Next Article in Journal
Endothelial Colony-Forming Cells Dysfunctions Are Associated with Arterial Hypertension in a Rat Model of Intrauterine Growth Restriction
Next Article in Special Issue
HOXC6-Mediated miR-188-5p Expression Induces Cell Migration through the Inhibition of the Tumor Suppressor FOXN2
Previous Article in Journal
Role of the Endocannabinoid/Endovanilloid System in the Modulation of Osteoclast Activity in Paget’s Disease of Bone
Previous Article in Special Issue
HOXB8 Counteracts MAPK/ERK Oncogenic Signaling in a Chicken Embryo Model of Neoplasia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of lncRNAs in Rare Tumors with a Focus on HOX Transcript Antisense RNA (HOTAIR)

by
Giuseppina Liguori
1,†,
Margherita Cerrone
1,†,
Annarosaria De Chiara
1,
Salvatore Tafuto
2,
Maura Tracey de Bellis
3,
Gerardo Botti
1,
Maurizio Di Bonito
1 and
Monica Cantile
1,*
1
Pathology Unit, Istituto Nazionale Tumori-Irccs-Fondazione G. Pascale, 80131 Naples, Italy
2
Sarcomas and Rare Tumors Unit, Istituto Nazionale Tumori-Irccs-Fondazione G. Pascale, 80131 Naples, Italy
3
Rehabilitation Medicine Unit, Istituto Nazionale Tumori-Irccs-Fondazione G. Pascale, 80131 Naples, Italy
*
Author to whom correspondence should be addressed.
Equally contributors.
Int. J. Mol. Sci. 2021, 22(18), 10160; https://doi.org/10.3390/ijms221810160
Submission received: 30 August 2021 / Revised: 17 September 2021 / Accepted: 17 September 2021 / Published: 21 September 2021
(This article belongs to the Special Issue HOX Genes in Development and Disease)

Abstract

:
Rare cancers are identified as those with an annual incidence of fewer than 6 per 100,000 persons and includes both epithelial and stromal tumors from different anatomical areas. The advancement of analytical methods has produced an accurate molecular characterization of most human cancers, suggesting a “molecular classification” that has allowed the establishment of increasingly personalized therapeutic strategies. However, the limited availability of rare cancer samples has resulted in very few therapeutic options for these tumors, often leading to poor prognosis. Long non coding RNAs (lncRNAs) are a class of non-coding RNAs mostly involved in tumor progression and drug response. In particular, the lncRNA HOX transcript antisense RNA (HOTAIR) represents an emergent diagnostic, prognostic and predictive biomarker in many human cancers. The aim of this review is to highlight the role of HOTAIR in rare cancers, proposing it as a new biomarker usable in the management of these tumors.

1. Introduction

Rare cancers are identified as diseases with an incidence of <15 cases per 100,000 people per year, as reported by the National Cancer Institute of the United States, or <6 per 100,000 people per year, as suggested by Surveillance of Rare Cancers in Europe (RARECARE) [1]. In Italy, the rare cancers group mainly includes rare epithelial tumors of the digestive system, followed by epithelial tumors of the head and neck, rare cancers of the female genital system, endocrine tumors, sarcomas, central nervous system tumors and rare epithelial tumors of the thoracic cavity [2].
Rare cancers account for approximately 20–25% of all cancer patients diagnosed each year, overall achieving a much higher incidence rate than any single common cancer. However, their low prevalence results in a lack of research funding, decreased awareness, late or misdiagnosis, few treatment options and limited clinical trials due to the small number of eligible patients.
Recently, new large-scale sequencing technologies have made it possible to molecularly characterize the majority of human cancers. However, most rare cancer studies have been limited to a small sample of patients [3,4].
All these factors result in poor prognosis for rare cancers causing a significant public health issue. Therefore, there is a need to establish large collections of rare tumor samples, preferably through the implementation of dedicated biobanks [5,6], and secondly, to define molecular characterization in order to identify new prognostic markers and therapeutic targets.
Only 2% of the human genome is transcribed and translated into proteins. About 70% of it is transcribed into ncRNA. LncRNAs represents a new class of RNA molecules, longer than 200 nucleotides, described as crucial biomarkers in cancer. They have secondary and three-dimensional structures which enable them to have both RNA- and protein-like functions [7]. It has been demonstrated that the majority of lncRNAs are localized in the nucleus [8] and in cytoplasm [9]. The change in cellular localization is associated with specific cellular functions in many cases. Nuclear lncRNA are mainly involved in chromatin regulation, transcription regulation and scaffolding, acting as platforms for the assembly of multiple-component complexes such as ribonucleoprotein (RNP) complexes. The role of cytoplasmic lncRNAs is mainly associated with post-transcriptional regulation, for example, sponging microRNAs; thereby, they reduce miRNA availability to target mRNA [10].
Recent studies have revealed that the deregulation of specific lncRNAs is widely involved in the development and progression of tumors, affecting molecular mechanisms associated with cell proliferation, migration, invasion, epithelial-to-mesenchymal transition (EMT) and apoptosis [11,12].
Between numerous cancer related lncRNAs, HOTAIR plays a main role in contributing to tumor development, metastatic progression and drug resistance. For its proven prognostic value, HOTAIR has also been suggested as a potential therapeutic target in human cancer [13].
Since little is known about the role of lncRNAs in low incidence cancers, and there is a need to identify new molecular markers and therapeutic targets for these neoplasms, in this review we will discuss the role of lncRNAs, with the focus on HOTAIR, in head and neck rare cancers, soft tissue tumors, neuroendocrine tumors, rare digestive system tumors and central nervous system tumors.

2. LncRNAs in Rare Tumors

An increasing number of ncRNAs, especially lncRNAs, were found to play crucial roles in the initiation and progression of rare cancers, suggesting that they could function as novel biomarkers and therapeutic targets [14,15] (Table 1).
Rare cancers of the head and neck are epithelial cancers of the larynx, hypopharynx, nasal cavity and sinuses, nasopharynx, major salivary glands and salivary-gland type tumors, oropharynx, oral cavity and lip, eye and adnexa and the middle ear. Apart from these tumors, other rare malignancies can be located in the head and neck region, such as soft tissue sarcoma, bone sarcoma and Merkel cell carcinoma [16].
The incidence of laryngeal carcinoma is relatively low, comprising between 2 and 5% of all malignant diseases diagnosed annually worldwide. More than 95% of laryngeal tumors are squamous cell carcinomas (LSCC) [16]. Different lncRNAs have been described as prognostic markers in this tumor [17]. Wu et al. [18] reported that the lncRNA H19 is necessary for the development and progression of LSCC. It is inversely correlated with the survival rate of LSCC patients, and its knockdown inhibits LSCC cells migration, invasion and proliferation. Moreover, H19 is able to promote LSCC progression via miR-148a-3p and DNA methyltransferase enzyme DNMT [18]. The lncRNA UCA1 (Urothelial Cancer Associated 1) expression in LSCC patients is significantly higher in tumor tissues compared with adjacent healthy tissues and its serum levels is increased in these patients compared to healthy controls. In vitro, UCA1 is able to promote cell proliferation, invasion and migration of LSCC cells by activating the Wnt/β-catenin signaling pathway [19]. Similarly, lncRNA small NF90-associated RNA (snaR) is upregulated in the plasma of patients with LSCC compared to healthy controls [20]. In LSCC patients with neck nodal metastasis, the lncRNA NEAT1 (nuclear paraspeckle assembly transcript1) is overexpressed and its gene silencing in vitro and in vivo models can inhibit tumor growth [21]. The overexpression of lncRNA PCAT19 (protocadherin 19) is strongly associated with decreased overall survival of LSCC patients and its silencing in cell lines decreases tumor growth in vivo by regulating the miR-182/PDK4 axis [22]. Many other lncRNAs are aberrantly expressed in LSCC tissues and are correlated with poor prognosis, such as LINC00668 [23], lncRNA TUG1 (taurine-upregulated gene 1) [24], HOXA11 antisense RNA (HOXA11-AS) [25], ATB [26], LINC02154 and MNX1 (motor neuron and pancreas homeobox 1)-AS1 [27].
Table 1. Main lncRNAs described in rare cancers.
Table 1. Main lncRNAs described in rare cancers.
LncRNAsTumor TypeExpressionReferences
AOC4PGISTupregulation[28]
ADAMTS9-AS2SACCupregulation[29]
AFAP1-AS1NPCupregulation[30]
ATBLSCCupregulation[26]
BCAR4Osteosarcomaupregulation[31]
CCDC26GISTupregulation[32,33]
DNM3OSGISTupregulation[34]
EWSAT1Ewing’s sarcomaupregulation[35]
H19LSCCupregulation[18]
NPCupregulation[36]
NENupregulation[37]
GISTupregulation[28]
Glioblastomaupregulation[38,39]
HNF1A-ASNPCupregulation[36]
Osteosarcomaupregulation[40]
HOTTIPOsteosarcomaupregulation[41]
HOXA11-ASLSCCupregulation[25]
NPCupregulation[42,43]
KCNQ1OT1NPCupregulation[44]
LINC00668LSCCupregulation[23]
LINC02154LSCCupregulation[27]
MAGI2-AS3NPCupregulation[45]
MALAT1Osteosarcomaupregulation[46]
GEP-NENupregulation[47]
GISTupregulation[48]
Glioblastomaupregulation[49,50,51,52,53]
MINCRNPCupregulation[54]
MRPL23-AS1SACCupregulation[55]
NEAT1LSCCupregulation[21]
Glioblastomaupregulation[56]
NONHSAT154433.1MECupregulation[57]
OIP5-AS1NPCupregulation[58]
PCA3Pulmonary NENupregulation[59]
PCAT19LSCCupregulation[22]
PILRLSLiposarcomaupregulation[60]
SNHG8NPCupregulation[61]
SNHG15NPCupregulation[62]
TP73-AS1NPCupregulation[63]
TUG1LSCCupregulation[24]
Osteosarcomaupregulation[64]
TUSC7Glioblastomaupregulation[65]
UCA1LSCCupregulation[19]
Osteosarcomaupregulation[66]
LETNPCdownregulation[67]
MEG3Osteosarcomadownregulation[68]
Pulmonary NENdownregulation[59]
TUS7Osteosarcomadownregulation[68,69]
Nasopharyngeal carcinoma (NPC) is a rare tumor arising from the epithelial cells that cover the surface and line the nasopharynx, with an annual incidence rate of 1/100,000 in Western countries. Although rare, NPC accounts for about one third of childhood nasopharyngeal tumors [16].
The lncRNA actin filament-associated protein 1 antisense RNA1 (AFAP1-AS1) is upregulated in NPC and associated with poor survival. Its silencing significantly inhibited NPC cells migration and invasion [30]. In the same manner, the lncRNA H19 and hepatocyte nuclear factor 1A-antisense RNA (HNF1A-AS) are overexpressed in NPC tissues and involved in the modulation of cell cycle progression, tumor cell proliferation, migration and epithelial to mesenchymal transition (EMT) [36,70]. On the contrary, the downregulation of lncRNA-low expression in tumor (lncRNA-LET) in NPC tissues is significantly correlated to advanced clinical stage, tumor size, lymph node metastases and poor survival of patients [67].
More recently, many other lncRNAs have been associated with NPC progression, with their capacity of sponging different microRNAs [58,61,62,63,71,72,73,74]. Additionally, in NPC patients, a large number of lncRNAs were also associated with cisplatin resistance [42,43,44,45,75,76,77,78] and radioresistance [54,79,80].
Salivary gland cancer (SGC) comprises a heterogeneous group of tumors with approximately 6.5% of the cases among the malignant tumors of the head and neck. They are considered rare cancers, having an annual incidence of less than 2/100,000 in most countries [81]. There are very few indications of the role of lncRNAs in salivary gland tumors.
An lncRNA microarray analysis highlighted that lncRNA ADAM metallopeptidase with thrombospondin type 1 motif, 9 (ADAMTS9) antisense RNA 2 (ADAMTS9-AS2) was significantly upregulated in salivary adenoid cystic carcinoma (SACC) and it is correlated with metastasis rate and poor prognosis in SACC patients [29]. Similarly, the long non-coding RNA (lncRNA) MRPL23 antisense RNA 1 (MRPL23-AS1) was highly expressed and correlated with lung metastasis and overall survival rate in patients with SACC [55]. More recently, the upregulation of lncRNA NONHSAT154433.1 and decreased expression of circ012342 have been closely related to the pathogenesis of mucoepidermoid carcinoma (MEC) [57].
Soft tissue sarcomas are relatively uncommon tumors, accounting for 1% of all malignancies. They are classified in about 80 histologic subtypes according to tissue components from which they are derived. Approximately 80% of sarcomas originate from soft tissues, while 20% from bone [82].
Many lncRNAs are involved in osteosarcoma (OS) progression. The lncRNA MALAT1 (metastasis-associated lung adenocarcinoma transcript 1) was closely correlated with lung metastasis in OS patients, and it is an independent prognostic factor of OS. Its knockdown affected the PI3K/AT signaling pathway and inhibited invasion and metastasis in vitro and in vivo [46]. In OS tissues, other lncRNAs are significantly upregulated and related with metastatic progression, such as HNF1A (HNF1 homeobox A)-AS1 [40], BCAR4 (breast cancer anti-estrogen resistance 4) [31] and HULC (highly upregulated in liver cancer RNA) [83]. The latter was strongly correlated with a shorter overall survival in OS patients [83]. In a same manner, HOTTIP (HOXA distal transcript antisense RNA) [41], UCA1 [66] and TUG1 [64] expression increased in osteosarcoma tissue and was associated with poorer overall survival. On the contrary, MEG3 (maternally expressed 3) and TUS7 are downregulated in human osteosarcoma tissue [68,69]. Emerging studies suggest that lncRNAs contribute to tumor cell growth and proliferation in Ewing’s sarcoma. The lncRNA EWS-AT1 (EWS RNA binding protein 1) was found to be induced and upregulated by EWS-FLI1 chimeric protein in primary pediatric human mesenchymal progenitor cells. EWSAT1 diminished cell viability in human Ewing sarcoma cell lines [35]. LncRNA PILRLS (Proliferation Interacting LncRNA in Retroperitoneal Liposarcoma) was overexpressed in retroperitoneal liposarcoma, and its silencing is able to significantly inhibit cell proliferation and colony formation of liposarcoma cells [60].
Neuroendocrine neoplasms (NENs) are a relatively rare and heterogeneous tumor types, accounting for about 0.5% of all newly diagnosed malignancies [84]. The most frequent primary sites are the gastrointestinal tract (62–67%) the lungs (22–27%), and more rarely the genitourinary tract [85].
Very little is known about the role of lncRNAs in epigenetic regulation of NENs development and progression. Expression levels of lncRNAs MALAT1 and HOTAIR analyzed by chromogenic in situ hybridization (ISH) were associated with tumor stages and development of metastases in GEP-NEN [47]. H19 was significantly upregulated in NEN tissues with malignant behaviors, and its upregulation is able to predict poor prognosis. In vitro and in vivo data showed that H19 overexpression promoted tumor growth and metastasis and revealed that H19 activated PI3K/AKT/CREB signaling and promoted pNEN progression by interacting with VGF (VGF nerve growth factor inducible) [37]. LncRNA-p21 is highly expressed in neuroendocrine prostate cancer patients and cells [86] while MEG3 and prostate cancer antigen 3 (PCA3) were aberrantly expressed in pulmonary NENs, including typical carcinoid tumors, atypical carcinoid tumors, small cell lung carcinoma (SCLC/NEC) and large cell neuroendocrine carcinoma (LCNEC/NEC) [59].
Digestive system tumors form in organs involved in digesting food and drinks and some of them belong to the category of rare tumors, such as cholangiocarcinoma and gastrointestinal stromal tumors (GISTs) [87]. Badalamenti et al. [48] analyzed the expression of H19 and MALAT1 in 40 metastatic GIST tissues, showing their upregulation in 50% of GIST patients. Both H19 and MALAT1 overexpression was significantly higher in patients with time to progression (TTP) < 6 months as compared to patients with TTP > 6 months. Moreover, MALAT1 expression levels seem to be correlated with c-KIT mutation status [48]. The aberrant expression of lncRNA AOC4P has been detected in high-risk GISTs compared with low- and intermediate-risk GISTs. In addition, its expression appeared closely associated with upregulation of epithelial–mesenchymal transition (EMT)-related proteins, such as TGF-β (transforming growth factor-beta), ZEB1 (Zinc finger E-Box binding homeobox 1), Vimentin, Snail, and E-cadherin. AOC4P silencing led to the decrease in cell proliferative migration and invasive ability of GIST cells [28]. Next-generation sequencing data of paired GIST and adjacent tissue samples were analyzed by a web-based lincRNA analysis, showing the deregulation of lncRNAs MALAT1, H19 and FENDRR (FOXF1 adjacent noncoding developmental regulatory RNA). Moreover, H19 upregulation appeared strongly related with different oncogenes, such as ETV1 (ETS variant transcription factor 1) and miR-455-3p [88]. The deregulation of other lncRNAs has been recently described in GIST. Cao et al. [32] described the role of the lncRNA coiled-coil domain-containing 26 (CCDC26) in imatinib resistance of GIST, highlighting that cells with lower CCDC26 expression were less sensitive to imatinib compared to those with higher CCDC26 expression. In addition, CCDC26 expression decreased in a time-dependent manner in the presence of imatinib and its silencing can upregulate c-KIT expression [32]. Recently, an oncomine analysis performed on a large series of low-risk and high-risk GISTs, revealed that the lncRNA DNM3OS was involved in the malignant transformation of GISTs and correlated with a worse prognosis. Finally, DNM3OS was involved in the Hippo signaling pathway by regulating the expression of GLUT4 (glucose transporter member 4) and CD36 [34].
Central nervous system (CNS) tumors are relatively rare, and they are associated with high morbidity and mortality. The most common glial tumors are glioblastoma multiform and anaplastic glioma, comprising more than 50 and 10%, respectively, of the total gliomas [89]. Abnormal expression of several lncRNAs have been detected in glioma/glioblastoma tumors and related cell lines. In particular, MALAT1 has been noted to be involved in the pathogenesis of glioblastoma. Vassallo et al. revealed that MALAT1 silencing is able to decrease glioblastoma cells migration, without affecting proliferation [49]. Most studies report the role of MALAT 1 as an important marker of chemoresistance to TMZ (temozolomide). It can enhance the resistance of glioma cells to TMZ by regulating ZEB1 [50]. Different molecular pathways are associated with resistance mechanism related to MALAT 1 in glioblastoma cells. Chen et al. reported that MALAT1 induces chemoresistance to TMZ through suppressing miR-203 expression and promoting the expression of thymidylate synthase [51]. Similarly, Cai et al. reported the upregulation of MALAT1 and its main role in TMZ-resistant glioblastoma cells by inhibiting the miR-101 signaling pathway [52]. More recently, NF-κB and p53 have been identified as regulators of the MALAT1 expression in induction of TMZ resistance in glioblastoma [53]. The serum levels of MALAT1 have also been associated with poor response to TMZ and lower survival rate of patients with glioblastoma [51]. H19 is another oncogenic lncRNA in glioblastoma whose aberrant expression is inversely correlated with the expression of NKD1 (NKD inhibitor of WNT signaling pathway 1), an inhibitor of the Wnt pathway [38]. Similar to MALAT1, H19 silencing is also able to modulate TMZ cytotoxicity in glioma cells by inhibiting EMT via the Wnt/β-catenin pathway and inactivating NF-κB signaling [39]. Other lncRNAs have been described as involved in TMZ resistance in glioblastoma cells, such as NEAT1 [56] and TUSC7 (tumor suppressor candidate 7) [65].

3. HOTAIR and Its Role in Human Cancers

HOTAIR is a lncRNA located within the intergenic region between HOXC11 and HOXC12 in the HOXC cluster on chromosome 12q13.13. Its principal transcript is 2364 bp RNA, transcribed from a 6449 bp gene locus and composed of six exons [90]. The human HOTAIR gene can be transcribed into several variants via alternative splicing and recently six major HOTAIR splicing variants have been described [91]. HOTAIR promoter contains binding sites for numerous transcription factors, which include AP1, Sp1, ERE elements, HRE elements and NF-κB [13]. HOTAIR, as well as many lncRNAs, is a key modulator of chromatin stability and is mainly involved in transcriptional silencing mechanisms [91]. Mechanistically, HOTAIR is able to bind the PRC2 (Polycomb repressive complex) at the 5′ end, and the LSD1 (lysine-specific histone demethylase 1A) at the 3′ end, acting as a molecular scaffold for the conjunction of the two complexes [90,91,92]. The HOTAIR-PRC2-LSD1 complex determines epigenetic changes contributing to targeted gene silencing and repressing their transcription via H3K27 trimethylation (PRC2 activity) and H3K4 demethylation (LSD1 activity) [91,92]. HOTAIR can also modulate gene expression at the post-transcriptional level describing that it could serve as a ubiquitination protein and subsequent degradation platform [93].
Physiologically, HOTAIR can be involved in the regulation of the cell cycle. It promotes the cell cycle that passes through the restriction point during the G1 phase by regulating CDK4/6-cyclin D and the Rb-E2F pathway [94]. During embryogenesis, HOTAIR is involved in the development of the lumbosacral region, through the repression of HOX D locus genes [90]. Furthermore, it was reported that HOTAIR possesses many miRNA recognition elements (MREs), and their functional interactions are able to modulate important cellular processes [95,96,97].
Numerous studies have shown that HOTAIR can be directly associated with tumor diseases being involved in tumor initiation, growth, angiogenesis, progression, recurrence and drug resistance mechanisms [92,98,99]. In addition, many clinical studies suggested HOTAIR as a fundamental biomarker associated with poor prognosis [100]. Early studies highlighted that the aberrant expression of HOTAIR in human tumors have been detected in breast cancer (BC) patients. HOTAIR appears to be a powerful predictor of BC tumor progression: its upregulation has been described in primary BC tumors with high metastatic potential and poor survival [101]. The deregulation of HOTAIR expression has been found in different molecular subtypes of BC often with conflicting data [102,103,104,105]. In BC, HOTAIR is also involved in the regulation of many different processes, mainly related with epithelial mesenchymal transition (EMT) [106]. BC cells are able to promote the EMT and metastasis processes, when treated with TGF-B1, through the upregulation of HOTAIR. The downregulation of HOTAIR results in the reduction of the ability to form colonies [106,107]. Similarly, the promotion of metastatic processes in BC is strongly influenced by the interaction of HOTAIR with a series of microRNAs. In BC, HOTAIR is able to interact with different miRNAs promoting tumor progression, such as miR-7 [108], miR-206 [95] and miR34a [96,109]. HOTAIR has been detected in the blood of BC patients, and its circulating DNA level significantly correlated with the clinical stage of the tumor [110]. Moreover, Tang et al. [111] showed that serum exosomal HOTAIR is a potent predictor of both poor survival and drug response in BC patients [111]. Different studies showed the crucial prognostic role of HOTAIR also in gastrointestinal tract tumors, especially in colorectal cancer [112] and gastric cancer [113]. In the latter, high level of circulating HOTAIR is associated with sensibility to fluorouracil and platinum-based combination therapy [113]. In liver cancer, HOTAIR upregulation correlates with clinical-pathological features and tumor progression [114] and its silencing increases chemotherapy sensitivity [115]. In urogenital cancers, the role of HOTAIR has been well documented, especially in prostate [116] and bladder cancer [117,118,119]. HOTAIR is overexpressed in ovarian cancer, and it is associated with stage, lymph node metastases and poor survival [120]; similarly, HOTAIR is overexpressed in cervical cancer [121,122] and in endometrial carcinoma [123] in which it is also associated with cisplatin resistance acquisition [124]. Finally, HOTAIR upregulation correlates with advanced stage, lymph nodes metastases, poor prognosis and drug resistance in non-small cell lung cancer (NSCLC) patients [125,126]

4. The Role of HOTAIR in Rare Tumors

A series of studies carried out with in vitro and in vivo models of rare cancers highlighted the fundamental role that HOTAIR has in these tumors and the complex network of molecular interactions in which it is involved. Its aberrant activity is capable of regulating the main molecular pathways associated with carcinogenesis, metastatic progression, angiogenesis and drug resistance (Figure 1). Furthermore, most of these studies highlighted the direct correlation between HOTAIR upregulation and the prognosis of patients with rare tumors. In Table 2, the main groups of rare tumors are summarized on the basis of their incidence (carcinomas and sarcomas).

4.1. Head and Neck Cancers

The aberrant expression of lncRNA HOTAIR has been abundantly described in head and neck tumors, especially in rare epithelial cancers such as laryngeal squamous cell carcinoma, nasopharyngeal carcinoma and salivary gland tumors.
HOTAIR was upregulated in primary LSCC, compared with adjacent noncancerous tissues and its overexpression was correlated with poor differentiation, lymph node metastasis and advanced clinical stages. HOTAIR silencing in LSCC cells leads to a significant decrease in invasive ability and promotes apoptosis. Furthermore, HOTAIR knockdown can effectively suppress the progression of LSCC in vivo xenografts mice [139]. The combined expression of HOTAIR and its interactor, which is an enhancer of zeste homolog 2 (EZH2, a regulator of epigenetic modification), were overexpressed in LSCC tissue. HOTAIR overexpression is significantly related to T phase, pathological grading and metastatic risk. Its silencing promoted cell proliferation and increased sensitivity to cis-platinum in the LSCC cells [127]. Moreover, a large case series of LSCC has been selected to evaluate the influences of cisplatin and paclitaxel on lncRNAs expression. HOTAIR was dramatically reduced with the increasing concentration of cisplatin and paclitaxel suggesting their target function on specific lncRNASs in LSCC patients [128]. Recently, a bioinformatics analysis to examine miRNAs, lncRNAs and mRNAs differentially expressed was performed on recurrent and non-recurrent LSCC sample datasets. Analysis showed that HOTAIR, HCG4 (HLA complex group 4) and EMX2OS (EMX2 opposite strand/antisense RNA) can represent a non-coding RNA signature in recurrent LSCC. Furthermore, the HOTAIR-miR-1-MAGEA2 (melanoma antigen A2 gene) interaction may be fundamental for the identification of recurrent LSCC [140].
HOTAIR’s role as a circulating marker has been extensively documented in the majority of solid cancers and a series of studies validated this role in rare tumors. A large study conducted on 52 LSCC patients and 49 patients with benign polyps of the vocal cords showed that the expression of exosomal HOTAIR was significantly higher only in patients with LSCC. In addition, patients with lymph node metastasis had higher serum exosomal HOTAIR expressions than those with no metastases, suggesting that circulating HOTAIR could be a valuable biomarker to screen and predict progression for LSCC patient [141]. Finally, exosome-mediated HOTAIR is able to act as ceRNA of miR-454-3p to regulate the tumor suppressor gene E2F2 (Eukaryotic elongation factor 2), negatively regulating the radiosensitivity of laryngeal cancer cells [142].
In nasopharyngeal carcinoma patients, HOTAIR expression levels increased with clinical stage progression, and it is associated with poor prognosis. The functional analysis in in vitro models showed that HOTAIR is able to modulate migration, invasion and proliferation of NPC cells [129]. Subsequent studies have validated the prognostic role of HOTAIR in NPC and highlighted a strong relationship with angiogenic pathways. Functional studies exhibited that silencing of HOTAIR by siHotair directly inactivated VEGFA transcriptional activity and suppressed the expression of glucose regulated protein 78 (GRP78); this suggests its main role in mediating tumorigenesis and angiogenesis in NPC [130]. Some other molecular pathways are related with aberrant expression of HOTAIR in NPC patients. The expression of fatty acid synthase (FASN) is positively correlated to HOTAIR and de novo synthesis of cellular free fatty acid in NPC cells is inhibited when HOTAIR was silenced [143]. HOTAIR is able to induce COX-2 (Cyclooxygenase-2) upregulation and promotes proliferation, migration and invasion in NPC cells. Moreover, miR-101 directly binds to the 3’-UTR of COX-2 and downregulates COX-2 expression, suggesting the importance of HOTAIR/miR-101/COX-2 axis in progression of nasopharyngeal carcinoma cells [144]. More recently, Yang et al. demonstrated that HOTAIR inhibits E-cadherin by stimulating the trimethylation of H3K27 to promote NPC cell progression through recruiting histone methylase EZH2 [145].
There is some evidence of the role of HOTAIR in salivary gland cancers. A recent study conducted on 86 patients with salivary adenoid cystic carcinoma (SACC) showed that HOTAIR expression in SACC tissue was higher than that in normal salivary gland tissue. Additionally, its expression in tissues of patients with TNM stages III or IV, nerve invasion, lymph node metastasis and poor survival rate is increased, suggesting that HOTAIR is a potential marker for prognostic assessment of patients with SACC [131].

4.2. Neuroendocrine Tumors

Numerous lncRNAs have been associated with neuroendocrine tumors pathogenesis especially in gastroenteropancreatic neuroendocrine tumors (GEP-NET); however, the role of HOTAIR is still poorly investigated.
Upregulation of HOTAIR has been described in GEP-NET and it is significantly associated with grade and aggressive phenotype [47]. More recently, this data has been confirmed. In fact, HOTAIR showed weak expression in low-grade GEP NENs and aberrant expression in NET G3 and NEC G3 categories. Furthermore, HOTAIR appeared inversely correlated with posterior HOX genes expression, highlighting that the combined expression can be useful in molecular stratification of GEP-NENs [146].
Regarding prostate cancer, Chang et al. [136] described that HOTAIR is upregulated in castration-resistant PCa (CRPC) with neuroendocrine differentiation (NEPC). Specifically, HOTAIR overexpression is sufficient to induce NED, whereas knockdown of HOTAIR suppressed it in PCa cells. In fact, HOTAIR upregulation induced the expression of some NEPC markers in prostate cancer cells. HOTAIR expression can be inhibited by the transcriptional repressor REST (RE1 silencing transcription factor), which is a master transcriptional repressor that restricts neuronal gene expression in stem cells and non-neuronal cell [136]. However, HOTAIR’s role in this process has recently been downsized. Mather et al. [147] found that, while REST is consistently downregulated in NEPC versus CRPC/adenocarcinoma samples, HOTAIR is expressed at very similar levels in the two groups, suggesting that the protein REST plays a pivotal role in inhibiting NEPC transdifferentiation, and that this effect is not mediated by HOTAIR [147].
Finally, a recent study analyzed HOTAIR expression in typical carcinoid tumors, atypical carcinoid tumors, small cell lung carcinoma (SCLC/NEC) and large cell neuroendocrine carcinoma (LCNEC/NEC), highlighting its upregulation only in SCLC/NEC patients [59].

4.3. Rare Digestive System Tumors

Among rare digestive system tumors, cholangiocarcinoma represents a heterogeneous group of highly aggressive malignancies originating from the biliary ducts with poor prognosis [148].
HOTAIR was highly expressed both in cholangiocarcinoma tissues and cell lines compared with corresponding normal bile duct tissues and intrahepatic biliary epithelial cells. Its overexpression is strongly correlated with tumor size, TNM stage and recurrence in cholangiocarcinoma patients. HOTAIR silencing significantly decreased the migration and invasion and increased apoptosis of cholangiocarcinoma cell models [149]. More recently, Lu et al. demonstrated that HOTAIR is able to promote cholangiocarcinoma progression by regulating HMGB1 to suppress cell apoptosis, autophagy and induce cell proliferation by sponging miR-204-5p [150]. The analysis of polymorphisms in the gene sequence of HOTAIR to evaluate the susceptibility to the development of cholangiocarcinoma has recently been evaluated. In a Greek cohort of patients, HOTAIR rs4759314 AG and GG genotypes were associated with a significantly increased cholangiocarcinoma risk [137].
GISTs are rare, making up less than 1% of all gastrointestinal tumors. HOTAIR appeared overexpressed in GISTs and the combined overexpression of miR-196a are strongly associated with high-risk grade, metastasis and poorer patient survival. Knockdown of HOTAIR altered the expression of GIST repressed cells invasiveness [138]. Lee et al. [151] described the upregulation of HOTAIR in surgically resected high-risk GISTs compared with low- and intermediate-risk GISTs. In GIST cell models, HOTAIR is able to repress apoptosis and to promote cell invasion and migration. Furthermore, HOTAIR induces methylation of PCDH10 (Protocadherin 10), a tumor suppressor gene, in GIST cells [151]. Several other studies confirmed the upregulation of HOTAIR in GISTs with a high risk of recurrence. Bure et al. highlighted that HOTAIR knockdown in GIST cells modulates the expression of genes involved in the organization and disassembly of the extracellular matrix and induces locus-specific alterations of DNA methylation patterns, especially in DPP4 (dipeptidyl peptidase 4), RASSF1 (Ras association domain family member 1) and ALDH1A3 (aldehyde dehydrogenase 1 family member A3) genes [152]. Furthermore, HOTAIR is described as a drug resistance-related lncRNA in GIST which is involved in imatinib resistance [153]. More recently, Zhang et al. [132] analyzed HOTAIR expression in GIST cells after imatinib treatments showed that HOTAIR is able to shift from nucleus to cytoplasm thereby modulating drug sensitivity via autophagy. In addition, HOTAIR, downregulating miRNA-130a and thereby activating the downstream target autophagy-related protein 2 homolog B (ATG2B), is able to modulate autophagy and imatinib sensitivity in GIST cells [132].

4.4. Central Nervous System Tumors

HOTAIR aberrant expression has been abundantly described in glioma tumors and closely associated with glioma grade and poor prognosis [154]. HOTAIR expression correlated with glioma molecular subtype and was preferentially expressed in the classical and mesenchymal subtypes compared with the neural and proneural subtypes. HOTAIR silencing induced colony formation suppression, cell cycle G0/G1 arrest and orthotopic tumor growth inhibition, acting as a crucial regulator of cell cycle progression in glioma cells [154]. The same authors described that EZH2 inhibition blocked cell cycle progression in glioma cells, consistent with the effects elicited by HOTAIR siRNA, suggesting that HOTAIR might regulate cell cycle progression through EZH2 [155]. Several molecular pathways have been regulated by HOTAIR in glioma. NLK (Nemo-like kinase), a negative regulator of the β-catenin pathway, was negatively correlated with HOTAIR expression. When the β-catenin pathway was inhibited, glioma cells became susceptible to cell cycle arrest and inhibition of invasion. HOTAIR expression induction in in vivo model of glioma upregulated β-catenin, while its silencing inhibited glioma cell migration/invasion [156]. Bromodomain and extraterminal (BET) proteins are important therapeutic targets in glioblastoma. Treatment of glioblastoma cells with the BET bromodomain inhibitor I-BET151 reduced levels of HOTAIR and restored the expression of several other glioblastoma downregulated lncRNAs. Moreover, bromodomain containing 4 (BRD4) is able to bind to HOTAIR, directly regulating its expression [133].
Different miRNAs can interact with HOTAIR during glioma progression. HOTAIR is the target of miR-326 and its silencing promotes their tumor-suppressive effects on glioma cell lines. Moreover, overexpressed miR-326 reduced the FGF1 expression, which played an oncogenic role in glioma by activating PI3K/AKT and MEK 1/2 pathways [157]. Bian et al. [158] showed that HOTAIR might act as an endogenous ‘sponge’ of miR-141, thereby regulating the de-repression of SKA2 (spindle and kinetochore associated complex subunit 2), a gene involved in cell cycle regulation. Both overexpression of miR-141 and knockdown of HOTAIR in a mouse model of human glioma resulted in significant reduction of tumor growth in vivo [158]. More recently, the interaction of HOTAIR and miR-219 has been described in glioma cells. HOTAIR silencing strongly induced the expression of miR-219 reducing cell proliferation and promoting apoptosis. Concomitantly, the protein expression level of Cyclin D1 declined significantly suggesting that HOTAIR can repress the proliferation and promote the apoptosis of glioblastoma cells by targeting miR-219 [33]. Circulating levels of HOTAIR have also been described in glioma patients and are associated with a poor prognosis [159]. Tan et al. detected high HOTAIR levels in serum samples from GBM patients and the serum-derived exosomes containing HOTAIR were significantly correlated with high grade brain tumors [134]. Regarding the analysis of polymorphisms in the HOTAIR sequence related with glioma susceptibility, Xavier-Magalhães et al. [160] reported a case-control study consisting of 177 Portuguese glioma patients and 199 cancer-free controls. HOTAIR SNPs rs920778 and rs12826786 do not play a significant role in glioma susceptibility but may be important prognostic factors in anaplastic oligodendroglioma [160].
Transcriptome analysis was carried out to evaluate the expression of the HOX genes and HOTAIR in several pediatric tumors such as teratoid rhabdoid tumors (ATRT), ependymomas, medulloblastomas, glioblastoma multiforme and juvenile pilocytic astrocytomas (JPAs). HOTAIR appeared overexpressed in ATRTs, medulloblastomas and JPAs, and downregulated in ependymomas [161]. Nevertheless, a more recent study analyzed HOTAIR expression in adult myxopapillary ependymoma (MPE), highlighting its overexpression compared with non-ependymoma spinal tumors [162].
HOTAIR upregulation has also been detected in medulloblastoma tissues and cell lines. In medulloblastoma cells, HOTAIR is able to negatively regulate miR-1 and miR-206 expression which can directly target YY1, a transcription factor described as a metastasis inducer. Finally, HOTAIR knockdown suppressed medulloblastoma cell proliferation, tumor growth, migration and invasion, and promoted cell apoptosis via the modulation of the miR-1/miR-206-YY1 axis, as well as EMT [163].

4.5. Soft Tissue Tumors

The role of HOTAIR in soft tissue sarcomas is mainly related to specific tumor types. Milhem et al. [164] selected primary and metastatic tumor samples from myxofibrosarcoma, synovial sarcoma, leiomyosarcoma and malignant fibrous histiocytoma sarcoma subtypes to analyze HOTAIR expression. In these tumor types, high levels of HOTAIR are correlated with a high probability of metastatic progression. In contrast, reduced expression of HOTAIR is correlated with a good response to treatment in terms of necrosis, suggesting that HOTAIR can be considered a useful predictor for metastatic risk and outcome of therapeutic treatments [164]. Many studies are focused on osteosarcoma (OS), highlighting the main role of HOTAIR as prognostic biomarker. HOTAIR is highly expressed in OS tumor samples and cells and its upregulation was closely correlated with advanced tumor stage along with highly histological grade. Furthermore, a high level of HOTAIR was significantly associated with shorter overall survival [165]. HOTAIR silencing leads to the downregulation of DNA methyltransferase 1 (DNMT1), promoting the decrease in global DNA methylation level. HOTAIR is able to induce the expression of DNMT1 through repressing miR-126, which is the negative regulator of DNMT1. Furthermore, HOTAIR silencing increases the sensibility of OS cells to DNMT1 inhibitor through regulating the viability and apoptosis of OS cells via HOTAIR-miR126-DNMT1-CDKN2A axis [135]. Wang et al. [166], using bioinformatics analysis, showed that HOTAIR can be targeted by the tumor suppressive gene miR-217. In addition, HOTAIR siRNA increased miR-217 expression and significantly repressed osteosarcoma cell growth, migration, invasion and induced cell apoptosis capacity. ZEB1 was identified as a downstream gene of miR-217, suggesting that HOTAIR can mediate osteosarcoma progression by upregulating ZEB1 expression via acting as a competitive endogenous RNA (ceRNA) via miR-217 [166]. Recently Wang et al. [167] described that LPS (a major component of Gram-negative bacteria) promotes tumor invasion, metastasis and EMT in osteosarcoma. HOTAIR expression markedly increases in LPS-induced EMT in osteosarcoma cells, such as TLR4 (Toll Like Receptor 4), which is the LPS receptor, suggesting that the effects of LPS on EMT in osteosarcoma cells is mediated via the TLR4/HOTAIR pathway [167]. Several studies reported that different polymorphisms, especially in intronic sequences as well as in promoter regions of HOTAIR, are often associated with its aberrant expression, patient prognosis and cancer susceptibility in different tumor types [99]. Zhou et al. [168] identified a SNP located in HOTAIR gene (rs7958904) that was significantly associated with decreased risk of OS. Furthermore, subjects with the rs7958904 CC genotype had significantly lower HOTAIR RNA levels than other genotypes [168].
Sporadic information on the role of HOTAIR has been associated with other types of sarcomas. In chondrosarcoma patients, the expression of HOTAIR is correlated with tumor stage and poor prognosis. HOTAIR knockdown led to growth inhibition via G0/G1 arrest and apoptosis in vitro and in vivo models of chondrosarcoma [169]. HOTAIR expression was increased in chondrosarcoma tissues compared with normal cartilage tissues and its aberrant expression was elevated in high-grade compared with low-grade chondrosarcoma tissues. In addition, overall survival time of patients with high expression of HOTAIR was significantly shorter than that of patients with low expression of HOTAIR [169]. HOTAIR is able to induce DNA methylation of miR-454-3p by recruiting EZH2 and DNMT1 in chondrosarcoma cells. Furthermore, signal transducer and activator of transcription 3 (STAT3) and autophagy-related gene 12 (ATG12) are targets of miR-454-3p, initiate HOTAIR deficiency-induced apoptosis and reduce autophagy [169]. More recently, Feng et al. [138] described the aberrant expression of HOTAIR also in synovial sarcoma (SS). Overexpression of HOTAIR correlates with histological grade, AJCC staging and distant metastasis [138]. In SS cells, HOTAIR silencing inhibited cellular proliferation, invasion and migration, promoting the G1/G0 phase of the cell cycle, and inhibiting the G2/S phase. Finally, HOTAIR knockdown increased miR-126 expression level and decreased the expression level of stromal cell-derived factor-1 (SDF-1) [170].

5. Conclusions

The understanding of the emerging role of lncRNAs in the main cell processes related with cancer development and progression represents a significant advance in oncology. LncRNAs are implicated in numerous biological processes such as cell cycle control, apoptosis, differentiation and epigenetic regulation of gene expression, becoming valid diagnostic and prognostic markers of human cancers [171,172]. Furthermore, the lncRNAs appear to be optimal biomarkers which are more stable in body fluids (urine, blood, saliva) and can be detected by using simple in situ (ISH) and molecular techniques [173]. The crucial role of lncRNAs, and in particular HOTAIR, is well documented in common human cancer such as breast, lung, urogenital and gastrointestinal cancer [172]. To date, the molecular characterization of the rarest cancers in the population has provided little information on potential new biomarkers and therapeutic molecular targets. However, much experimental evidence highlights the crucial role that HOTAIR plays in these tumors. The implementation of targeted functional studies could help to better understand how to interfere/block the aberrant activity of this biomarker, providing a new tool for the management of rare tumors. More recently, the design of small molecules able to specifically interfere with conserved RNA structures and to block HOTAIR protein complexes have proved more useful [174,175]. The direct or indirect block/inhibition of HOTAIR may represent a new and effective therapeutic strategy for rare cancer and tumors.

Funding

This research was funded by the Italian Ministry of Health.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

HOTAIRHOX transcript antisense RNA
EMTEpithelial-to-mesenchymal transition
LSCCLaryngeal tumors are squamous cell carcinomas
DNMTDNA methyltransferase enzyme
UCA1Urothelial Cancer Associated 1
NEAT1Nuclear paraspeckle assembly transcript1
PCAT19Protocadherin 19
TUG1Taurine-upregulated gene 1
HOXA11-ASHOXA11 antisense RNA
MNX1Motor Neuron And Pancreas Homeobox 1)-AS1
NPCNasopharyngeal carcinoma
AFAP1-AS1Actin filament-associated protein 1 antisense RNA1 ()
LETLow Expression in Tumor
SGCSalivary gland cancer
ADAMTS9AS2ADAM metallopeptidase thrombospondin type 1, 9 antisense RNA 2
SACCSalivary adenoid cystic carcinoma
MRPL23-AS1MRPL23 antisense RNA 1
MECMucoepidermoid carcinoma
OSosteosarcoma
MALAT1Metastasis Associated Lung Adenocarcinoma Transcript 1
HNF1AHNF1 Homeobox A-AS1
BCAR4Breast Cancer Anti-Estrogen Resistance 4
HULCHighly upregulated in liver cancer RNA
HOTTIPHOXA Distal Transcript Antisense RNA
MEG3Maternally expressed 3
EWS-AT1EWS RNA Binding Protein 1
PILRLSProliferation Interacting LncRNA in Retroperitoneal Liposarcoma
NENNeuroendocrine neoplasm
ISHIn situ hybridization
VGFVGF Nerve Growth Factor Inducible
PCA3Prostate cancer antigen 3
GISTsGastrointestinal stromal tumors
TGF-βTransforming growth factor-beta
ZEB1Zinc Finger E-Box Binding Homeobox 1
FENDRRFOXF1 Adjacent Non-Coding Developmental Regulatory RNA
CCDC26Coiled-coil domain-containing 26
GLUT4Glucose transporter member 4
TMZTemozolomide
NKD1NKD Inhibitor of WNT Signaling Pathway 1
TUSC7Tumor Suppressor Candidate 7
PRC2Polycomb repressive complex
LSD1Lysine-specific histone demethylase 1A
MREmiRNA recognition element
EZH2Zeste homolog 2
HCG4HLA Complex Group 4
EMX2OSEMX2 Opposite Strand/Antisense RNA
MAGEA2Melanoma Antigen A2 gene
E2F2Eukaryotic elongation factor 2
GRP78Glucose regulated protein 78
FASNFatty acid synthase
COX-2Cyclooxygenase-2
TLR4Toll Like Receptor 4
STAT3Signal transducer and activator of transcription 3
ATG12Autophagy-related gene 12
SDF-1Stromal cell-derived factor-1
GEP-NETGastroenteropancreatic Neuroendocrine Tumors
RESTRE1 Silencing Transcription Factor
PCDH10Protocadherin 10
DPP4Dipeptidyl peptidase 4
RASSF1Ras Association Domain Family Member 1
ALDH1A3Aldehyde Dehydrogenase 1 Family Member A3
ATG2BAutophagy-related protein 2 homolog B
BRD4Bromodomain Containing 4
SKA2Spindle And Kinetochore Associated Complex Subunit 2
ATRTTeratoid rhabdoid tumors
MPEMyxopapillary ependymoma

References

  1. Botta, L.; Gatta, G.; Trama, A.; Bernasconi, A.; Sharon, E.; Capocaccia, R.; Mariotto, A.B.; RARECARE Net Working Group. Incidence and survival of rare cancers in the US and Europe. Cancer Med. 2020, 9, 5632–5642. [Google Scholar] [CrossRef]
  2. AIRTUM Working Group; Busco, S.; Buzzoni, C.; Mallone, S.; Trama, A.; Castaing, M.; Bella, F.; Amodio, R.; Bizzoco, S.; Cassetti, T.; et al. Italian cancer figures—Report 2015: The burden of rare cancers in Italy. Epidemiol. Prev. 2016, 40, 1–120. [Google Scholar]
  3. International Cancer Genome Consortium. International network of cancer genome projects. Nature 2010, 464, 993–998. [Google Scholar] [CrossRef] [Green Version]
  4. Hutter, C.; Zenklusen, J.C. The Cancer genome atlas: Creating lasting value beyond its data. Cell 2018, 173, 283–285. [Google Scholar] [CrossRef]
  5. Botti, G.; Franco, R.; Cantile, M.; Ciliberto, G.; Ascierto, P.A. Tumor biobanks in translational medicine. J. Transl. Med. 2012, 10, 204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Botti, G.; Di Bonito, M.; Cantile, M. Organoid biobanks as a new tool for pre-clinical validation of candidate drug efficacy and safety. Int. J. Physiol. Pathophysiol. Pharmacol. 2021, 13, 17–21. [Google Scholar]
  7. Novikova, I.V.; Hennelly, S.P.; Sanbonmatsu, K.Y. Tackling Structures of Long Noncoding RNAs. Int. J. Mol. Sci. 2013, 14, 23672–23684. [Google Scholar] [CrossRef] [PubMed]
  8. Derrien, T.; Johnson, R.; Bussotti, G.; Tanzer, A.; Djebali, S.; Tilgner, H.; Guernec, G.; Martín, 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]
  9. Cabili, M.N.; Dunagin, M.C.; McClanahan, P.D.; Biaesch, A.; Padovan-Merhar, O.; Regev, A.; Rinn, J.L.; Raj, A. Localization and abundance analysis of human lncRNAs at single-cell and single-molecule resolution. Genome Biol. 2015, 16, 145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Statello, L.; Guo, C.J.; Chen, L.L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef]
  11. Xu, Y.; Zhang, X.; Hu, X.; Zhou, W.; Zhang, P.; Zhang, J.; Yang, S.; Liu, Y. The effects of lncRNA MALAT1 on proliferation, invasion and migration in colorectal cancer through regulating SOX9. Mol. Med. 2018, 24, 52. [Google Scholar] [CrossRef]
  12. Peng, F.; Wang, R.; Zhang, Y.; Zhao, Z.; Zhou, W.; Chang, Z.; Liang, H.; Zhao, W.; Qi, L.; Guo, Z.; et al. Differential expression analysis at the individual level reveals a lncRNA prognostic signature for lung adenocarcinoma. Mol. Cancer 2017, 16, 98. [Google Scholar] [CrossRef]
  13. Cantile, M.; Di Bonito, M.; Cerrone, M.; Collina, F.; De Laurentiis, M.; Botti, G. Long Non-Coding RNA HOTAIR in Breast Cancer Therapy. Cancers 2020, 12, 1197. [Google Scholar] [CrossRef]
  14. Kupcinskas, J. Small Molecules in Rare Tumors: Emerging Role of MicroRNAs in GIST. Int. J. Mol. Sci. 2018, 19, 397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Sarver, A.E.; Subramanian, S. MicroRNAs in the pathobiology of sarcomas. Lab. Investig. 2015, 95, 987–994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. WHO/IARC. Classification of Tumours, 4th ed.; El Naggar, A.K., Chan, J.K.C., Grandis, J.R., Takata, T., Slootweg, P., Eds.; WHO: Geneva, Switzerland, 2020; Volume 9. [Google Scholar]
  17. Li, W.; Chen, Y.; Niem, X. Regulatory Mechanisms of lncRNAs and Their Target Gene Signaling Pathways in Laryngeal Squamous Cell Carcinoma. Front. Pharmacol. 2020, 11, 1140. [Google Scholar] [CrossRef] [PubMed]
  18. Wu, T.; Qu, L.; He, G.; Tian, L.; Li, L.; Zhou, H.; Jin, Q.; Ren, J.; Wang, Y.; Wang, J.; et al. Regulation of laryngeal squamous cell cancer progression by the lncRNA H19/miR-148a-3p/DNMT1 axis. Oncotarget 2016, 7, 11553–11566. [Google Scholar] [CrossRef] [Green Version]
  19. Sun, S.; Gong, C.; Yuan, K. LncRNA UCA1 promotes cell proliferation, invasion and migration of laryngeal squamous cell carcinoma cells by activating Wnt/β-catenin signaling pathway. Exp. Ther. Med. 2019, 17, 1182–1189. [Google Scholar] [CrossRef] [Green Version]
  20. Liang, K.; Yang, Y.; Zha, D.; Yue, B.; Qiu, J.; Zhang, C. Overexpression of lncRNAsnaR is correlated with progression and predicts poor survival of laryngeal squamous cell carcinoma. J. Cell Biochem. 2019, 120, 8492–8498. [Google Scholar] [CrossRef]
  21. Wang, P.; Wu, T.; Zhou, H.; Jin, Q.; He, G.; Yu, H.; Xuan, L.; Wang, X.; Tian, L.; Sun, Y.; et al. Long noncoding RNA NEAT1 promotes laryngeal squamous cell cancer through regulating miR-107/CDK6 pathway. J. Exp. Clin. Cancer Res. 2016, 35, 22. [Google Scholar] [CrossRef] [Green Version]
  22. Xu, S.; Guo, J.; Zhang, W. LncRNA PCAT19 promotes the proliferation of laryngocarcinoma cells via modulation of the miR-182/PDK4 axis. J. Cell. Biochem. 2019, 120, 12810–12821. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, L.; Cao, H.; Chi, W.; Meng, W.; Cui, W.; Guo, W.; Wang, B. Expression profile analysis identifies the long non-coding RNA landscape and the potential carcinogenic functions of LINC00668 in laryngeal squamous cell carcinoma. Gene 2019, 687, 47–55. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, Z.; Wang, X.; Cao, S.; Han, X.; Wang, Z.; Zhao, X.; Liu, X.; Li, G.; Pan, X.; Lei, D. The Long Noncoding RNA TUG1 Promotes Laryngeal Cancer Proliferation and Migration. Cell. Physiol. Biochem. 2018, 49, 2511–2520. [Google Scholar] [CrossRef]
  25. Qu, L.; Jin, M.; Yang, L.; Sun, C.; Wang, P.; Li, Y.; Tian, L.; Liu, M.; Sun, Y. Expression of long non-coding RNA HOXA11-AS is correlated with progression of laryngeal squamous cell carcinoma. Am. J. Transl. Res. 2018, 10, 573–580. [Google Scholar]
  26. Huang, D.N.; Liu, H.W.; Li, Z.D. Expression of lncRNA-ATB in laryngeal carcinoma and its relationship with prognosis. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 11148–11153. [Google Scholar] [PubMed]
  27. Zhang, G.; Fan, E.; Zhong, Q.; Feng, G.; Shuai, Y.; Wu, M.; Chen, Q.; Gou, X. Identificationand potential mechanisms of a 4-lncRNA signature that predicts prognosis in patients with laryngeal cancer. Hum. Genom. 2019, 13, 36. [Google Scholar] [CrossRef] [PubMed]
  28. Hu, J.C.; Wang, Q.; Jiang, L.X.; Cai, L.; Zhai, H.Y.; Yao, Z.W.; Zhang, M.L.; Feng, Y. Effect of long non-coding RNA AOC4P on gastrointestinal stromal tumor cells. OncoTargets Ther. 2018, 11, 6259–6269. [Google Scholar] [CrossRef] [Green Version]
  29. Xie, S.; Yu, X.; Li, Y.; Ma, H.; Fan, S.; Chen, W.; Pan, G.; Wang, W.; Zhang, H.; Li, J.; et al. Upregulation of lncRNA ADAMTS9-AS2 Promotes Salivary Adenoid Cystic Carcinoma Metastasis via PI3K/Akt and MEK/Erk Signaling. Mol. Ther. 2018, 26, 2766–2778. [Google Scholar] [CrossRef] [Green Version]
  30. Bo, H.; Gong, Z.; Zhang, W.; Li, X.; Zeng, Y.; Liao, Q.; Chen, P.; Shi, L.; Lian, Y.; Jing, Y.; et al. Upregulated long non-coding RNA ATB expression is associated with progression and poor prognosis of nasopharyngeal carcinoma. Oncotarget 2015, 6, 20404–20418. [Google Scholar] [CrossRef] [Green Version]
  31. Chen, F.; Mo, J.; Zhang, L. Long noncoding RNA BCAR4 promotes osteosarcoma progression through activating GLI2-dependent gene transcription. Tumour Biol. 2016, 37, 13403–13412. [Google Scholar] [CrossRef]
  32. Cao, K.; Li, M.; Miao, J.; Lu, X.; Kang, X.; Zhu, H.; Du, S.; Li, X.; Zhang, Q.; Guan, W.; et al. CCDC26 knockdown enhances resistance of gastrointestinal stromal tumor cells to imatinib by interacting with c-KIT. Am. J. Transl Res. 2018, 10, 274–282. [Google Scholar] [PubMed]
  33. Li, H.; Guan, C. HOTAIR inhibits the proliferation of glioblastoma cells by targeting miR-219. Cancer Biomark. 2020, 28, 41–47. [Google Scholar] [CrossRef] [PubMed]
  34. Yin, X.; Yin, Y.; Dai, L.; Shen, C.; Chen, N.; Li, J.; Cai, Z.; Jiang, Z.; Wang, J.; Zhao, Z.; et al. Integrated analysis of long non-coding RNAs and mRNAs associated with malignant transformation of gastrointestinal stromal tumors. Cell Death Dis. 2021, 12, 669. [Google Scholar] [CrossRef] [PubMed]
  35. Marques Howarth, M.; Simpson, D.; Ngok, S.P.; Nieves, B.; Chen, R.; Siprashvili, Z.; Vaka, D.; Breese, M.R.; Crompton, B.D.; Alexe, G.; et al. Long noncoding RNA ewsat1-mediated gene repression facilitates ewing sarcoma oncogenesis. J. Clin. Investig. 2014, 124, 5275–5290. [Google Scholar] [CrossRef] [PubMed]
  36. Zhuang, K.; Wu, Q.; Jin, C.S.; Yuan, H.J.; Cheng, J.Z. Long non-coding RNA HNF1A-AS is upregulated and promotes cell proliferation and metastasis in nasopharyngeal carcinoma. Cancer Biomark. 2016, 16, 291–300. [Google Scholar] [CrossRef]
  37. Ji, M.; Yao, Y.; Liu, A.; Shi, L.; Chen, D.; Tang, L.; Yang, G.; Liang, X.; Peng, J.; Shao, C. lncRNA H19 binds VGF and promotes pNEN progression via PI3K/AKT/CREB signalling. Endocr. Relat. Cancer 2019, 26, 643–658. [Google Scholar] [CrossRef]
  38. Fazi, B.; Garbo, S.; Toschi, N.; Mangiola, A.; Lombari, M.; Sicari, D. The Lncrna H19 Positively Affects the Tumorigenic Properties of Glioblastoma Cells and Contributes to NKD1 Repression Through the Recruitment of EZH2 on Its Promoter. Oncotarget 2018, 9, 15512. [Google Scholar] [CrossRef] [PubMed]
  39. Duan, S.; Li, M.; Wang, Z.; Wang, L.; Liu, Y. H19 Induced by Oxidative Stress Confers Temozolomide Resistance in Human Glioma Cells Via Activating NF-κb Signaling. OncoTargets Ther. 2018, 11, 6395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Zhao, H.; Hou, W.; Tao, J.; Zhao, Y.; Wan, G.; Ma, C.; Xu, H. Upregulation of lncRNA HNF1A-AS1 promotes cell proliferation and metastasis in osteosarcoma through activation of the Wnt/β-catenin signaling pathway. Am. J. Transl. Res. 2016, 8, 3503–3512. [Google Scholar]
  41. Li, F.; Cao, L.; Hang, D.; Wang, F.; Wang, Q. Long non-coding RNA HOTTIP is upregulated and associated with poor prognosis in patients with osteosarcoma. Int. J. Clin. Exp. Pathol. 2015, 8, 11414–11420. [Google Scholar]
  42. Lin, F.J.; Lin, X.D.; Xu, L.Y.; Zhu, S.Q. Long Noncoding RNA HOXA11-AS Modulates the Resistance of Nasopharyngeal Carcinoma Cells to Cisplatin via miR-454-3p/c-Met. Mol. Cells 2020, 43, 856–869. [Google Scholar]
  43. Li, H.; Huang, J.; Yu, S.; Li, H.; Zhou, Y.; Wu, Q. HOXA11-AS induces cisplatin resistance by modulating the microRNA-98/PBX3 axis in nasopharyngeal carcinoma. Oncol. Lett. 2021, 21, 493. [Google Scholar] [CrossRef]
  44. Yuan, F.; Lou, Z.; Zhou, Z.; Yan, X. Long non-coding RNA LET promotes nasopharyngeal carcinoma cell cisplatin resistance via the miR-454/USP47 axis. Int. J. Mol. Med. 2021, 47, 54. [Google Scholar] [CrossRef]
  45. Cao, C.; Zhou, S.; Hu, J. Long noncoding RNA MAGI2-AS3/miR-218-5p/GDPD5/SEC61A1 axis drives cellular proliferation and migration and confers cisplatin resistance in nasopharyngeal carcinoma. Int. Forum Allergy Rhinol. 2020, 10, 1012–1023. [Google Scholar] [CrossRef]
  46. Dong, Y.; Liang, G.; Yuan, B.; Yang, C.; Gao, R.; Zhou, X. MALAT1 promotes the proliferation and metastasis of osteosarcoma cells by activating the PI3k/Akt pathway. Tumour Biol. 2015, 36, 1477–1486. [Google Scholar] [CrossRef] [PubMed]
  47. Chu, Y.H.; Hardin, H.; Jens Eickhoff, J.; Lloyd, R.V. In Situ Hybridization Analysis of Long Non-coding RNAs MALAT1 and HOTAIR in Gastroenteropancreatic Neuroendocrine Neoplasms. Endocr. Pathol. 2019, 30, 56–63. [Google Scholar] [CrossRef] [PubMed]
  48. Badalamenti, G.; Barraco, N.; Incorvaia, L.; Galvano, A.; Fanale, D.; Cabibi, D.; Calò, V.; Currò, G.; Bazan, V.; Russo, A. Are Long Noncoding RNAs New Potential Biomarkers in Gastrointestinal Stromal Tumors (GISTs)? The Role of H19 and MALAT1. J. Oncol. 2019, 2019, 5458717. [Google Scholar] [CrossRef] [Green Version]
  49. Vassallo, I.; Zinn, P.; Lai, M.; Rajakannu, P.; Hamou, M.F.; Hegi, M.E. WIF1 re-expression in glioblastoma inhibits migration through attenuation of non-canonical WNT signaling by downregulating the lncRNA MALAT1. Oncogene 2016, 35, 12–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Li, H.; Yuan, X.; Yan, D.; Li, D.; Guan, F.; Dong, Y. Long Non-Coding RNA MALAT1 Decreases the Sensitivity of Resistant Glioblastoma Cell Lines to Temozolomide. Cell Physiol. Biochem. 2017, 42, 1192–1201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Chen, W.; Xu, X.-K.; Li, J.-L.; Kong, K.-K.; Li, H.; Chen, C. MALAT1 is a Prognostic Factor in Glioblastoma Multiforme and Induces Chemoresistance to Temozolomide Through Suppressing miR-203 and Promoting Thymidylate Synthase Expression. Oncotarget 2017, 8, 22783. [Google Scholar] [CrossRef] [Green Version]
  52. Cai, T.; Liu, Y.; Xiao, J. Long Noncoding RNA MALAT 1 Knockdown Reverses Chemoresistance to Temozolomide Via Promoting Micro RNA-101 in Glioblastoma. Cancer Med. 2018, 7, 1404–1415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Voce, D.J.; Bernal, G.M.; Wu, L.; Crawley, C.D.; Zhang, W.; Mansour, N.M. Temozolomide Treatment Induces Lncrna MALAT1 in an NF-κb and p53 Codependent Manner in Glioblastoma. Cancer Res. 2019, 79, 2536–2548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zhong, Q.; Chen, Y.; Chen, Z. LncRNA MINCR regulates irradiation resistance in nasopharyngeal carcinoma cells via the microRNA-223/ZEB1 axis. Cell Cycle 2020, 19, 53–66. [Google Scholar] [CrossRef]
  55. Chen, C.W.; Fu, M.; Du, Z.H.; Zhao, F.; Yang, W.W.; Xu, L.H.; Li, S.L.; Ge, X.Y. Long Noncoding RNA NEAT1 Promotes Adenoid Cystic Carcinoma Lung Metastasis. Cancer Res. 2020, 80, 2273–2285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Xu, K.-L.; Xu, Q.-S.; Wu, Z.-X.; Xu, S.-J.; Shen, J. Lncrna NEAT1 is Involved in Temozolomide Resistance by Regulating MGMT in Glioblastoma Multiforme. Clin. Surg. Res. Commun. 2018, 2, 24–30. [Google Scholar]
  57. Lu, H.; Han, N.; Xu, W.; Zhu, Y.; Liu, L.; Liu, S.; Yang, W. Screening and bioinformatics analysis of mRNA, long non-coding RNA and circular RNA expression profiles in mucoepidermoid carcinoma of salivary gland. Biochem. Biophys. Res. Commun. 2019, 508, 66–71. [Google Scholar] [CrossRef]
  58. Tang, J.; Fu, C.; Li, Y.; Chen, S.; Jiang, X.; Xu, W.; Xie, H. Long Noncoding RNA PCA3 Promotes the Disease Progression in Nasopharyngeal Carcinoma by Targeting miR-203. Biomed. Res. Int. 2021, 2021, 9850928. [Google Scholar] [CrossRef]
  59. Narayanan, D.; Mandal, R.; Hardin, H.; Chanana, V.; Schwalbe, M.; Rosenbaum, J.; Buehler, D.; Lloyd, R.V. Long Non-coding RNAs in Pulmonary Neuroendocrine Neoplasms. Endocr. Pathol. 2020, 31, 254–263. [Google Scholar] [CrossRef]
  60. Shao, Y.; Zhang, Y.; Hou, Y.; Tong, H.; Zhuang, R.; Ji, Z.; Wang, B.; Zhou, Y.; Lu, W. A novel long noncoding RNA PILRLS promote proliferation through TCL1A by activing MDM2 in retroperitoneal liposarcoma. Oncotarget 2017, 8, 13971–13978. [Google Scholar] [CrossRef] [Green Version]
  61. Tian, X.; Liu, Y.; Wang, Z.; Wu, S. lncRNA SNHG8 promotes aggressive behaviors of nasopharyngeal carcinoma via regulating miR-656-3p/SATB1 axis. Biomed. Pharmacother. 2020, 131, 110564. [Google Scholar] [CrossRef]
  62. Yi, T.W.; Lv, X.X.; Fan, H.; Zan, N.; Su, X.D. LncRNA SNHG15 promotes the proliferation of nasopharyngeal carcinoma via sponging miR-141-3p to upregulate KLF9. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 6744–6751. [Google Scholar]
  63. Hu, J.; Pan, J.; Luo, Z.; Duan, Q.; Wang, D. Long non-coding RNA FOXD3-AS1 silencing exerts tumor suppressive effects in nasopharyngeal carcinoma by downregulating FOXD3 expression via microRNA-185-3p upregulation. Cancer Gene Ther. 2021, 28, 602–618. [Google Scholar] [CrossRef] [PubMed]
  64. Ma, B.; Li, M.; Zhang, L.; Huang, M.; Lei, J.B.; Fu, G.H.; Liu, C.X.; Lai, Q.W.; Chen, Q.Q.; Wang, Y.L. Upregulation of long non-coding RNA TUG1 correlates with poor prognosis and disease status in osteosarcoma. Tumour Biol. 2016, 37, 4445–4455. [Google Scholar] [CrossRef] [PubMed]
  65. Shang, C.; Tang, W.; Pan, C.; Hu, X.; Hong, Y. Long Non-Coding RNA TUSC7 Inhibits Temozolomide Resistance by Targeting miR-10a in Glioblastoma. Cancer Chemother. Pharmacol. 2018, 81, 671–678. [Google Scholar] [CrossRef] [PubMed]
  66. Li, W.; Xie, P.; Ruan, W.H. Overexpression of lncRNA UCA1 promotes osteosarcoma progression and correlates with poor prognosis. J. Bone Oncol. 2016, 5, 80–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Sun, Q.; Liu, H.; Li, L.; Zhang, S.; Liu, K.; Liu, Y.; Yang, C. Long noncoding RNA-LET, which is repressed by EZH2, inhibits cell proliferation and induces apoptosis of nasopharyngeal carcinoma cell. Med. Oncol. 2015, 32, 226. [Google Scholar] [CrossRef] [PubMed]
  68. Tian, Z.Z.; Guo, X.J.; Zhao, Y.M.; Fang, Y. Decreased expression of long non-coding RNA MEG3 acts as a potential predictor biomarker in progression and poor prognosis of osteosarcoma. Int. J. Clin. Exp. Pathol. 2015, 8, 15138–15142. [Google Scholar] [PubMed]
  69. Cong, M.; Li, J.; Jing, R.; Li, Z. Long non-coding RNA tumor suppressor candidate 7 functions as a tumor suppressor and inhibits proliferation in osteosarcoma. Tumour Biol. 2016, 37, 9441–9450. [Google Scholar] [CrossRef]
  70. Li, X.; Lin, Y.; Yang, X.; Wu, X.; He, X. Long noncoding RNA H19 regulates EZH2 expression by interacting with miR-630 and promotes cell invasion in nasopharyngeal carcinoma. Biochem. Biophys. Res. Commun. 2016, 13, 913–919. [Google Scholar] [CrossRef] [PubMed]
  71. Han, Y.E.; Tao, J.M.; Wang, S.X.; Ju, X.; Song, Z.Y. Long non-coding RNA HEIH modulates CDK8 expression by inhibiting miR-193a-5p to accelerate nasopharyngeal carcinoma progression. Eur. Rev. Med. Pharmacol. Sci. 2021, 25, 770–778. [Google Scholar]
  72. Dai, B.Q.; Jiang, X.; Feng, L.C. LncRNA TUG1 regulates miR-495 expression to promote migration and invasion of nasopharyngeal carcinoma cells through junctional adhesion molecule A. Kaohsiung J. Med. Sci. 2021, 37, 361–370. [Google Scholar] [CrossRef]
  73. Feng, M.B.; Li, G.H.; Dou, F.F. Long-chain non-coding RNA LOC554202 promotes proliferation, migration, and invasion of nasopharyngeal carcinoma cells by binding to microRNA-31 expression and regulating RhoA expression. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 10550–10556. [Google Scholar] [PubMed]
  74. Yue, W.J.; Wang, Y.; Li, W.Y.; Wang, Z.D. LINC00887 regulates the proliferation of nasopharyngeal carcinoma via targeting miRNA-203b-3p to upregulate NUP205. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 8863–8870. [Google Scholar]
  75. Li, H.; Huang, J.; Yu, S.; Lou, Z. Long Non-Coding RNA DLEU1 Up-Regulates BIRC6 Expression by Competitively Sponging miR-381-3p to Promote Cisplatin Resistance in Nasopharyngeal Carcinoma. OncoTargets Ther. 2020, 13, 2037–2045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Cui, Z.; Pu, T.; Zhang, Y.; Wang, J.; Zhao, Y. Long non-coding RNA LINC00346 contributes to cisplatin resistance in nasopharyngeal carcinoma by repressing miR-342-5p. Open Biol. 2020, 10, 190286. [Google Scholar] [CrossRef] [PubMed]
  77. Zheng, Z.Q.; Li, Z.X.; Guan, J.L.; Liu, X.; Li, J.Y.; Chen, Y.; Lin, L.; Kou, J.; Lv, J.W.; Zhang, L.L.; et al. Long Noncoding RNA TINCR-Mediated Regulation of Acetyl-CoA Metabolism Promotes Nasopharyngeal Carcinoma Progression and Chemoresistance. Cancer Res. 2020, 80, 5174–5188. [Google Scholar] [CrossRef] [PubMed]
  78. Zhu, X.; Liu, L.; Wang, Y.; Cong, J.; Lin, Z.; Wang, Y.; Liu, Q.; Wang, L.; Yang, B.; Li, T. lncRNA MIAT/HMGB1 Axis Is Involved in Cisplatin Resistance via Regulating IL6-Mediated Activation of the JAK2/STAT3 Pathway in Nasopharyngeal Carcinoma. Front. Oncol. 2021, 11, 651693. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, H.; Zheng, W.; Chen, Q.; Zhou, Y.; Pan, Y.; Zhang, J.; Bai, Y.; Shao, C. lncRNA CASC19 Contributes to Radioresistance of Nasopharyngeal Carcinoma by Promoting Autophagy via AMPK-mTOR Pathway. Int. J. Mol. Sci. 2021, 22, 1407. [Google Scholar] [CrossRef]
  80. Peng, J.; Liu, F.; Zheng, H.; Wu, Q.; Liu, S. IncRNA ZFAS1 contributes to the radioresistance of nasopharyngeal carcinoma cells by sponging hsa-miR-7-5p to upregulate ENO2. Cell Cycle 2021, 20, 126–141. [Google Scholar] [CrossRef]
  81. Speight, P.M.; Barrett, A.W. Salivary gland tumours: Diagnostic challenges and an update on the latest WHO classification. Diagn. Histopathol. 2020, 26, 147–158. [Google Scholar] [CrossRef]
  82. Kallen, M.E.; Hornick, J.L. The 2020 WHO Classification: What’s New in Soft Tissue Tumor Pathology? Am. J. Surg. Pathol. 2021, 45, e1–e23. [Google Scholar] [CrossRef] [PubMed]
  83. Sun, X.H.; Yang, L.B.; Geng, X.L.; Wang, R.; Zhang, Z.C. Increased expression of lncRNA HULC indicates a poor prognosis and promotes cell metastasis in osteosarcoma. Int. J. Clin. Exp. Pathol. 2015, 8, 2994–3000. [Google Scholar] [PubMed]
  84. Taal, B.G.; Visser, O. Epidemiology of neuroendocrine tumours. Neuroendocrinology 2004, 8, 3–7. [Google Scholar] [CrossRef] [PubMed]
  85. Oronsky, B.; Ma, P.C.; Morgensztern, D.; Carter, C.A. Nothing But NET: A Review of Neuroendocrine Tumors and Carcinomas. Neoplasia 2017, 19, 991–1002. [Google Scholar] [CrossRef]
  86. Luo, J.; Wang, K.; Yeh, S.; Sun, Y.; Liang, L.; Xiao, Y.; Xu, W.; Niu, Y.; Liang Cheng, L.; Maity, S.N.; et al. LncRNA-p21 alters the antiandrogen enzalutamide-induced prostate cancer neuroendocrine differentiation via modulating the EZH2/STAT3 signaling. Nat. Commun. 2019, 10, 2571. [Google Scholar] [CrossRef] [Green Version]
  87. Al-Share, B.; Alloghbi, A.; Al Hallak, M.N.; Uddin, H.; Azmi, A.; Mohammad, R.M.; Kim, S.H.; Shields, A.F.; Philip, P.A. Gastrointestinal stromal tumor: A review of current and emerging therapies. Cancer Metastasis Rev. 2021, 40, 625–641. [Google Scholar] [CrossRef]
  88. Gyvyte, U.; Kupcinskas, J.; Juzenas, S.; Inciuraite, R.; Poskiene, L.; Salteniene, V.; Link, A.; Fassan, M.; Franke, A.; Kupcinskas, L.; et al. Identification of long intergenic non-coding RNAs (lincRNAs) deregulated in gastrointestinal stromal tumors (GISTs). PLoS ONE 2018, 13, 0209342. [Google Scholar]
  89. Reni, M.; Mazza, E.; Zanon, S.; Gatta, G.; Vecht, C.J. Central nervous system gliomas. Crit. Rev. Oncol. Hematol. 2017, 113, 213–234. [Google Scholar] [CrossRef]
  90. Rinn, J.L.; Kertesz, M.; Wang, J.K.; Squazzo, S.L.; Xu, X.; Brugmann, S.A.; Goodnough, L.H.; Helms, J.A.; Farnham, P.J.; Segal, E. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell 2007, 129, 1311–1323. [Google Scholar] [CrossRef] [Green Version]
  91. Ding, A.; Mandal, S.S. LncRNA HOTAIR: A master regulator of chromatin dynamics and cancer. Biochim. Biophys. Acta 2015, 1856, 151–164. [Google Scholar]
  92. Majello, B.; Gorini, F.; Saccà, C.D.; Amente, S. Expanding the role of the histone lysine-specific demethylase LSD1 in cancer. Cancers 2019, 11, 324. [Google Scholar] [CrossRef] [Green Version]
  93. Yoon, J.H.; Abdelmohsen, K.; Kim, J.; Yang, X.; Martindale, J.L.; Tominaga-Yamanaka, K.; White, E.J.; Orjalo, A.V.; Rinn, J.L.; Kreft, S.G. Scaffold function of long non-coding RNA HOTAIR in protein ubiquitination. Nat. Commun. 2013, 4, 1–14. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, M.; Zhang, H.; Li, Y.; Wang, R.; Li, Y.; Zhang, H.; Ren, D.; Liu, H.; Kang, C.; Chen, J. HOTAIR, a long noncoding RNA, is a marker of abnormal cell cycle regulation in lung cancer. Cancer Sci. 2018, 109, 2717–2733. [Google Scholar] [CrossRef] [Green Version]
  95. Ding, W.; Ren, J.; Ren, H.; Wang, D. Long Noncoding RNA HOTAIR Modulates MiR-206-mediated Bcl-w Signaling to Facilitate Cell Proliferation in Breast Cancer. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Han, B.; Peng, X.; Cheng, D.; Zhu, Y.; Du, J.; Li, J.; Yu, X. Delphinidin suppresses breast carcinogenesis through the HOTAIR/microRNA-34a axis. Cancer Sci. 2019, 110, 3089–3097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Cantile, M.; Di Bonito, M.; Tracey De Bellis, M.; Botti, G. Functional Interaction among lncRNA HOTAIR and MicroRNAs in Cancer and Other Human Diseases. Cancers 2021, 13, 570. [Google Scholar] [CrossRef] [PubMed]
  98. Xu, S.; Kong, D.; Chen, Q.; Ping, Y.; Pang, D. Oncogenic long noncoding RNA landscape in breast cancer. Mol. Cancer 2017, 16, 129. [Google Scholar] [CrossRef] [Green Version]
  99. Botti, G.; Collina, F.; Scognamiglio, G.; Aquino, G.; Cerrone, M.; Liguori, G.; Gigantino, V.; Malzone, M.G.; Cantile, M. LncRNA HOTAIR Polymorphisms Association with Cancer Susceptibility in Different Tumor Types. Curr. Drug Targets 2018, 19, 1220–1226. [Google Scholar] [CrossRef]
  100. Qu, X.; Alsager, S.; Zhuo, Y.; Shan, B. HOX transcript antisense RNA (HOTAIR) in cancer. Cancer Lett. 2019, 454, 90–97. [Google Scholar] [CrossRef] [PubMed]
  101. Gupta, R.A.; Shah, N.; Wang, K.C.; Kim, J.; Horlings, H.M.; Wong, D.J.; Tsai, M.C.; Hung, T.; Argani, P.; Rinn, J.L. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature 2010, 464, 1071–1076. [Google Scholar] [CrossRef] [PubMed]
  102. Gökmen-Polar, Y.; Vladislav, I.T.; Neelamraju, Y.; Janga, S.C.; Badve, S. Prognostic impact of HOTAIR expression is restricted to ER-negative breast cancers. Sci. Rep. 2015, 5, 8765. [Google Scholar] [CrossRef] [Green Version]
  103. Sørensen, K.P.; Thomassen, M.; Tan, Q.; Bak, M.; Cold, S.; Burton, M.; Larsen, M.J.; Kruse, T.A. Long non-coding RNA HOTAIR is an independent prognostic marker of metastasis in estrogen receptor-positive primary breast cancer. Breast Cancer Res. Treat. 2013, 142, 529–536. [Google Scholar] [CrossRef] [PubMed]
  104. Zhuang, Y.; Nguyen, H.T.; Burow, M.E.; Zhuo, Y.; El-Dahr, S.S.; Yao, X.; Cao, S.; Flemington, E.K.; Nephew, K.P.; Fang, F. Elevated expression of long intergenic non-coding RNA HOTAIR in a basal-like variant of MCF-7 breast cancer cells. Mol. Carcinog. 2015, 54, 1656–1667. [Google Scholar] [CrossRef] [Green Version]
  105. Collina, F.; Aquino, G.; Brogna, M.; Cipolletta, S.; Buonfanti, G.; De Laurentiis, M.; Di Bonito, M.; Cantile, M.; Botti, G. LncRNA HOTAIR upregulation is strongly related with lymph nodes metastasis and LAR subtype of Triple Negative Breast Cancer. J. Cancer 2019, 10, 2018–2024. [Google Scholar] [CrossRef] [Green Version]
  106. Pádua Alves, C.; Fonseca, A.S.; Muys, B.R.; de Barros, E.; Lima Bueno, R.; Bürger, M.C.; de Souza, J.E.; Valente, V.; Zago, M.A.; Silva, W.A., Jr. Brief report: The lincRNA Hotair is required for epithelial-to-mesenchymal transition and stemness maintenance of cancer cell lines. Stem Cells 2013, 31, 2827–2832. [Google Scholar] [CrossRef]
  107. Ren, Y.; Jia, H.H.; Xu, Y.Q.; Zhou, X.; Zhao, X.H.; Wang, Y.F.; Song, X.; Zhu, Z.Y.; Sun, T.; Dou, Y. Paracrine and epigenetic control of CAF-induced metastasis: The role of HOTAIR stimulated by TGF-ß1 secretion. Mol. Cancer 2018, 17, 5. [Google Scholar] [CrossRef]
  108. Zhang, H.; Cai, K.; Wang, J.; Wang, X.; Cheng, K.; Shi, F.; Jiang, L.; Zhang, Y.; Dou, J. MiR-7, inhibited indirectly by lincRNA HOTAIR, directly inhibits SETDB1 and reverses the EMT of breast cancer stem cells by downregulating the STAT3 pathway. Stem Cells 2014, 32, 2858–2868. [Google Scholar] [CrossRef] [PubMed]
  109. Deng, J.; Yang, M.; Jiang, R.; An, N.; Wang, X.; Liu, B. Long Non-Coding RNA HOTAIR Regulates the Proliferation, Self-Renewal Capacity, Tumor Formation and Migration of the Cancer Stem-Like Cell (CSC) Subpopulation Enriched from Breast Cancer Cells. PLoS ONE 2017, 12, e0170860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Zhang, L.; Song, X.; Wang, X.; Xie, Y.; Wang, Z.; Xu, Y.; You, X.; Liang, Z.; Cao, H. Circulating DNA of HOTAIR in serum is a novel biomarker for breast cancer. Breast Cancer Res. Treat. 2015, 152, 199–208. [Google Scholar] [CrossRef]
  111. Tang, S.; Zheng, K.; Tang, Y.; Li, Z.; Zou, T.; Liu, D. Overexpression of serum exosomal HOTAIR is correlated with poor survival and poor response to chemotherapy in breast cancer patients. J. Biosci. 2019, 44, 37. [Google Scholar] [CrossRef]
  112. Luo, Z.F.; Zhao, D.; Li, X.Q.; Cui, Y.X.; Ma, N.; Lu, C.X.; Liu, M.Y.; Zhou, Y. Clinical significance of HOTAIR expression in colon cancer. World J. Gastroenterol. 2016, 22, 5254–5259. [Google Scholar] [CrossRef] [PubMed]
  113. Zhao, W.; Dong, S.; Duan, B.; Chen, P.; Shi, L.; Gao, H.; Qi, H. HOTAIR is a predictive and prognostic biomarker for patients with advanced gastric adenocarcinoma receiving fluorouracil and platinum combination chemotherapy. Am. J. Transl. Res. 2015, 7, 1295–1302. [Google Scholar] [PubMed]
  114. Ishibashi, M.; Kogo, R.; Shibata, K.; Sawada, G.; Takahashi, Y.; Kurashige, J.; Akiyoshi, S.; Sasaki, S.; Iwaya, T. Clinical significance of the expression of long non-coding RNA HOTAIR in primary hepatocellular carcinoma. Oncol. Rep. 2013, 29, 946–950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Yang, Z.; Zhou, L.; Wu, L.M.; Lai, M.C.; Xie, H.Y.; Zhang, F.; Zheng, S.S. Overexpression of long non-coding RNA HOTAIR predicts tumor recurrence in hepatocellular carcinoma patients following liver transplantation. Ann. Surg. Oncol. 2011, 18, 1243–1250. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, A.; Zhao, J.C.; Kim, J.; Fong, K.W.; Yang, Y.A.; Chakravarti, D.; Mo, Y.Y.; Yu, J. LncRNA HOTAIR Enhances the Androgen-Receptor-Mediated Transcriptional Program and Drives Castration-Resistant Prostate Cancer. Cell Rep. 2015, 13, 209–221. [Google Scholar] [CrossRef] [Green Version]
  117. Yan, T.H.; Lu, S.W.; Huang, Y.Q.; Que, G.B.; Chen, J.H.; Chen, Y.P.; Zhang, H.B.; Liang, X.L.; Jiang, J.H. Upregulation of the long noncoding RNA HOTAIR predicts recurrence in stage Ta/T1 bladder cancer. Tumour Biol. 2014, 35, 10249–10257. [Google Scholar] [CrossRef]
  118. Shang, C.; Guo, Y.; Zhang, H.; Xue, Y.X. Long noncoding RNA HOTAIR is a prognostic biomarker and inhibits chemosensitivity to doxorubicin in bladder transitional cell carcinoma. Cancer Chemother. Pharmacol. 2016, 77, 507–513. [Google Scholar] [CrossRef]
  119. Berrondo, C.; Flax, J.; Kucherov, V.; Siebert, A.; Osinski, T.; Rosenberg, A.; Fucile, C.; Richheimer, S.; Beckham, C.J. Expression of the Long Non-Coding RNA HOTAIR Correlates with Disease Progression in Bladder Cancer and Is Contained in Bladder Cancer Patient Urinary Exosomes. PLoS ONE 2016, 11, e0147236. [Google Scholar]
  120. Qiu, J.J.; Lin, Y.Y.; Ye, L.C.; Ding, J.X.; Feng, W.W.; Jin, H.Y.; Zhang, Y.; Li, Q.; Hua, K.Q. Overexpression of long non-coding RNA HOTAIR predicts poor patient prognosis and promotes tumor metastasis in epithelial ovarian cancer. Gynecol. Oncol. 2014, 134, 121–128. [Google Scholar] [CrossRef]
  121. Huang, L.; Liao, L.M.; Liu, A.W.; Wu, J.B.; Cheng, X.L.; Lin, J.X.; Zheng, M. Overexpression of long noncoding RNA HOTAIR predicts a poor prognosis in patients with cervical cancer. Arch. Gynecol. Obstet. 2014, 290, 717–723. [Google Scholar] [CrossRef] [PubMed]
  122. Li, J.; Wang, Y.; Yu, J.; Dong, R.; Qiu, H. A high level of circulating HOTAIR is associated with progression and poor prognosis of cervical cancer. Tumour Biol. 2015, 36, 1661–1665. [Google Scholar] [CrossRef] [PubMed]
  123. He, X.; Bao, W.; Li, X.; Chen, Z.; Che, Q.; Wang, H.; Wan, X.P. The long non-coding RNA HOTAIR is upregulated in endometrial carcinoma and correlates with poor prognosis. Int. J. Mol. Med. 2014, 33, 325–333. [Google Scholar] [CrossRef] [Green Version]
  124. Sun, M.Y.; Zhu, J.Y.; Zhang, C.Y.; Zhang, M.; Song, Y.N.; Rahman, K.; Zhang, L.J.; Zhang, H. Autophagy regulated by lncRNA HOTAIR contributes to the cisplatin-induced resistance in endometrial cancer cells. Biotechnol. Lett. 2017, 39, 1477–1484. [Google Scholar] [CrossRef] [PubMed]
  125. Nakagawa, T.; Endo, H.; Yokoyama, M.; Abe, J.; Tamai, K.; Tanaka, N.; Sato, I.; Takahashi, S.; Kondo, T.; Satoh, K. Large noncoding RNA HOTAIR enhances aggressive biological behavior and is associated with short disease-free survival in human non-small cell lung cancer. Biochem. Biophys. Res. Commun. 2013, 436, 319–324. [Google Scholar] [CrossRef]
  126. Liu, M.Y.; Li, X.Q.; Gao, T.H.; Cui, Y.; Ma, N.; Zhou, Y.; Zhang, G.J. Elevated HOTAIR expression associated with cisplatin resistance in non-small cell lung cancer patients. J. Thorac. Dis. 2016, 8, 3314–3322. [Google Scholar] [CrossRef] [Green Version]
  127. Zheng, J.; Xiao, X.; Wu, C.; Huang, J.; Zhang, Y.; Xie, M.; Zhang, M.; Zhou, L. The role of long non-coding RNA HOTAIR in the progression and development of laryngeal squamous cell carcinoma interacting with EZH2. Acta Otolaryngol. 2017, 137, 90–98. [Google Scholar] [CrossRef] [PubMed]
  128. Chen, H.; Xin, Y.; Zhou, L.; Huang, J.M.; Tao, L.; Cheng, L.; Tian, J. Cisplatin and paclitaxel target significant long noncoding RNAs in laryngeal squamous cell carcinoma. Med. Oncol. 2014, 31, 246. [Google Scholar] [CrossRef]
  129. Nie, Y.; Liu, X.; Qu, S.; Song, E.; Zou, H.; Gong, C. Long non-coding RNA HOTAIR is an independent prognostic marker for nasopharyngeal carcinoma progression and survival. Cancer Sci. 2013, 104, 458–464. [Google Scholar] [CrossRef]
  130. Fu, W.M.; Lu, Y.F.; Hu, B.G.; Liang, W.C.; Zhu, X.; Yang, H.D.; Li, G.; Zhang, J.F. Long noncoding RNA Hotair mediated angiogenesis in nasopharyngeal carcinoma by direct and indirect signaling pathways. Oncotarget 2016, 7, 4712–4723. [Google Scholar] [CrossRef] [Green Version]
  131. Fu, H.B.; Luo, L. Expression of HOX transcript antisense intergenic RNA in salivary adenoid cystic carcinoma and its influence on prognosis. Hua Xi Kou Qiang Yi Xue Za Zhi West China J. Stomatol. 2020, 38, 509–512. [Google Scholar]
  132. Zhang, J.; Chen, K.; Tang, Y.; Luan, X.; Zheng, X.; Lu, X.; Mao, J.; Hu, L.; Zhang, S.; Zhang, X.; et al. LncRNA-HOTAIR activates autophagy and promotes the imatinib resistance of gastrointestinal stromal tumor cells through a mechanism involving the miR-130a/ATG2B pathway. Cell Death Dis. 2021, 12, 367. [Google Scholar] [CrossRef]
  133. Pastori, C.; Kapranov, P.; Penas, C.; Peschansky, V.; Volmar, C.H.; Sarkaria, J.N.; Bregy, A.; Komotar, R.; St Laurent, G.; Ayad, N.G.; et al. The Bromodomain protein BRD4 controls HOTAIR, a long noncoding RNA essential for glioblastoma proliferation. Proc. Natl. Acad. Sci. USA 2015, 112, 8326–8331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Tan, S.K.; Pastori, C.; Penas, C.; Komotar, R.J.; Ivan, M.E.; Wahlestedt, C.; Ayad, N.G. Serum long noncoding RNA HOTAIR as a novel diagnostic and prognostic biomarker in glioblastoma multiforme. Mol. Cancer 2018, 17, 74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Li, X.; Lu, H.; Fan, G.; He, M.; Sun, Y.; Xu, K.; Shi, F. A novel interplay between HOTAIRn and DNA methylation in osteosarcoma cells indicates a new therapeutic strategy. J. Cancer Res. Clin. Oncol. 2017, 143, 2189–2200. [Google Scholar] [CrossRef] [PubMed]
  136. Chang, Y.T.; Lin, T.P.; Tang, J.T.; Campbell, M.; Luo, Y.L.; Lu, S.Y.; Yang, C.P.; Cheng, T.Y.; Chang, C.H.; Liu, T.T.; et al. HOTAIR is a REST-regulated lncRNA that promotes neuroendocrine differentiation in castration resistant prostate cancer. Cancer Lett. 2018, 433, 43–52. [Google Scholar] [CrossRef] [PubMed]
  137. Lampropoulou, D.I.; Laschos, K.; Aravantinos, G.; Georgiou, K.; Papiris, K.; Theodoropoulos, G.; Gazouli, M.; Filippou, D. Association between homeobox protein transcript antisense intergenic ribonucleic acid genetic polymorphisms and cholangiocarcinoma. World J. Clin. Cases 2021, 9, 1785–1792. [Google Scholar] [CrossRef] [PubMed]
  138. Niinuma, T.; Suzuki, H.; Nojima, M.; Nosho, K.; Yamamoto, H.; Takamaru, H.; Yamamoto, E.; Maruyama, R. Upregulation of miR-196a and HOTAIR drive malignant character in gastrointestinal stromal tumors. Cancer Res. 2012, 72, 1126–1136. [Google Scholar] [CrossRef] [Green Version]
  139. Li, D.; Feng, J.; Wu, T.; Wang, Y.; Sun, Y.; Ren, J.; Liu, M. Long intergenic noncoding RNA HOTAIR is overexpressed and regulates PTEN methylation in laryngeal squamous cell carcinoma. Am. J. Pathol. 2013, 182, 64–70. [Google Scholar] [CrossRef]
  140. Tang, Z.; Wei, G.; Zhang, L.; Xu, Z. Signature microRNAs and long noncoding RNAs in laryngeal cancer recurrence identified using a competing endogenous RNA network. Mol. Med. Rep. 2019, 19, 4806–4818. [Google Scholar] [CrossRef] [Green Version]
  141. Wang, J.; Zhou, Y.; Lu, J.; Sun, Y.; Xiao, H.; Liu, M.; Tian, L. Combined detection of serum exosomal miR-21 and HOTAIR as diagnostic and prognostic biomarkers for laryngeal squamous cell carcinoma. Med. Oncol. 2014, 31, 148. [Google Scholar] [CrossRef]
  142. Cui, X.; Xiao, D.; Cui, Y.; Wang, X. Exosomes-Derived Long Non-Coding RNA HOTAIR Reduces Laryngeal Cancer Radiosensitivity by Regulating microRNA-454-3p/E2F2 Axis. OncoTargets Ther. 2019, 12, 10827–10839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Ma, D.D.; Yuan, L.L.; Lin, L.Q. LncRNA HOTAIR contributes to the tumorigenesis of nasopharyngeal carcinoma via up-regulating FASN. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 5143–5152. [Google Scholar] [PubMed]
  144. Hu, W.; Xu, W.; Shi, Y.; Dai, W. lncRNA HOTAIR upregulates COX-2 expression to promote invasion and migration of nasopharyngeal carcinoma by interacting with miR-101. Biochem. Biophys. Res. Commun. 2018, 505, 1090–1096. [Google Scholar] [CrossRef] [PubMed]
  145. Yang, F.L.; Wei, Y.; Liao, B.Y.; Wei, G.J.; Qin, H.M.; Pang, X.X.; Wang, J.L. LncRNA HOTAIR regulates the expression of E-cadherin to affect nasopharyngeal carcinoma progression by recruiting histone methylase EZH2 to mediate H3K27 trimethylation. Genomics 2021, 113, 2276–2289. [Google Scholar] [CrossRef] [PubMed]
  146. Di Mauro, A.; Scognamiglio, G.; Aquino, G.; Cerrone, M.; Liguori, G.; Clemente, O.; Di Bonito, M.; Cantile, M.; Botti, G.; Tafuto, S.; et al. Aberrant Expression of Long Non Coding RNA HOTAIR and De-Regulation of the Paralogous 13 HOX Genes Are Strongly Associated with Aggressive Behavior of Gastro-Entero-Pancreatic Neuroendocrine Tumors. Int. J. Mol. Sci. 2021, 22, 7049. [Google Scholar] [CrossRef] [PubMed]
  147. Mather, R.L.; Wang, Y.; Crea, F. Is HOTAIR really involved in neuroendocrine prostate cancer differentiation? Epigenomics 2018, 10, 1259–1261. [Google Scholar] [CrossRef] [Green Version]
  148. Blechacz, B.; Komuta, M.; Roskams, T.; Gores, G.J. Clinical diagnosis and staging of cholangiocarcinoma. Nat. Rev. Gastroenterol. Hepatol. 2011, 8, 512–522. [Google Scholar] [CrossRef] [Green Version]
  149. Qin, W.; Kang, P.; Xu, Y.; Leng, K.; Li, Z.; Huang, L.; Gao, J.; Cui, Y.; Zhong, X. Long non-coding RNA HOTAIR promotes tumorigenesis and forecasts a poor prognosis in cholangiocarcinoma. Sci. Rep. 2018, 8, 12176. [Google Scholar] [CrossRef]
  150. Lu, M.; Qin, X.; Zhou, Y.; Li, G.; Liu, Z.; Yue, H.; Geng, X. LncRNA HOTAIR suppresses cell apoptosis, autophagy and induces cell proliferation in cholangiocarcinoma by modulating the miR-204-5p/HMGB1 axis. Biomed. Pharmacother. 2020, 130, 110566. [Google Scholar] [CrossRef] [PubMed]
  151. Lee, N.K.; Lee, J.H.; Kim, W.K.; Yun, S.; Youn, Y.H.; Park, C.H.; Choi, Y.Y.; Kim, H.; Lee, S.K. Promoter methylation of PCDH10 by HOTAIR regulates the progression of gastrointestinal stromal tumors. Oncotarget 2016, 7, 75307–75318. [Google Scholar] [CrossRef]
  152. Bure, I.; Geer, S.; Knopf, J.; Roas, M.; Henze, S.; Ströbel, P.; Agaimy, A.; Wiemann, S.; Hoheisel, J.D.; Hartmann, A.; et al. Long noncoding RNA HOTAIR is upregulated in an aggressive subgroup of gastrointestinal stromal tumors (GIST) and mediates the establishment of gene-specific DNA methylation patterns. Genes Chromosomes Cancer 2018, 57, 584–597. [Google Scholar] [CrossRef]
  153. Yan, J.; Chen, D.; Chen, X.; Sun, X.; Dong, Q.; Hu, C.; Zhou, F.; Chen, W. Downregulation of lncRNA CCDC26 contributes to imatinib resistance in human gastrointestinal stromal tumors through IGF-1R upregulation. Braz. J. Med. Biol. Res. 2019, 52, e8399. [Google Scholar] [CrossRef] [PubMed]
  154. Zhang, J.X.; Han, L.; Bao, Z.S.; Wang, Y.Y.; Chen, L.Y.; Yan, W.; Yu, S.Z.; Pu, P.Y.; Liu, N.; You, Y.P.; et al. Chinese Glioma Cooperative Group. HOTAIR, a cell cycle-associated long noncoding RNA and a strong predictor of survival, is preferentially expressed in classical and mesenchymal glioma. Neuro-Oncology 2013, 15, 1595–1603. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, K.; Sun, X.; Zhou, X.; Han, L.; Chen, L.; Shi, Z.; Zhang, A.; Ye, M.; Wang, Q.; Liu, C.; et al. Long non-coding RNA HOTAIR promotes glioblastoma cell cycle progression in an EZH2 dependent manner. Oncotarget 2015, 6, 537–546. [Google Scholar] [CrossRef] [Green Version]
  156. Zhou, X.; Ren, Y.; Zhang, J.; Zhang, C.; Zhang, K.; Han, L.; Kong, L.; Wei, J.; Chen, L.; Yang, J.; et al. HOTAIR is a therapeutic target in glioblastoma. Oncotarget 2015, 6, 8353–8365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Ke, J.; Yao, Y.L.; Zheng, J.; Wang, P.; Liu, Y.H.; Ma, J.; Li, Z.; Liu, X.B.; Li, Z.Q.; Wang, Z.H.; et al. Knockdown of long non-coding RNA HOTAIR inhibits malignant biological behaviors of human glioma cells via modulation of miR-326. Oncotarget 2015, 6, 21934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Bian, E.B.; Ma, C.C.; He, X.J.; Wang, C.; Zong, G.; Wang, H.L.; Zhao, B. Epigenetic modification of miR-141 regulates SKA2 by an endogenous ‘sponge’ HOTAIR in glioma. Oncotarget 2016, 7, 30610–30625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Shen, J.; Hodges, T.R.; Song, R.; Gong, Y.; Calin, G.A.; Heimberger, A.B.; Zhao, H. Serum HOTAIR and GAS5 levels as predictors of survival in patients with glioblastoma. Mol. Carcinog. 2018, 57, 137–141. [Google Scholar] [CrossRef]
  160. Xavier-Magalhães, A.; Oliveira, A.I.; de Castro, J.V.; Pojo, M.; Gonçalves, C.S.; Lourenço, T.; Viana-Pereira, M.; Costa, S.; Linhares, P.; Vaz, R.; et al. Effects of the functional HOTAIR rs920778 and rs12826786 genetic variants in glioma susceptibility and patient prognosis. J. Neuro-Oncol. 2017, 132, 27–34. [Google Scholar] [CrossRef]
  161. Chakravadhanula, M.; Ozols, V.V.; Hampton, C.N.; Zhou, L.; Catchpoole, D.; Bhardwaj, R.D. Expression of the HOX genes and HOTAIR in atypical teratoid rhabdoid tumors and other pediatric brain tumors. Cancer Genet. 2014, 207, 425–428. [Google Scholar] [CrossRef]
  162. Zheng, H.; Baranova, K.; Song, J.; Yan, L.; Biswas, S.; Chakrabarti, S.; Zhang, Q. Overexpression of Long Noncoding RNA HOTAIR Is a Unique Epigenetic Characteristic of Myxopapillary Ependymoma. J. Neuropathol. Exp. Neurol. 2020, 79, 1193–1202. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, J.; Li, N.; Fu, J.; Zhou, W. Long noncoding RNA HOTAIR promotes medulloblastoma growth, migration and invasion by sponging miR-1/miR-206 and targeting YY1. Biomed. Pharmacother. 2020, 124, 109887. [Google Scholar] [CrossRef] [PubMed]
  164. Milhem, M.M.; Knutson, T.; Yang, S.; Zhu, D.; Wang, X.; Leslie, K.K.; Meng, X. Correlation of MTDH/AEG-1 and HOTAIR Expression with Metastasis and Response to Treatment in Sarcoma Patients. J. Cancer Sci. Ther. 2011, 5, 004. [Google Scholar] [CrossRef] [Green Version]
  165. Wang, B.; Su, Y.; Yang, Q.; Lv, D.; Zhang, W.; Tang, K.; Wang, H.; Zhang, R.; Liu, Y. Overexpression of Long Non-Coding RNA HOTAIR Promotes Tumor Growth and Metastasis in Human Osteosarcoma. Mol. Cells 2015, 38, 432–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Wang, B.; Qu, X.L.; Liu, J.; Lu, J.; Zhou, Z.Y. HOTAIR promotes osteosarcoma development by sponging miR-217 and targeting ZEB1. J. Cell Physiol. 2019, 234, 6173–6181. [Google Scholar] [CrossRef]
  167. Wang, N.; Meng, X.; Liu, Y.; Chen, Y.; Liang, Q. LPS promote Osteosarcoma invasion and migration through TLR4/HOTAIR. Gene 2019, 680, 1–8. [Google Scholar] [CrossRef] [PubMed]
  168. Zhou, Q.; Chen, F.; Fei, Z.; Zhao, J.; Liang, Y.; Pan, W.; Liu, X.; Zheng, D. Genetic variants of lncRNA HOTAIR contribute to the risk of osteosarcoma. Oncotarget 2016, 7, 19928–19934. [Google Scholar] [CrossRef] [Green Version]
  169. Bao, X.; Ren, T.; Huang, Y.; Sun, K.; Wang, S.; Liu, K.; Zheng, B.; Guo, W. Knockdown of long non-coding RNA HOTAIR increases miR-454-3p by targeting Stat3 and Atg12 to inhibit chondrosarcoma growth. Cell Death Dis. 2017, 8, e2605. [Google Scholar] [CrossRef]
  170. Feng, Q.; Wang, D.; Guo, P.; Zhang, Z.; Feng, J. Long non-coding RNA HOTAIR promotes the progression of synovial sarcoma through microRNA-126/stromal cell-derived factor-1 regulation. Oncol. Lett. 2021, 21, 444. [Google Scholar] [CrossRef]
  171. Yang, G.; Lu, X.; Yuan, L. LncRNA: A link between RNA and cancer. Biochim. Biophys. Acta 2014, 1839, 1097–1109. [Google Scholar] [CrossRef]
  172. Peng, W.X.; Koirala, P.; Mo, Y.Y. LncRNA-mediated regulation of cell signaling in cancer. Oncogene 2017, 36, 5661–5667. [Google Scholar] [CrossRef] [PubMed]
  173. Botti, G.; Marra, L.; Malzone, M.G.; Anniciello, A.; Botti, C.; Franco, R.; Cantile, M. LncRNA HOTAIR as Prognostic Circulating Marker and Potential Therapeutic Target in Patients with Tumor Diseases. Curr. Drug Targets 2017, 18, 27–34. [Google Scholar] [CrossRef] [PubMed]
  174. Ren, Y.; Wang, Y.F.; Zhang, J.; Wang, Q.X.; Han, L.; Mei, M.; Kang, C.S. Targeted design and identification of AC1NOD4Q to block activity of HOTAIR by abrogating the scaffold interaction with EZH2. Clin. Epigenet. 2019, 11, 29. [Google Scholar] [CrossRef] [PubMed]
  175. Li, Y.; Ren, Y.; Wang, Y.; Tan, Y.; Wang, Q.; Cai, J.; Zhou, J.; Yang, C.; Zhao, K.; Yi, K. A Compound AC1Q3QWB Selectively Disrupts HOTAIR-Mediated Recruitment of PRC2 and Enhances Cancer Therapy of DZNep. Theranostics 2019, 9, 4608–4623. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the main molecular pathways related to HOTAIR deregulation in rare cancers. In rare tumors of the head and neck rare (green box), HOTAIR promotes cell growth, proliferation, migration and invasion through the direct interaction and consequent inhibition of different miRNAs: HOTAIR/miR-1 functional interaction promotes MAGEA2 expression, which is able to suppress p53-dependent apoptosis in response to drugs, decrease cellular senescence and increase cell proliferation in LSCC cells; HOTAIR inhibits miR-454-3p that target EF2, the principal target of the tumor suppressor pRB, reducing tumor suppression in LSCC cells; HOTAIR/miR-101 functional interaction in NPC cells induces COX2 expression, a regulator of tumor metabolism, angiogenesis and tumor microenvironment. In LSCC cells, HOTAIR interacts with VEGFA promoting angiogenic processes. In soft tissue tumors (orange box), HOTAIR interacts with different miRNAs leading to their inhibition and promoting cell proliferation, migration and invasion of tumor cells: HOTAIR/miR-454-3p functional interaction induces (i) ATG12 expression, a positive regulator of autophagic vesicle formation, and (ii) STAT3, involved in regulation of cancer inflammation and metastasis, in chondrosarcoma cells. HOTAIR/miR-126 functional interaction promotes (i) SDF-1 expression, the main receptor of the metastasis inducer gene CXCR4 able to modulate proliferation/survival and EMT of synovial sarcoma cells, and (ii) DNMT1, involved in cancer stem cell maintenance and tumorigenesis, in osteosarcoma cells. HOTAIR inhibits miR-217 upregulating ZEB1, a potent inducer of EMT, in osteosarcoma cells. In rare digestive system tumors (yellow box), HOTAIR interacts with miR-204 to upregulate HMGB1 gene, a crucial regulator of inflammation and cancer progression, in cholangiosarcoma cells, and with miR-130a to induce ATG2B expression promoting autophagy in GIST cells. In central nervous system tumors (blue box), HOTAIR is able to induce b-catenin expression, modulating Wnt signal pathways in glioma cells, and establishes functional interaction with different miRNAs to promote cell proliferation invasion and migration: HOTAIR/miR-141 interaction induce upregulation of SKA2, a gene involved in cell cycle regulation, in glioma cells; HOTAIR/miR-326 interaction leads to the upregulation of FGF1, involved in the repairing progress of damaged, in glioma cells; HOTAIR inhibits miR-219 upregulating Cyclin D1, a key regulator of cell cycle progression, in glioma cells; HOTAIR/miR206 functional interaction induces YY1 expression promoting metastatic progression in medulloblastoma cells. Green arrow connects related pathways and function; Red arrow indicates induction of expression; T indicates inhibition of expression.
Figure 1. Schematic representation of the main molecular pathways related to HOTAIR deregulation in rare cancers. In rare tumors of the head and neck rare (green box), HOTAIR promotes cell growth, proliferation, migration and invasion through the direct interaction and consequent inhibition of different miRNAs: HOTAIR/miR-1 functional interaction promotes MAGEA2 expression, which is able to suppress p53-dependent apoptosis in response to drugs, decrease cellular senescence and increase cell proliferation in LSCC cells; HOTAIR inhibits miR-454-3p that target EF2, the principal target of the tumor suppressor pRB, reducing tumor suppression in LSCC cells; HOTAIR/miR-101 functional interaction in NPC cells induces COX2 expression, a regulator of tumor metabolism, angiogenesis and tumor microenvironment. In LSCC cells, HOTAIR interacts with VEGFA promoting angiogenic processes. In soft tissue tumors (orange box), HOTAIR interacts with different miRNAs leading to their inhibition and promoting cell proliferation, migration and invasion of tumor cells: HOTAIR/miR-454-3p functional interaction induces (i) ATG12 expression, a positive regulator of autophagic vesicle formation, and (ii) STAT3, involved in regulation of cancer inflammation and metastasis, in chondrosarcoma cells. HOTAIR/miR-126 functional interaction promotes (i) SDF-1 expression, the main receptor of the metastasis inducer gene CXCR4 able to modulate proliferation/survival and EMT of synovial sarcoma cells, and (ii) DNMT1, involved in cancer stem cell maintenance and tumorigenesis, in osteosarcoma cells. HOTAIR inhibits miR-217 upregulating ZEB1, a potent inducer of EMT, in osteosarcoma cells. In rare digestive system tumors (yellow box), HOTAIR interacts with miR-204 to upregulate HMGB1 gene, a crucial regulator of inflammation and cancer progression, in cholangiosarcoma cells, and with miR-130a to induce ATG2B expression promoting autophagy in GIST cells. In central nervous system tumors (blue box), HOTAIR is able to induce b-catenin expression, modulating Wnt signal pathways in glioma cells, and establishes functional interaction with different miRNAs to promote cell proliferation invasion and migration: HOTAIR/miR-141 interaction induce upregulation of SKA2, a gene involved in cell cycle regulation, in glioma cells; HOTAIR/miR-326 interaction leads to the upregulation of FGF1, involved in the repairing progress of damaged, in glioma cells; HOTAIR inhibits miR-219 upregulating Cyclin D1, a key regulator of cell cycle progression, in glioma cells; HOTAIR/miR206 functional interaction induces YY1 expression promoting metastatic progression in medulloblastoma cells. Green arrow connects related pathways and function; Red arrow indicates induction of expression; T indicates inhibition of expression.
Ijms 22 10160 g001
Table 2. Principal roles of HOTAIR in patients with rare tumors.
Table 2. Principal roles of HOTAIR in patients with rare tumors.
Tumor TypesClinical StudiesReferences
Head and Neck rare cancersLSCCHOTAIR overexpression correlates with poor differentiation, pathological grade, metastatic risk and chemotherapy sensibility. HOTAIR is a circulating markers in LSCC patients.[127,128]
NPCHOTAIR overexpression correlates with clinical stage and poor prognosis.[129,130]
SACCHOTAIR overexpression correlates with clinical stage, nerve invasion, metastatic risk and poor survival.[131]
Neuroendocrine tumorsGEP NENHOTAIR overexpression correlates with tumor grade and tumor stage.[47]
Rare digestive system tumorsCholangiocarcinomaHOTAIR overexpression correlates with tumor size, TNM stage and metastatic risk. [132]
GISTHOTAIR overexpression correlates with tumor grade, recurrence and metastatic risk, poor survival and drug resistance.[133]
Central Nervous System tumorsGlioma/GlioblastomaHOTAIR overexpression correlates with tumor grade, molecular subtype and poor prognosis. HOTAIR is a circulating markers in glioma patients.[134,135]
Soft tissue tumorsOsteosarcomaHOTAIR overexpression correlates with advanced tumor stage, high histological grade and shorter overall survival.[136]
ChondrosarcomaHOTAIR overexpression correlates with tumor grade and shorter overall survival.[137]
Synovial sarcomaHOTAIR overexpression correlates with histological grade, advanced tumor stage and metastatic risk.[138]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liguori, G.; Cerrone, M.; De Chiara, A.; Tafuto, S.; Tracey de Bellis, M.; Botti, G.; Di Bonito, M.; Cantile, M. The Role of lncRNAs in Rare Tumors with a Focus on HOX Transcript Antisense RNA (HOTAIR). Int. J. Mol. Sci. 2021, 22, 10160. https://doi.org/10.3390/ijms221810160

AMA Style

Liguori G, Cerrone M, De Chiara A, Tafuto S, Tracey de Bellis M, Botti G, Di Bonito M, Cantile M. The Role of lncRNAs in Rare Tumors with a Focus on HOX Transcript Antisense RNA (HOTAIR). International Journal of Molecular Sciences. 2021; 22(18):10160. https://doi.org/10.3390/ijms221810160

Chicago/Turabian Style

Liguori, Giuseppina, Margherita Cerrone, Annarosaria De Chiara, Salvatore Tafuto, Maura Tracey de Bellis, Gerardo Botti, Maurizio Di Bonito, and Monica Cantile. 2021. "The Role of lncRNAs in Rare Tumors with a Focus on HOX Transcript Antisense RNA (HOTAIR)" International Journal of Molecular Sciences 22, no. 18: 10160. https://doi.org/10.3390/ijms221810160

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