, 12, 7095–7109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Li, H.L.; Han, P.H.; Pan, D.Q.; Chen, G.; Lu, X.H.; Li, J. LncRNA XIST regulates cell proliferation, migration and invasion of glioblastoma via regulating miR-448 and ROCK1. J. Biol. Regul. Homeost. Agents 2020, 34, 2049–2058. [Google Scholar] [CrossRef] [PubMed]
  123. Du, Y.L.; Liang, Y.; Cao, Y.; Liu, L.; Li, J.; Shi, G.Q. LncRNA XIST Promotes Migration and Invasion of Papillary Thyroid Cancer Cell by Modulating MiR-101-3p/CLDN1 Axis. Biochem. Genet. 2021, 59, 437–452. [Google Scholar] [CrossRef] [PubMed]
  124. Shi, J.; Tan, S.; Song, L.; Song, L.; Wang, Y. LncRNA XIST knockdown suppresses the malignancy of human nasopharyngeal carcinoma through XIST/miRNA-148a-3p/ADAM17 pathway in vitro and in vivo. Biomed. Pharmacother. 2020, 121, 109620. [Google Scholar] [CrossRef]
  125. Zhao, C.H.; Bai, X.F.; Hu, X.H. Knockdown of lncRNA XIST inhibits hypoxia-induced glycolysis, migration and invasion through regulating miR-381-3p/NEK5 axis in nasopharyngeal carcinoma. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 2505–2517. [Google Scholar] [CrossRef]
  126. Sun, L.; Zhang, M.; Qu, H. lncRNA XIST regulates cell proliferation, migration and invasion via regulating miR-30b and RECK in nasopharyngeal carcinoma. Oncol. Lett. 2021, 21, 256. [Google Scholar] [CrossRef]
  127. Xiao, D.; Cui, X.; Wang, X. Long noncoding RNA XIST increases the aggressiveness of laryngeal squamous cell carcinoma by regulating miR-124-3p/EZH2. Exp. Cell Res. 2019, 381, 172–178. [Google Scholar] [CrossRef]
  128. Cui, C.L.; Li, Y.N.; Cui, X.Y.; Wu, X. lncRNA XIST promotes the progression of laryngeal squamous cell carcinoma by sponging miR-144 to regulate IRS1 expression. Oncol. Rep. 2020, 43, 525–535. [Google Scholar] [CrossRef]
  129. Liu, C.; Lu, Z.; Liu, H.; Zhuang, S.; Guo, P. LncRNA XIST promotes the progression of laryngeal squamous cell carcinoma via sponging miR-125b-5p to modulate TRIB2. Biosci. Rep. 2020, 40, BSR20193172. [Google Scholar] [CrossRef] [Green Version]
  130. Chen, Z.; Hu, X.; Wu, Y.; Cong, L.; He, X.; Lu, J.; Feng, J.; Liu, D. Long non-coding RNA XIST promotes the development of esophageal cancer by sponging miR-494 to regulate CDK6 expression. Biomed. Pharmacother. 2019, 109, 2228–2236. [Google Scholar] [CrossRef]
  131. Sun, X.; Wei, B.; Peng, Z.H.; Fu, Q.L.; Wang, C.J.; Zheng, J.C.; Sun, J.C. Knockdown of lncRNA XIST suppresses osteosarcoma progression by inactivating AKT/mTOR signaling pathway by sponging miR-375-3p. Int. J. Clin. Exp. Pathol. 2019, 12, 1507–1517. [Google Scholar]
  132. Liu, W.; Long, Q.; Zhang, L.; Zeng, D.; Hu, B.; Zhang, W.; Liu, S.; Deng, S.; Chen, L. Long non-coding RNA X-inactive specific transcript promotes osteosarcoma metastasis via modulating microRNA-758/Rab16. Ann. Transl. Med. 2021, 9, 841. [Google Scholar] [CrossRef]
  133. Zhang, M.; Wang, F.; Xiang, Z.; Huang, T.; Zhou, W.B. LncRNA XIST promotes chemoresistance of breast cancer cells to doxorubicin by sponging miR-200c-3p to upregulate ANLN. Clin. Exp. Pharmacol. Physiol. 2020, 47, 1464–1472. [Google Scholar] [CrossRef]
  134. Hao, Y.Q.; Liu, K.W.; Zhang, X.; Kang, S.X.; Zhang, K.; Han, W.; Li, L.; Li, Z.H. GINS2 was regulated by lncRNA XIST/miR-23a-3p to mediate proliferation and apoptosis in A375 cells. Mol. Cell Biochem. 2021, 476, 1455–1465. [Google Scholar] [CrossRef]
  135. Wang, C.; Li, L.; Li, M.; Wang, W.; Liu, Y.; Wang, S. Silencing long non-coding RNA XIST suppresses drug resistance in acute myeloid leukemia through down-regulation of MYC by elevating microRNA-29a expression. Mol. Med. 2020, 26, 114. [Google Scholar] [CrossRef]
  136. Liu, Q.; Ran, R.; Wu, Z.; Li, X.; Zeng, Q.; Xia, R.; Wang, Y. Long Non-coding RNA X-Inactive Specific Transcript Mediates Cell Proliferation and Intrusion by Modulating the miR-497/Bcl-w Axis in Extranodal Natural Killer/T-cell Lymphoma. Front. Cell Dev. Biol. 2020, 8, 599070. [Google Scholar] [CrossRef]
  137. Mo, Y.; Lu, Y.; Wang, P.; Huang, S.; He, L.; Li, D.; Li, F.; Huang, J.; Lin, X.; Li, X.; et al. Long non-coding RNA XIST promotes cell growth by regulating miR-139-5p/PDK1/AKT axis in hepatocellular carcinoma. Tumour. Biol. 2017, 39, 1010428317690999. [Google Scholar] [CrossRef] [Green Version]
  138. Ning, D.; Chen, J.; Du, P.; Liu, Q.; Cheng, Q.; Li, X.; Zhang, B.; Chen, X.; Jiang, L. The crosstalk network of XIST/miR-424-5p/OGT mediates RAF1 glycosylation and participates in the progression of liver cancer. Liver Int. 2021, 41, 1933–1944. [Google Scholar] [CrossRef]
  139. Yang, L.; Xie, F.; Xu, W.; Xu, T.; Ni, Y.; Tao, X.; Zang, Y.; Jin, J. Long non-coding RNA XIST accelerates hepatic carcinoma progression by targeting the microRNA-320a/PIK3CA axis. Oncol. Lett. 2021, 22, 801. [Google Scholar] [CrossRef]
  140. Sun, K.; Jia, Z.; Duan, R.; Yan, Z.; Jin, Z.; Yan, L.; Li, Q.; Yang, J. Long non-coding RNA XIST regulates miR-106b-5p/P21 axis to suppress tumor progression in renal cell carcinoma. Biochem. Biophys. Res. Commun. 2019, 510, 416–420. [Google Scholar] [CrossRef]
  141. Xu, G.; Mo, L.; Wu, C.; Shen, X.; Dong, H.; Yu, L.; Pan, P.; Pan, K. The miR-15a-5p-XIST-CUL3 regulatory axis is important for sepsis-induced acute kidney injury. Ren. Fail. 2019, 41, 955–966. [Google Scholar] [CrossRef] [Green Version]
  142. Tang, B.; Li, W.; Ji, T.; Li, X.; Qu, X.; Feng, L.; Zhu, Y.; Qi, Y.; Zhu, C.; Bai, S. Downregulation of XIST ameliorates acute kidney injury by sponging miR-142-5p and targeting PDCD4. J. Cell Physiol. 2020, 235, 8852–8863. [Google Scholar] [CrossRef]
  143. He, X.; Luo, X.; Dong, J.; Deng, X.; Liu, F.; Wei, G. Long Non-Coding RNA XIST Promotes Wilms Tumor Progression Through the miR-194-5p/YAP Axis. Cancer Manag. Res. 2021, 13, 3171–3180. [Google Scholar] [CrossRef]
  144. Yang, J.; Shen, Y.; Yang, X.; Long, Y.; Chen, S.; Lin, X.; Dong, R.; Yuan, J. Silencing of long noncoding RNA XIST protects against renal interstitial fibrosis in diabetic nephropathy via microRNA-93-5p-mediated inhibition of CDKN1A. Am. J. Physiol. Renal. Physiol. 2019, 317, F1350–F1358. [Google Scholar] [CrossRef]
  145. Jin, L.W.; Pan, M.; Ye, H.Y.; Zheng, Y.; Chen, Y.; Huang, W.W.; Xu, X.Y.; Zheng, S.B. Down-regulation of the long non-coding RNA XIST ameliorates podocyte apoptosis in membranous nephropathy via the miR-217-TLR4 pathway. Exp. Physiol. 2019, 104, 220–230. [Google Scholar] [CrossRef] [Green Version]
  146. Xia, W.P.; Chen, X.; Ru, F.; He, Y.; Liu, P.H.; Gan, Y.; Zhang, B.; Li, Y.; Dai, G.Y.; Jiang, Z.X.; et al. Knockdown of lncRNA XIST inhibited apoptosis and inflammation in renal fibrosis via microRNA-19b-mediated downregulation of SOX6. Mol. Immunol. 2021, 139, 87–96. [Google Scholar] [CrossRef]
  147. Xu, X.; Ma, C.; Liu, C.; Duan, Z.; Zhang, L. Knockdown of long noncoding RNA XIST alleviates oxidative low-density lipoprotein-mediated endothelial cells injury through modulation of miR-320/NOD2 axis. Biochem. Biophys. Res. Commun. 2018, 503, 586–592. [Google Scholar] [CrossRef]
  148. Xu, J.; Li, H.; Lv, Y.; Zhang, C.; Chen, Y.; Yu, D. Silencing XIST mitigated lipopolysaccharide (LPS)-induced inflammatory injury in human lung fibroblast WI-38 cells through modulating miR-30b-5p/CCL16 axis and TLR4/NF-κB signaling pathway. Open Life Sci. 2021, 16, 108–127. [Google Scholar] [CrossRef]
  149. Wang, S.; Cao, F.; Gu, X.; Chen, J.; Xu, R.; Huang, Y.; Ying, L. LncRNA XIST, as a ceRNA of miR-204, aggravates lipopolysaccharide-induced acute respiratory distress syndrome in mice by upregulating IRF2. Int. J. Clin. Exp. Pathol. 2019, 12, 2425–2434. [Google Scholar]
  150. Chen, P.; Jiang, P.; Chen, J.; Yang, Y.; Guo, X. XIST promotes apoptosis and the inflammatory response in CSE-stimulated cells via the miR-200c-3p/EGR3 axis. BMC Pulm. Med. 2021, 21, 215. [Google Scholar] [CrossRef]
  151. Li, J.; Wei, L.; Han, Z.; Chen, Z.; Zhang, Q. Long non-coding RNA X-inactive specific transcript silencing ameliorates primary graft dysfunction following lung transplantation through microRNA-21-dependent mechanism. EBioMedicine 2020, 52, 102600. [Google Scholar] [CrossRef] [Green Version]
  152. Yuan, W.; Liu, X.; Zeng, L.; Liu, H.; Cai, B.; Huang, Y.; Tao, X.; Mo, L.; Zhao, L.; Gao, C. Silencing of Long Non-Coding RNA X Inactive Specific Transcript (Xist) Contributes to Suppression of Bronchopulmonary Dysplasia Induced by Hyperoxia in Newborn Mice via microRNA-101-3p and the transforming growth factor-beta 1 (TGF-β1)/Smad3 Axis. Med. Sci. Monit. 2020, 26, e922424. [Google Scholar] [CrossRef]
  153. Wei, M.; Li, L.; Zhang, Y.; Zhang, Z.J.; Liu, H.L.; Bao, H.G. LncRNA X inactive specific transcript contributes to neuropathic pain development by sponging miR-154-5p via inducing toll-like receptor 5 in CCI rat models. J. Cell Biochem. 2018, 120, 1271–1281. [Google Scholar] [CrossRef]
  154. Liu, B.Y.; Li, L.; Bai, L.W.; Xu, C.S. Long Non-coding RNA XIST Attenuates Diabetic Peripheral Neuropathy by Inducing Autophagy Through MicroRNA-30d-5p/sirtuin1 Axis. Front. Mol. Biosci. 2021, 8, 655157. [Google Scholar] [CrossRef]
  155. Wang, Z.Q.; Xiu, D.H.; Jiang, J.L.; Liu, G.F. Long non-coding RNA XIST binding to let-7c-5p contributes to rheumatoid arthritis through its effects on proliferation and differentiation of osteoblasts via regulation of STAT3. J. Clin. Lab. Anal. 2020, 34, e23496. [Google Scholar] [CrossRef]
  156. Zhang, M.; Yang, H.; Chen, Z.; Hu, X.; Wu, T.; Liu, W. Long Noncoding RNA X-Inactive-Specific Transcript Promotes the Secretion of Inflammatory Cytokines in LPS Stimulated Astrocyte Cell Via Sponging miR-29c-3p and Regulating Nuclear Factor of Activated T cell 5 Expression. Front. Endocrinol. 2021, 12, 573143. [Google Scholar] [CrossRef]
  157. Lv, P.; Liu, H.; Ye, T.; Yang, X.; Duan, C.; Yao, X.; Li, B.; Tang, K.; Chen, Z.; Liu, J. XIST Inhibition Attenuates Calcium Oxalate Nephrocalcinosis-Induced Renal Inflammation and Oxidative Injury via the miR-223/NLRP3 Pathway. Oxid. Med. Cell Longev. 2021, 2021, 1676152. [Google Scholar] [CrossRef]
  158. Lian, L.P.; Xi, X.Y. Long non-coding RNA XIST protects chondrocytes ATDC5 and CHON-001 from IL-1β-induced injury via regulating miR-653-5p/SIRT1 axis. J. Biol. Regul. Homeost. Agents 2020, 34, 379–391. [Google Scholar] [CrossRef]
  159. Niu, S.; Xiang, F.; Jia, H. Downregulation of lncRNA XIST promotes proliferation and differentiation, limits apoptosis of osteoblasts through regulating miR-203-3p/ZFPM2 axis. Connect. Tissue Res. 2020, 24, 381–392. [Google Scholar] [CrossRef]
  160. Chen, W.; Li, S.; Zhang, F. Role of lncRNA XIST/microRNA-19/PTEN network in autophagy of nucleus pulposus cells in intervertebral disc degeneration via the PI3K/Akt signaling pathway. Cell Cycle 2021, 20, 1629–1641. [Google Scholar] [CrossRef]
  161. Liao, X.; Tang, D.; Yang, H.; Chen, Y.; Chen, D.; Jia, L.; Yang, L.; Chen, X. Long Non-coding RNA XIST May Influence Cervical Ossification of the Posterior Longitudinal Ligament Through Regulation of miR-17-5P/AHNAK/BMP2 Signaling Pathway. Calcif. Tissue Int. 2019, 105, 670–680. [Google Scholar] [CrossRef]
  162. Yue, D.; Guanqun, G.; Jingxin, L.; Sen, S.; Shuang, L.; Yan, S.; Minxue, Z.; Ping, Y.; Chong, L.; Zhuobo, Z.; et al. Silencing of long noncoding RNA XIST attenuated Alzheimer’s disease-related BACE1 alteration through miR-124. Cell Biol. Int. 2020, 44, 630–636. [Google Scholar] [CrossRef]
  163. Zhou, Q.; Zhang, M.M.; Liu, M.; Tan, Z.G.; Qin, Q.L.; Jiang, Y.G. LncRNA XIST sponges miR-199a-3p to modulate the Sp1/LRRK2 signal pathway to accelerate Parkinson’s disease progression. Aging 2021, 13, 4115–4137. [Google Scholar] [CrossRef]
  164. Weng, S.; Wang, S.; Jiang, J. Long Noncoding RNA X-Inactive Specific Transcript Regulates Neuronal Cell Apoptosis in Ischemic Stroke Through miR-98/BACH1 Axis. DNA Cell Biol. 2021, 40, 979–987. [Google Scholar] [CrossRef]
  165. Wang, Y.; Li, Y.; Ma, C.; Zhou, T.; Lu, C.; Ding, L.; Li, L. LncRNA XIST Promoted OGD-Induced Neuronal Injury Through Modulating/miR-455-3p/TIPARP Axis. Neurochem. Res. 2021, 46, 1447–1456. [Google Scholar] [CrossRef]
  166. Zhang, H.; Xia, J.; Hu, Q.; Xu, L.; Cao, H.; Wang, X.; Cao, M. Long non-coding RNA XIST promotes cerebral ischemia/reperfusion injury by modulating miR-27a-3p/FOXO3 signaling. Mol. Med. Rep. 2021, 24, 566. [Google Scholar] [CrossRef]
  167. Zhang, M.; Yang, J.K.; Ma, J. Regulation of the long noncoding RNA XIST on the inflammatory polarization of microglia in cerebral infarction. Exp. Ther. Med. 2021, 22, 924. [Google Scholar] [CrossRef]
  168. Wang, C.; Dong, J.; Sun, J.; Huang, S.; Wu, F.; Zhang, X.; Pang, D.; Fu, Y.; Li, L. Silencing of lncRNA XIST impairs angiogenesis and exacerbates cerebral vascular injury after ischemic stroke. Mol. Ther. Nucleic Acids 2021, 26, 148–160. [Google Scholar] [CrossRef]
  169. Cao, W.; Feng, Y. LncRNA XIST promotes extracellular matrix synthesis, proliferation and migration by targeting miR-29b-3p/COL1A1 in human skin fibroblasts after thermal injury. Biol. Res. 2019, 52, 52. [Google Scholar] [CrossRef]
  170. Xie, Z.Y.; Wang, F.F.; Xiao, Z.H.; Liu, S.F.; Lai, Y.L.; Tang, S.L. Long noncoding RNA XIST enhances ethanol-induced hepatic stellate cells autophagy and activation via miR-29b/HMGB1 axis. IUBMB Life 2019, 71, 1962–1972. [Google Scholar] [CrossRef]
  171. Peng, H.; Luo, Y.; Ying, Y. lncRNA XIST attenuates hypoxia-induced H9c2 cardiomyocyte injury by targeting the miR-122-5p/FOXP2 axis. Mol. Cell. Probes 2020, 50, 101500. [Google Scholar] [CrossRef] [PubMed]
  172. Wang, X.; Li, X.L.; Qin, L.J. The lncRNA XIST/miR-150-5p/c-Fos axis regulates sepsis-induced myocardial injury via TXNIP-modulated pyroptosis. Lab. Investig. 2021, 101, 1118–1129. [Google Scholar] [CrossRef]
  173. Zhang, X.; Wu, H.; Mai, C.; Qi, Y. Long Noncoding RNA XIST/miR-17/PTEN Axis Modulates the Proliferation and Apoptosis of Vascular Smooth Muscle Cells to Affect Stanford Type A Aortic Dissection. J. Cardiovasc. Pharmacol. 2020, 76, 53–62. [Google Scholar] [CrossRef]
  174. Liang, K.; Cui, M.; Fu, X.; Ma, J.; Zhang, K.; Zhang, D.; Zhai, S. LncRNA Xist induces arterial smooth muscle cell apoptosis in thoracic aortic aneurysm through miR-29b-3p/Eln pathway. Biomed. Pharmacother. 2021, 137, 111163. [Google Scholar] [CrossRef]
  175. Yan, B.; Liu, T.; Yao, C.; Liu, X.; Du, Q.; Pan, L. LncRNA XIST shuttled by adipose tissue-derived mesenchymal stem cell-derived extracellular vesicles suppresses myocardial pyroptosis in atrial fibrillation by disrupting miR-214-3p-mediated Arl2 inhibition. Lab. Investig. 2021, 101, 1427–1438. [Google Scholar] [CrossRef]
  176. Yang, Y.; Zhang, J.; Chen, X.; Xu, X.; Cao, G.; Li, H.; Wu, T. LncRNA FTX sponges miR-215 and inhibits phosphorylation of vimentin for promoting colorectal cancer progression. Gene Ther. 2018, 25, 321–330. [Google Scholar] [CrossRef]
  177. Chen, G.Q.; Liao, Z.M.; Liu, J.; Li, F.; Huang, D.; Zhou, Y.D. LncRNA FTX Promotes Colorectal Cancer Cells Migration and Invasion by miRNA-590-5p/RBPJ Axis. Biochem. Genet. 2021, 59, 560–573. [Google Scholar] [CrossRef]
  178. Li, H.; Yao, G.; Zhai, J.; Hu, D.; Fan, Y. LncRNA FTX Promotes Proliferation and Invasion of Gastric Cancer via miR-144/ZFX Axis. Onco Targets Ther. 2019, 12, 11701–11713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Zhang, F.; Wang, X.S.; Tang, B.; Li, P.A.; Wen, Y.; Yu, P.W. Long non-coding RNA FTX promotes gastric cancer progression by targeting miR-215. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 3037–3048. [Google Scholar] [CrossRef] [PubMed]
  180. Zhang, W.; Bi, Y.; Li, J.; Peng, F.; Li, H.; Li, C.; Wang, L.; Ren, F.; Xie, C.; Wang, P.; et al. Long noncoding RNA FTX is upregulated in gliomas and promotes proliferation and invasion of glioma cells by negatively regulating miR-342-3p. Lab Investig. 2017, 97, 447–457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Liu, B.; Ma, X.; Liu, Q.; Xiao, Y.; Pan, S.; Jia, L. Aberrant mannosylation profile and FTX/miR-342/ALG3-axis contribute to development of drug resistance in acute myeloid leukemia. Cell Death Dis. 2018, 9, 688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Liu, L.; Li, X.; Shi, Y.; Chen, H. The long noncoding RNA FTX promotes a malignant phenotype in bone marrow mesenchymal stem cells via the miR-186/c-Met axis. Biomed. Pharmacother. 2020, 131, 110666. [Google Scholar] [CrossRef] [PubMed]
  183. Liu, F.; Yuan, J.H.; Huang, J.F.; Yang, F.; Wang, T.T.; Ma, J.Z.; Zhang, L.; Zhou, C.C.; Wang, F.; Yu, J.; et al. Long noncoding RNA FTX inhibits hepatocellular carcinoma proliferation and metastasis by binding MCM2 and miR-374a. Oncogene 2016, 35, 5422–5434. [Google Scholar] [CrossRef] [PubMed]
  184. Li, X.; Giri, V.; Cui, Y.; Yin, M.; Xian, Z.; Li, J. LncRNA FTX inhibits hippocampal neuron apoptosis by regulating miR-21-5p/SOX7 axis in a rat model of temporal lobe epilepsy. Biochem. Biophys. Res. Commun. 2019, 512, 79–86. [Google Scholar] [CrossRef]
  185. Long, B.; Li, N.; Xu, X.X.; Li, X.X.; Xu, X.J.; Guo, D.; Zhang, D.; Wu, Z.H.; Zhang, S.Y. Long noncoding RNA FTX regulates cardiomyocyte apoptosis by targeting miR-29b-1-5p and Bcl2l2. Biochem. Biophys. Res. Commun. 2018, 495, 312–318. [Google Scholar] [CrossRef] [PubMed]
  186. Li, L.; Li, L.; Zhang, Y.Z.; Yang, H.Y.; Wang, Y.Y. Long non-coding RNA FTX alleviates hypoxia/reoxygenation-induced cardiomyocyte injury via miR-410-3p/Fmr1 axis. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 396–408. [Google Scholar] [CrossRef]
  187. Yu, Y.; Yao, P.; Wang, Z.; Xie, W. Down-regulation of FTX promotes the differentiation of osteoclasts in osteoporosis through the Notch1 signaling pathway by targeting miR-137. BMC Musculoskelet. Disord. 2020, 21, 456. [Google Scholar] [CrossRef] [PubMed]
  188. Jin, M.; Ren, J.; Luo, M.; You, Z.; Fang, Y.; Han, Y.; Li, G.; Liu, H. Long non-coding RNA JPX correlates with poor prognosis and tumor progression in non-small-cell lung cancer by interacting with miR-145-5p and CCND2. Carcinogenesis 2020, 41, 634–645. [Google Scholar] [CrossRef]
  189. Pan, J.; Fang, S.; Tian, H.; Zhou, C.; Zhao, X.; Tian, H.; He, J.; Shen, W.; Meng, X.; Jin, X.; et al. lncRNA JPX/miR-33a-5p/Twist1 axis regulates tumorigenesis and metastasis of lung cancer by activating Wnt/β-catenin signaling. Mol. Cancer 2020, 19, 9. [Google Scholar] [CrossRef]
  190. Han, X.; Liu, Z. Long non-coding RNA JPX promotes gastric cancer progression by regulating CXCR6 and autophagy via inhibiting miR-197. Mol. Med. Rep. 2021, 23, 60. [Google Scholar] [CrossRef]
  191. Yao, Y.; Chen, S.; Lu, N.; Yin, Y.; Liu, Z. LncRNA JPX overexpressed in oral squamous cell carcinoma drives malignancy via miR-944/CDH2 axis. Oral Dis. 2021, 27, 924–933. [Google Scholar] [CrossRef]
  192. Yang, H.; Wang, G.; Liu, J.; Lin, M.; Chen, J.; Fang, Y.; Li, Y.; Cai, W.; Zhan, D. LncRNA JPX regulates proliferation and apoptosis of nucleus pulposus cells by targeting the miR-18a-5p/HIF-1α/Hippo-YAP pathway. Biochem. Biophys. Res. Commun. 2021, 566, 16–23. [Google Scholar] [CrossRef] [PubMed]
  193. Kwok, Z.H.; Zhang, B.; Chew, X.H.; Chan, J.J.; Teh, V.; Yang, H.; Kappei, D.; Tay, Y. Systematic Analysis of Intronic miRNAs Reveals Cooperativity within the Multicomponent FTX Locus to Promote Colon Cancer Development. Cancer Res. 2021, 81, 1308–1320. [Google Scholar] [CrossRef]
  194. Jin, S.; He, J.; Zhou, Y.; Wu, D.; Li, J.; Gao, W. LncRNA FTX activates FOXA2 expression to inhibit non-small-cell lung cancer proliferation and metastasis. J. Cell Mol. Med. 2020, 24, 4839–4849. [Google Scholar] [CrossRef] [Green Version]
  195. Torre, L.A.; Bray, F.; Siegel, R.L.; Ferlay, J.; Lortet-Tieulent, J.; Jemal, A. Global cancer statistics, 2012. CA Cancer J. Clin. 2015, 65, 87–108. [Google Scholar] [CrossRef] [Green Version]
  196. El-Serag, H.B.; Rudolph, K.L. Hepatocellular carcinoma: Epidemiology and molecular carcinogenesis. Gastroenterology 2007, 132, 2557–2576. [Google Scholar] [CrossRef]
  197. Guo, X.; Su, B.; Zhou, Z.; Sha, J. Rapid evolution of mammalian X-linked testis microRNAs. BMC Genom. 2009, 10, 97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Di Palo, A.; Siniscalchi, C.; Salerno, M.; Russo, A.; Gravholt, C.H.; Potenza, N. What microRNAs could tell us about the human X chromosome. Cell Mol. Life Sci. 2021, 77, 4069–4080. [Google Scholar] [CrossRef]
  199. Berglund, A.; Stochholm, K.; Gravholt, C.H. The epidemiology of sex chromosome abnormalities. Am. J. Med. Genet C Semin. Med. Genet. 2020, 184, 202–215. [Google Scholar] [CrossRef] [PubMed]
  200. Lin, A.E.; Prakash, S.K.; Andersen, N.H.; Viuff, M.H.; Levitsky, L.L.; Rivera-Davila, M.; Crenshaw, M.L.; Hansen, L.; Colvin, M.K.; Hayes, F.J.; et al. Recognition and management of adults with Turner syndrome: From the transition of adolescence through the senior years. Am. J. Med. Genet. A 2019, 179, 1987–2033. [Google Scholar] [CrossRef] [Green Version]
  201. Skakkebaek, A.; Viuff, M.; Nielsen, M.M.; Gravholt, C.H. Epigenetics and genomics in Klinefelter syndrome. Am. J. Med. Genet. C Semin. Med. Genet. 2020, 184, 216–225. [Google Scholar] [CrossRef] [PubMed]
  202. Stochholm, K.; Juul, S.; Gravholt, C.H. Mortality and incidence in women with 47, XXX and variants. Am. J. Med. Genet. A 2010, 152, 367–372. [Google Scholar] [CrossRef] [PubMed]
  203. Rao, E.; Weiss, B.; Fukami, M.; Rump, A.; Niesler, B.; Mertz, A.; Muroya, K.; Binder, G.; Kirsch, S.; Winkelmann, M.; et al. Pseudoautosomal deletions encompassing a novel homeobox gene cause growth failure in idiopathic short stature and Turner syndrome. Nat. Genet. 1997, 16, 54–63. [Google Scholar] [CrossRef]
  204. Ellison, J.W.; Wardak, Z.; Young, M.F.; Gehron Robey, P.; Laig-Webster, M.; Chiong, W. PHOG, a candidate gene for involvement in the short stature of Turner syndrome. Hum. Mol. Genet. 1997, 6, 1341–1347. [Google Scholar] [CrossRef] [Green Version]
  205. Ottesen, A.M.; Aksglaede, L.; Garn, I.; Tartaglia, N.; Tassone, F.; Gravholt, C.H.; Bojesen, A.; Sørensen, K.; Jørgensen, N.; Rajpert-De Meyts, E.; et al. Increased number of sex chromosomes affects height in a nonlinear fashion: A study of 305 patients with sex chromosome aneuploidy. Am. J. Med. Genet. A 2010, 152, 1206–1212. [Google Scholar] [CrossRef] [Green Version]
  206. Skakkebæk, A.; Nielsen, M.M.; Trolle, C.; Vang, S.; Hornshøj, H.; Hedegaard, J.; Wallentin, M.; Bojesen, A.; Hertz, J.M.; Fedder, J.; et al. DNA hypermethylation and differential gene expression associated with Klinefelter syndrome. Sci. Rep. 2018, 8, 13740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Trolle, C.; Nielsen, M.M.; Skakkebæk, A.; Lamy, P.; Vang, S.; Hedegaard, J.; Nordentoft, I.; Ørntoft, T.F.; Pedersen, J.S.; Gravholt, C.H. Widespread DNA hypomethylation and differential gene expression in Turner syndrome. Sci. Rep. 2016, 6, 34220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Nielsen, M.M.; Trolle, C.; Vang, S.; Hornshøj, H.; Skakkebaek, A.; Hedegaard, J.; Nordentoft, I.; Pedersen, J.S.; Gravholt, C.H. Epigenetic and transcriptomic consequences of excess X-chromosome material in 47, XXX syndrome-A comparison with Turner syndrome and 46,XX females. Am. J. Med. Genet. C Semin. Med. Genet. 2020, 184, 279–293. [Google Scholar] [CrossRef]
  209. Raznahan, A.; Parikshak, N.N.; Chandran, V.; Blumenthal, J.D.; Clasen, L.S.; Alexander-Bloch, A.F.; Zinn, A.R.; Wangsa, D.; Wise, J.; Murphy, D.G.M.; et al. Sex-chromosome dosage effects on gene expression in humans. Proc. Natl. Acad. Sci. USA 2018, 115, 7398–7403. [Google Scholar] [CrossRef] [Green Version]
  210. Schouten, P.C.; Vollebergh, M.A.; Opdam, M.; Jonkers, M.; Loden, M.; Wesseling, J.; Hauptmann, M.; Linn, S.C. High XIST and Low 53BP1 Expression Predict Poor Outcome after High-Dose Alkylating Chemotherapy in Patients with a BRCA1-like Breast Cancer. Mol. Cancer Ther. 2016, 15, 190–198. [Google Scholar] [CrossRef] [Green Version]
  211. Bojesen, A.; Juul, S.; Birkebaek, N.H.; Gravholt, C.H. Morbidity in Klinefelter syndrome: A Danish register study based on hospital discharge diagnoses. J. Clin. Endocrinol. Metab. 2006, 91, 1254–1260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Viuff, M.H.; Stochholm, K.; Lin, A.; Berglund, A.; Juul, S.; Gravholt, C.H. Cancer occurrence in Turner syndrome and the effect of sex hormone substitution therapy. Eur. J. Endocrinol. 2021, 184, 79–88. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Human X chromosome inactivation center. Representation of a portion of the X chromosome, with zoomed-in view of the region that controls XCI. XIC harbors many lncRNA genes and one protein-coding gene, ZCCHC13 (gray arrow); a few other protein-coding genes are located upstream of TSIX. The most well-characterized lncRNAs are indicated in bold font and with thicker arrows; all the others are inferred pseudogenes and indicated by thinner arrows. The direction of the arrows indicates the direction of transcription. Positions of the miRNA clusters on FTX are also indicated. Data were retrieved from Ensembl and NCBI databases.
Figure 1. Human X chromosome inactivation center. Representation of a portion of the X chromosome, with zoomed-in view of the region that controls XCI. XIC harbors many lncRNA genes and one protein-coding gene, ZCCHC13 (gray arrow); a few other protein-coding genes are located upstream of TSIX. The most well-characterized lncRNAs are indicated in bold font and with thicker arrows; all the others are inferred pseudogenes and indicated by thinner arrows. The direction of the arrows indicates the direction of transcription. Positions of the miRNA clusters on FTX are also indicated. Data were retrieved from Ensembl and NCBI databases.
Ijms 23 00611 g001
Figure 2. Multifaceted role of lncRNAs from the XIC. XIST is the essential molecule for the X chromosome inactivation (Xi); its ceRNA activity is invoked in different pathways, such as those indicated, with specific examples in brackets. JPX, FTX and TSIX are other XIC lncRNAs acting as positive or negative regulators of XIST, thus contributing to Xi; they are also endowed with ceRNET activity, to date less well characterized. Figure created with BioRender.com.
Figure 2. Multifaceted role of lncRNAs from the XIC. XIST is the essential molecule for the X chromosome inactivation (Xi); its ceRNA activity is invoked in different pathways, such as those indicated, with specific examples in brackets. JPX, FTX and TSIX are other XIC lncRNAs acting as positive or negative regulators of XIST, thus contributing to Xi; they are also endowed with ceRNET activity, to date less well characterized. Figure created with BioRender.com.
Ijms 23 00611 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Siniscalchi, C.; Di Palo, A.; Russo, A.; Potenza, N. The lncRNAs at X Chromosome Inactivation Center: Not Just a Matter of Sex Dosage Compensation. Int. J. Mol. Sci. 2022, 23, 611. https://doi.org/10.3390/ijms23020611

AMA Style

Siniscalchi C, Di Palo A, Russo A, Potenza N. The lncRNAs at X Chromosome Inactivation Center: Not Just a Matter of Sex Dosage Compensation. International Journal of Molecular Sciences. 2022; 23(2):611. https://doi.org/10.3390/ijms23020611

Chicago/Turabian Style

Siniscalchi, Chiara, Armando Di Palo, Aniello Russo, and Nicoletta Potenza. 2022. "The lncRNAs at X Chromosome Inactivation Center: Not Just a Matter of Sex Dosage Compensation" International Journal of Molecular Sciences 23, no. 2: 611. https://doi.org/10.3390/ijms23020611

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

Article Metrics

Back to TopTop