Skip to main content

New insights into long noncoding RNAs and their roles in glioma

Abstract

Glioma is one of the most prevalent types of primary intracranial carcinoma with varying malignancy grades I–IV and histological subtypes, including astrocytomas, glioblastoma multiform (GBM), oligodendrogliomas and mixed tumors. Glioma is characterized by rapid cell proliferation and angiogenesis, and the WHO grade IV glioblastoma, which is highly malignant with poor prognosis because GBM stem-like cells (GSCs) are resistant to conventional therapy and easily recrudescent, accounts for the majority of gliomas. Consequently, investigations exploring the accurate molecular mechanisms and reliable therapeutic targets for gliomas have drawn extensive attention.

Based on the increasing amount of functional lncRNAs aberrantly expressed in glioma tissues and cell lines, lncRNAs might be critical for glioma initiation, progression and other malignant phenotypes. This review summarizes the latest insights into the lncRNA field and their functional roles in glioma, therefore evaluating the potential clinical applications of lncRNAs as prospective novel biomarkers and therapeutic targets.

Background

LncRNAs comprise a wide variety of RNA transcripts of sizes greater than 200 nucleotides (nt) that lack significant protein-coding capacity [1], typically with 5’m7G caps and 3′ poly(A) in tails, similar to the structure of mRNAs [2], but are more tissue-specific and dynamic than mRNAs, suggesting that these molecules have distinct biological roles [3]. LncRNAs were initially implicated in the epigenetic regulation of X chromosome inactivation during embryogenesis [4]. The last several years has seen a steady increasing interest in identifying lncRNAs and understanding their regulatory functions in almost all aspects from gene expression to protein translation and stability [5], either in cis (regulate neighboring genes) or in trans (regulate distant genes) [6]. (Table 1).

Table 1 Examples of lncRNA functional mechanisms

In accordance with their significant roles in extensive biological processes, lncRNAs have been implicated in the onset and progression of cancer malignancy, including glioma, such as stemness, proliferation, angiogenesis and drug resistance [7, 8]. Moreover, aberrant lncRNA expression profiles in clinical glioma specimens correlates with malignancy grade and histological differentiation, which have important clinical implications in glioma diagnosis of sub-classification [9] and prognostication [10, 11]. In addition, lncRNAs transmitted from cell to cell participate in intercellular communications for maintaining microenvironment homeostasis or mediating tumor metastasis. This review summarizes the most up-to-date knowledge regarding how lncRNAs function at the molecular level and their implications in the areas of glioma research and therapy.

LncRNA categories and structures

LncRNA classification based on genomic location

LncRNAs can be classified into the following five categories according to their genomic location relative to neighboring protein-coding genes [12]: sense, antisense, bidirectional [13], intronic [14] and intergenic (lincRNAs) [1] (more detail in Fig. 1). Moreover, the promoter upstream transcripts (PROMPTs) and enhancer-associated RNAs (eRNAs), transcribed from promoters or enhancers, show some similarities in functional mechanism with regulatory DNA elements [12]. For instance, eRNAs likely facilitate enhancer and promoter interactions and thereby activate target genes [15].

Fig. 1
figure 1

Schematic representation of lncRNA species and structures. a LncRNAs transcribed from functional DNA element regions: promoter upstream transcript (PROMPTs), enhancer-associated RNA(eRNAs); (b) Mainstream five categories based on genomic location and transcript orientation of lncRNA: (1) sense lncRNAs and (2) antisense lncRNAs, overlapping and transcribed from the same or opposite strand of protein-coding genes, (3) bidirectional lncRNAs, typically transcribed in the opposite direction of its neighboring protein-coding gene (less than 1 kb away) [13], (4) intronic lncRNAs, transcribed entirely from introns [14], (5) intergenic lncRNAs (lincRNAs), transcribed from genomic interval [1]; (c, d) LncRNAs characterized by snoRNA caps: Sno-lncRNAs have snoRNA caps at both ends, while SPA-lncRNAs possess 5’ snoRNA only; (e) LncRNAs characterized by tRNA-like extremities, with the 3′ end alternatively processed by ribonuclease P (RNase P), e.g., MALAT1 and NEAT1_2; (f, g) LncRNAs with close circular structures: derived from back-splicing of exons named circular RNAs, and ciRNAs are produced from intron lariats.

LncRNA categories based on subcellular localization

Since the functional mechanism of lncRNAs depends on their subcellular localization [16], lncRNAs can also be grouped into nuclear or cytoplasmic categories [5]. LncRNAs that exhibit distinct nuclear localization patterns typically alter gene transcription via transcriptional interference [17] and chromatin remodeling [15], whereas lncRNAs exported to the cytoplasm perform regulatory roles by mediating RNA processing, affecting mRNA stability or directly regulating protein function [18].

New LncRNA species based on unique structures

Given that lncRNAs structures are associated with their stability and functional mechanisms [16], the identification of lncRNA secondary structures will contribute to their applications in clinical medicine [19]. Most linear lncRNAs have 5’m7G caps and 3′ poly(A) tails, similar to mRNAs, while some noncanonical terminal lncRNAs processed with unique transcription patterns can be defined as new RNA species [20]. Sno-lncRNAs and SPA-lncRNAs [21] are characterized by snoRNA caps at both ends or only at the 5′ terminus, having longer half-lives than other lncRNAs. In addition, the RNase P-mediated processing of lncRNAs results in tRNA-like structures at the extremity and increased stability in the cytoplasm, for which a typical example is MALAT1 [22]. As for lncRNAs that exist in loop form, there are circular RNAs [23] and circular intronic RNAs (ciRNAs) [24] produced from the back-splicing of exons or introns respectively (Fig. 1).

Functional mechanisms of lncRNAs in glioma and other cancers

LncRNAs regulate genome activity

Unsurprisingly, many lncRNAs participate in the epigenetic regulation of gene expression by recruiting chromatin modifiers (e.g., EZH2/PRC2 [25] and WDR5/TrxG [26]) to a specific genomic location as a Scaffold, segregating chromatin modifiers away from their regulatory targets as a Decoy [17] or mediating the 3D organization of chromosomes [15].

In glioma, NEAT1 mediates the trimethylation of H3K27 in the promoter region of WNT/β-catenin pathway negatively regulated factors (Axin2, ICAT and GSK3B) by physically interacting with the PRC2 subunit EZH2 as a Scaffold, thereby resulting in WNT/β-catenin pathway activation [25]. During hepatocarcinogenesis, lncRNA HOTTIP directly interacts with the trithorax group (TrxG) protein WDR5 and induces an open DNA-chromatin configuration to target WDR5/MLL complexes, driving H3K4 trimethylation and thus promoting the transcription of HOXA locus genes [26]. Increased by estradiol (E2) stimulation in breast cancer, FOXC1 contributes to E2-dependent gene activation by promoting the 3D contacts between the specific promoter and enhancer cluster and increasing the strength of enhancer–promoter looping [15].

LncRNAs related to posttranscriptional regulation

Aside from regulating the alternative splicing [21] and stability [27] of mRNAs, the competing endogenous RNAs (ceRNAs) or miRNA sponge mechanism are also implicated in lncRNAs post-transcriptional regulation. (Detailed in Table 2).

Table 2 Deregulated lncRNAs in glioma

During mRNA alternative splicing, lncRNA BC200 recruits splicing factors hnRNP A2/B1 to Bcl-x pre-mRNA and contributes to breast cancer pathogenesis [28]. From the perspective of regulating mRNA stability, lncRNAs may either form protective “lncRNA-mRNA” duplex to stabilize mRNAs [27] or, conversely, facilitate mRNA degradation by recruiting RNA-binding proteins, such as PTBP1 to target pre-mRNA [29]. Moreover, LncRNAs can regulate the efficiency of mRNA translation. Upon energy stress, FILNC1 interacts with AUF1, a c-Myc mRNA-binding protein, and sequesters AUF1 from binding c-Myc mRNA, leading to the downregulation of c-Myc protein. Cytoplasmic lncRNAs can participate in the ceRNA regulatory network and act as endogenous miRNA sponges through direct base pairing. In high-grade GBM, the up-regulation of CRNDE competitively binds to miR-384 and decreases the availability of miR-384 to its downstream target genes (STAT3, cyclin D1 and Bcl-2), thus abrogating the suppression of miR-384 downstream genes [30].

LncRNAs in protein modification and anchoring

LncRNAs can also directly bind proteins essential for a signaling pathway and modulate their functions. A novel mechanism by which lncRNA plays an oncogenic role through modulating the phosphorylation status of its interaction protein is demonstrated in many cancers. LINK-A recruits breast tumor kinase (BRK) to phosphorylate Tyr565 of hypoxia-inducible factorα (HIF1α), and the activated HIF1α signaling promotes breast cancer glycolysis reprogramming [18]. HULC specifically binds to YB-1 (translationally inactive ribonucleoprotein particle) and promotes YB-1 phosphorylation, which leads to the release of YB-1 from its bound mRNA. Consequently, the translation of silenced oncogenic mRNAs (cyclin D1, cyclin E1, and MMP3) in hepatocellular carcinoma would be activated [31].

Moreover, lncRNA SNHG5 can trap MTA2 in the cytoplasm, preventing MTA2 translocation into the nucleus and interfering with nucleosome remodeling. The significantly downregulated SNHG5 in gastric cancer, thereby promoting gastric cancer progression [32].

LncRNAs encode functional micropeptides

Recently, some lncRNAs have been shown to encode functional micropeptides based on small-ORFs [33]. For instance, micropeptides translated from lncRNAs regulate skeletal muscle regeneration or contractility by interacting with sarcoplasmic reticulum Ca2+-ATPase (SERCA) [34], lysosomal v-ATPase [35], or displacing SERCA inhibitors [36]. Moreover, the micropeptides encoded by lncRNA HOXB-AS3 competitively bind to hnRNPs, suppressing glucose metabolism reprogramming in colon cancer [37].

Exosome-transmitted lncRNAs function as intercellular communicator

LncRNAs play roles in intercellular communication, wherein the release of lncRNAs into the recipient cells can modulate cell phenotypes through the regulation of target signaling pathways [38]. Studies have shown that lncRNAs secreted from drug-resistant renal carcinoma cells via exosomes can transform adjacent drug-sensitive cells into a resistant phenotype [39]. Moreover, exosome-transmitted lncRNAs can regulate the tumor microenvironment and mediate the communication of heterogeneous cells [40, 41]. Through the release of exosomes containing lncRNA H19, CD90+ liver cancer cells can modulate the phenotypes of endothelial cells, promoting angiogenesis and cell-to-cell adhesion [40].

LncRNAs regulate glioma malignant phenotypes

Accumulating evidence has demonstrated that many dysregulated lncRNAs in glioma are closely related to tumorigenesis, metastasis and prognosis or therapies [7, 8] (Fig. 2 and Table 2). LncRNAs can function as molecular signaling mediators, modulating the expression of a certain set of genes and corresponding signaling pathways (e.g., the NEAT1-WNT/β catenin pathway [25] and CRNDE- mTOR signaling [42]).

Fig. 2
figure 2

Functional roles of lncRNAs in glioma malignancy and chemoresistance. LncRNAs contribute to five major malignant phenotypes of glioma by targeting different cell species: general glioma cells, glioma-associated endothelial cells, GBM cells, and GBM stem-like cells. Selected examples of lncRNAs and their molecular partners or genomic targets are shown for stemness, drug resistance, BTB permeability, angiogenesis and motility cancer phenotypes.

The majority of the glioma-associated lncRNAs serve as “miRNA sponges” to quench miRNA activity (e.g., HOTAIR/miR-326 [43], Gas5/miR-196a-5p [44], CASC2/ miR-181a [45] and ECONEXIN /miR-411-5p [46]).

LncRNAs related to glioma stemness

Recent studies have shown that lncRNA TUG1 maintains the stemness of GSCs through epigenetically suppressing multiple neuronal differentiation-associated genes. TUG1 physically interacts with PRC2 components (EZH2 and SUZ12) and transcription factor YY1 as a protein scaffold, promoting the locus-specific methylation of histone H3K27 [47]. LncRNA GAS5 suppresses GSC proliferation, migration and invasion by binding to onco-miR-196a-5p and upregulating the downstream FOXO1. Interestingly, the increased FOXO1 promotes GAS5 transcription, thus forming a positive feedback loop [44]. LncRNA XIST promotes the proliferation and inhibits the apoptosis of GSCs. More than directly binding to miR-152 as a sponge, XIST is likely in the same RNA-induced silencing complex (RISC) with miR-152, thus abrogating miR-152 downstream pathways [48].

Post-transcriptional mRNA modification provides an additional regulation of mRNA stability, and N6-methyladenosine (m6A) is the most prevalent internal modification on mRNAs [49]. LncRNA FOXM1-AS promotes the interaction of FOXM1 nascent transcripts with RNAm6A demethylase ALKBH5, leading to the demethylation and elevated expression of FOXM1. The elevated expression of transcription factor FOXM1 then facilitates GSC self-renewal and tumorigenesis [50].

LncRNAs related to glioma proliferation and migration

Regulated by EGFR pathway activity, NEAT1 is critical for glioma cell growth and invasion through the WNT/β-catenin pathway. NEAT1 mediates the H3K27 trimethylation of three WNT/β-catenin pathway negatively regulated factors, thereby increasing β-catenin nuclear transport and rekindling the WNT/β-catenin pathway [25]. Considering that NBAT-1 is expressed at low levels in high-risk neuroblastoma, NBAT-1 acts as a scaffold for EZH2 recruitment and represses NBAT-1/EZH2 target genes, which are involved in neuroblastoma progression [51]. It was demonstrated that the 5′ domain of HOTAIR can bind to the EZH2 and promote glioblastoma cell cycle progression [52].

The oncogene ECONEXIN is predominantly located in the cytoplasm and interacts with miR-411-5p via two binding sites, regulating TOP2A by sponging miR-411-5p in early glioma tumorigenesis [46]. Furthermore, lncRNA HCP5 sponge miR-139 to up-regulate RUNX1, and then RUNX1 feedback promoted HCP5 expression by binding to HCP5 promoters, creating a HCP5/miR-139/RUNX1 positive feedback loop to regulate the malignant behavior of glioma cells [53]. As a glioma-suppressor, MALAT1 inhibits cell viability by downregulating miR-155 expression [54], and a similar mechanism was also observed in the “TUSC7-miR-23b” axis [55]. The knockdown of CRNDE decreases the protein level of PIWIL4, a target of miR-384, which leads to glioma regression in vivo [30].

MALAT1 play a tumor-suppressive role in glioma by attenuating ERK/MAPK-mediated growth and MMP2-mediated invasiveness. The decreased level of MALAT1 in glioma significantly increased tumorigenicity in both subcutaneous and intracranial human glioma xenograft models [56].

Angiogenesis is a universal characteristic of blood vessel-rich glioma progression that facilitates proliferation and migration. In glioma microvessels, the overexpression of H19 induces glioma-associated endothelial cells (GECs) proliferation, migration and tube formation via increasing the expression of angiogenic factor VASH2 as a miR-29a sponge [57]. Up-regulated XIST can facilitate glioma angiogenesis by increasing the promoter activity of chemokine receptor 7b (CXCR7) [58], or down-regulating miR-429 as a molecular sponge [59]. LncRNA HULC regulates ESM-1 via the PI3K/Akt/mTOR signaling pathway and hence plays a pro-angiogenesis role in human gliomas [60].

LncRNAs related to invalid chemotherapy

Conventional treatment of glioma includes surgical resection followed by radiotherapy and temozolomide (TMZ) chemotherapy. Invalid drug administration in glioma is mainly caused by either the poor permeability of the blood–tumor barrier (BTB), which invalidates adjuvant chemotherapy [61] or the acquired chemo-resistance of tumor cells.

Modulating the biological behaviors of glioma vascular GECs is an important aspect of lncRNAs influencing therapeutic effect. For example, HOTAIR decreases BTB permeability via binding to miR-148b-3p, which further increasing tight junction-related proteins (ZO-1, occludin and claudin-5) in GECs by regaining upstream stimulating factor 1 (USF1) [62]. Similarly, the aberrant high expression of lncRNA XIST in GECs reduces BTB permeability by increasing tight junction-related proteins (ZO-1, ZO-2 and occludin) [58]. LncRNA TUG1 is also up-regulated in endothelial cells and regulates BTB permeability through targeting miR-144 [63]. Some researches demonstrated that the knockdown of MALAT1 resulted in the increased permeability of BTB, which might contribute to the reworking of chemotherapeutics [61].

TMZ is a first-line chemotherapeutic drug most widely used for treating gliomas, whereas drug resistance is largely responsible for the failure of glioma chemotherapy [64]. Research has shown that cells with epithelial-to-mesenchymal transition (EMT) characteristics tend to be more resistant to anticancer drugs. LncRNA MALAT1 is up-regulated in multidrug-resistant glioblastoma cell lines and decreases the sensitivity of glioblastoma cells to TMZ by up-regulating EMT-related proteins (ZEB1, Snail and SLUG) [65]. LncRNA CASC2 up-regulates the tumor suppressor PTEN through directly inhibiting miR-181a, thus amplify TMZ sensibility of glioma cells [45].

Promising biomarker: lncRNAs in diagnostic or prognostic applications

The aberrant lncRNA expression profiles in clinical specimens correlates with malignancy grade and histological differentiation, which have important clinical implications in the glioma diagnosis of sub-classification [9] and prognostication [10, 11]. One of the distinctive features of lncRNAs is their highly tissue- and cell type-specific expression patterns [3], which could accurately classify different subtypes of glioma and predict responses to treatments.

One-way analysis of variance (ANOVA) indicates a markedly significant difference in HOXA11-AS expression between the four GBM subtypes in CGGA mRNA microarray datasets, and HOXA11-AS expression in classical and mesenchymal subtypes is higher than that in neural and proneural subtypes, suggesting that HOXA11-AS might serve as a biomarker for identifying glioma molecular subtypes [9]. The up-regulated oncogene CRNDE is correlated with larger tumor size (p = 0.011), higher WHO grade (p = 0.001), and recurrence (p = 0.008). Additionally, survival analysis demonstrated that up-regulated CRNDE expression is associated with poor overall survival of glioma patients (p < 0.001) [10]. Cox regression analysis revealed that HOXA-AS3 is an independent prognostic factor, and its high-expression is associated with tumor grade and poor prognosis in glioma patients [11].

As previously described, exosome-transmitted lncRNAs reflect the progress of the tumor microenvironment, which are stable and easily detectable in plasma or other body fluids and enable accurate diagnosis as desirable markers [66]. A potential application for minimally invasive liquid biopsy lies in the use of circulating lncRNA levels, which has been well documented in other cancers [67]. By comparing the expression of ten candidate circulating serum lncRNAs in a case-control study, three lncRNAs, LINC00152, RP11-160H22.5 and XLOC014172, have been identified as useful in the early diagnosis of hepatocellular carcinoma [68]. Moreover, lncRNAs with low expression levels in cells may be enriched in secreted exosomes for the regulation of cancer metastasis, thereby showing higher clinical utility [69].

Therapeutics targeting tumor-specific lncRNA abnormalities

Since the modulation of lncRNA networks mediates anti-proliferative effects in malignant tumors and triggers therapeutic efficacy, targeting tumor-specific lncRNA abnormalities using small interfering RNAs (siRNAs) may be an effective strategy [70].

Many epigenetic inhibitors are currently used in clinical trials. For instance, bromodomain and extraterminal (BET) domain proteins reduce the levels of several oncogenes in glioblastoma, including lncRNA HOTAIR, and the use of BET inhibitor can restore the malignancy caused by oncogenic lncRNAs [71].

Antisense oligonucleotides (ASOs) can correct aberrant lncRNA networks by inducing RNase H-dependent degradation or sterically blocking lncRNA activity. For example, intravenous treatment with ASOs targeting lncRNA TUG1 coupled with a drug delivery system induces glioma stem cell differentiation and efficiently represses tumor growth in vivo [47]. Furthermore, chemical modifications of ASOs, such as the addition of 2′-omethyl and locked nucleic acids at both 5′ and 3′ ends, can protect ASOs from degradation and maintain sufficient concentrations. A preclinical study demonstrated that the use of modified ASOs targeting MALAT1 in a breast cancer mouse model promotes cystic differentiation and decreases tumor growth [72].

To make “anti-ncRNA therapy” more thermally stable in the tumor microenvironment, specific nucleic acids conjugated with a neutral-charged peptide backbone, called peptide nucleic acids (PNAs), are modified for in vivo applications [73]. PNAs-based strategy blocks the ability of lncRNA HOTAIR to interact with EZH2, subsequently reduces HOTAIR-EZH2 activity and inhibits ovarian and breast cancer cell invasion, which may be suitable for other solid cancers, such as glioma [74].

With a better understanding of the lncRNA binding domain, the use of small molecule inhibitors to disrupt the interactions of lncRNAs with their binding partners, particularly the chromatin modification complex, is an alternative strategy, which has been demonstrated for miRNAs [75]. However, given the length of some lncRNAs, it is possible that extensive secondary structures will restrict the accessibility of small molecule inhibitors to crucial parts of target lncRNAs, again emphasizing the importance of structural mapping for lncRNAs [19].

Conclusion

At present, tens of thousands of identified lncRNAs have been associated with imbalanced gene regulation and aberrant biological processes in cancers [3, 6], and many lncRNAs play indispensable roles in the onset and progression of glioma malignancy, including the stemness, proliferation, angiogenesis and drug resistance [7, 8].

However, only a few functional mechanisms of lncRNAs have been well characterized, and some recent insights into lncRNAs identified in other diseases and cancers, such as molecules encoding tumor-related micropeptides, changes in the activities or conformation of proteins and the regulation of tumor microenvironment homeostasis, have not yet been clearly demonstrated in glioma pathogenesis. The majority of studies of glioma-associated lncRNAs are typically focused on ceRNA regulatory networks; thus, there is an urgent need to elucidate the detailed mechanisms of the roles of lncRNAs in glioma. Additionally, whether lncRNA expression in glioma is dysregulated and abnormal remains elusive, and the underlying mechanisms are still unknown.

Since lncRNAs mainly exert their functions by interacting with other biomolecules, the secondary structures of the interaction domains are of particular importance, and we anticipate that further investigations will focus on resolving lncRNA-binding motifs, which could yield new RNA-based targets for the prevention and treatment of this disease.

The lncRNA-based therapeutic strategies currently remain limited in vivo application, due to the relatively poor stability and intracellular uptake of drugs. Exosome delivery systems can greatly increase their bioavailability by preserving integrity, and the engineering of ligand-dependent exosome membranes that target specific brain tissues is also drawing interest, making therapy more efficient and target-specific.

Abbreviations

Akt:

Protein kinase B

ALKBH5:

A-ketoglutarate-dependent dioxygenase alkB homolog 5

ASO:

Antisense oligonucleotide

Bcl-x:

B-cell lymphoma x

BET:

Bromodomain and extraterminal

BRK:

Breast tumor kinase

BTB:

Blood–tumor barrier

CASC2:

Cancer susceptibility candidate 2

circRNA:

Circular RNA

ciRNA:

Circular intronic RNA

CRNDE:

Colorectal neoplasia differentially expressed

CSC:

Cancer stem-cell-like

CXCR7:

Chemokine receptor 7b

EMT:

Epithelial-to-mesenchymal transition

EOC:

Epithelial ovarian cancer cell

eRNA:

Enhancer-associated RNA

ERV:

Endogenous retrovirus

ESM-1:

Endothelial cell-specific molecule 1

EZH2:

Enhancer of zeste homolog 2

FLC:

FLOWERING LOCUS C

FOXM1-AS:

Forkhead box protein M1 antisense transcript

GBM:

Glioblastoma multiform

GECs:

Glioma-associated endothelial cells

GSCs:

GBM stem-like cells

H3K27me3:

Histone 3 lysine 27 trimethylation

H3K56ac:

Histone 3 lysine 56 acetylation

HIF1α:

Hypoxia-inducible factorα

hnRNP:

Heterogeneous nuclear ribonucleoprotein

HOTAIR:

HOX transcript antisense RNA

HOXA11-AS:

Homeobox protein Hox-A11 antisense RNA 1

HULC:

Human universal load carrier

lincRNAs:

Intergenic lncRNAs

LINK-A:

Long intergenic noncoding RNA for kinase activation

LncRNAs:

Long non-coding RNAs

MALAT1:

Metastasisassociated lung adenocarcinoma transcript 1

MEG3:

Maternally expressed gene 3

MMP 3:

Matrix metalloproteinase 3

MTA2:

Metastasis-associated protein 2

mTOR:

Mechanistic target of rapamycin

NEAT1_2:

Nuclear enriched abundant transcript 1 the long isoform

ORF:

Open reading frame

PI3K:

Phosphatidylinositol-4,5-bisphosphate 3-kinase

PRC2:

Polycomb repressive complex 2

PROMPT:

Promoter upstream transcript

PTBP1:

Polypyrimidine tract-binding protein 1

RISC:

RNA-induced silencing complex

RNase P:

Ribonuclease P

SERCA:

Sarcoplasmic reticulum Ca2+-ATPase

SHP:

Small heterodimer partner

m6A

N6-Methyladenosine

siRNA:

Small interfering RNA

SIRT6:

Sirtuin-6,

Snail:

Snail1

SNHG5:

Small nucleolar RNA host gene 5

sno-lncRNA:

SnoRNA-ended lncRNA

snoRNA:

Small nucleolar RNA

SPA:

5’ SnoRNA-ended and 3′-polyadenylated lncRNA

SPAR:

Small regulatory polypeptide of amino acid response

TAM:

Tumor associated macrophage

TERRA:

Telomeric repeat-containing RNA

TMZ:

Temozolomide

TrxG:

Trithorax Group proteins

TUG1:

Taurine upregulated gene 1

TUSC7:

Tumor suppressor candidate 7

USF:

Stimulating factor

XIST:

X-inactive specific transcript

Reference

  1. Ulitsky I, Bartel DP. lincRNAs: genomics, evolution, and mechanisms. Cell. 2013;154(1):26–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Lagarde J, et al. Extension of human lncRNA transcripts by RACE coupled with long-read high-throughput sequencing (RACE-Seq). Nat Commun. 2016;7:12339.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Deveson IW, et al. The dimensions, dynamics, and relevance of the mammalian noncoding transcriptome. Trends Gen Tig. 2017;33(7):464–78.

    Article  CAS  Google Scholar 

  4. Almeida M, et al. PCGF3/5-PRC1 initiates Polycomb recruitment in X chromosome inactivation. Science. 2017;356(6342):1081–84.

    Article  CAS  PubMed  Google Scholar 

  5. Chen LL. Linking long noncoding RNA localization and function. Trends Biochem Sci. 2016;41(9):761–72.

    Article  CAS  PubMed  Google Scholar 

  6. Hu X, et al. The role of long noncoding RNAs in cancer: the dark matter matters. Curr Opin Genet Dev. 2017;48:8–15.

    Article  PubMed  Google Scholar 

  7. Shi J, et al. Long non-coding RNA in glioma: signaling pathways. Oncotarget. 2017;8(16):27582–92.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Yan Y, et al. An insight into the increasing role of LncRNAs in the pathogenesis of gliomas. Front Mol Neurosci. 2017;10(24516):53.

    PubMed  PubMed Central  Google Scholar 

  9. Wang Q, et al. A novel cell cycle-associated lncRNA, HOXA11-AS, is transcribed from the 5-prime end of the HOXA transcript and is a biomarker of progression in glioma. Cancer Lett. 2016;373(2):251–9.

    Article  CAS  PubMed  Google Scholar 

  10. Jing SY, et al. Expression of long non-coding RNA CRNDE in glioma and its correlation with tumor progression and patient survival. Eur Rev Med Pharmacol Sci. 2016;20(19):3992–6.

    PubMed  Google Scholar 

  11. Wu F, et al. Upregulation of long noncoding RNA HOXA-AS3 promotes tumor progression and predicts poor prognosis in glioma. Oncotarget. 2017;8(32):53110–23.

    PubMed  PubMed Central  Google Scholar 

  12. St Laurent G, Wahlestedt C, Kapranov P. The landscape of long noncoding RNA classification. Trends Genet. 2015;31(5):239–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Rinn JL, Chang HY. Genome regulation by long noncoding RNAs. Annu Rev Biochem. 2012;81(1):145–66.

    Article  CAS  PubMed  Google Scholar 

  14. Ma L, Bajic VB, Zhang Z. On the classification of long non-coding RNAs. RNA Biol. 2013;10(6):925–33.

    Article  PubMed  Google Scholar 

  15. Li W, et al. Functional roles of enhancer RNAs for oestrogen-dependent transcriptional activation. Nature. 2013;498(7455):516–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Quinn JJ, Chang HY. Unique features of long non-coding RNA biogenesis and function. Nat Rev Genet. 2016;17(1):47–62.

    Article  CAS  PubMed  Google Scholar 

  17. Jain AK, et al. LncPRESS1 is a p53-regulated LncRNA that safeguards pluripotency by disrupting SIRT6-mediated de-acetylation of histone H3K56. Mol Cell. 2016;64(5):967–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Lin A, et al. TheLINK-AlncRNA activates normoxic HIF1α signaling in triple-negative breast cancer. Nat Cell Biol. 2016;18(2):213–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Somarowthu S, et al. HOTAIR forms an intricate and modular secondary structure. Mol Cell. 2015;58(2):353–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Wu H, Yang L, Chen LL. The diversity of long noncoding RNAs and their generation. Trends Genet. 2017;33(8):540–52.

    Article  CAS  PubMed  Google Scholar 

  21. Wu H, et al. Unusual processing generates SPA LncRNAs that sequester multiple RNA binding proteins. Mol Cell. 2016;64(3):534–48.

    Article  CAS  PubMed  Google Scholar 

  22. Zhang Y, Yang L, Chen LL. Life without a tail: new formats of long noncoding RNAs. Int J Biochem Cell Biol. 2014;54(1):338–49.

    Article  CAS  PubMed  Google Scholar 

  23. Meng S, et al. CircRNA: functions and properties of a novel potential biomarker for cancer. Mol Cancer. 2017;16(1):94.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Zhang Y, et al. Circular intronic long noncoding RNAs. Mol Cell. 2013;51(6):792–806.

    Article  CAS  PubMed  Google Scholar 

  25. Chen Q, et al. Long non-coding RNA NEAT1, regulated by the EGFR pathway, contributes to glioblastoma progression through the WNT/beta-catenin pathway by scaffolding EZH2. Clin Cancer Res. 2017;24(3):684–95.

    Article  PubMed  Google Scholar 

  26. Quagliata L, et al. Long noncoding RNA HOTTIP/HOXA13 expression is associated with disease progression and predicts outcome in hepatocellular carcinoma patients. Hepatology. 2014;59(3):911–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Huang B, et al. Long non-coding antisense RNA KRT7-AS is activated in gastric cancers and supports cancer cell progression by increasing KRT7 expression. Oncogene. 2016;35(37):4927–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Singh R, et al. Regulation of alternative splicing of Bcl-x by BC200 contributes to breast cancer pathogenesis. Cell Death Dis. 2016;7(6):e2262.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Zhang L, et al. Long noncoding RNA MEG3 induces cholestatic liver injury by interaction with PTBP1 to facilitate shp mRNA decay. Hepatology. 2017;65(2):604–15.

    Article  CAS  PubMed  Google Scholar 

  30. Zheng J, et al. CRNDE promotes malignant progression of glioma by attenuating miR-384/PIWIL4/STAT3 axis. Mol Ther J Am Soc Gene Ther. 2016;24(7):1199–215.

    Article  CAS  Google Scholar 

  31. Li D, et al. Long noncoding RNA HULC modulates the phosphorylation of YB-1 through serving as a scaffold of extracellular signal-regulated kinase and YB-1 to enhance hepatocarcinogenesis. Hepatology. 2017;65(5):1612–27.

    Article  CAS  PubMed  Google Scholar 

  32. Zhao L, et al. Long non-coding RNA SNHG5 suppresses gastric cancer progression by trapping MTA2 in the cytosol. Oncogene. 2016;35(44):5770–80.

    Article  CAS  PubMed  Google Scholar 

  33. Andrews SJ, Rothnagel JA. Emerging evidence for functional peptides encoded by short open reading frames. Nat Rev Genet. 2014;15(3):193–204.

    Article  CAS  PubMed  Google Scholar 

  34. Anderson DM, et al. A micropeptide encoded by a putative long noncoding RNA regulates muscle performance. Cell. 2015;160(4):595–606.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Matsumoto A, et al. mTORC1 and muscle regeneration are regulated by the LINC00961-encoded SPAR polypeptide. Nature. 2017;541(7636):228–32.

    Article  CAS  PubMed  Google Scholar 

  36. Nelson BR, et al. A peptide encoded by a transcript annotated as long noncoding RNA enhances SERCA activity in muscle. Science. 2016;351(6270):271–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Huang JZ, et al. A peptide encoded by a putative lncRNA HOXB-AS3 suppresses colon cancer growth. Mol Cell. 2017;68(1):171–84.

    Article  CAS  PubMed  Google Scholar 

  38. Barile L, Vassalli G. Exosomes: therapy delivery tools and biomarkers of diseases. Pharmacol Ther. 2017;​174:63–78.

  39. Qu L, et al. Exosome-transmitted lncARSR promotes Sunitinib resistance in renal cancer by acting as a competing endogenous RNA. Cancer Cell. 2016;29(5):653–68.

    Article  CAS  PubMed  Google Scholar 

  40. Conigliaro A, et al. CD90+ liver cancer cells modulate endothelial cell phenotype through the release of exosomes containing H19 lncRNA. Mol Cancer. 2015;14(1):155.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Wu Q, et al. Suppression of endothelial cell migration by tumor associated macrophage-derived exosomes is reversed by epithelial ovarian cancer exosomal lncRNA. Cancer Cell Int. 2017;17(1):62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Wang Y, et al. CRNDE, a long-noncoding RNA, promotes glioma cell growth and invasion through mTOR signaling. Cancer Lett. 2015;367(2):122–8.

    Article  CAS  PubMed  Google Scholar 

  43. Ke J, 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(26):21934–49.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Zhao X, et al. GAS5 suppresses malignancy of human glioma stem cells via a miR-196a-5p/FOXO1 feedback loop . Biochim Biophys Acta. 2017;1864(10):1605–17.

    Article  CAS  PubMed  Google Scholar 

  45. Liao Y, et al. LncRNA CASC2 interacts with miR-181a to modulate glioma growth and resistance to TMZ through PTEN pathway.J Cell Biochem. 2017;118(7):1889–99. 

  46. Deguchi S, et al. Oncogenic effects of evolutionarily conserved noncoding RNA ECONEXIN on gliomagenesis. Oncogene. 2017;36(32):4629–40.

    Article  CAS  PubMed  Google Scholar 

  47. Katsushima K, et al. Targeting the notch-regulated non-coding RNA TUG1 for glioma treatment. Nat Commun. 2016;7:13616.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Yao Y, et al. Knockdown of long non-coding RNA XIST exerts tumor-suppressive functions in human glioblastoma stem cells by up-regulating miR-152. Cancer Lett. 2015;359(1):75–86.

    Article  CAS  PubMed  Google Scholar 

  49. Dixit D, et al. Messenger RNA methylation regulates glioblastoma tumorigenesis. Cancer Cell. 2017;31(4):474–5.

    Article  CAS  PubMed  Google Scholar 

  50. Zhang S, et al. m6A demethylase ALKBH5 maintains Tumorigenicity of glioblastoma stem-like cells by sustaining FOXM1 expression and cell proliferation program. Cancer Cell. 2017;31(4):591–606. e6

    Article  CAS  PubMed  Google Scholar 

  51. Pandey GK, et al. The risk-associated long noncoding RNA NBAT-1 controls neuroblastoma progression by regulating cell proliferation and neuronal differentiation. Cancer Cell. 2014;26(5):722–37.

    Article  CAS  PubMed  Google Scholar 

  52. Zhang K, et al. Long non-coding RNA HOTAIR promotes glioblastoma cell cycle progression in an EZH2 dependent manner. Oncotarget. 2015;6(1):537–46.

    PubMed  Google Scholar 

  53. Teng H, et al. Role of HCP5-miR-139-RUNX1 feedback loop in regulating malignant behavior of glioma cells. Mol Ther. 2016;24(10):1806–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Cao S, et al. Tumor-suppressive function of long noncoding RNA MALAT1 in glioma cells by suppressing miR-155 expression and activating FBXW7 function. Am J Cancer Res. 2016;6(11):2561–74.

    PubMed  PubMed Central  Google Scholar 

  55. Shang C, et al. Long non-coding RNA TUSC7, a target of miR-23b, plays tumor-suppressing roles in human gliomas. Front Cell Neurosci. 2016;10(6):235.

    PubMed  PubMed Central  Google Scholar 

  56. Han Y, et al. Tumor-suppressive function of long noncoding RNA MALAT1 in glioma cells by downregulation of MMP2 and inactivation of ERK/MAPK signaling. Cell Death Dis. 2016;7(3):e2123.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Jia P, et al. Long non-coding RNA H19 regulates glioma angiogenesis and the biological behavior of glioma-associated endothelial cells by inhibiting microRNA-29a. Cancer Lett. 2016;381(2):359–69.

    Article  CAS  PubMed  Google Scholar 

  58. Yu H, et al. Knockdown of long non-coding RNA XIST increases blood-tumor barrier permeability and inhibits glioma angiogenesis by targeting miR-137. Oncogene. 2017;6(3):e303.

    Article  CAS  Google Scholar 

  59. Cheng Z, et al. Long non-coding RNA XIST promotes glioma Tumorigenicity and angiogenesis by acting as a molecular sponge of miR-429. J Cancer. 2017;8(19):4106–16.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Zhu Y, et al. HULC long noncoding RNA silencing suppresses angiogenesis by regulating ESM-1 via the PI3K/Akt/mTOR signaling pathway in human gliomas. Oncotarget. 2016;7(12):14429–40.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Ma J, et al. Knockdown of long non-coding RNA MALAT1 increases the blood-tumor barrier permeability by up-regulating miR-140. Biochim Biophys Acta. 2016;1859(2):324–38.

    Article  CAS  PubMed  Google Scholar 

  62. Sa L, et al. The role of HOTAIR/miR-148b-3p/USF1 on regulating the permeability of BTB. Front Mol Neurosci. 2017;10:194.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Cai H, et al. The long noncoding RNA TUG1 regulates blood-tumor barrier permeability by targeting miR-144. Oncotarget. 2015;6(23):19759–79.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Messaoudi K, Clavreul A, Lagarce F. Toward an effective strategy in glioblastoma treatment. Part I: resistance mechanisms and strategies to overcome resistance of glioblastoma to temozolomide. Drug Discov Today. 2015;20(7):899–905.

    Article  CAS  PubMed  Google Scholar 

  65. Li H, et al. Long non-coding RNA MALAT1 decreases the sensitivity of resistant glioblastoma cell lines to Temozolomide. Cell Physiol Biochem. 2017;42(3):1192–1201.

    Article  CAS  PubMed  Google Scholar 

  66. Huarte M. The emerging role of lncRNAs in cancer. Nat Med. 2015;21(11):1253–61.

    Article  CAS  PubMed  Google Scholar 

  67. Di Meo A, et al. Liquid biopsy: a step forward towards precision medicine in urologic malignancies. Mol Cancer. 2017;16(1):80.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Yuan W, et al. Circulating LncRNAs serve as diagnostic markers for hepatocellular carcinoma. Cell Physiol Biochem. 2017;44(1):125–32.

    Article  PubMed  Google Scholar 

  69. Gourlay J, et al. The emergent role of exosomes in glioma. J Clin Neurosci. 2017;35:13–23.

    Article  CAS  PubMed  Google Scholar 

  70. Matsui M, Corey DR. Non-coding RNAs as drug targets. Nat Rev Drug Discov. 2017;16(3):167–79.

    Article  CAS  PubMed  Google Scholar 

  71. Pastori C, et al. The Bromodomain protein BRD4 controls HOTAIR, a long noncoding RNA essential for glioblastoma proliferation. Proc Natl Acad Sci U S A. 2015;112(27):8326–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Arun G, et al. Differentiation of mammary tumors and reduction in metastasis upon Malat1 lncRNA loss. Genes Dev. 2016;30(1):34–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Cheng CJ, et al. MicroRNA silencing for cancer therapy targeted to the tumour microenvironment. Nature. 2015;518(7537):107–10.

    Article  CAS  PubMed  Google Scholar 

  74. Özeş AR, et al. Therapeutic targeting using tumor specific peptides inhibits long non-coding RNA HOTAIR activity in ovarian and breast cancer. Sci Rep. 2017;7(1):894.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Monroig Pdel C, et al. Small molecule compounds targeting miRNAs for cancer therapy. Adv Drug Deliv Rev. 2015;81:104–16.

    Article  PubMed  Google Scholar 

  76. Zhang R, Jin H, Lou F. The long non-coding RNA TP73-AS1 interacted with miR-142 to modulate brain glioma growth through HMGB1/RAGE pathway. J Cell Biochem. 2017. [Epub ahead of print].

  77. Li DX, et al. The long non-coding RNA CRNDE acts as a ceRNA and promotes glioma malignancy by preventing miR-136-5p-mediated downregulation of Bcl-2 and Wnt2. Oncotarget. 2017;8(50):88163–78.

    PubMed  PubMed Central  Google Scholar 

  78. Li C, et al. H19 derived microRNA-675 regulates cell proliferation and migration through CDK6 in glioma. Am J Transl Res. 2015;7(10):1747–64.

    PubMed  PubMed Central  Google Scholar 

  79. Su R, et al. Knockdown of SOX2OT inhibits the malignant biological behaviors of glioblastoma stem cells via up-regulating the expression of miR-194-5p and miR-122. Mol Cancer. 2017;16(1):171.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Liu C, et al. CASC2c as an unfavorable prognosis factor interacts with miR-101 to mediate astrocytoma tumorigenesis. Cell Death Dis. 2017;8(3):e2639.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Li J, et al. LncRNA TUG1 acts as a tumor suppressor in human glioma by promoting cell apoptosis. Exp Biol Med (Maywood). 2016;241(6):644–9.

    Article  CAS  Google Scholar 

  82. Zhao X, et al. Gas5 exerts tumor-suppressive functions in human glioma cells by targeting miR-222. Mol Ther. 2015;23(12):1899–911.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Funding

This work was supported by National Key Technology Research and Development program of the Ministry of Science and Technology of China (2014BAI04B02) and Graduate Research and Innovation Projects of Central South University (Grant 2017zzts012).

Author information

Authors and Affiliations

Authors

Contributions

ZP wrote the manuscript and drew the figures. CL and MW made corrections for the manuscript. MW gave the idea of the review and made the final corrections for the manuscript. All authors read and approved the final manuscript under the supervision of MW before the submission.

Corresponding author

Correspondence to Minghua Wu.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Peng, Z., Liu, C. & Wu, M. New insights into long noncoding RNAs and their roles in glioma. Mol Cancer 17, 61 (2018). https://doi.org/10.1186/s12943-018-0812-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12943-018-0812-2

Keywords