Next Article in Journal
Clinical Effectiveness of Faecal Immunochemical Test in the Early Detection of Colorectal Cancer—An Umbrella Review
Previous Article in Journal
Does Vaginal Cuff Creation and Avoidance of a Uterine Manipulator Improve the Prognosis of Total Laparoscopic Radical Hysterectomy for Early Cervical Cancer? A Retrospective Multicenter Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Long and the Short of It: NEAT1 and Cancer Cell Metabolism

by
Nadine E. Smith
1,
Phaedra Spencer-Merris
1,
Archa Hannah Fox
2,
Janni Petersen
1,* and
Michael Z. Michael
1,3,*
1
Flinders Health and Medical Research Institute, Cancer Program, Flinders University, Bedford Park, SA 5042, Australia
2
School of Human Sciences, School of Molecular Sciences, The University of Western Australia, Crawley, WA 6009, Australia
3
Department of Gastroenterology and Hepatology, Flinders Centre for Innovation in Cancer, Flinders Medical Centre, Bedford Park, SA 5042, Australia
*
Authors to whom correspondence should be addressed.
Cancers 2022, 14(18), 4388; https://doi.org/10.3390/cancers14184388
Submission received: 29 August 2022 / Accepted: 5 September 2022 / Published: 9 September 2022
(This article belongs to the Topic Cancer Cell Metabolism)

Abstract

:

Simple Summary

Altered metabolism is a hallmark of most cancers. The way that cancer cells regulate their energy production to fuel constant proliferation has been of interest with the hope that it may be exploited therapeutically. The long noncoding RNA, NEAT1, is often dysregulated in tumours. NEAT1 RNA can be transcribed as two isoforms with different lengths, with each variant responsible for different functions. This review explores how the isoforms contribute to cancer metabolism.

Abstract

The long noncoding RNA NEAT1 is known to be heavily dysregulated in many cancers. A single exon gene produces two isoforms, NEAT1_1 and NEAT1_2, through alternative 3′-end processing. As the longer isoform, NEAT1_2 is an essential scaffold for nuclear paraspeckle formation. It was previously thought that the short NEAT1_1 isoform only exists to keep the NEAT1 locus active for rapid paraspeckle formation. However, a recent glycolysis-enhancing function for NEAT1_1, contributing to cancer cell proliferation and the Warburg effect, has been demonstrated. Previous studies have mainly focused on quantifying total NEAT1 and NEAT1_2 expression levels. However, in light of the NEAT1_1 role in cancer cell metabolism, the contribution from specific NEAT1 isoforms is no longer clear. Here, the roles of NEAT1_1 and NEAT1_2 in metabolism and cancer progression are discussed.

1. Background

In healthy cells, glucose uptake stimulates cell growth through the activation of intracellular signalling pathways, including glycolysis [1]. In cancer, genetic and epigenetic alterations switch these pathways to a permanently “on” state, promoting continuous growth, which depletes key metabolites. As tumour size increases, oxygen availability decreases, limiting the oxygen-dependent final step of the electron transport chain (ETC). This shift from oxidative phosphorylation (OXPHOS) to aerobic glycolysis is termed the Warburg effect [2,3,4]. Although there is an increase in both glycolytic enzymes and activity in many cancers, the Warburg effect alone does not explain why cancer cells maintain enhanced glycolysis in normoxic cell culture nor in circulating blood [4].
In general, the functions of long noncoding RNAs (lncRNAs) are not well-understood due to their generally low expression levels and high tissue specificity [5]. An exception to this is nuclear enriched abundant transcript 1 (NEAT1) dysregulation in many diseases [6]. The highly conserved, single exon, intergenic lncRNA is transcribed near the multiple endocrine neoplasia locus on human chromosome 11q13.1 [7]. Altered 3′-end processing results in two transcripts: 3.7 kb NEAT1_1 and 22.7 kb NEAT1_2; the latter is a well-documented and essential architectural scaffold for subnuclear paraspeckles (see review by McCluggage and Fox, 2021) (Figure 1) [8,9]. A triple-helix structure, processed by RNase P at the 3′-end of NEAT1_2, stabilises the transcript to protect it from degradation [9,10]. Less studied is the shorter, polyadenylated, paraspeckle-independent NEAT1_1 isoform, whose function, until recently, remained elusive. As previously reported, abnormal NEAT1 expression is correlated with several malignancies, and the distinction between the two isoforms has often been overlooked [11]. Hence, the relative contribution of each isoform to metabolic homeostasis, or pathology, requires attention.

2. NEAT1_1 Enhances the Warburg Effect by Accelerating Glycolytic Metabolite Flux

Although both isoforms are abundant in the nucleus, NEAT1_1 is also exported to the cytoplasm [12]. In 2019, Adriaens et al. suggested that NEAT1_1 is potentially nonfunctional and serves to only keep the transcriptional locus active, making the switch to NEAT1_2 rapidly available during stress induction [13]. However, a recent study has discovered a novel mechanism of action for NEAT1_1 in the cytoplasm, in both in vitro and in vivo breast cancer (BC) models [12]. The authors demonstrated that the translocation of NEAT1_1 from the nucleus to the cytoplasm, through binding with the nuclear speckle-associated protein pinin (encoded by the PNN gene), occurs in a glucose-dependent manner [12,14] (Figure 2). Interestingly, the depletion of pinin reduces cytoplasmic NEAT1_1 even after glucose stimulation, indicating the importance of pinin in the nucleocytoplasmic transport of NEAT1_1 [12]. Importantly, in the cytoplasm, NEAT1_1 interacts with the glycolytic enzymes phosphoglycerate kinase (PGK1), phosphoglycerate mutase (PGAM1), and alpha enolase (ENO1) to promote glycolysis, enhancing growth, proliferation, invasion, and metastasis (Figure 2) [12,15].
PGK1 is constitutively expressed in all somatic and premeiotic cells and is essential in the glycolysis pathway, but it is implicated in cancer as it encompasses many characteristics of an oncogene [16]. This bilobed enzyme has both nucleotide-binding and catalytic domains and is involved in the conversion of 1,3-bisphosphoglycerate (1,3-BPG) to 3-phosphoglycerate (3-PG) in glycolysis (Figure 2A). It also catalyses the first ATP of anerobic glycolysis [16,17]. This substrate-level phosphorylation is of great significance in the continuous production of cellular energy under hypoxic conditions [16]. PGK1 also acts as a protein kinase after translocation to the mitochondria, where it directly phosphorylates pyruvate dehydrogenase kinase isozyme 1 (PDHK1) [17]. Phosphorylated PDHK enhances pyruvate dehydrogenase E1α, which inactivates pyruvate dehydrogenase, preventing the conversion of pyruvate to coenzyme A (CoA), thus suppressing mitochondrial pyruvate metabolism and increasing lactate production [18]. This rate-limiting enzyme plays a role in the promotion of tumourigenesis through the activation of oncogenic pathways, such as Akt/mTOR, Myc, and ß-catenin, and post-translational modifications, such as phosphorylation, acetylation, ubiquitination, and succinylation (as reviewed by Liu et al., 2022) [19]. In vitro research by Gou et al. [20] and Wang et al. [21] have shown that the small-molecule inhibitor of PGK1, NG52, had a dose-dependent effect on the proliferation of ovarian cancer and glioma cells, respectively.
PGAM1 is a highly conserved glycolytic isomerase enzyme involved in the conversion of 3-PG to 2-phosphoglycerate (2-PG), whilst also supporting antioxidative defences by reducing mitochondrial reactive oxygen species (ROS) (Figure 2B) [22,23]. Several studies have linked PGAM1 to cancer progression. Earlier works report PGAM1 knockdown by short hairpin RNA (shRNA) results in an increase in 3-PG and a subsequent decrease in 2-PG, whilst also decreasing glycolysis, pentose phosphate pathway flux, biosynthesis, and cell proliferation in diverse solid and leukaemia cell lines [24]. Investigation of dysregulated PGAM1 levels in human urothelial bladder cancer tissues found a positive correlation with histological-grade tumours, when compared to adjacent normal tissue [25]. Loss of functional tumour-suppressor protein p53, encoded by tumour protein 53 (TP53), is common in cancer, and it has been found to upregulate both NEAT1 and PGAM1 [23]. Similar to many metabolic enzymes, PGAM1 asserts a nonenzymatic function, as the physical interaction with checkpoint kinase 1 (Chk1) increases proliferation, specifically in RAS-driven cancer cells [26].
Alpha enolase (ENO1), encoded by ENO1, is another tumour-related, multifunctional protein, which is responsible for the conversion of 2-PG to phosphoenolpyruvate (PEP) in glycolysis (Figure 2C). An increase in ENO1 expression has been reported in human diseases (i.e., systemic sclerosis, type II diabetes mellitus, lupus, Alzheimer disease) and many cancers, as well as being involved in cell growth and hypoxia tolerance [27,28,29,30,31,32,33,34]. Additionally, silencing ENO1 reduces the rate of glycolysis in cell lines, favouring OXPHOS even when glucose influx remains high [35]. ENO1 expression is correlated with colorectal cancer (CRC) progression, and the newly identified protein translational modification, lysine crotonylation, has been identified at lysine residue 420 in CRC cell lines [33]. Interestingly, although ENO1 is reportedly overexpressed in many human diseases and cancers, in non-small-cell lung cancer (NSCLC), ENO1 is downregulated at the protein level even though its mRNA levels remain elevated, suggesting post-transcriptional regulation [36,37].
Given the cancer-specific roles for each of these enzymes, the role that NEAT1_1 plays, either in glycolysis or in sequestering enzymes from other activities, requires further investigation.

3. NEAT1_2 and Paraspeckle Abundance Increase following Stress

First described in 2002 by Fox et al., paraspeckles are discrete, subnuclear bodies, measuring approximately 360 nm in diameter [38,39,40]. Paraspeckle formation relies solely on the generation of NEAT1_2 in the nucleus, which sets them apart from cytoplasmic stress granules and P bodies, which require multiple proteins and RNA elements to form [8]. Paraspeckle formation occurs only following the recruitment of proteins, such as non-POU-domain-containing octamer-binding protein (NONO), splicing factor proline and glutamine-rich (SFPQ) and fused in sarcoma (FUS) proteins, among others, to the mid-region of NEAT1_2 transcript. Once localised, the high concentration of molecules aggregates to form a distinct spheroid with spatial organisation [41] (Figure 1). Hydrophilic proteins bind 3′ and 5′ regions of the NEAT1_2 transcript to form the paraspeckle shell, whilst the middle segment of the transcript forms the hydrophobic core [8,10,38,42]. Individual paraspeckles are spheroidal, but during stress conditions, they can be linked together to generate elongated paraspeckle structures [15]. In HeLa cells, there are ~5–20 paraspeckles per nucleus [43]. However, in 2020, Grosch et al. reported that the size of the nuclei likely influences paraspeckle abundance in human pluripotent stem cells [43]. Regardless of basal paraspeckle abundance, their numbers seem to increase during stress, suggesting that NEAT1_2 accumulates.

4. What Is Driving the NEAT1 Isoform Switch?

Since the formation of paraspeckles is dependent on the transcriptional read-through of NEAT1_1 to NEAT1_2, understanding the mechanistic control of this isoform switch is crucial. Polyadenylation (polyA) signals terminate the primary NEAT1 transcript, resulting in canonical processing of the 3′ polyA tail and consequently a short NEAT1_1 lncRNA [44]. The long NEAT1_2 isoform is generated when heterogeneous nuclear ribonucleoprotein K (hnRNPK) competes with cleavage-and-polyadenylation-specific factor 6 (CPSF6) for nudix hydrolase 21 (NUDT21) binding, inhibiting the CPSF6–NUDT21 (CFIm) complex from forming, and facilitating the 3′-end polyA (Figure 2) [44,45]. In vitro binding assays have demonstrated that inhibiting the formation of the CFIm complex prevents the polyadenylation of NEAT1_1 and increases the nuclear levels of NEAT1_2 and, thereby, paraspeckles [44,46]. Additionally, recent work has reported that arsenic resistance protein 2 (ARS2) acts as a chaperone, guiding CFIm to the NEAT1 transcript to facilitate the polyadenylation of NEAT1_1 in osteosarcoma cell lines [47]. Furthermore, the knockdown of ARS2 leads to an increase in, and the preferential stabilisation of, NEAT1_2 transcripts [47]. Moreover, RNA binding protein (RBP) transactive response (TAR) DNA binding protein 43 kDa (TDP-43) directly represses the formation of paraspeckles, but it increases NEAT1_1 transcription by binding the NEAT1_1 GU-rich motifs upstream of the polyA site [48,49]. Although TDP-43 has been thoroughly investigated in amyotrophic lateral sclerosis and has been linked to altered miRNA expression, the understanding of its involvement in a cancer context remains limited [49,50,51,52,53,54,55]. In summary, the isoform switch from NEAT1_1 to NEAT1_2 may involve several factors with context-specific roles; their oncogenic relevance requires further clarification.

5. NEAT1 and Paraspeckles Alter Metabolism via Mitochondria

Mitochondrial function reaches beyond the established role in energy generation. In addition to ATP production, mitochondria generate macromolecules which alleviate mitochondrial stress [56]. Interestingly, the mitochondrial stressors, FCCP, rotenone, and doxycycline, all increase NEAT1 levels, in part through ATF2-induced NEAT1 expression [10]. Mito–nuclear communication is crucial for ensuring cellular homeostasis during mitochondrial stress, and recently it has been hypothesised that NEAT1 and paraspeckles may play a role in mitochondrial homeostasis [57,58]. Mitochondrial fusion and fission are controlled by dynamin-related GTPases MFN1/MFN2 and DRP1, respectively [58]. NEAT1-depletion in HeLa and HEK293 cells resulted in mitochondrial elongation and the enhanced retention of mito-mRNAs encoding functional mitochondrial proteins, such as cytochrome c, subunits of NADH dehydrogenases, and carnitine o-palmitoyl transferase [10]. This was further supported by reduced DRP1 but stable MFN1 and MFN2 expression, increased mitochondrial mRNAs (mito-mRNAs) exported from the nucleus, reduced respiration capacity, ATP generation, extracellular acidification rate (ECAR), and proliferation [10]. On the contrary, NEAT1 overexpression showed an increase in DRP1 expression and DRP1 phosphorylation, corresponding to fragmented mitochondria and increased mitochondrial DNA (mtDNA), and this was phenocopied by deleting the NEAT1_1 polyadenylation signal [10].

6. Alternative Processing of lncRNAs in Cancer

Many lncRNAs are implicated in cancer, either through direct or indirect processing, as previously reviewed [59,60]. Similarly to NEAT1, the noncoding product of a neighbouring locus metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) is retained in the nucleus and is involved in the nuclear architecture as well as in RNA splicing and gene-regulation [61]. Whilst MALAT1 has been found to be overexpressed in 14% of breast tumour samples, an alternatively spliced variant of MALAT1 ( Δ sv-MALAT1) showed decreased expression in a subset of tumours and shows potential as an individual prognostic factor for BC [62]. Additionally, similar to NEAT1, MALAT1 contains a tRNA-like clover-leaf structure near the 3′ terminus; however, this structure is cleaved [63] and accumulates as MALAT1-associated small cytoplasmic RNA (mascRNA) in the cytoplasm, where it promotes global translation [64] and hepatocellular cancer cell proliferation [65]. Additionally, the lncRNA ANRIL recruits polycomb proteins to regulate target-gene expression, and the overexpression of truncated isoforms has been reported in bladder cancers [66]. Clearly, the function of many lncRNAs is nuanced and not limited to the full-length transcript. Understanding the activities of truncated forms and processed derivatives of lncRNAs will enlighten our understanding of these complex regulatory molecules.

7. Dysregulation of Both Short and Long NEAT1 in Cancer

NEAT1 (without distinguishing between the long and short isoform) is expressed at similar levels in most healthy tissues, but it is reported to have a relatively low expression in the brain, heart, and whole blood [6]. On the contrary, NEAT1 is upregulated in many solid cancers (reviewed in [6,67,68,69,70]), and in most cases, it is associated with aggressive disease and poor outcomes [67,68,69]. However, only a handful of recent studies have established the relative abundance of the two isoforms, despite many reports suggesting that NEAT1 acts as either a tumour suppressor or oncogene, depending on this isoform ratio (summarised in Table 1) [70,71,72,73,74]. In some haematological malignancies, NEAT1 functions as a tumour suppressor, enhancing the expression of NEAT1_2, and thus paraspeckle formation, which may counteract oncogene-induced stressors in cancer [75,76]. NEAT1 levels are upregulated in multiple myeloma (MM) when compared to healthy controls; however, no correlation with patient prognosis has been found [77]. In chronic lymphocytic leukaemia (CLL), the total NEAT1 expression levels remained similar to those of healthy controls, but the expression of NEAT1_2 was found to be significantly higher [77]. On the other hand, in acute and chronic myeloid leukaemia (AML and CML, respectively) and acute lymphoblastic leukaemia (ALL), total NEAT1 levels decrease in patients’ peripheral blood and bone marrow, and this was found to be an essential mediator of apoptosis induced by imatinib in BCR-ABL-expressing cells [75,77,78,79,80].
Tumour suppressor p53 is regarded as the guardian of the genome, but it is mutated in over 50% of malignancies, enabling cells to escape apoptotic signalling, bypass cell-cycle arrest, and inhibit senescence [112,113]. It is well-established that NEAT1 is induced by p53 binding to the NEAT1 promotor to activate expression [76,112,114]. Interestingly, NEAT1 also promotes p53 and Chk1 through ATR signalling in response to replication stress [76]. Furthermore, in CRC cells, the induction of both NEAT1 isoforms were p53-dependent when exposed to a chemotherapeutic agent and the topoisomerase 2 inhibitor, doxorubicin [112]. In CML, p53 mutations are uncommon [78]; instead, MYC binds to the NEAT1 promotor to enhance expression [78]. Accordingly, Ronchetti et al. reported an increase in total NEAT1 and NEAT1_2 expression in CML patients when compared to normal B-cells [77].
A positive correlation was identified for increased NEAT1 expression and higher histological grades of BC, and NEAT1 levels were elevated in patient plasma and peripheral blood although no relative abundance of NEAT1 isoforms was reported [11]. The same study reported higher expression of NEAT1 in estrogen receptor positive (ER+) breast cancers, when compared to estrogen-receptor-negative- (ER-) BC [11]. Similarly, correlation to lymph node positivity was seen with ER+, but not with ER- [11]. Evidence suggests that NEAT1 point mutations are drivers for breast and prostate cancers, regardless of little change in NEAT1 transcription levels [115,116,117]. However, a more recent study suggested these mutations were likely caused by transcription errors instead of cancer-specific selection pressures [118]. Regardless, NEAT1 downregulation has been reported in invasive breast carcinoma, oesophageal cancer, pheochromocytoma, and paraganglioma, suggesting NEAT1 plays a tumour-suppressor role in these malignancies [6]. In summary, it is apparent that NEAT1 expression varies greatly between malignancies and that relative isoform abundance may play a key role in disease progression.

8. Considerations for Isoform Detection of NEAT1

Most recent NEAT1 studies have concentrated on the paraspeckle-associated long isoform NEAT1_2, at the risk of overlooking essential roles for the shorter NEAT1_1, paraspeckle-independent isoform. In addition, an analysis of the literature (Table 1) shows the reliance placed upon glyceraldehyde-3-phosphate dehydrogenase (GAPDH) transcripts levels for normalizing NEAT1 expression in RT-PCR studies. As new light has been shed on the involvement of NEAT1_1 in glycolysis, a glycolytic housekeeping transcript, such as GAPDH, may not be ideal for normalizing NEAT1 expression in studies related to metabolism.
Inconsistent polyadenylation and contextual processing can make lncRNAs difficult to quantify using standard RT-PCR protocols. Quantifying the relative abundance of NEAT1 isoforms is not straightforward. Kolenda et al. [119] compared cDNA synthesis protocols for various cancer-associated lncRNAs; however, isoform differentiation was not a primary outcome. Similarly, the RNA purification method used can significantly influence the relative abundance of isoforms, with a heating step liberating NEAT1_2 from paraspeckle complexes and increasing yields [120]. While oligo-dT primed cDNA might be used to specifically amplify NEAT1_1 sequences, the presence of poly-A stretches downstream in the NEAT1_2 sequence necessitates the careful calibration of reverse transcription conditions, which is rarely reported. Validated oligo-dT clamp cDNA protocols would be advantageous.
It may be expected that RNA-seq should enable accurate comparison of isoforms; however, many studies report oligo-dT primed libraries, such that the long isoform is under-represented or ignored, and even total RNA-seq data may be influenced by the aforementioned bias in isoform ratios as a result of RNA isolation methods. Newer, long-read direct RNA sequencing methods promise improved qualitative and quantitative data [121].
Visualisation of NEAT1_2, employing RNA-fluorescence in situ hybridization, is often used to quantify paraspeckle abundance and can also be directed to detect NEAT1_1 [13]. Similarly, dCas13 tagging has recently been used to detect both isoforms of NEAT1 in living cells [122].

9. Conclusions and Future Directions

In the context of cancer, NEAT1 may have either a protective, tumour-suppressive role, or a tumour-promoting oncogenic role, depending upon the type of cancer and, most likely, also upon the specific NEAT1 isoform expressed. Many previous studies have concentrated on the total NEAT1 transcription level, rather than on the isoform ratio; hence, isoform-specific contributions are unclear. Improving detection of specific NEAT1 isoforms is crucial in understanding the tumour-promoting vs. the tumour-protective roles of NEAT1 in cancer. Future directions will compare the relevant contributions of paraspeckle-mediated sequestration and epigenetic regulation against the glycolytic influence of the shorter isoform on tumour progression. Whether one or both isoforms are found to be necessary for specifically maintaining neoplasia, that will impact their relative value as therapeutic targets.

Author Contributions

Conceptualization, N.E.S., P.S.-M., A.H.F., J.P., and M.Z.M.; writing—original draft preparation, N.E.S. and J.P.; writing—review and editing, N.E.S., P.S.-M., A.H.F., J.P., and M.Z.M.; administration, J.P. and M.Z.M. All authors have read and agreed to the submission of this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are supported by the Australian National Health and Medical Research Council (GNT2012373), Tour de Cure Australia, and the Flinders Foundation. N.E.S. is supported by a Flinders Health and Medical Research Institute post-graduate award.

Conflicts of Interest

The authors declare no conflict of interest. The funding agencies had no role in the planning or writing of the manuscript, nor in the decision to publish.

Abbreviations

ALLAcute lymphoblastic leukaemia
AMLAcute myeloid leukaemia
APLAcute promyelocytic leukaemia
ATCAnaplastic thyroid carcinoma
ATRAtaxia telangiectasia and Rad3-related
BCBreast cancer
CBMsCirculating blood monocytes
CCCervical cancer
CFImCPSF6-NUDT21 complex
Chk1Checkpoint kinase 1
CLLChronic lymphocytic leukaemia
CMLChronic myeloid leukaemia
CoACo-enzyme A
CPSF6Cleavage and polyadenylation-specific factor 6
CRCColorectal cancer
DRP1Dynamin-related protein 1
ECARExtracellular acidification rate
ENO1Alpha enolase
ETCElectron transport chain
FCCpCarbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone
FUSFused in sarcoma
GAPDHGlyceraldehyde 3-phosphate dehydrogenase
GCGastric cancer
HCCHepatocellular carcinoma
KDKnockdown
KOKnockout
LCLaryngeal cancer
lncRNALong noncoding RNA
LSCCLaryngeal squamous cell cancer
LUADLung adenocarcinoma
mascRNAMALAT1-associated cytoplasmic RNA
MFN1/2Mitofusion protein
MMMultiple myeloma
mtDNAMitochondrial DNA
NADHNicotinamide adenine dinucleotide (NAD) + hydrogen
NEAT1Nuclear enriched abundant transcript 1
NONONon-POU-domain-containing octamer-binding protein
NPCNasopharyngeal carcinoma
NSCLCNon-small-cell lung cancer
NUDT21Nudix hydrolase 21
OCOvarian cancer
OEOver-expression
OSCCOesophageal squamous cell carcinoma
OXPHOSOxidative phosphorylation
p53Tumour suppressor protein 53
PCProstate cancer
PDHK1Pyruvate dehydrogenase kinase isozyme 1
PGAM1Phosphoglycerate mutase 1
PGK1Phosphoglycerate kinase 1
PTCPapillary thyroid carcinoma
RASRat sarcoma virus oncogene
RBPRNA binding protein
ROSReactive oxygen species
SFPQSplicing factor proline and glutamine-rich
shRNAShort hairpin RNA
TAMsTumour associated macrophages
TARTransactive response
TCThyroid carcinoma
TDP-43TAR DNA binding protein 43 kDa

References

  1. Pavlova, N.N.; Zhu, J.; Thompson, C.B. The hallmarks of cancer metabolism: Still emerging. Cell Metab. 2022, 34, 355–377. [Google Scholar] [CrossRef] [PubMed]
  2. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg Effect: The Metabolic Requirements of Cell Proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef]
  3. Warburg, O. On the Origin of Cancer Cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef]
  4. Koppenol, W.H.; Bounds, P.L.; Dang, C.V. Otto Warburg’s contributions to current concepts of cancer metabolism. Nat. Rev. Cancer 2011, 11, 325–337. [Google Scholar] [CrossRef]
  5. Derrien, T.; Johnson, R.; Bussotti, G.; Tanzer, A.; Djebali, S.; Tilgner, H.; Guernec, G.; Martin, D.; Merkel, A.; Knowles, D.G.; et al. The GENCODE v7 Catalog of Human Long Noncoding RNAs: Analysis of Their Gene Structure, Evolution, and Expression. Genome Res. 2012, 22, 1775–1789. [Google Scholar] [CrossRef] [PubMed]
  6. Li, S.; Li, J.; Chen, C.; Zhang, R.; Wang, K. Pan-cancer analysis of long non-coding RNA NEAT1 in various cancers. Gene Funct. Dis. 2017, 5, 27–35. [Google Scholar] [CrossRef]
  7. Bu, F.; Wang, A.; Zhu, Y.; You, H.; Zhang, Y.; Meng, X.; Huang, C.; Li, J.; Zhu, Y. LncRNA NEAT1: Shedding light on mechanisms and opportunities in liver diseases. Liver Int. 2020, 40, 2612–2626. [Google Scholar] [CrossRef] [PubMed]
  8. McCluggage, F.; Fox, A.H. Paraspeckle nuclear condensates: Global sensors of cell stress? BioEssays 2021, 43, e2000245. [Google Scholar] [CrossRef]
  9. Wilusz, J.E.; JnBaptiste, C.K.; Lu, L.Y.; Kuhn, C.-D.; Joshua-Tor, L.; Sharp, P.A. A triple helix stabilizes the 3′ ends of long noncoding RNAs that lack poly(A) tails. Genes Dev. 2012, 26, 2392–2407. [Google Scholar] [CrossRef] [PubMed]
  10. Wang, Y.; Hu, S.-B.; Wang, M.-R.; Yao, R.W.; Wu, D.; Yang, L.; Chen, L.-L. Genome-wide screening of NEAT1 regulators reveals cross-regulation between paraspeckles and mitochondria. Nat. Cell Biol. 2018, 20, 1145–1158. [Google Scholar] [CrossRef]
  11. Li, W.; Zhang, Z.; Liu, X.; Cheng, X.; Zhang, Y.; Han, X.; Zhang, Y.; Liu, S.; Yang, J.; Xu, B.; et al. The FOXN3-NEAT1-SIN3A repressor complex promotes progression of hormonally responsive breast cancer. J. Clin. Investig. 2017, 127, 3421–3440. [Google Scholar] [CrossRef] [PubMed]
  12. Park, M.K.; Zhang, L.; Min, K.-W.; Cho, J.-H.; Yeh, C.-C.; Moon, H.; Hormaechea-Agulla, D.; Mun, H.; Ko, S.; Lee, J.W.; et al. NEAT1 is essential for metabolic changes that promote breast cancer growth and metastasis. Cell Metab. 2021, 33, 2380–2397.e9. [Google Scholar] [CrossRef]
  13. Adriaens, C.; Rambow, F.; Bervoets, G.; Silla, T.; Mito, M.; Chiba, T.; Asahara, H.; Hirose, T.; Nakagawa, S.; Jensen, T.H.; et al. The long noncoding RNA NEAT1_1 is seemingly dispensable for normal tissue homeostasis and cancer cell growth. RNA 2019, 25, 1681–1695. [Google Scholar] [CrossRef]
  14. Alpatov, R.; Munguba, G.C.; Caton, P.; Joo, J.H.; Shi, Y.; Shi, Y.; Hunt, M.E.; Sugrue, S.P. Nuclear Speckle-Associated Protein Pnn/DRS Binds to the Transcriptional Corepressor CtBP and Relieves CtBP- Mediated Repression of the E-Cadherin Gene. Mol. Cell. Biol. 2004, 24, 10223–10235. [Google Scholar] [CrossRef] [PubMed]
  15. Liang, J.; Liu, C.; Xu, D.; Xie, K.; Li, A. LncRNA NEAT1 facilitates glioma progression via stabilizing PGK1. J. Transl. Med. 2022, 20, 1–13. [Google Scholar] [CrossRef]
  16. Fu, Q.; Yu, Z. Phosphoglycerate kinase 1 (PGK1) in cancer: A promising target for diagnosis and therapy. Life Sci. 2020, 256, 117863. [Google Scholar] [CrossRef] [PubMed]
  17. Lu, Z.; Hunter, T. Metabolic Kinases Moonlighting as Protein Kinases. Trends Biochem. Sci. 2018, 43, 301–310. [Google Scholar] [CrossRef] [PubMed]
  18. Li, X.; Jiang, Y.; Meisenhelder, J.; Yang, W.; Hawke, D.H.; Zheng, Y.; Xia, Y.; Aldape, K.; He, J.; Hunter, T.; et al. Mitochondria-Translocated PGK1 Functions as a Protein Kinase to Coordinate Glycolysis and the TCA Cycle in Tumorigenesis. Mol. Cell 2016, 61, 705–719. [Google Scholar] [CrossRef]
  19. Liu, H.; Wang, X.; Shen, P.; Ni, Y.; Han, X. The basic functions of phosphoglycerate kinase 1 and its roles in cancer and other diseases. Eur. J. Pharmacol. 2022, 920, 174835. [Google Scholar] [CrossRef] [PubMed]
  20. Gou, R.; Hu, Y.; Liu, O.; Dong, H.; Gao, L.; Wang, S.; Zheng, M.; Li, X.; Lin, B. PGK1 Is a Key Target for Anti-Glycolytic Therapy of Ovarian Cancer: Based on the Comprehensive Analysis of Glycolysis-Related Genes. Front. Oncol. 2021, 11. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, W.-L.; Jiang, Z.-R.; Hu, C.; Chen, C.; Hu, Z.-Q.; Wang, A.-L.; Wang, L.; Liu, J.; Liu, Q.-S. Pharmacologically inhibiting phosphoglycerate kinase 1 for glioma with NG52. Acta Pharmacol. Sin. 2020, 42, 633–640. [Google Scholar] [CrossRef] [PubMed]
  22. Kondoh, H.; Lleonart, M.E.; Nakashima, Y.; Yokode, M.; Tanaka, M.; Bernard, D.; Gil, J.; Beach, D. A High Glycolytic Flux Supports the Proliferative Potential of Murine Embryonic Stem Cells. Antioxidants Redox Signal. 2007, 9, 293–299. [Google Scholar] [CrossRef] [PubMed]
  23. Kondoh, H.; Lleonart, M.E.; Gil, J.; Wang, J.; Degan, P.; Peters, G.; Martinez, D.; Carnero, A.; Beach, D. Glycolytic Enzymes Can Modulate Cellular Life Span. Cancer Res. 2005, 65, 177–185. [Google Scholar] [CrossRef] [PubMed]
  24. Hitosugi, T.; Zhou, L.; Elf, S.; Fan, J.; Kang, H.-B.; Seo, J.H.; Shan, C.; Dai, Q.; Zhang, L.; Xie, J.; et al. Phosphoglycerate Mutase 1 Coordinates Glycolysis and Biosynthesis to Promote Tumor Growth. Cancer Cell 2012, 22, 585–600. [Google Scholar] [CrossRef]
  25. Peng, X.; Gong, F.; Chen, Y.; Qiu, M.; Cheng, K.; Tang, J.; Ge, J.; Chen, N.; Zeng, H.; Liu, J. Proteomics identification of PGAM1 as a potential therapeutic target for urothelial bladder cancer. J. Proteom. 2015, 132, 85–92. [Google Scholar] [CrossRef] [PubMed]
  26. Mikawa, T.; Shibata, E.; Shimada, M.; Ito, K.; Ito, T.; Kanda, H.; Takubo, K.; Lleonart, M.E.; Inagaki, N.; Yokode, M.; et al. Phosphoglycerate Mutase Cooperates with Chk1 Kinase to Regulate Glycolysis. iScience 2020, 23, 101306. [Google Scholar] [CrossRef] [PubMed]
  27. Cancemi, P.; Buttacavoli, M.; Roz, E.; Feo, S. Expression of Alpha-Enolase (ENO1), Myc Promoter-Binding Protein-1 (MBP-1) and Matrix Metalloproteinases (MMP-2 and MMP-9) Reflect the Nature and Aggressiveness of Breast Tumors. Int. J. Mol. Sci. 2019, 20, 3952. [Google Scholar] [CrossRef]
  28. Perconti, G.; Pratesi, F.; Angelotti, F.; Manca, L.; Puxeddu, I.; Rubino, P.; Maranto, C.; Giallongo, A.; Migliorini, P. Fingerprinting of anti-alpha enolase antibodies in systemic sclerosis. Clin. Exp. Rheumatol. 2020, 38 (Suppl. 125), 115–119. [Google Scholar]
  29. Li, M.; Li, J.; Wang, J.; Li, Y.; Yang, P. Serum level of anti-α-enolase antibody in untreated systemic lupus erythematosus patients correlates with 24-hour urine protein and D-dimer. Lupus 2018, 27, 139–142. [Google Scholar] [CrossRef] [PubMed]
  30. Cho, H.; Um, J.; Lee, J.-H.; Kim, W.-H.; Kang, W.S.; Kim, S.H.; Ha, H.-H.; Kim, Y.-C.; Ahn, Y.-K.; Jung, D.-W.; et al. ENOblock, a unique small molecule inhibitor of the non-glycolytic functions of enolase, alleviates the symptoms of type 2 diabetes. Sci. Rep. 2017, 7, srep44186. [Google Scholar] [CrossRef]
  31. Butterfield, D.A.; Lange, M.L.B. Multifunctional roles of enolase in Alzheimer’s disease brain: Beyond altered glucose metabolism. J. Neurochem. 2009, 111, 915–933. [Google Scholar] [CrossRef] [Green Version]
  32. Altenberg, B.; Greulich, K. Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics 2004, 84, 1014–1020. [Google Scholar] [CrossRef] [PubMed]
  33. Hou, J.-Y.; Cao, J.; Gao, L.-J.; Zhang, F.-P.; Shen, J.; Zhou, L.; Shi, J.-Y.; Feng, Y.-L.; Yan, Z.; Wang, D.-P.; et al. Upregulation of α enolase (ENO1) crotonylation in colorectal cancer and its promoting effect on cancer cell metastasis. Biochem. Biophys. Res. Commun. 2021, 578, 77–83. [Google Scholar] [CrossRef]
  34. Hippner, M.; Majkowski, M.; Biecek, P.; Szkudlarek, T.; Simiczyjew, A.; Pieniazek, M.; Nowak, D.; Miazek, A.; Donizy, P. Alpha-Enolase (ENO1) Correlates with Invasiveness of Cutaneous Melanoma—An In Vitro and a Clinical Study. Diagnostics 2022, 12, 254. [Google Scholar] [CrossRef]
  35. Capello, M.; Ferri-Borgogno, S.; Riganti, C.; Chattaragada, M.S.; Principe, M.; Roux, C.; Zhou, W.; Petricoin, E.F.; Cappello, P.; Novelli, F. Targeting the Warburg effect in cancer cells through ENO1 knockdown rescues oxidative phosphorylation and induces growth arrest. Oncotarget 2015, 7, 5598–5612. [Google Scholar] [CrossRef] [PubMed]
  36. Chang, Y.S.; Wu, W.; Walsh, G.; Hong, W.K.; Mao, L. Enolase-alpha is frequently down-regulated in non-small cell lung cancer and predicts aggressive biological behavior. Clin. Cancer Res. 2003, 9, 3641–3644. [Google Scholar]
  37. Fu, Q.-F.; Liu, Y.; Fan, Y.; Hua, S.-N.; Qu, H.-Y.; Dong, S.-W.; Li, R.-L.; Zhao, M.-Y.; Zhen, Y.; Yu, X.-L.; et al. Alpha-enolase promotes cell glycolysis, growth, migration, and invasion in non-small cell lung cancer through FAK-mediated PI3K/AKT pathway. J. Hematol. Oncol. 2015, 8, 22. [Google Scholar] [CrossRef] [PubMed]
  38. Hirose, T.; Yamazaki, T.; Nakagawa, S. Molecular anatomy of the architectural NEAT1 noncoding RNA: The domains, interactors, and biogenesis pathway required to build phase-separated nuclear paraspeckles. Wiley Interdiscip. Rev. RNA 2019, 10, e1545. [Google Scholar] [CrossRef] [PubMed]
  39. Fox, A.H.; Lamond, A. Paraspeckles. Cold Spring Harb. Perspect. Biol. 2010, 2, a000687. [Google Scholar] [CrossRef] [PubMed]
  40. Fox, A.H.; Lam, Y.W.; Leung, A.K.; Lyon, C.E.; Andersen, J.; Mann, M.; Lamond, A.I. Paraspeckles: A Novel Nuclear Domain. Curr. Biol. 2002, 12, 13–25. [Google Scholar] [CrossRef]
  41. Fox, A.H.; Nakagawa, S.; Hirose, T.; Bond, C.S. Paraspeckles: Where Long Noncoding RNA Meets Phase Separation. Trends Biochem. Sci. 2018, 43, 124–135. [Google Scholar] [CrossRef] [PubMed]
  42. Souquere, S.; Beauclair, G.; Harper, F.; Fox, A.; Pierron, G. Highly Ordered Spatial Organization of the Structural Long Noncoding NEAT1 RNAs within Paraspeckle Nuclear Bodies. Mol. Biol. Cell 2010, 21, 4020–4027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Grosch, M.; Ittermann, S.; Rusha, E.; Greisle, T.; Ori, C.; Truong, D.-J.J.; O’Neill, A.C.; Pertek, A.; Westmeyer, G.G.; Drukker, M. Nucleus size and DNA accessibility are linked to the regulation of paraspeckle formation in cellular differentiation. BMC Biol. 2020, 18, 42. [Google Scholar] [CrossRef]
  44. Naganuma, T.; Nakagawa, S.; Tanigawa, A.; Sasaki, Y.; Goshima, N.; Hirose, T. Alternative 3′-end processing of long noncoding RNA initiates construction of nuclear paraspeckles. EMBO J. 2012, 31, 4020–4034. [Google Scholar] [CrossRef]
  45. Kim, S.; Yamamoto, J.; Chen, Y.; Aida, M.; Wada, T.; Handa, H.; Yamaguchi, Y. Evidence that cleavage factor Im is a heterotetrameric protein complex controlling alternative polyadenylation. Genes Cells 2010, 15, 1003–1013. [Google Scholar] [CrossRef] [PubMed]
  46. Naganuma, T.; Hirose, T. Paraspeckle formation during the biogenesis of long non-coding RNAs. RNA Biol. 2013, 10, 456–461. [Google Scholar] [CrossRef] [PubMed]
  47. Machitani, M.; Taniguchi, I.; Ohno, M. ARS2 Regulates Nuclear Paraspeckle Formation through 3′-End Processing and Stability of NEAT1 Long Noncoding RNA. Mol. Cell. Biol. 2020, 40, e00269-19. [Google Scholar] [CrossRef]
  48. Modic, M.; Grosch, M.; Rot, G.; Schirge, S.; Lepko, T.; Yamazaki, T.; Lee, F.C.; Rusha, E.; Shaposhnikov, D.; Palo, M.; et al. Cross-Regulation between TDP-43 and Paraspeckles Promotes Pluripotency-Differentiation Transition. Mol. Cell 2019, 74, 951–965.e13. [Google Scholar] [CrossRef]
  49. Tollervey, J.R.; Curk, T.; Rogelj, B.; Briese, M.; Cereda, M.; Kayikci, M.; König, J.; Hortobágyi, T.; Nishimura, A.L.; Župunski, V.; et al. Characterizing the RNA targets and position-dependent splicing regulation by TDP-43. Nat. Neurosci. 2011, 14, 452–458. [Google Scholar] [CrossRef]
  50. Shelkovnikova, T.A.; Kukharsky, M.S.; An, H.; DiMasi, P.; Alexeeva, S.; Shabir, O.; Heath, P.R.; Buchman, V.L. Protective paraspeckle hyper-assembly downstream of TDP-43 loss of function in amyotrophic lateral sclerosis. Mol. Neurodegener. 2018, 13, 30. [Google Scholar] [CrossRef] [PubMed]
  51. Nishimoto, Y.; Nakagawa, S.; Okano, H. NEAT1 lncRNA and amyotrophic lateral sclerosis. Neurochem. Int. 2021, 150, 105175. [Google Scholar] [CrossRef] [PubMed]
  52. Guerrero, E.N.; Wang, H.; Mitra, J.; Hegde, P.M.; Stowell, S.E.; Liachko, N.; Kraemer, B.C.; Garruto, R.M.; Rao, K.; Hegde, M.L. TDP-43/FUS in motor neuron disease: Complexity and challenges. Prog. Neurobiol. 2016, 145–146, 78–97. [Google Scholar] [CrossRef] [Green Version]
  53. Suzuki, H.; Shibagaki, Y.; Hattori, S.; Matsuoka, M. C9-ALS/FTD-linked proline–arginine dipeptide repeat protein associates with paraspeckle components and increases paraspeckle formation. Cell Death Dis. 2019, 10, 1–16. [Google Scholar] [CrossRef]
  54. Ma, X.; Ying, Y.; Xie, H.; Liu, X.; Wang, X.; Li, J. The Regulatory Role of RNA Metabolism Regulator TDP-43 in Human Cancer. Front. Oncol. 2021, 11, 1–10. [Google Scholar] [CrossRef] [PubMed]
  55. Chen, X.; Fan, Z.; McGee, W.; Chen, M.; Kong, R.; Wen, P.; Xiao, T.; Chen, X.; Liu, J.; Zhu, L.; et al. TDP-43 regulates cancer-associated microRNAs. Protein Cell 2018, 9, 848–866. [Google Scholar] [CrossRef]
  56. Spinelli, J.B.; Haigis, M.C. The multifaceted contributions of mitochondria to cellular metabolism. Nat. Cell Biol. 2018, 20, 745–754. [Google Scholar] [CrossRef] [PubMed]
  57. Quiros, P.M.; Mottis, A.; Auwerx, J. Mitonuclear communication in homeostasis and stress. Nat. Rev. Mol. Cell Biol. 2016, 17, 213–226. [Google Scholar] [CrossRef]
  58. Rizzuto, R.; De Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as sensors and regulators of calcium signalling. Nat. Rev. Mol. Cell Biol. 2012, 13, 566–578. [Google Scholar] [CrossRef] [PubMed]
  59. Ouyang, J.; Zhong, Y.; Zhang, Y.; Yang, L.; Wu, P.; Hou, X.; Xiong, F.; Li, X.; Zhang, S.; Gong, Z.; et al. Long non-coding RNAs are involved in alternative splicing and promote cancer progression. Br. J. Cancer 2022, 126, 1113–1124. [Google Scholar] [CrossRef] [PubMed]
  60. Wilson, C.; Kanhere, A. The Missing Link Between Cancer-Associated Variants and LncRNAs. Trends Genet. 2021, 37, 410–413. [Google Scholar] [CrossRef] [PubMed]
  61. Gutschner, T.; Baas, M.; Diederichs, S. Noncoding RNA gene silencing through genomic integration of RNA destabilizing elements using zinc finger nucleases. Genome Res. 2011, 21, 1944–1954. [Google Scholar] [CrossRef] [PubMed]
  62. Meseure, D.; Vacher, S.; Lallemand, F.; Alsibai, K.D.; Hatem, R.; Chemlali, W.; Nicolas, A.; De Koning, L.; Pasmant, E.; Callens, C.; et al. Prognostic value of a newly identified MALAT1 alternatively spliced transcript in breast cancer. Br. J. Cancer 2016, 114, 1395–1404. [Google Scholar] [CrossRef] [Green Version]
  63. Wilusz, J.E.; Freier, S.M.; Spector, D.L. 3′ End Processing of a Long Nuclear-Retained Noncoding RNA Yields a tRNA-like Cytoplasmic RNA. Cell 2008, 135, 919–932. [Google Scholar] [CrossRef] [PubMed]
  64. Lu, X.; Huang, J.; Wu, S.; Zheng, Q.; Liu, P.; Feng, H.; Su, X.; Fu, H.; Xi, Q.; Wang, G. The tRNA -like small noncoding RNA masc RNA promotes global protein translation. EMBO Rep. 2020, 21, e49684. [Google Scholar] [CrossRef]
  65. Xie, S.-J.; Diao, L.-T.; Cai, N.; Zhang, L.-T.; Xiang, S.; Jia, C.-C.; Qiu, D.-B.; Liu, C.; Sun, Y.-J.; Lei, H.; et al. mascRNA and its parent lncRNA MALAT1 promote proliferation and metastasis of hepatocellular carcinoma cells by activating ERK/MAPK signaling pathway. Cell Death Discov. 2021, 7, 1–14. [Google Scholar] [CrossRef]
  66. Hoffmann, M.J.; Dehn, J.; Droop, J.; Niegisch, G.; Niedworok, C.; Szarvas, T.; Schulz, W.A. Truncated Isoforms of lncRNA ANRIL Are Overexpressed in Bladder Cancer, But Do Not Contribute to Repression of INK4 Tumor Suppressors. Non-Coding RNA 2015, 1, 266–284. [Google Scholar] [CrossRef] [PubMed]
  67. Fang, J.; Qiao, F.; Tu, J.; Xu, J.; Ding, F.; Liu, Y.; Akuo, B.A.; Hu, J.; Shao, S. High expression of long non-coding RNA NEAT1 indicates poor prognosis of human cancer. Oncotarget 2017, 8, 45918–45927. [Google Scholar] [CrossRef]
  68. Yang, C.; Li, Z.; Li, Y.; Xu, R.; Wang, Y.; Tian, Y.; Chen, W. Long non-coding RNA NEAT1 overexpression is associated with poor prognosis in cancer patients: A systematic review and meta-analysis. Oncotarget 2017, 8, 2672–2680. [Google Scholar] [CrossRef] [PubMed]
  69. Chakravarty, D.; Sboner, A.; Nair, S.S.; Giannopoulou, E.G.; Li, R.; Hennig, S.; Mosquera, J.M.; Pauwels, J.; Park, K.; Kossai, M.; et al. The oestrogen receptor alpha-regulated lncRNA NEAT1 is a critical modulator of prostate cancer. Nat. Commun. 2014, 5, 5383. [Google Scholar] [CrossRef]
  70. Wu, Y.; Yang, L.; Zhao, J.; Li, C.; Nie, J.; Liu, F.; Zhuo, C.; Zheng, Y.; Li, B.; Wang, Z.; et al. Nuclear-enriched abundant transcript 1 as a diagnostic and prognostic biomarker in colorectal cancer. Mol. Cancer 2015, 14, 191. [Google Scholar] [CrossRef] [PubMed]
  71. Xianguo, C.; Zongyao, H.; Jun, Z.; Song, F.; Guangyue, L.; Ligang, Z.; Kaiping, Z.; Yangyang, Z.; ChaoZhao, L. Promoting progression and clinicopathological significance of NEAT1 over-expression in bladder cancer. Oncotarget 2016, 5. [Google Scholar] [CrossRef]
  72. Lu, Y.; Li, T.; Wei, G.; Liu, L.; Chen, Q.; Xu, L.; Zhang, K.; Zeng, D.; Liao, R. The long non-coding RNA NEAT1 regulates epithelial to mesenchymal transition and radioresistance in through miR-204/ZEB1 axis in nasopharyngeal carcinoma. Tumor Biol. 2016, 37, 11733–11741. [Google Scholar] [CrossRef] [PubMed]
  73. Li, Y.; Li, Y.; Chen, W.; He, F.; Tan, Z.; Zheng, J.; Wang, W.; Zhao, Q.; Li, J. NEAT expression is associated with tumor recurrence and unfavorable prognosis in colorectal cancer. Oncotarget 2015, 6, 27641–27650. [Google Scholar] [CrossRef]
  74. Vancura, A.; Lanzós, A.; Bosch-Guiteras, N.; Esteban, M.T.; Gutierrez, A.H.; Haefliger, S.; Johnson, R. Cancer LncRNA Census 2 (CLC2): An enhanced resource reveals clinical features of cancer lncRNAs. NAR Cancer 2021, 3, zcab013. [Google Scholar] [CrossRef]
  75. Zeng, C.; Xu, Y.; Xu, L.; Yu, X.; Cheng, J.; Yang, L.; Chen, S.; Li, Y. Inhibition of long non-coding RNA NEAT1 impairs myeloid differentiation in acute promyelocytic leukemia cells. BMC Cancer 2014, 14, 693. [Google Scholar] [CrossRef] [PubMed]
  76. Adriaens, C.; Standaert, L.; Barra, J.; Latil, M.; Verfaillie, A.; Kalev, P.; Boeckx, B.; Wijnhoven, P.W.G.; Radaelli, E.; Vermi, W.; et al. p53 induces formation of NEAT1 lncRNA-containing paraspeckles that modulate replication stress response and chemosensitivity. Nat. Med. 2016, 22, 861–868. [Google Scholar] [CrossRef] [PubMed]
  77. Ronchetti, D.; Favasuli, V.; Monti, P.; Cutrona, G.; Fabris, S.; Silvestris, I.; Agnelli, L.; Colombo, M.; Menichini, P.; Matis, S.; et al. NEAT1 Long Isoform Is Highly Expressed in Chronic Lymphocytic Leukemia Irrespectively of Cytogenetic Groups or Clinical Outcome. Non-Coding RNA 2020, 6, 11. [Google Scholar] [CrossRef]
  78. Zeng, C.; Liu, S.; Lu, S.; Yu, X.; Lai, J.; Wu, Y.; Chen, S.; Wang, L.; Yu, Z.; Luo, G.; et al. The c-Myc-regulated lncRNA NEAT1 and paraspeckles modulate imatinib-induced apoptosis in CML cells. Mol. Cancer 2018, 17, 130. [Google Scholar] [CrossRef]
  79. Duan, M.-Y.; Li, M.; Tian, H.; Tang, G.; Yang, Y.-C.; Peng, N.-C. Down-regulation of lncRNA NEAT1 regulated by miR-194-5p/DNMT3A facilitates acute myeloid leukemia. Blood Cells Mol. Dis. 2020, 82, 102417. [Google Scholar] [CrossRef] [PubMed]
  80. Feng, S.; Liu, N.; Chen, X.; Liu, Y.; An, J. Long non-coding RNA NEAT1/miR-338-3p axis impedes the progression of acute myeloid leukemia via regulating CREBRF. Cancer Cell Int. 2020, 20, 112. [Google Scholar] [CrossRef]
  81. Taiana, E.; Ronchetti, D.; Favasuli, V.; Todoerti, K.; Manzoni, M.; Amodio, N.; Tassone, P.; Agnelli, L.; Neri, A. Long non-coding RNA NEAT1 shows high expression unrelated to molecular features and clinical outcome in multiple myeloma. Haematologica 2019, 104, e72–e76. [Google Scholar] [CrossRef]
  82. Blume, C.J.; Hotz-Wagenblatt, A.; Hüllein, J.; Sellner, L.; Jethwa, A.; Stolz, T.; Slabicki, M.; Lee, K.; Sharathchandra, A.; Benner, A.; et al. p53-dependent non-coding RNA networks in chronic lymphocytic leukemia. Leukemia 2015, 29, 2015–2023. [Google Scholar] [CrossRef]
  83. Zhao, L.; Zhou, N.; Zhao, P. Expression level of NEAT1 differentiates benign and malignant thyroid nodules by regulating NEAT1/miR-9/PTEN and NEAT1/miR-124/PDCD6 signalling. Int. J. Mol. Med. 2020, 46, 1661–1670. [Google Scholar] [CrossRef]
  84. Li, J.-H.; Zhang, S.-Q.; Qiu, X.-G.; Zhang, S.-J.; Zheng, S.-H.; Zhang, D.-H. Long non-coding RNA NEAT1 promotes malignant progression of thyroid carcinoma by regulating miRNA-214. Int. J. Oncol. 2016, 50, 708–716. [Google Scholar] [CrossRef] [PubMed]
  85. Tan, X.; Wang, P.; Lou, J.; Zhao, J. Knockdown of lncRNA NEAT1 suppresses hypoxia-induced migration, invasion and glycolysis in anaplastic thyroid carcinoma cells through regulation of miR-206 and miR-599. Cancer Cell Int. 2020, 20, 132. [Google Scholar] [CrossRef]
  86. Zhang, H.; Cai, Y.; Zheng, L.; Zhang, Z.; Lin, X.; Jiang, N. Long noncoding RNA NEAT1 regulate papillary thyroid cancer progression by modulating miR-129-5p/ KLK7 expression. J. Cell. Physiol. 2018, 233, 6638–6648. [Google Scholar] [CrossRef] [PubMed]
  87. Naveed, A.; Cooper, J.A.; Li, R.; Hubbard, A.; Chen, J.; Liu, T.; Wilton, S.D.; Fletcher, S.; Fox, A.H. NEAT1 polyA-modulating antisense oligonucleotides reveal opposing functions for both long non-coding RNA isoforms in neuroblastoma. Cell. Mol. Life Sci. 2021, 78, 2213–2230. [Google Scholar] [CrossRef]
  88. Jiang, X.; Zhou, Y.; Sun, A.-J.; Xue, J.-L. NEAT1 contributes to breast cancer progression through modulating miR-448 and ZEB1. J. Cell. Physiol. 2018, 233, 8558–8566. [Google Scholar] [CrossRef] [PubMed]
  89. Knutsen, E.; Oslo Breast Cancer Research Consortium (OSBREAC); Lellahi, S.M.; Aure, M.R.; Nord, S.; Fismen, S.; Larsen, K.B.; Gabriel, M.T.; Hedberg, A.; Bjørklund, S.S.; et al. The expression of the long NEAT1_2 isoform is associated with human epidermal growth factor receptor 2-positive breast cancers. Sci. Rep. 2020, 10, 1–14. [Google Scholar] [CrossRef]
  90. Choudhry, H.; Albukhari, A.; Morotti, M.; Haider, S.; Moralli, D.; Smythies, J.; Schödel, J.; Green, C.; Camps, C.; Buffa, F.; et al. Tumor hypoxia induces nuclear paraspeckle formation through HIF-2α dependent transcriptional activation of NEAT1 leading to cancer cell survival. Oncogene 2015, 34, 4482–4490. [Google Scholar] [CrossRef]
  91. Bhatt, U.; Kretzmann, A.L.; Guédin, A.; Ou, A.; Kobelke, S.; Bond, C.S.; Evans, C.W.; Hurley, L.H.; Mergny, J.-L.; Iyer, K.S.; et al. The role of G-Quadruplex DNA in Paraspeckle formation in cancer. Biochimie 2021, 190, 124–131. [Google Scholar] [CrossRef] [PubMed]
  92. Li, R.; Harvey, A.R.; Hodgetts, S.I.; Fox, A.H. Functional dissection of NEAT1 using genome editing reveals substantial localization of the NEAT1_1 isoform outside paraspeckles. RNA 2017, 23, 872–881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Tan, H.; Zhao, L. lncRNA nuclear-enriched abundant transcript 1 promotes cell proliferation and invasion by targeting miR-186-5p/HIF-1α in osteosarcoma. J. Cell. Biochem. 2019, 120, 6502–6514. [Google Scholar] [CrossRef]
  94. Xu, H.; Sun, X.; Huang, Y.; Si, Q.; Li, M. Long non-coding RNA NEAT1 modifies cell proliferation, colony formation, apoptosis, migration and invasion via the miR-4500/BZW1 axis in ovarian cancer. Mol. Med. Rep. 2020, 22, 3347–3357. [Google Scholar] [CrossRef]
  95. An, J.; Lv, W.; Zhang, Y. LncRNA NEAT1 contributes to paclitaxel resistance of ovarian cancer cells by regulating ZEB1 expression via miR-194. OncoTargets Ther. 2017, 10, 5377–5390. [Google Scholar] [CrossRef] [PubMed]
  96. Li, W.; Xu, W.; Sun, K.; Wang, F.; Wong, T.W.; Kong, A. Identification of novel biomarkers in prostate cancer diagnosis and prognosis. J. Biochem. Mol. Toxicol. 2022. epub ahead of print. [Google Scholar] [CrossRef] [PubMed]
  97. Wen, S.; Wei, Y.; Zen, C.; Xiong, W.; Niu, Y.; Zhao, Y. Long non-coding RNA NEAT1 promotes bone metastasis of prostate cancer through N6-methyladenosine. Mol. Cancer 2020, 19, 171. [Google Scholar] [CrossRef] [PubMed]
  98. Shaker, O.G.; Abdelwahed, M.Y.; Ahmed, N.A.; Hassan, E.A.; Ahmed, T.I.; Abousarie, M.A.; Ayoub, S.E. Evaluation of serum long noncoding RNA NEAT and MiR-129-5p in hepatocellular carcinoma. IUBMB Life 2019, 71, 1571–1578. [Google Scholar] [CrossRef]
  99. Wang, Z.; Zou, Q.; Song, M.; Chen, J. NEAT1 promotes cell proliferation and invasion in hepatocellular carcinoma by negative regulating miR-613 expression. Biomed. Pharmacother. 2017, 94, 612–618. [Google Scholar] [CrossRef]
  100. Zheng, X.; Zhang, Y.; Liu, Y.; Fang, L.; Li, L.; Sun, J.; Pan, Z.; Xin, W.; Huang, P. HIF-2α activated lncRNA NEAT1 promotes hepatocellular carcinoma cell invasion and metastasis by affecting the epithelial-mesenchymal transition. J. Cell. Biochem. 2018, 119, 3247–3256. [Google Scholar] [CrossRef]
  101. Fu, J.-W.; Kong, Y.; Sun, X. Long noncoding RNA NEAT1 is an unfavorable prognostic factor and regulates migration and invasion in gastric cancer. J. Cancer Res. Clin. Oncol. 2016, 142, 1571–1579. [Google Scholar] [CrossRef] [PubMed]
  102. Jen, J.; Tang, Y.-A.; Lu, Y.-H.; Lin, C.-C.; Lai, W.-W.; Wang, Y.-C. Oct4 transcriptionally regulates the expression of long non-coding RNAs NEAT1 and MALAT1 to promote lung cancer progression. Mol. Cancer 2017, 16, 104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Liu, J.; Li, K.; Wang, R.; Chen, S.; Wu, J.; Li, X.; Ning, Q.; Yang, G.; Pang, Y. The interplay between ATF2 and NEAT1 contributes to lung adenocarcinoma progression. Cancer Cell Int. 2020, 20, 1–13. [Google Scholar] [CrossRef]
  104. Wu, F.; Mo, Q.; Wan, X.; Dan, J.; Hu, H. NEAT1/hsa-mir-98-5p/MAPK6 axis is involved in non–small-cell lung cancer development. J. Cell. Biochem. 2019, 120, 2836–2846. [Google Scholar] [CrossRef]
  105. Zhang, M.; Weng, W.; Zhang, Q.; Wu, Y.; Ni, S.; Tan, C.; Xu, M.; Sun, H.; Liu, C.; Wei, P.; et al. The lncRNA NEAT1 activates Wnt/β-catenin signaling and promotes colorectal cancer progression via interacting with DDX5. J. Hematol. Oncol. 2018, 11, 1–13. [Google Scholar] [CrossRef] [PubMed]
  106. Zhu, Z.; Du, S.; Yin, K.; Ai, S.; Yu, M.; Liu, Y.; Shen, Y.; Liu, M.; Jiao, R.; Chen, X.; et al. Knockdown long noncoding RNA nuclear paraspeckle assembly transcript 1 suppresses colorectal cancer through modulating miR-193a-3p/KRAS. Cancer Med. 2019, 8, 261–275. [Google Scholar] [CrossRef] [PubMed]
  107. Luo, Y.; Chen, J.-J.; Lv, Q.; Qin, J.; Huang, Y.-Z.; Yu, M.-H.; Zhong, M. Long non-coding RNA NEAT1 promotes colorectal cancer progression by competitively binding miR-34a with SIRT1 and enhancing the Wnt/β-catenin signaling pathway. Cancer Lett. 2019, 440, 11–22. [Google Scholar] [CrossRef]
  108. Liu, Z.; Wu, K.; Wu, J.; Tian, D.; Chen, Y.; Yang, Z.; Wu, A. NEAT1 is a potential prognostic biomarker for patients with nasopharyngeal carcinoma. J. Cell. Biochem. 2019, 120, 9831–9838. [Google Scholar] [CrossRef]
  109. Gao, C.; Zhang, Y.; Sun, H. Mechanism of miR-340–5p in laryngeal cancer cell proliferation and invasion through the lncRNA NEAT1/MMP11 axis. Pathol. Res. Pract. 2022, 236, 153912. [Google Scholar] [CrossRef]
  110. 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]
  111. Li, Y.; Chen, D.; Gao, X.; Li, X.; Shi, G. LncRNA NEAT1 Regulates Cell Viability and Invasion in Esophageal Squamous Cell Carcinoma through the miR-129/CTBP2 Axis. Dis. Markers 2017, 2017, 1–11. [Google Scholar] [CrossRef] [PubMed]
  112. Mello, S.S.; Sinow, C.; Raj, N.; Mazur, P.K.; Bieging-Rolett, K.; Broz, D.K.; Imam, J.F.C.; Vogel, H.; Wood, L.D.; Sage, J.; et al. Neat1 is a p53-inducible lincRNA essential for transformation suppression. Genes Dev. 2017, 31, 1095–1108. [Google Scholar] [CrossRef] [Green Version]
  113. Olivier, M.; Hollstein, M.; Hainaut, P. TP53 Mutations in Human Cancers: Origins, Consequences, and Clinical Use. Cold Spring Harb. Perspect. Biol. 2010, 2, a001008. [Google Scholar] [CrossRef] [PubMed]
  114. Botcheva, K.; McCorkle, S.; McCombie, W.R.; Dunn, J.J.; Anderson, C.W. Distinct p53 genomic binding patterns in normal and cancer-derived human cells. Cell Cycle 2011, 10, 4237–4249. [Google Scholar] [CrossRef] [PubMed]
  115. Nik-Zainal, S.; Davies, H.; Staaf, J.; Ramakrishna, M.; Glodzik, D.; Zou, X.; Martincorena, I.; Alexandrov, L.B.; Martin, S.; Wedge, D.C.; et al. Landscape of somatic mutations in 560 breast cancer whole-genome sequences. Nature 2016, 534, 47–54. [Google Scholar] [CrossRef] [PubMed]
  116. Rheinbay, E.; Parasuraman, P.; Grimsby, J.; Tiao, G.; Engreitz, J.M.; Kim, J.; Lawrence, M.S.; Taylor-Weiner, A.; Rodriguez-Cuevas, S.; Rosenberg, M.; et al. Recurrent and functional regulatory mutations in breast cancer. Nature 2017, 547, 55–60. [Google Scholar] [CrossRef]
  117. Wedge, D.C.; CAMCAP Study Group; Gundem, G.; Mitchell, T.; Woodcock, D.J.; Martincorena, I.; Ghori, M.; Zamora, J.; Butler, A.; Whitaker, H.; et al. Sequencing of prostate cancers identifies new cancer genes, routes of progression and drug targets. Nat. Genet. 2018, 50, 682–692. [Google Scholar] [CrossRef] [PubMed]
  118. Rheinbay, E.; Nielsen, M.M.; Abascal, F.; Wala, J.A.; Shapira, O.; Tiao, G.; Hornshøj, H.; Hess, J.M.; Juul, R.I.; Lin, Z.; et al. Analyses of non-coding somatic drivers in 2,658 cancer whole genomes. Nature 2020, 578, 102–111. [Google Scholar] [CrossRef] [PubMed]
  119. Kolenda, T.; Ryś, M.; Guglas, K.; Teresiak, A.; Bliźniak, R.; Mackiewicz, J.; Lamperska, K. Quantification of long non-coding RNAs using qRT-PCR: Comparison of different cDNA synthesis methods and RNA stability. Arch. Med Sci. 2021, 17, 1006–1015. [Google Scholar] [CrossRef] [PubMed]
  120. Chujo, T.; Yamazaki, T.; Kawaguchi, T.; Kurosaka, S.; Takumi, T.; Nakagawa, S.; Hirose, T. Unusual semi-extractability as a hallmark of nuclear body-associated architectural noncoding RNAs. EMBO J. 2017, 36, 1447–1462. [Google Scholar] [CrossRef]
  121. Gleeson, J.; Leger, A.; Prawer, Y.D.J.; A Lane, T.; Harrison, P.J.; Haerty, W.; Clark, M.B. Accurate expression quantification from nanopore direct RNA sequencing with NanoCount. Nucleic Acids Res. 2022, 50, e19. [Google Scholar] [CrossRef] [PubMed]
  122. Yang, L.-Z.; Wang, Y.; Li, S.-Q.; Yao, R.-W.; Luan, P.-F.; Wu, H.; Carmichael, G.G.; Chen, L.-L. Dynamic Imaging of RNA in Living Cells by CRISPR-Cas13 Systems. Mol. Cell 2019, 76, 981–997.e7. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The NEAT1 gene gives rise to two isoforms with identical 5′ sequences. The paraspeckle-independent NEAT1_1 undergoes canonical 3′ polyadenylation whilst the blocking of 3′ polyadenylation via competitive binding of hnRNPK to the CFIm complex yields the paraspeckle-essential NEAT1_2. The recruitment of paraspeckle proteins to the hydrophilic and hydrophobic regions of NEAT1_2 leads to phase-separated paraspeckles.
Figure 1. The NEAT1 gene gives rise to two isoforms with identical 5′ sequences. The paraspeckle-independent NEAT1_1 undergoes canonical 3′ polyadenylation whilst the blocking of 3′ polyadenylation via competitive binding of hnRNPK to the CFIm complex yields the paraspeckle-essential NEAT1_2. The recruitment of paraspeckle proteins to the hydrophilic and hydrophobic regions of NEAT1_2 leads to phase-separated paraspeckles.
Cancers 14 04388 g001
Figure 2. Nucleocytoplasmic transport of NEAT1_1 enhances the Warburg effect via the binding to glycolytic enzymes. TDP-43, ARS2, and CFIm promote the canonical 3′ polyadenylation of NEAT1_1, which can then bind to pinin for nuclear export to the cytoplasm. Once in the cytoplasm, NEAT1_1 can bind with the glycolytic enzymes PGK1 (A), PGAM1 (B), and ENO1 (C) to enhance glycolytic flux, and hence the Warberg effect, in transformed cells.
Figure 2. Nucleocytoplasmic transport of NEAT1_1 enhances the Warburg effect via the binding to glycolytic enzymes. TDP-43, ARS2, and CFIm promote the canonical 3′ polyadenylation of NEAT1_1, which can then bind to pinin for nuclear export to the cytoplasm. Once in the cytoplasm, NEAT1_1 can bind with the glycolytic enzymes PGK1 (A), PGAM1 (B), and ENO1 (C) to enhance glycolytic flux, and hence the Warberg effect, in transformed cells.
Cancers 14 04388 g002
Table 1. Summary of NEAT1 expression studies, as determined by RT-PCR, in various human cancer subtypes, published between 2014 and 2022.
Table 1. Summary of NEAT1 expression studies, as determined by RT-PCR, in various human cancer subtypes, published between 2014 and 2022.
Cancer Typen/Cell Line/Sample TypeNEAT1
Isoforms
Investigated
Major FindingsRT-PCR Control GeneYearRef
Multiple
myeloma (MM)
n = 46 MM, n = 14 plasma cell leukaemia (PCL)
n = 628 from publicly available datasets: #GSE5900 (44 MM, 12 MGUS, 22 healthy donors) #GSE2658 and #GSE24080
Total NEAT1 and NEAT1_2
NEAT1_1 and NEAT1_2 with RNAseq
RNAseq allowed estimated isoform abundance was based on unambiguously mapped reads.
NEAT1 in tumour samples when compared to healthy controls
90% of total NEAT1 was NEAT1_1.
A negative correlation was found between—4NEAT1 and UPR.
Neither total NEAT1 nor NEAT1_2 correlated with overall survival or time-to-next-treatment.
NEAT1 was not found to be differentially expressed in diverse cell types, i.e., primary vs. secondary cell leukaemia.
Undisclosed2019[81]
n = 82 blood samplesTotal NEAT1 and NEAT1_2 NEAT1_1, inferred by the difference in Ct value between total NEAT1 and NEAT1_2.GAPDH2020[77]
B-cell acute lymphoblastic leukaemia (ALL)n = 16 blood samples NEAT1_1, inferred by the difference in Ct value between total NEAT1 and NEAT1_2.
NEAT1_1 and NEAT1_2.
Acute myeloid leukaemia (AML)n = 20 blood samples NEAT1_1, inferred by the difference in Ct value between total NEAT1 and NEAT1_2.
NEAT1_1 and NEAT1_2.
Chronic lymphocytic
leukaemia (CLL)
n = 310 blood samples NEAT1_1, inferred by the difference in Ct value between total NEAT1 and NEAT1_2.
Stable total NEAT1 but NEAT1_2 (40% of total NEAT1).
NEAT1_2 in patients with IGHV gene mutations.
NEAT1_2 in patients with Trisomy 12.
n = 72 peripheral blood samplesTotal NEAT1p53 binds to the NEAT1 promotor in CLL and lymphoma.
p21 and NEAT1 expression levels significantly correlated after irradiation.
Nutlin-3 induced NEAT1 expression 2.3-fold in WT p53 primary CLL cells, compared to 1.2-fold in p53 mutant cells.
Lamin B12015[82]
Chronic myeloid leukaemia (CML)n = 26 peripheral blood samplesTotal NEAT1 and NEAT1_2 Total NEAT1 and NEAT1_2.
Silencing BCR-ABL expression total NEAT1 and NEAT1_2 in CML cell line K562, suggesting NEAT1 may regulate BCR-ABL mediated pathways.
c-myc represses NEAT1 transcription by binding to promotor.
ACTB2018[78]
Acute promyelocytic
leukaemia (APL)
n = 31 APL and n = 12 normal blood samples
NB4, NB4-R2, and U937-PR9 cell lines
Total NEAT1 and NEAT1_2 Total NEAT1 and NEAT1_2 in APL patient samples when compared to normal granulocytes.
NEAT1 expression is repressed by PML-RAR α fusion gene.
NEAT1 expression is involved in the differentiation of APL cells.
ACTB2014[75]
Thyroid
carcinoma (TC)
n = 98 Peripheral blood and thyroid tissue samples (malignant n = 52, benign n = 46)NEAT1_2 NEAT1_2 in benign vs. malignant thyroid nodules.GAPDH2020[83]
Circulating blood monocytes (CBMs) and tumour-associated macrophages (TAMs)
n = undisclosed
TPC-1 cell line
Bone marrow-derived macrophages and macrophages
Undefined NEAT1 expression in TAMs, compared to CBMs.
NEAT1 is a direct target of miR-214 in TC cell lines.
Knockdown of NEAT1 impairs malignant progression of thyroid papillary carcinoma and tumour growth in vivo.
GAPDH2017[84]
Anaplastic
thyroid carcinoma (ATC)
n = 25 matched samples
SW1736 and KAT-18 cell lines
Total NEAT1 NEAT1 in ATC tissues and cells exposed to hypoxic conditions.GAPDH2020[85]
Papillary
thyroid
carcinoma (PTC)
n = 20 matched samples
NPA87, TPC-1, KAT-5, and HT-ori3 (control) cell lines
Total NEAT1 NEAT1 expression in patient PTC samples when compared to adjacent normal tissues.
NEAT1 expression in PTC cell lines compared to control cells.
GAPDH2018[86]
NeuroblastomaPublicly available datasets (total n = 1062): Versteeg (n = 88), Kocak (n = 476), and SEQC (n = 498)Total NEAT1 and NEAT1_2
(RT-PCR and RNA-FISH)
NEAT1_1 abundance inferred by subtracting NEAT1_2 levels from total NEAT1 levels.
NEAT1_1:NEAT1_2 in aggressive neuroblastoma.
NEAT1_2 and paraspeckles in nonaggressive neuroblastoma.
RPLP02021[87]
Breast cancer (BC)MCF-7, MDA-MB-453, MDA-MB-231, SKBR3, and MCF-10A (control) cell linesTotal NEAT1 NEAT1 in all cancer cell lines when compared to control cell line. NEAT1 was negatively correlated with miR-448.GAPDH2018[88]
n = 1065 post-data filtering of TCGA (n = 526), Oslo2 (n = 378), and METABRIC (n = 1904) cohorts.
BT474, BT549, HCC1569, Hs578T, MDA-MB-231, MDA-MB-468, MCF7, SK-BR-3, and T-47D cell lines
n = 74 BC biopsies and n = 27 non-malignant biopsies
NEAT1_2
NEAT1_1 with RNAseq
NEAT1_2 with RNA-FISH
NEAT1_1 expression level determined from polyA-selected RNAseq data from TCGA cohort.
NEAT1_1 expression is highest in ER-positive luminal A and B breast cancer.
NEAT1_2 and paraspeckle abundance correlate with high-grade disease (RNA-FISH).
NEAT1_2 in HER2-enriched and luminal B BC in all three cohorts.
NEAT1_2 is not expressed at RNA-FISH-detectable levels in normal breast tissue.
Geometric mean of GAPDH, B2M, and RPLP02020[89]
MCF-7, MDA-MB-231, and MDA-MB-468 cell lines
Gene expression data n = 2000
NEAT1_1 and NEAT1_2NEAT1_1 transcription was analysed by using a polyA primer for cDNA generation before RT-qPCR, using primers targeting total NEAT1.
NEAT1_2 transcription was analysed by using random primers for cDNA generation before primers specifically targeting the NEAT1_2 region of the transcript.
NEAT1 associated with poor patient prognosis.
RPL112015[90]
MDA-MB-231 and MCF-10A (control) cell linesNEAT1_2 with RNA-FISH Paraspeckle formation in MCF-7 cell lines when compared to MCF-10A cells.
NEAT1_2 expression after G-quadruplex (G4)-specific stabilization with small molecules.
NEAT1_2 expression could be regulated by a G4s.
GAPDH and ACTB2021[91]
Osteosarcoma (OS)U2OS cell lineTotal NEAT1 and NEAT1_2 (RT-qPCR and RNA-FISH).
NEAT1_1 in NEAT1_2 KO cells
NEAT1 isoform-specific KO cell lines were achieved using CRISPR-Cas9 technologies.
NEAT1_1 levels were unaltered or increased in some NEAT1_2−/− lines.
NEAT1_1 localises to nuclear speckles, independent of paraspeckles.
RPLP02017[92]
n = 47 biopsies and adjacent matched tissues
HOS, SaOS2, MG63, U2OS, and hFOB1.19 (control) cell lines
Total NEAT1 NEAT1 expression HIF-1 α in MG63 cells, and this NEAT1-mediated HIF-1 α expression was reversed by miR-186-5p in HOS cells.
NEAT1 in OS tissues and cell lines.
NEAT1 associated with advanced clinicopathologic features and poor overall survival.
NEAT1 promotes proliferation, invasion, and EMT in cell lines.
NEAT1 promoted growth in vivo.
miR-186-5p is a downstream target of NEAT1 in osteosarcoma.
GAPDH2019[93]
U2OS cell lineTotal NEAT1 and NEAT1_2Total NEAT1 levels were slightly higher in CBP80-KD and ARS2-KD cells when compared to control KD cells.
NEAT1_2 alone 5-fold in ARS2-KD cells, but not in CBP80- or PHAX-KD cells.
ARS2 suppresses the formation of paraspeckles.
GAPDH2020[47]
Ovarian cancer (OC)n = 30 paired tissue samples; CAOV3, ES-2, and IOSE80 (control) cell linesTotal NEAT1 NEAT1 in patient samples and OC cell lines.
NEAT1 knockdown with siRNA increased apoptosis and decreased proliferation, colony formation, migration, invasion, and glycolysis.
GAPDH2020[94]
ovarian carcinoma patient specimens (n = 18 responsive, n = 14 resistant)
SKOV3 HeyA-8, PTX-resistant, SKOV3/PTX, and HeyA-8/PTX cell lines.
n = 10 BALB/c athymic mice
Total NEAT1 NEAT1 in treatment-resistant patients when compared to treatment-responsive patients.
NEAT1 knockdown enhanced PTX sensitivity in PTX-resistant OC cells.
NEAT1 negatively regulates miR-194 expression.
NEAT1 sponges miR-194, leading to upregulation of ZEB1 expression.
NEAT1 knockdown improved sensitivity to PTX in OC in vivo.
GAPDH2017[95]
Prostate cancer (PC)LNCaP, DU145, and RWPE-1 (control) cell linesTotal NEAT1 NEAT1 in PCa cells.
NEAT1 negatively regulates hsa-miR-218-5p and has-miR-483-3p when compared to normal prostate epithelial cells.
GAPDH2022[96]
Explant cultures from primary, patient-derived bone metastatic prostate panel
Primary prostate and bone metastatic tissues
Patient-derived xenograft
TCGA datasets
Total NEAT1
Total NEAT1_1 (RNA-FISH)
NEAT1 in prostate cancer when compared to normal tissues (from TCGA datasets).
NEAT1_1 predicts poor patient prognosis.
NEAT1_1 enhances prostate-patient-derived xenograft growth through the post-transcriptional RNA modification N6-methyladenosine (m6A).
m6A level of NEAT1_1 correlated to prostate cancer progression and bone metastasis, and negatively correlated to patient survival.
GAPDH2020[97]
Hepatocellular carcinoma (HCC)blood samples (n = 36 HCC, n = 36 controls)Total NEAT1 NEAT1 in HCC patient samples when compared to healthy controls.
miR-129-5p negatively correlated to NEAT1 levels.
GAPDH2019[98]
n = 62 matched biopsies
MHCC97H, MHCC97L, SMCC7721 and LO2 (control) cell lines
Total NEAT1 NEAT1 in HCC tissues compared to adjacent tissues.
NEAT1 correlated with tumour size and vascular invasion.
NEAT1 knockdown inhibits proliferation, colony formation, and cell invasion in HCC.
miR-613 is a target of NEAT1 in HCC.
GAPDH2017[99]
n = 28 biopsies and adjacent tissues
HepG2, MHCC97L, MHCC97H, and LO2 (control) cell lines
Total NEAT1 NEAT1 expression compared to matched tumour samples.
Patients with NEAT1 expression had HIF-2 α expression, whilst patients with—NEAT1 expression (though still significantly higher than matched samples) had ¯ HIF-2 α expression.
GAPDH2018[100]
Gastric cancer (GC)n = 140 samples and n = 20 adjacent tissues
NCI-N87, SGC-7901, MKN-45, AGS, and GES-1 (control) cell lines
Total NEAT1 NEAT1 expression in GC cell lines compared to control cell line.
NEAT1 regulates expression of EMT-associated genes in GC cells; in vimentin and N-cadherin, in Zo-1 and E-cadherin; suggests KD of NEAT1 may inhibit EMT.
GAPDH2016[101]
Lung adenocarcinoma (LUAD)n = 124 biopsies and adjacent tissues
A549, CL1-0, and
BEAS-2B (control) cell lines
Total NEAT1Overexpression rate of NEAT1 in lung cancer samples was 90.3%.
Significant positive correlations found between NEAT1 and Oct4 mRNA expression levels.
Oct4 directly binds to NEAT1 promoter.
Lung cancer cell lines A549 and CL1-0 transiently overexpressing Oct4 induced NEAT1 promoter activity.
GAPDH2017[102]
A549, H460, H1650, H1975, H1299, and
NHBE (control) cell lines
TCGA database: n = 687
Total NEAT1 NEAT1 expression in all cell lines and patient samples when compared to normal tissue and control cell lines.
Positive correlation between ATF2 and NEAT1 expression in LUAD tissues.
ACTB2020[103]
Non-small-cell lung cancer (NSCLC)A549, H1299, H460, H1975, and BES-2B (control) cell linesTotal NEAT1 NEAT1 expression in all carcinoma cell lines when compared to control cell lines.
NEAT1 promotes growth, migration, and invasion of A549 and H460 cells.
NEAT1 directly targets hsa-miR-98-5p, and its expression was significantly downregulated in NSCLC cell lines when compared to normal lung epithelial cell line.
MAPK6 is a direct target of hsa-miR-98-5p in NSCLC cells.
GAPDH2019[104]
CRCn = 30 blood samples, n = 30 controls; validation in n = 100 patients, n = 100 controls.
n = 29 matched tissue samples, n = 19, whole blood and tissue samples.
HCT116 and LOVO cell lines
Total NEAT1 and NEAT1_2Details as to how NEAT1_1 expression was measured were not disclosed.
NEAT1 in whole blood of CRC patients when compared to normal controls.
Total NEAT1 and NEAT1_2 expression found to be highly accurate in distinguishing CRC patients from normal controls.
KD of NEAT1_1 inhibits proliferation and invasion.
KD of NEAT1_2 promoted growth.
NEAT1 expression was elevated in neutrophils in CRC patients.
NEAT1_2 correlated with better overall survival.
ACTB2015[71]
n = 71 tissue samples and n = 61 normal tissue samples from publicly available dataset,
RKO, CACO2, SW1116, LOVO, SW480, SW620, HT29, and HCT116 cell lines
BALB/c nude mice
Total NEAT1 NEAT1 associated with poor prognosis in CRC patients.
NEAT1 mediates cell proliferation in vitro and tumorigenicity in vivo.
KD of NEAT1 significantly inhibited flattening and spreading abilities of HCT116 and SW1116 cells, whilst overexpressing NEAT1 strongly promoted these abilities in HT29 cells.
E-cadherin and—N-cadherin expressed at both mRNA and protein levels in NEAT1 KD cells.
NEAT1 OE recovered proliferation potential of CRC cell lines which were impaired by simultaneous downregulation of DDX5.
DDX5 correlated with NEAT1 expression in 71 CRC samples.
GAPDH2018[105]
n = 12 paired patient samples
SW480, HT29, and Caco2 cell lines
Nude mice (n = 5–7 per group)
NEAT1_2 NEAT1 in CRC tissues is negatively correlated with miR-193a-3p.
NEAT1 KD miR-193a-3p expression and attenuates CRC cells.
KRAS acts as a target of miR-193a-3p.
GAPDH2019[106]
GEO databases GSE20916 and GSE9348
n = 100 and adjacent tissue samples
SW620, SW480, HCT116, HT29, CaCo-2, LOVO, and Colo205 cell lines
Total NEAT1 NEAT1 in tumour tissue, when compared to normal tissue, in both the independent datasets and in the matched tissue samples.
NEAT1 expression correlated with carcinoembryonic antigen (CEA) levels, tumour size, and distant metastasis.
NEAT1 predicts overall survival in CRC patients.
NEAT1 regulates cell proliferation and invasion through miR-34a.
GAPDH2019[107]
Nasopharyngeal carcinoma (NPC)n = 96 NPC and n = 32 nasopharyngeal epilethium tissuesTotal NEAT1 NEAT1 expression in patient samples when compared to normal tissues.
NEAT1 expression was negatively correlated with overall survival of NPC patients.
ACTB2019[108]
Laryngeal
cancer (LC)
n = 50 paired patient samples
TU686, TU177, AMC-HN-8, and 16HBE (control) cell lines
Total NEAT1miR-340-5p OE NEAT1 stability via direct binding and consequently NEAT1 expression in LC cells.
NEAT1 OE reversed repression of miR-340-5p OE on LC cell proliferation and invasion.
GAPDH2022[109]
Laryngeal
squamous cell cancer (LSCC)
n = 52 paired tissue samples
Hep-2 cell line
Total NEAT1 NEAT1 expression in LSCC tumour tissues compared to nonneoplastic tissues.
NEAT1 expression correlated with T grade, neck nodal metastasis, and clinical stages of LSCC.
NEAT1 knockdown inhibited the growth of LSCC xenografts in mice.
NEAT1 knockdown induced apoptosis in LSCC cells in vivo.
ACTB2016[110]
Oesophageal squamous cell carcinoma (OSCC)EC109, EC9706, and HET-1A (control) cell linesTotal NEAT1 NEAT1 expression in EC109 and EC9706 cell lines.
NEAT1 functions as an endogenous sponge for miR-129.
GAPDH2017[111]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Smith, N.E.; Spencer-Merris, P.; Fox, A.H.; Petersen, J.; Michael, M.Z. The Long and the Short of It: NEAT1 and Cancer Cell Metabolism. Cancers 2022, 14, 4388. https://doi.org/10.3390/cancers14184388

AMA Style

Smith NE, Spencer-Merris P, Fox AH, Petersen J, Michael MZ. The Long and the Short of It: NEAT1 and Cancer Cell Metabolism. Cancers. 2022; 14(18):4388. https://doi.org/10.3390/cancers14184388

Chicago/Turabian Style

Smith, Nadine E., Phaedra Spencer-Merris, Archa Hannah Fox, Janni Petersen, and Michael Z. Michael. 2022. "The Long and the Short of It: NEAT1 and Cancer Cell Metabolism" Cancers 14, no. 18: 4388. https://doi.org/10.3390/cancers14184388

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