Next Article in Journal
CD147 Is a Promising Target of Tumor Progression and a Prognostic Biomarker
Previous Article in Journal
Efficacy of Cancer Immunotherapy: An Umbrella Review of Meta-Analyses of Randomized Controlled Trials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Long Non-Coding RNA: Dual Effects on Breast Cancer Metastasis and Clinical Applications

1
Department of Pathophysiology, Basic Medical College, Nanchang University, Nanchang 330006, China
2
Second Clinical Medical College, Nanchang University, Nanchang 330006, China
3
First Clinical Medical College, Nanchang University, Nanchang 330006, China
4
Jiangxi Province Key Laboratory of Tumor Pathogenesis and Molecular Pathology, Nanchang 330006, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2019, 11(11), 1802; https://doi.org/10.3390/cancers11111802
Submission received: 19 September 2019 / Revised: 10 November 2019 / Accepted: 12 November 2019 / Published: 16 November 2019

Abstract

:
As a highly heterogeneous malignancy, breast cancer (BC) has become the most significant threat to female health. Distant metastasis and therapy resistance of BC are responsible for most of the cases of mortality and recurrence. Distant metastasis relies on an array of processes, such as cell proliferation, epithelial-to-mesenchymal transition (EMT), mesenchymal-to-epithelial transition (MET), and angiogenesis. Long non-coding RNA (lncRNA) refers to a class of non-coding RNA with a length of over 200 nucleotides. Currently, a rising number of studies have managed to investigate the association between BC and lncRNA. In this study, we summarized how lncRNA has dual effects in BC metastasis by regulating invasion, migration, and distant metastasis of BC cells. We also emphasize that lncRNA has crucial regulatory effects in the stemness and angiogenesis of BC. Clinically, some lncRNAs can regulate chemotherapy sensitivity in BC patients and may function as novel biomarkers to diagnose or predict prognosis for BC patients. The exact impact on clinical relevance deserves further study. This review can be an approach to understanding the dual effects of lncRNAs in BC, thereby linking lncRNAs to quasi-personalized treatment in the future.

1. Introduction

Nowadays, breast cancer (BC) is one of the most common cancers on a global scale, and it is the main cause of cancer death in women. In 2018, the number of newly diagnosed female BC cases worldwide reached 2.1 million, accounting for nearly a quarter of female cancer cases [1]. Metastasis is responsible for most cases of cancer mortality and relies on an array of processes, including the bilateral transition between epithelial-to-mesenchymal transition (EMT) and mesenchymal-to-epithelial transition (MET), promotion of cancer cell invasion, migration, stemness and metastasis, inhibition of cancer cell proliferation, anoikis, and post metastatic angiogenesis. Therefore, molecular mechanisms that affect any of these processes may be involved in regulating tumor metastasis.
Ever since the definition of four subtypes of BC (luminal A, luminal B, human epidermal growth factor receptor 2 positive (HER-2+), and basal-like), the clinical treatment for BC patients has become more specific and individualized. An increasing amount of mechanism research has made it possible to individualize treatment and achieve a better prognosis for patients. For example, in HER-2 overexpression patients, the successful application of anti-HER-2 drugs commendably demonstrates the advantages of the latest achievements in molecular science of basic medicine [2]. Triple negative breast cancer (TNBC) is defined by the negative expression of estrogen receptors (ER), progesterone receptors (PR), and the lack of expression of HER-2 [3]. Clinically, TNBC has high invasiveness and the metastasis rate is also abnormally high [4]. A lack of receptor expression leads to a low chemotherapy response rate and poor efficiency of targeted therapy. Until now, there has been no clear and effective molecular targeted therapy for TNBC [5]. Previous studies have demonstrated the crosstalk between tumor metastasis and therapy resistance in various cancers, including BC; among them, EMT regulation has a crucial role [6]. Additionally, EMT programming in cancer cells enables the remodeling of the extracellular matrix to break the dormancy of relapse-initiating cancer stem cells (CSCs) [7].
According to the encyclopedia of DNA elements (ENCODE) project, more than 80% of the human genome is transcribed into biochemically functional non-coding RNAs (ncRNAs) [8]. Transfer RNAs (tRNAs), ribosomal RNAs (rRNAs), microRNAs (miRNAs), small nuclear RNAs (snRNAs), and long non-coding RNAs (lncRNAs) are all types of ncRNAs [9]. LncRNAs are a class of ncRNA with a length over 200 nucleotides and can be divided into five categories according to the position of their DNA fragments in the genome, including bidirectional, antisense, intergenic, intronic, and sense lncRNAs [10].
As part of the ENCODE consortium, GENCODE (version 32) annotated 17,910 lncRNA genes and 48,351 lncRNA transcripts [11]. Among them, previous studies have indicated as many as 60,000 lncRNAs in humans and other mammals [12]. The discovery of numerous lncRNA transcripts has dramatically altered our understanding of cell biology, especially the biology of underexamined diseases such as cancer. Currently, many studies have investigated the association between lncRNAs and cardiovascular disease [13,14], neurological diseases [15,16], diabetes [17], cancer [18], and more. LncRNAs function through a variety of molecular mechanisms, such as acting as scaffolds for ribonucleoprotein (RNP) complexes, decoys for transcriptional factors or miRNAs, RNA interference, targeting of transcriptional factors or chromatin modifier proteins to specific genomic loci, and transcriptional regulation in cis or trans [19]. Interestingly, a number of studies have proposed that lncRNAs could inactivate miRNAs through a “sponge” effect, namely by sequestering miRNAs from their target mRNAs as a competing endogenous RNA (ceRNA) [20,21]. This interaction also exists during cancer progression [22,23]. In addition to basic research, the clinical application of lncRNAs is also an emerging research field [24]. LncRNAs are ideal diagnostic biomarkers and therapeutic targets. As of yet, due to the lack of basic clinical research, lncRNAs have not been applied in commercial diagnostic tests.
In this study, we reviewed the latest progress in this area and found that lncRNAs not only have a role in regulating the aggressiveness of BC cells but also have an effect on the distant metastasis and therapy resistance in BC patients. To our surprise, we found that lncRNAs also have a role in regulating the stemness of cancer stem cells and tumor angiogenesis, affecting cancer metastasis. In addition, by reviewing a number of studies, we discuss whether and how lncRNAs facilitate the diagnosis and prognosis prediction and improve the specificity for therapy of BC by acting as biomarkers. A considerable number of studies have uncovered the effect of lncRNAs on BC metastasis via various signaling pathways, which is also a focus of this review.

2. LncRNAs That Enhance Aggressiveness of BC Cells

2.1. LncRNAs That Promote Invasion and Migration of BC Cells

2.1.1. DANCR

Previous studies have identified lncRNA differentiation antagonizing non-protein-coding RNA (lncRNA DANCR) associated with hepatocellular carcinoma, colorectal cancer, and BC, especially in TNBC [25,26,27]. Knockdown of DANCR inhibited BC cell invasion and migration. Mechanistically, DANCR downregulation impaired the expression of CD44 and ATP binding cassette subfamily G member 2 (ABCG2) through enhancement of the zeste homolog 2 (EZH2) [27]. EZH2 is a part of the polycomb repressive complex 2 (PRC2) that facilitates target gene silencing through trimethylation of histone 3 lysine 27 (H3K27) residues [28]. DANCR suppression could have a negative effect on EZH2 and PRC2-mediated H3K27-trimethylation at the epigenetic level [29]. Another potential mechanism was identified by bioinformatics analysis, which confirmed the DANCR/miRNA-216a-5p axis in TNBC cells, and DANCR knockdown resulted in inhibited cell invasion and decreased expression of Nanog, SRY-box containing gene 2 (SOX2), and octamer-binding transcription factor-4 (OCT4) [30].

2.1.2. H19

LncRNA H19 has been identified to be functionally associated with many biological processes, such as cell proliferation, invasion, and apoptosis of tumors, including in BC [31,32]. DNA hypermethylation is involved in BC carcinogenesis and cell survival, and is initiated by abnormal expression of DNA methyltransferases (DNMTs), such as DNMT1, DNMT3a, and DNMT3b [33,34,35]. Evidence has indicated that H19 is aberrantly upregulated in BC and promotes invasion of BC cells via the miR-152/DNMT1 axis, and such outcomes can be reversed by miR-152 overexpression and DNMT1 knockdown [36]. H19 and insulin-like growth factor 2 (IGF2) usually jointly accelerate cancer development of mammary and pulmonary tumors [37,38]. At the H19/IGF2 locus, Vennin et al. discovered a newly conserved lncRNA antisense to H19 gene—H91 [39]. H91 promotes IGF2 gene expression through a novel promoter named Pm [40]. In-vitro and in-vivo studies have established that over-expressed H91 disturbs epigenetic modifications of H19 and IGF2 to promote BC development as an oncogene [39].

2.1.3. HOTAIR

Homeobox transcript antisense RNA (HOTAIR) is closely related to BC size, advancement, and extent of metastasis [41,42]. Years ago, it was speculated that HOTAIR might promote tumor aggressiveness through the upregulation of vascular endothelial growth factor (VEGF) and matrix metalloproteinase (MMP) and EMT-related genes [43]. Knockdown of HOTAIR inhibits the migration and invasion ability of BC cells via the P53/protein kinase B (AKT)/c-Jun N-terminal kinase (JNK)/MMPs signaling pathway [44] and the AKT/mammalian target of the rapamycin (mTOR) signaling pathway [45]. Currently, based on the bidirectional interaction between lncRNAs and miRNAs, researchers are trying to investigate more signaling networks involving miRNAs. HOTAIR could act as a sponge for miR-20a-5p and significantly influences migration and invasion through the HOTAIR/miR-20a-5p/high-mobility group AT-hook 2 (HMGA2) axis [46].

2.1.4. LINC00152

LncRNA LINC00152, an 828-bp lncRNA that is located on chromosome 2p11.2, was initially defined as differentially hypomethylated during hepatocarcinogenesis [47]. LINC00152 is highly expressed in various types of BC tissues, promoting invasion and migration of cancer cells [48,49,50]. With the knockdown of LINC00152, many studies observed inhibited migration and invasion of MDA-MB-231 cells and MCF-7 cells [49,50]. In agreement with findings in TNBC cell lines, LINC00152 is highly expressed in TNBC tissues and mechanistically LINC000152 induces tumorigenesis by inactivation of the tumor-suppressive breast cancer gene 1 (BRCA1)/gene of phosphate and tension homology deleted on chromosome ten (PTEN) signaling [48]. In ER-positive cancer cells, Hu et al. cast light not only on the positive effect of LINC00152 on one of the cellular preconditions for cancer metastasis, EMT, but also on chemo-resistance to doxorubicin (DOX) [49].

2.1.5. LINC00461

LncRNA LINC00461 is a novel tumor promoter in BC and is transcribed from a gene located at an intergenic region of human chromosome 5 [51]. Over-expression of LINC00461 results in upregulated expression of vimentin, E-cadherin, and ZEB1 [52,53,54]. Additionally, up-regulation of LINC00461 accelerates BC cell migration and invasion through the miR-30a-5p/integrin β3 axis [53]. These findings provide a basis for the association of LINC000461, EMT, and aggressive phenotypes of human cancer.

2.1.6. NEAT1

Transcriptome sequencing and next-generation sequencing files have shown that nuclear-enriched abundant transcript 1 (NEAT1) is one of the highest expressed lncRNAs in BC [55,56]. Across the four subtype-related ceRNA networks, NEAT1 has specific roles in each subtype through competing with diverse mRNAs [56]. These results hint at a strong connection between oncogenic effects of NEAT1 and BC. The latest studies suggest NEAT1 has a crucial role in promoting the growth, migration and invasion capacity of BC cells [57]. Despite the lack of in-vivo research with clinical BC samples, NEAT1 has been verified as an miR-448 sponge to enhance zinc finger E-Box binding homeobox 1 (ZEB1) expression, and this phenomenon can be reversed by an miR-448 mimic in vitro [58]. In triple-negative breast cancer, NEAT1 confers oncogenic effects through modulating chemoresistance and cancer stemness [59].

2.1.7. LINC01857

As a sharply up-regulated lncRNA in BC tissue, previous studies have noted that LINC01857 is usually associated with poor prognosis of BC patients [60]. The cAMP response element-binding protein binding protein (CREBBP) is an acetyltransferase that is commonly considered a tumor suppressor in BC [61,62]. Knockdown of LINC01857 inhibits H3K27Ac and cAMP response element-binding protein 1 (CREB1) transcription via reducing the enrichment of CREBBP in the CREB1 promoter region. Overexpression of CREB1 in MCF-7 cells reverses the inhibition effect of LINC01857 knockdown, leading to larger quantities of migration and invasion cells [60].

2.2. LncRNAs That Promote Distant Metastasis of BC Cells

2.2.1. H19

In 2017, Zhou et al. isolated BC cells from the primary mammary tumor, circulation, and metastatic lesions in the lungs of TA2 mice and found that H19 is an essential key factor in several tumor metastasis procedures. In cells isolated from plasma blood, H19 acts as an endogenous sponge by directly binding to let-7b. Subsequently, the binding of let-7b and cytohesin 3 (CYTH3; one of ArfGEFs) activates adenosine 5′-diphosphate (ADP) ribosylation factor (ARF) and EMT [63]. Contrarily, in cells isolated from the primary tumor and lung metastasis tumor, H19 regulates GIT2 via sponging miR-200b/c and promotes G protein-coupled receptor kinase interacting protein 2 (GTI2; one of ArfGAPs) expression through inactivating ARF [63]; thus, MET took place in order to accelerate successful metastatic colonization of a secondary organ [64,65]. Matouk et al. revealed in their research that hypoxia-induced EMT is associated with strong induction of both H19 and miR-675, and the H19 gene is highly expressed in common metastatic sites regardless of the tumor’s primary origin [66]. Therefore, we suggest H19 is a tumor-promoting lncRNA that endows distant metastatic potential to BC cells.

2.2.2. HOTAIR

In recent years, HOTAIR has been one of the most well-studied lncRNAs. Alongside invasion and migration of BC cells, HOTAIR participates in metastasis of BC as well. Mounting evidence has revealed the positive correlation between high circulating HOTAIR and lymph node metastasis, as well as distant metastasis of BC [67].
Epigenetic signaling from the tumor microenvironment regulates the switch from dormancy to metastatic growth [68]. Apart from studies focusing on cross talks in the BC cell itself, the latest emerging studies have investigated the roles of lncRNAs in the tumor microenvironment. By establishing an orthotopic mouse model of MDA-MB-231 cells, previous studies have explored a novel transforming growth factor-beta1 (TGF-β1)/cyclin-dependent kinase 5 (CDK5)/HOTAIR/H3K27 signaling that could promote tumor growth, lymph node metastasis, and breast-to-lung metastasis [69]. Additionally, emerging evidence from laminin-rich extracellular matrix-based three-dimensional organotypic culture (lrECM 3D) has suggested the presence of epigenetic regulation of gene expression by extracellular matrix (ECM) signaling in cancer cells [70,71]. Interestingly, HOTAIR expression exhibits robust induction in lrECM 3D over that in 2D culture, and such an induction partly relies on bromodomain containing 4 (BRD4) and the canonical ECM signaling pathway, namely integrins and Src kinase [72]. This validates the significant observations of Gupta et al., who suggested conventional 2D cultures which lacked critical factors that stimulated the expression of HOTAIR and promoted cancer metastasis very early [73].

2.2.3. HIF1A-AS2

Hypoxia-inducible factor 1 alpha-antisense RNA 2 (HIF1A-AS2) may be a latent tumor promotor in TNBC patients. In vitro, HIF1A-AS2 obviously depresses TNBC cell migration and invasion; in 86 TNBC cases, high expression of HIF1A-AS2 was associated with more lymph node metastasis, distant metastasis, and unfavorable histological grade [74]. Previous studies have found even higher plasma levels of HIF1A-AS2 expression in BC cases with lymph node metastasis [75].

2.2.4. RP1

Jia et al. first identified a novel lncRNA on chromosome 22 and named it lncRNA RP1. They found that RP1 was significantly upregulated in BC tissue, and RP1 could enhance the metastasis of BC in vitro and in vivo through the Kruppel-like factor 5 (KLF5)/RP1/p27kip1 signaling pathway. Mechanistically, RP1 attenuated p27kip1 translation through interacting with complex p-4E-BP1/eIF4E and restored p27kip1-downregulated Snail1, thus promoting BC progression [76].

2.2.5. MALAT1

The nuclear lncRNA metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) is among the most conserved and highly abundant lncRNAs in normal tissues, suggesting that it may have vital biological implications [77]. When it comes to BC metastasis, overwhelming previous findings indicate its increased expression is associated with relapse and metastatic progression in BC. Up-regulation of MALAT1 in the tissue and serum of BC, generated by different strategies, has been reported by independent groups [78,79,80]. This is one novel epigenetic mechanism by which MALAT1 facilitates a pro-metastatic phenotype in BC by trans-regulating EEF1A1 [81]. Li et al. overexpressed MALAT1 and found this promoted lipopolysaccharide (LPS)-induced invasion and metastasis of human and mouse mammary tumor cells [82]. Moreover, a novel regulation mechanism of MALAT1 in BC cells was presented by more recent research. They suggested that hypoxia might induce the specific chromatin interactions and increases MALAT1 expression as well as its antisense strand TALAM1 in BC cells [83]. An in-vivo study showed that Malat1−/− mice exhibited a significant reduction in lung macrometastases and a significantly decreased metastatic burden. Moreover, knockdown of Malat1 results in mammary tumor differentiation [84]. Similar conclusions can be drawn from in-vitro and in-vivo studies on TNBC. Knockdown of MALAT1 impaired TNBC growth and metastasis via the miR-1/slug axis [80] and MALAT1/miR-129-5p axis [85].

3. LncRNAs That Attenuate Aggressiveness of BC Cells

3.1. LncRNAs That Inhibit Invasion and Migration of BC Cells

3.1.1. GAS5

Growth arrest-specific 5 (GAS5) is a noncoding gene that was initially isolated from mouse NIH 3T3 cells [86]. The expression level of GAS5 is significantly lower in BC in younger cases [87]. Additionally, studies have demonstrated the negative correlation between GAS5 expression and malignancy of different cell lines of BC [88]. GAS5 could function as an endogenous sponge that sequesters miR-21, resulting in reciprocal repression of miR-21 and GAS5. This interaction relies on the putative miR-21-binding site at exon 4 of GAS5 and the RNA-induced silencing complex (RISC) [88]. Additionally, GAS5 inhibits TNBC cell invasion through the miR-196a-5p/forkhead box protein O1 (FOXO1)/phosphatidylinositol 3-kinase (PI3K)/AKT axis, which starts with competitive binding of GAS5 to miR-196a-5p [89].

3.1.2. MT1JP

The lncRNA metallothionein 1J pseudogene (MT1JP) is located on chromosome 16 in a cluster that consists of several homologous protein-coding genes in the metallothionein family. It was first reported as a tumor suppressor that modulates the p53 protein expression level, thereby regulating the p53-related signaling pathway [90]. Decreased MT1JP expression is significantly associated with more lymphatic metastasis and advanced Tumor Node Metastasis (TNM) stage for gastric cancer patients, and in-vitro and in-vivo studies have shown the suppressive role of MT1JP on the migration and invasion of gastric cancer [91,92].
In terms of BC, one recent study presented the first evidence that MT1JP overexpression markedly inhibits invasion and enhances cisplatin sensitivity of BC cells, which is coincident with gastric cancer results. Mechanistically, MT1JP competitively binds to miR-24-3p and inhibits the Wnt/β-catenin signaling pathway [93]. More importantly, down-regulated MT1JP expression in BC patients indicates a better prognosis.

3.1.3. NEF

A considerable number of studies have observed the down-regulated expression of lncRNA-neighboring enhancer of FOXA2 (NEF) as a tumor suppressor in various types of cancers, for instance, cholangiocarcinoma [94], non-small-cell lung cancer [95], and hepatocellular carcinoma [96]. BC is no exception. Previous in-vitro cell experiment results have indicated that NEF might inhibit TNBC cell migration and invasion by downregulating miRNA-155, and overexpression of miRNA-155 significantly attenuates the inhibitory effects of NEF overexpression on cancer cell migration [97].

3.1.4. NKILA

In various cancer types, evidence indicates that TGF-β-induced EMT is nuclear factor kappa-B (NF-κB)-dependent. LncRNA NKILA is a newly discovered lncRNA that functions as a tumor suppressor in BC metastasis [98,99]. In BC, recent research has linked the negative feedback regulation of lncRNA NKILA with NF-κB activation and TGF-β-induced EMT, thus indicating NKILA may inhibit the breast EMT process of BC cells, thus retarding metastasis. Mechanistically, NKILA can suppress several NF-κB target genes by directly binding to the NF-κB/IκB complex [100]. Furthermore, there is a positive correlation between NKILA and E-cadherin expression level, and the long-term survival of BC patients [99].

3.1.5. LET

LncRNA LET is a newly discovered molecule involved in the invasion and migration of BC. Despite the overall expression of LET in BC tissue and cell lines being down-regulated, LET expression differs in various BC cell lines. LET expression is lower in MDA-MB-231 (BC cell line, triple negative) than MCF-10 (non-tumorigenic epithelial breast cell line), which suggests the negative association between the LET expression level and the acquisition of tumorigenesis abilities of BC cells. Additionally, over-expressing LET increases E-cadherin and decreases N-cadherin and vimentin expression, and thereby represses EMT of BC cells [101].

3.1.6. TFAP2A-AS1

Transcription factor AP-2 alpha-antisense RNA 1 (TFAP2A-AS1) may be a novel tumor suppressor in BC. Transwell assay showed a remarkable decrease of invasive cell numbers in MCF-7 and MDA-MB-231 cells transfected with TFAP2A-AS1. Mechanistically, TFAP2A-AS1 sequesters the miR-933-inhibiting Smad signaling pathway by downregulating drosophila mothers against decapentaplegic family member 2 (SMAD2). An in-vivo study showed that TFAP2A-AS1-overexpressed MCF-7 cells injected in nude mice ended up with a significantly smaller tumor volume and weight [102].

3.1.7. LncKLHDC7B

With knockdown of LncKLHDC7B, Beltran-Anaya et al. confirmed an inverse association between the unique LncKLHDC7B overexpression and migration and invasion in the immunomodulatory subtype of TNBC [103]. However, they did not investigate the potential mechanism further. To understand the clinical implications of LncKLHDC7B expression, they analyzed public data in 2012 and 2016 [104,105,106] and found that lower LncKLHDC7B expression was related to lower survival and an increased risk of a recurrent or metastatic event.

3.2. LncRNAs That Inhibit Distant Metastasis of BC Cells

3.2.1. MALAT1

On the one hand, as we discussed before, increased expression of MALAT1 is associated with relapse and metastatic progression in breast cancer. On the other hand, a few studies surprisingly provide new insight into MALAT1 as a metastasis-suppressing lncRNA in BC. Several studies have found down-regulated expression of MALAT1 in BC [107,108]. Kwok et al. suggested that the PTEN-microRNA-MALAT1 axis may promote tumorigenesis and demonstrated the first evidence that MALAT1 possesses novel tumor-suppressive capabilities in BC [109]. A similar case was found by establishing both genetically engineered mouse models and xenograft models; Kim et al. observed that metastasis was specifically induced by somatic knockout of Malat1. Mechanistically, MALAT1 prevents prometastatic transcription factor TEA domain family member 2 (TEAD) from associating with its co-activator Yes-associated protein (YAP) [107].
Collectively, we cannot conclude whether MALAT1 is a tumor-promoting or suppressing lncRNA; numerous looping events remain to be discovered to explore the possible explanation for the dual effects of MALAT1 in BC.

3.2.2. MEG3

LncRNA maternally expressed gene 3 (MEG3) has been implicated in tumorigenesis and progression of BC, and the mechanisms in most cases are associated with DNMT of MEG3. It has been demonstrated that miR-506 could target DNMT1 and DNMT3b to retard tumor development [110]. Downregulating miR-506 increases the methylation level of MEG3 promoter and inhibits MEG3 expression via the miR-506/SP3/SP1/DNMT1/MEG3 axis in human BC cell lines, resulting in attenuation of metastasis of MCF-7 and MDA-MB-231 cells [111]. Accordingly, it is plausible to speculate that both miR-506 and MEG3 function as tumor suppressors in BC, which is coincident with previous results [112,113]. Additionally, treatment of DNA methylation inhibitor (5′-Aza-2′-deoxycytidine) may partly reverse the signaling and retard tumor development [114].

3.2.3. NLIPMT

Jiang et al. identified a hitherto uncharacterized lncRNA called novel lncRNA, inhibiting proliferation and metastasis (NLIPMT) in BC. An in-vitro study and xenograft study showed that over-expression of NLIPMT inhibited cell motility, growth, and metastasis of BC. Mechanistically, downregulated NLIPMT accelerated proliferation and restored motility of BC cells by promoting the expression of glycogen synthase kinase 3β (GSK3β) and EMT proteins [115]. Interestingly, according to previous studies, GSK3β has dual effects in different types of cancer. On the one hand, evidence indicates that GSK3β functions as a tumor suppressor in the breast [116,117,118]. On the other hand, a tumor-promoting effect of GSK3β has been indicated in colon [119] and pancreatic cancers [120]. Therefore, the conflict of GSK3β’s functions might disturb the effect of NLIPMT in different types of cancer.

3.2.4. XIST

The lncRNA X inactive specific transcript (XIST) is a potential tumor suppressor in cancer [121,122]. A luciferase reporter assay verified that XIST targets the miR-155/CDX1 axis, and overexpression of XIST remarkably inhibits BC cell migration and invasion [123]. Xing et al. demonstrated how the overexpression of XIST in BC cells significantly attenuates its metastatic ability to the brain from two different aspects. Firstly, a pathway screening based on gene set enrichment analysis (GSEA) uncovered that XIST activates the c-Met signaling pathway by activation of moesin (MSN) on the X chromosome. MSN, a member of the ERM (ezrin, radixin, and moesin) protein family, participates in the maintenance of epithelial integrity [124]. Loss of c-Met decreases the transmigration abilities of BC cells and suppresses the breast-to-brain metastatic abilities in vivo [125]. Secondly, considering activated microglia can modulate neurodegenerative diseases and tumor progression under pathological conditions [126], they also investigated the role of XIST in microglia. Interestingly, XIST-downregulated BC cells could secrete exosomal miR-503 to reprogram microglia from the M1 (tumor-suppressive) into the M2 (tumor-promoting) phenotype, causing local immune suppression [125].

4. The Regulation of LncRNAs in the Stemness of BC Cells

The cancer stem cell theory indicates that cancers are sustained by tumor-initiating cells, and researchers initially termed these cancer stem cells, with distinct phenotypes and a high tumorigenic potential [127]. CSCs are the “seeds” of tumor metastasis and recurrence, and have a limitless and high invasion and migration capacity [128]. The subpopulation of CSCs will express specific biomarkers, giving rise to metastasis; therapy resistance; and the recurrence of cancers such as CD 26, CD133 and CD44v6 [129,130].
Currently, with further research emerging on lncRNAs, many researchers have suggested that lncRNAs are required for stemness maintenance of cancer cell lines [131]. However, whether and how lncRNAs have a role in breast cancer stem-like cells (BCSCs) remains unclear. Discussing the potential mechanism of lncRNA function may contribute to developing novel therapeutic approaches. Herein, we review our current understanding of the cell-biological mechanisms that regulate BC metastasis, with BCSCs and lncRNAs as the pivotal objects.
EMT is a cellular process that is not only strongly correlated with tumor metastasis, but also with CSC [132]. LncRNAs have been demonstrated to play vital roles during EMT, and some of these molecules also have regulatory roles in the proliferation of CSC [133], although their functions are still unclear in BC. Current research indicates that T-cell leukemia 1 upstream neural differentiation-associated RNA (TUNAR, LINC00617) functions as a vital promotor of EMT and causes generation of stem cell properties via activating the transcription of SOX2, thus promoting BC progression and metastasis [134]. LncRNA ES1 (LINC01108) and a hedgehog pathway-associated lncRNA called lncRNA-Hh can enhance the stem cell properties of BC via regulating SOX2 as well [135,136]. ES1 controls the expression of stemness transcription factors in BC cells by regulating the OCT4/SOX2/miR-302 axis [135]; lncRNA-Hh directly targets GAS1 to increase the SOX2 and OCT4 expression, thus promoting the CSCs-like characteristics of Twist-driven EMT cells and Twist-positive BC cells [136]. LncRNAs could also attenuate stemness in BC cells via inhibiting EMT. The study on brain metastasis of BC found that the lncRNA XIST suppressed EMT and the MSN/c-Met axis via MSN-mediated protein stabilization, which leads to the attenuation of stemness in the tumor cells. The negative correlation between RNA of XIST and c-Met expression was also observed in brain metastasis samples from BC patients [125].
LncRNAs can regulate oncogenic transcription factor in stemness properties via sponging miRNAs as a ceRNA. HOTAIR could suppress miR-34a, resulting in remarkably up-regulated SOX2, P53, self-renewal capacity, and migration of CSC-MCF7 [137]. Apart from HOTAIR, FEZ family zinc finger 1-antisense 1 (FEZF1-AS1) and LINC00511 can promote the BCSC properties via FEZF1-AS1/miR-30a/Nanog and the LINC00511/miR-185-3p/E2F1/Nanog signaling pathway, respectively [138,139]. Contrarily, the tumor-suppressive lncRNA fibroblast growth factor 13-antisense RNA 1 (FGF13-AS1) could act as an endogenous sponge by directly binding to Myc mRNA via binding insulin-like growth factor 2 mRNA binding proteins (IGF2BP), followed by the obstructive combination of IGF2BPs and Myc mRNA [140]. Myc is an oncogenic transcription factor that plays vital roles in glycolysis and stemness properties [141]. Furthermore, a feedback loop between Myc and FGF13-AS1 also participates in regulating these novel suppressive effects of FGF13-AS1 [141]. Impaired function of Myc may explain how FGF13-AS1-inhibited BC cells achieve migration and invasion.
In addition to the effect of lncRNA dysregulation on EMT proteins and oncogenic transcription factors in stemness properties, the latest studies point out that lncRNAs play a role in BCSC maintenance under hypoxia via the tricarboxylic acid cycle [142]. Mechanistically, H19 sequesters miRNA let-7 as a ceRNA, leading to hypoxia-inducible factor-1α (HIF-1α) release and an increase in pyruvate dehydrogenase kinase 1 (PDK1) expression [143]. PDK1, an essential glycolytic enzyme, is correlated with tumor proliferation, metastasis, and poor prognosis [144,145]. Therefore, it is plausible to conclude that H19 may activate PDK1 in BCSC via the let-7/HIF-1 axis under hypoxia, and this mechanism may be required in the development of primary breast carcinomas.

5. The Regulation of LncRNAs in the Angiogenesis in BC

Angiogenesis refers to the process of neovessels sprouting from preexisting vessels, as opposed to the vasculogenesis of embryonic fibroblasts and their consequently activating proliferation, invasion, and metastasis of cancer cells [146]. It is well known that the more pathological angiogenesis occurs, the more chances for cancer cells to migrate into the circulatory system and achieve distant metastasis. However, during the last few years, despite many reviews aimed at organizing the mechanism underneath BC angiogenesis, only a few reviews have associated BC angiogenesis with lncRNAs. Here, we review how lncRNAs participate in the angiogenesis of BC, thus further supporting the understanding of their involvement in promoting or inhibiting BC metastasis.
LncRNA MALAT1 is considered an oncogenic regulator of BC, and we have discussed its role in regulating the aggressiveness of BC cells. Vascular endothelial growth factor (VEGF) is the crucial signaling molecule for angiogenesis, acting by regulating proliferation, survival, and migration of the cancer [147]. In terms of BC angiogenesis, up-regulated MALAT1 in BC tissue also shows oncogenic effects in vitro and in vivo through VEGF regulation. Mechanistically, knockdown of MALAT1 significantly promotes angiogenesis of BC through upregulating the expression level of miR-145 in MCF-7 cells and BC tissue [148].
Some tumor-suppressive lncRNAs also exhibit negative regulatory effects on BC angiogenesis. For example, MEG3 not only inhibits the aggressiveness of BC cells, but also inhibits BC angiogenesis via regulating the VEGF or AKT pathway [149,150].

6. The Regulation of LncRNAs in Chemotherapy Resistance in BC

The current clinical treatments for BC patients include surgery and postoperative chemotherapy, targeted therapy, or radiation therapy [151]. It is critical to select the chemotherapy regimens wisely, because different types of BC will respond to different regiments and specific drugs. For example, ribociclib plus fulvestrant (hormone receptor-positive/HER-2+) [152]; trastuzumab, neratinib and lapatinib (HER-2+) [153,154]; or ipatasertib plus paclitaxel (TNBC) [155]. Working on distinct molecular mechanisms, these chemotherapies act synergistically and eliminate most tumor cells during the initial treatment phase. However, the frequent development of drug resistance is still a major obstacle for medical personnel and BC patients. BC recurrences continued to occur steadily throughout the study period from 5 to 20 years, with risks ranging from 10% to 41% [156]. This emphasizes the urgent need for deeper insight into the molecular mechanisms leading to therapy resistance, especially chemotherapy resistance.
It is well established that EMT is not only associated with tumor metastasis and cancer cell stemness, but also with resistance to conventional therapies [6]. Chemotherapy-resistant cells treated by chemotherapeutics (e.g., oxaliplatin, 5-fluorouracil) undergo EMT, and this applies to cells treated with monoclonal antibodies (e.g., trastuzumab) as well [157]. As outlined in previous sections, it is well established that several lncRNAs play crucial roles in the aggressiveness of BC. Next, we are going to review lncRNAs in participating signaling pathways that modulate both the aggressiveness and chemotherapy resistance of BC cells.
BC cells that undergo EMT usually show resistance to 5-fluoroutacil (5-FU) treatment [158]. Knockdown of NEAT1 sensitizes TNBC cells to chemotherapy, indicating its involvement in chemotherapy resistance [59]. Mechanistically, in-vitro and in-vivo studies have shown that NEAT1 promotes distant metastasis and 5-FU resistance through the miR-129/ZEB2 axis and miR-211/HMGA2 axis in BC [159,160]. Similarly, up-regulated terminal differentiation-induced non-coding RNA (TINCR) significantly promotes EMT-associated protein expression and HER-2 expression in trastuzumab-resistant BC cells via the miR-125b-HER-2/Snail-1 signaling pathway [157].
Wnt/β-catenin signaling is one of the well-established signaling pathways that regulates EMT in various cancer types. LncRNAs regulating aggressive cytology of cancer cells via the Wnt/β-catenin signaling pathway have been reviewed before [161]. Several novel lncRNAs may regulate chemotherapy resistance through regulating Wnt/β-catenin signaling. LINC00968 is a newly discovered lncRNA that has suppressive roles in cancer progression [162]. LINC00968 can attenuate drug resistance of MCF-7/ADM and KPL-4/ADM cells to adriamycin, taxel, and vincristine. Mechanistically, LINC00968 inhibits the Wnt2/β-catenin signaling pathway via hairy/enhancer-of-split related with YRPW motif 1 (HEY1) [163]. Additionally, knockdown of lncRNA UCA1 increases the tamoxifen sensitivity of BC cells via inhibition of the Wnt/β-catenin pathway.
Differently to the lncRNAs above that were correlated with EMT, the lncRNA in the non-homologous end-joining pathway 1 (LINP1) is overexpressed in drug-resistant cells and mediates its oncogenic role in BC by decreased apoptosis-related proteins (such as caspase-8 and caspase-9), thus promoting cell growth, metastasis, and chemotherapy resistance to 5-FU and DOX in BC [164]. 5-FU and DOX are two effective and crucial chemotherapeutics against BC [165], which can interact with DNA by hindering macro-molecular biosynthesis [166].
Taken together, the studies above have revealed a new role for lncRNAs in elucidating the epigenetic mechanism of chemo-resistance. Mechanistically, most lncRNAs are recognized as involved in EMT to amplify or attenuate BC cell response to chemotherapy, and among them, Wnt/β-catenin signaling is a crucial common pathway. LncRNA-regulated apoptosis of BC cells also has an additional effect on chemotherapy resistance to 5-FU and DOX. However, more tissue samples and animal models are required to further investigate the correlation between associated lncRNA expression and patient clinicopathological features, in order to evaluate the potential of lncRNAs as independent biomarkers in BC.

7. LncRNAs That Function as Prognostic Biomarkers for BC Patients

7.1. LncRNAs as Prognostic Biomarkers for BC Patients

As we reviewed above, MALAT1 could be either a metastasis promoter or suppressor in BC patients. In terms of prognosis, current research approves of the association between over-expression of MALAT1 and poor prognosis in BC patients. Meta-analysis indicated that upregulated expression of MALAT1 in BC tissues is significantly associated with more lymph node metastasis, shorter 5-year adverse disease-free survival (DFS) and shorter overall survival (OS) [167]. Interestingly, MALAT1 was also an important pro-inflammatory factor regulating lipopolysaccharide-induced inflammatory responses in endothelial cells of BC. Li et al. found that MALAT1 was elevated in BC patients with postoperative fever, and high expression of MALAT1 predicted adverse short-term recurrence-free survival (RFS) [82]. Additionally, an upregulated MALAT1 expression level was related to positive progesterone receptor (PR) status. This might be coincident with the dual role of MALAT1 in BC [167].
Many studies have illustrated the tumor-promoting role of HOTAIR in BC progression and the association between over-expression of HOTAIR in tissues and shorter survival in BC patients [168]. Of particular note, the Kaplan-Meier survival curve showed that patients with high circulating HOTAIR expression had a worse DFS than those with low circulating HOTAIR [67]. Similarly, plasma H19 levels were significantly correlated with lymph node metastasis, and high plasma H19 levels were significantly reversed in postoperative samples [169]. In addition, analysis in the preoperative and postoperative plasma samples showed the postoperative levels of circulating GAS5 and H19 significantly decreased, and circulating GAS5 levels in the patients with a positive lymph node metastasis state decreased after surgery [170]. Therefore, circulating HOTAIR, H19 and GAS5 could be potential biomarkers for BC early screening and prognosis monitoring.
In addition, as shown in Figure 1, the expression of LINC00473 [171], TINCR [157], LINP1 [172] and lncRNA breast cancer anti-estrogen resistance 4 (BCAR4) [173,174] were all up-regulated in BC tissues and increased further during metastasis, and they were all associated with poor prognosis of BC patients. Of particular note, the over-expression of BCAR4 was not only negatively correlated with OS in BC patients, but was also highly correlated with the incidence of BC, so it might be applied as a major negative and unfavorable prognostic marker for BC [173,174].

7.2. LncRNAs as Biomarkers for Diagnosis and Prognosis of TNBC Patients

Due to the lack of effective targeted therapies and high recurrence rate in response to chemotherapy, the prognosis of TNBC is the most unfavorable among all types of BC. It is therefore urgent to identify the prognostic biomarkers for the diagnosis of TNBC, or its individualized therapy targets.
In 2016, through detecting the expression levels of lncRNAs in TNBC and non-TNBC tissues separately, Lv et al. identified four lncRNAs (RP11-434D9.1, LINC00052, BC016831, and immunoglobulin kappa variable (IGKV)) as biomarkers to differentiate TNBC cancer from non-TNBC [175]. As shown in Figure 2, we reviewed the latest studies that revealed lncRNAs as novel prognostic biomarkers for TNBC patients in particular.
Some tumor-promoting lncRNAs in TNBC may function as biomarkers for diagnosis and prognosis of TNBC patients. Collation of research on 63 TNBC tissues showed DANCR expression was abruptly upregulated and correlated with worse OS and TNM stages [27]. By utilizing The Cancer Genome Atlas (TCGA) database and analyzing clinical TNBC samples, Zhang et al. found that the expression of nicotinamide phosphoribosyltransferase antisense RNA (NAMPT-AS) was up-regulated, which epigenetically activates NAMPT to promote tumor progression and metastasis [176]. In addition, according to Kaplan-Meier survival curve analysis, up-regulated expression levels of lncRNA-ATB, NAMPT-AS and HIF1A-AS2 are negatively correlated with OS and DFS of TNBC patients [74,176,177].
Further, some tumor-suppressive lncRNAs in TNBC may function as biomarkers for diagnosis and prognosis of TNBC patients as well. For example, the latest study observed that TNBC patients with low miR-503 host gene (MIR503HG) expression had a significantly worse prognosis compared with those with high MIR503HG expression, and low MIR503HG expression was a poor independent prognostic factor for OS in TNBC patients [178].
In summary, the clinical function of lncRNA is a topic of interest nowadays, including how they can be used for predicting prognosis or serving as therapeutic targets. However, existing research is not sufficient to support clinical applications, and the underlining clinical relevance of lncRNAs and BC still awaits for future validation. Our review aims to provide direction for further research in the future to promote better clinical practice applications.

8. Discussion

BC is an aggressive malignant disease in women worldwide, with a high tendency to metastasize [1]. Novel lncRNA discoveries and the increasing quality of lncRNA function studies are providing new ideas to improve understanding of the mechanism of BC pathogenesis and progression. It has been demonstrated that many lncRNAs have crucial roles in the cell proliferation of BC [179]. Despite the fact that BC is at the forefront of cancer research, very few studies had investigated the association between lncRNAs and BC distant metastasis until the last 5 years.
Here, we have reviewed the latest findings of how lncRNAs participate in BC progression, especially metastasis ability. Many researchers have tried to cast light on signaling pathways involved in the metastasis of BC. As shown in Table 1, we summarize the mechanisms of different processes that are necessary for metastasis to take place.
  • LncRNAs participate in regulating the aggressiveness of BC cells (such as HOTAIR, H19, DANCR, RP1, GAS5, LNC00152, NEF, and NKILA).
  • LncRNAs can regulate the stemness properties of BC cells (such as HOTAIR, H19, ES1, TUNAR, lncRNA-Hh, XIST, and FEZF1-AS1).
  • LncRNAs can regulate angiogenesis of BC (such as MALAT1, MEG3, and LINC00968).
In the next section, we briefly review lncRNAs that can regulate the sensitivity to chemotherapy in vitro and in vivo, and some may function as biomarkers to diagnosis or predict prognosis for BC patients (such as HOTAIR, H19, MALAT1, TINCR, LINC00473, LINP1, BCAR4, and SNHG6).
Interestingly, we found that some particular lncRNAs have dual effects in BC, indicated by opposite previous findings and conclusions; for example, different research provided almost opposite evidence for the roles of MALAT1 in BC. Of note, there exist some differences in the materials and design of these experiments; for example, Kim et al. suggested MALAT1 was a suppressor by conducting experiments on MCF10A and MDA-MB-231 cells [107]. Arun et al. report that genetic loss or systemic knockdown of Malat1 results in slower tumor growth accompanied by significant differentiation into cystic tumors and a reduction in metastasis [84]. Additionally, Li et al. suggested MALAT1 might promote recurrence and lung metastasis of BC patients with early postoperative fever by analyzing plasma levels of MALAT1 in 258 cases of patients with primary breast carcinoma, and knock down of MALAT1 in 4T1 xenograft mice [82]. Therefore, the foremost cause of the discrepancy may be due to the tumor microenvironment (TME) and metabolism of lncRNA in vivo, and clearly further work needs to be performed to establish whether MALAT1 does have dual effects in BC.
The intimate association between EMT and CSCs has provided further insights into tumor progression and the contribution of CSCs to tumor metastasis and colonization. Interestingly, most studies on CSC and EMT center on the epigenetic differences between CSCs and non-stem cancer cells, while irreversible mutations in genes can also endow cells with metastasis ability. For example, gene-specific cell cloning in pancreatic cancer and medulloblastoma causes tumor metastasis [181,182]. This unexpected phenomenon raises a question: Is it that only CSC can lead to metastasis through genetic changes, or can non-stem cancer cells acquire tumorigenesis and metastasis at the same time via genetic changes? This question still needs in-situ cancer formation experiments to find an answer.
Autophagy is a rather conservative biological process, regulated by autophagy-related genes (ATGs) [183]. In this process, eukaryotic organelles and various substances form autophagosomes in the phospholipid bilayer; subsequently, autophagosomes go through degradation and recycling after transporting to the lysosomes. Some state that autophagy is a double-edge-sword for tumor metastasis. On the one hand, it reduces tumor necrosis induced by hypoxia, inhibits the infiltration of inflammatory cells, and thus inhibits tumor metastasis [184]. On the other hand, in the early stages of metastasis, autophagy accelerates tumor metastasis through promoting the survival of cells under metabolic stress and oxygen-deficient conditions [185]. Some have investigated whether lncRNAs function as regulators in BC cells autophagy. Both in-vitro and in-vivo studies indicate that the lncRNA GAS5 facilitates autophagy of BC cells. GAS5 can regulate the expression of ATG3 via the GAS5-miR-23a-ATG3 axis, not only in MDA-MB-231 and MCF-7 cells, but also in patients with BC [186].
Some consider that autophagy can promote tumor metastasis by reducing anoikis of cancer cells [187]. Anoikis refers to programmed apoptosis after separation from ECM [188]. The anoikis prevention of detached malignant cancer cells is the precondition for metastasis, and an elevated threshold for anoikis usually leads to heightened metastatic potential [189]. Currently, more and more studies are investigating the role of lncRNAs in cancer cell anoikis. For example, in thyroid carcinoma, lncRNA FOXD2 adjacent opposite strand RNA 1(FOXD2-AS1) promotes the cancer stem cell features and anoikis resistance in thyroid cancer cells via inhibiting the miR-7-5p/telomerase reverse transcriptase (TERT) axis [190]. In ovarian cancer, caspase 3/7 assay showed that MALAT1 knockdown resulted in increased anoikis [191]. Similarly, in bladder cancer, knockdown of LINC00958 attenuated resistance to anoikis [192]. However, as of the time of writing, no research has revealed how lncRNAs function in the anoikis of BC cells. We consider that this may be a promising area for future research.
In five breast cancer subtypes (luminal A, luminal B, HER2-overexpressing, basal-like, and normal-like), the pathological characteristics and prognostics exhibit significant differences [193,194]. In order to elucidate the mechanism and improve personalized clinical treatment, systematical analysis of these common and specific lncRNAs is significant.
HOTAIR acts as a vital cancer-promoting lncRNA in the BC process, and it was found to be up-regulated in the three described subtypes (ER-positive, HER2-positive and TNBC) of the disease [195,196]. After assessing the expression profile of BC subtypes, Mathias et al. found down-regulation of LINC0051 in the ER-positive subtypes and upregulation in HER2-positive and TNBC subtypes. Additionally, MEG3, LINC00152, DANCR, and MALAT1 were linked to ER-positive and TNBC subtypes. When it comes to TNBC and HER2-positive subtypes, only GAS5 was commonly expressed [195]. By conducting RNAseq of seven pairs of HER2-positive tumor vs non-tumor tissues, Yang et al. demonstrated that LOC100288637 is the highest positive correlative lncRNA with HER2, while RPL13P5 is the highest negative correlative one [197].
As is commonly known, the description of basal-like BC usually regards cell aggressiveness, stem-like phenotype, poor prognosis, and high rates of metastasis [198]. In the basal-like subtype-related network, three significantly dysregulated lncRNAs, including NEAT1, FAM83H-AS1 and XIST, were further investigated. Liu et al. demonstrated that combined use of docetaxel and carboplatin in ovarian cancer could rescue the down-regulated XIST expression [199]. Thus, we consider that XIST may be a proper example to explain drug repositioning.
One possible mechanism that links the regulation on lncRNAs to some specific subtype may be associated with specific hormone receptors. In BC cells, under hormone stimulation or deprivation, estrogen could regulate lncRNAs expression via estrogen binding or independent of estrogen binding mechanism (Apo-ER) [200]. The regulation on SNHG3, TINCR, HOTAIR, LINC00067, etc. expression is ligand dependent; the regulation on DANCR, LINC01016, DSCAM-AS1, etc. expression is ligand independent [195]. Among them, HOTAIR’s expression is induced after 17β-estradiol (E2) treatment in BC cells, indicating that HOTAIR’ expression may be associated with estrogen receptors [196]. Achievements in clinical treatment strategies have improved the survival and prognosis of BC patients. Based on the mechanisms of how lncRNAs promote tumor invasion and metastasis, researchers have begun to use drugs to regulate the expression of lncRNAs in order to impede tumor metastasis. Zhao et al. showed that a high concentration of 17b-estriol could impair the promotion of MALAT1 in the growth, invasion, and metastasis of BC cell lines by reducing lncRNA MALAT-1 expression in a dose-dependent manner [201]. In addition, the presence of acquired and de-novo resistance is still a serious concern. Many researchers have illustrated crosstalk between resistance and metastasis, which relies on EMT, plasticity acquisition, tumor heterogeneity, and so forth [202,203,204,205]. At present, research uncovering whether and how lncRNAs function during the treatment of BC is relatively sufficient, and some of them, as we discussed above, have helped to gain a better understanding of the molecular mechanisms in chemotherapy resistance. In terms of radiotherapy, the relationship between lncRNAs and radiosensitization was first demonstrated in nasopharyngeal carcinoma cells. Huang et al. showed that the lncRNA Cur enhanced radiosensitization in nasopharyngeal carcinoma cells [206]. To the best of our knowledge, in BC, HOTAIR and LINP1 were the initially investigated lncRNAs that may regulate BC radiosensitization. Knockdown of HOTAIR sensitizes BC cells to ionizing radiation through activating miR-218 [207]. Overexpression of LINP1 enhances the survival of BC cells exposed to radiation [208].
An understanding of the networks between lncRNAs, target miRNAs and genes, and BC metastasis is very useful. Besides the expression profile of lncRNAs, more functional and mechanistic investigations of lncRNA regulation of BCSC, cell autophagy and anoikis are required. Happily, lncRNAs have been in the center of public discussion since 5 years ago. This review helps to systematically understandi the molecular underpinning for finding more clinical biomarkers and treatment targets for BC in the near future.

Author Contributions

All authors contributed to the study conception and design. Material preparation was performed by Q.-Y.H. (Qi-Yuan Huang) and G.-F.L. Data collection and analysis were performed by Q.-Y.H. (Qi-Yuan Huang), G.-F.L., X.-L.Q., and L.-B.T. The first draft of the manuscript was written by Q.-Y.H. (Qi-Yuan Huang), L.-B.T., and Q.-Y.H. (Qing-Yun Huang). X.-L.Q. and L.-X.X. critically revised the work. All authors commented on previous versions of the manuscript. All authors read and approved the final manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (nos. 31860317).

Acknowledgments

We would like to thank the technical support by Xing-Hua Xiao, and the support from our families and friends.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.; Torre, L.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed]
  2. Pondé, N.; Brandão, M.; El-Hachem, G.; Werbrouck, E.; Piccart, M. Treatment of advanced HER2-positive breast cancer: 2018 and beyond. Cancer Treat. Rev. 2018, 67, 10–20. [Google Scholar] [CrossRef] [PubMed]
  3. Suba, Z. Triple-negative breast cancer risk in women is defined by the defect of estrogen signaling: Preventive and therapeutic implications. Onco Targets Ther. 2014, 7, 147–164. [Google Scholar] [CrossRef] [PubMed]
  4. Gadi, V.K.; Davidson, N.E. Practical Approach to Triple-Negative Breast Cancer. J. Oncol. Pract. 2017, 13, 293–300. [Google Scholar] [CrossRef] [PubMed]
  5. Rabanal, C.; Ruiz, R.; Neciosup, S.; Gomez, H. Metronomic chemotherapy for non-metastatic triple negative breast cancer: Selection is the key. World J. Clin. Oncol. 2017, 8, 437–446. [Google Scholar] [CrossRef] [PubMed]
  6. Pastushenko, I.; Brisebarre, A.; Sifrim, A.; Fioramonti, M.; Revenco, T.; Boumahdi, S.; Van Keymeulen, A.; Brown, D.; Moers, V.; Lemaire, S.; et al. Identification of the tumour transition states occurring during EMT. Nature 2018, 556, 463–468. [Google Scholar] [CrossRef]
  7. Mitra, A.; Mishra, L.; Li, S. EMT, CTCs and CSCs in tumor relapse and drug-resistance. Oncotarget 2015, 6, 10697–10711. [Google Scholar] [CrossRef]
  8. The ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef]
  9. Morris, K.V.; Mattick, J.S. The rise of regulatory RNA. Nat. Rev. Genet. 2014, 15, 423–437. [Google Scholar] [CrossRef]
  10. Ponting, C.P.; Oliver, P.L.; Reik, W. Evolution and functions of long noncoding RNAs. Cell 2009, 136, 629–641. [Google Scholar] [CrossRef]
  11. GENECODE. Available online: https://www.gencodegenes.org/ (accessed on 19 September 2019).
  12. Evans, J.R.; Feng, F.Y.; Chinnaiyan, A.M. The bright side of dark matter: lncRNAs in cancer. J. Clin. Investig. 2016, 126, 2775. [Google Scholar] [CrossRef] [PubMed]
  13. Congrains, A.; Kamide, K.; Oguro, R.; Yasuda, O.; Miyata, K.; Yamamoto, E.; Kawai, T.; Kusunoki, H.; Yamamoto, H.; Takeya, Y.; et al. Genetic variants at the 9p21 locus contribute to atherosclerosis through modulation of ANRIL and CDKN2A/B. Atherosclerosis 2012, 220, 449–455. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, Y.; Cai, Y.; Wu, G.; Chen, X.; Liu, Y.; Wang, X.; Yu, J.; Li, C.; Chen, X.; Jose, P.; et al. Plasma long non-coding RNA, CoroMarker, a novel biomarker for diagnosis of coronary artery disease. Clin. Sci. 2015, 129, 675–685. [Google Scholar] [CrossRef] [PubMed]
  15. Ashraf, G.; Ganash, M.; Athanasios, A. Computational analysis of non-coding RNAs in Alzheimer‘s disease. Bioinformation 2019, 15, 351–357. [Google Scholar] [CrossRef]
  16. Huang, X.; Luo, Y.L.; Mao, Y.S.; Ji, J.L. The link between long noncoding RNAs and depression. Prog. Neuropsychopharmacol. Biol. Psychiatry 2017, 73, 73–78. [Google Scholar] [CrossRef]
  17. Mirza, A.H.; Kaur, S.; Pociot, F. Long non-coding RNAs as novel players in β cell function and type 1 diabetes. Hum. Genom. 2017, 11, 17. [Google Scholar] [CrossRef]
  18. Lorenzi, L.; Avila Cobos, F.; Decock, A.; Everaert, C.; Helsmoortel, H.; Lefever, S.; Verboom, K.; Volders, P.; Speleman, F.; Vandesompele, J.; et al. Long noncoding RNA expression profiling in cancer: Challenges and opportunities. GenesChromosomes Cancer 2019, 58, 191–199. [Google Scholar] [CrossRef]
  19. Angrand, P.; Vennin, C.; Le Bourhis, X.; Adriaenssens, E. The role of long non-coding RNAs in genome formatting and expression. Front. Genet. 2015, 6, 165. [Google Scholar] [CrossRef]
  20. Fan, C.; Ma, L.; Liu, N. Systematic analysis of lncRNA-miRNA-mRNA competing endogenous RNA network identifies four-lncRNA signature as a prognostic biomarker for breast cancer. J. Transl. Med. 2018, 16, 264. [Google Scholar] [CrossRef]
  21. Huang, Y. The novel regulatory role of lncRNA-miRNA-mRNA axis in cardiovascular diseases. J. Cell. Mol. Med. 2018, 22, 5768–5775. [Google Scholar] [CrossRef]
  22. Lian, Y.; Xiong, F.; Yang, L.; Bo, H.; Gong, Z.; Wang, Y.; Wei, F.; Tang, Y.; Li, X.; Liao, Q.; et al. Long noncoding RNA AFAP1-AS1 acts as a competing endogenous RNA of miR-423-5p to facilitate nasopharyngeal carcinoma metastasis through regulating the Rho/Rac pathway. J. Exp. Clin. Cancer Res. CR 2018, 37, 253. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, X.; Cheng, T.; He, Q.; Lei, Z.; Chi, J.; Tang, Z.; Liao, Q.; Zhang, H.; Zeng, L.; Cui, S. LINC01133 as ceRNA inhibits gastric cancer progression by sponging miR-106a-3p to regulate APC expression and the Wnt/β-catenin pathway. Mol. Cancer 2018, 17, 126. [Google Scholar] [CrossRef] [PubMed]
  24. Leucci, E. Cancer development and therapy resistance: Spotlights on the dark side of the genome. Pharmacol. Ther. 2018, 189, 22–30. [Google Scholar] [CrossRef] [PubMed]
  25. Yuan, S.; Wang, J.; Yang, F.; Tao, Q.; Zhang, J.; Wang, L.; Yang, Y.; Liu, H.; Wang, Z.; Xu, Q.; et al. Long noncoding RNA DANCR increases stemness features of hepatocellular carcinoma by derepression of CTNNB1. Hepatology 2016, 63, 499–511. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, Y.; Zhang, M.; Liang, L.; Li, J.; Chen, Y. Over-expression of lncRNA DANCR is associated with advanced tumor progression and poor prognosis in patients with colorectal cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 11480–11484. [Google Scholar] [PubMed]
  27. Sha, S.; Yuan, D.; Liu, Y.; Han, B.; Zhong, N. Targeting long non-coding RNA DANCR inhibits triple negative breast cancer progression. Biol. Open 2017, 6, 1310–1316. [Google Scholar] [CrossRef]
  28. Hock, H. A complex Polycomb issue: The two faces of EZH2 in cancer. Genes Dev. 2012, 26, 751–755. [Google Scholar] [CrossRef]
  29. Vaidya, A.M.; Sun, Z.; Ayat, N.; Schilb, A.; Liu, X.; Jiang, H.; Sun, D.; Scheidt, J.; Qian, V.; He, S.; et al. Systemic Delivery of Tumor-Targeting siRNA Nanoparticles against an Oncogenic LncRNA Facilitates Effective Triple-Negative Breast Cancer Therapy. Bioconj. Chem. 2019. [Google Scholar] [CrossRef]
  30. Tao, W.; Wang, C.; Zhu, B.; Zhang, G.; Pang, D. LncRNA DANCR contributes to tumor progression via targetting miR-216a-5p in breast cancer: lncRNA DANCR contributes to tumor progression. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef]
  31. Muller, V.; Ferrer, L.O.; Steinbach, B.; Pantel, K.; Schwarzenbach, H. Interplay of lncRNA H19/miR-675 and lncRNA NEAT1/miR-204 in breast cancer. Mol. Oncol. 2019. [Google Scholar] [CrossRef]
  32. Lottin, S.; Adriaenssens, E.; Dupressoir, T.; Berteaux, N.; Montpellier, C.; Coll, J.; Dugimont, T.; Curgy, J.J. Overexpression of an ectopic H19 gene enhances the tumorigenic properties of breast cancer cells. Carcinogenesis 2002, 23, 1885–1895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kastl, L.; Brown, I.; Schofield, A.C. Altered DNA methylation is associated with docetaxel resistance in human breast cancer cells. Int. J. Oncol. 2010, 36, 1235–1241. [Google Scholar] [PubMed] [Green Version]
  34. Rajabi, H.; Tagde, A.; Alam, M.; Bouillez, A.; Pitroda, S.; Suzuki, Y.; Kufe, D. DNA methylation by DNMT1 and DNMT3b methyltransferases is driven by the MUC1-C oncoprotein in human carcinoma cells. Oncogene 2016, 35, 6439–6445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhang, W.; Chang, Z.; Shi, K.E.; Song, L.; Cui, L.I.; Ma, Z.; Li, X.; Ma, W.; Wang, L. The correlation between DNMT1 and ERalpha expression and the methylation status of ERalpha, and its clinical significance in breast cancer. Oncol. Lett. 2016, 11, 1995–2000. [Google Scholar] [CrossRef] [Green Version]
  36. Li, Z.; Li, Y.; Li, Y.; Ren, K.; Li, X.; Han, X.; Wang, J. Long non-coding RNA H19 promotes the proliferation and invasion of breast cancer through upregulating DNMT1 expression by sponging miR-152. J. Biochem. Mol. Toxicol. 2017, 31. [Google Scholar] [CrossRef]
  37. Bates, P.; Fisher, R.; Ward, A.; Richardson, L.; Hill, D.J.; Graham, C.F. Mammary cancer in transgenic mice expressing insulin-like growth factor II (IGF-II). Br. J. Cancer 1995, 72, 1189–1193. [Google Scholar] [CrossRef] [Green Version]
  38. Min, H.; Lee, S.; Woo, J.; Jung, H.; Park, K.; Jeong, H.; Hyun, S.; Cho, J.; Lee, W.; Park, J.; et al. Essential role of DNA methyltransferase 1-mediated transcription of insulin-like growth factor 2 in resistance to histone deacetylase inhibitors. Clin. Cancer Res. 2017, 23, 1299–1311. [Google Scholar] [CrossRef] [Green Version]
  39. Vennin, C.; Spruyt, N.; Robin, Y.M.; Chassat, T.; Le Bourhis, X.; Adriaenssens, E. The long non-coding RNA 91H increases aggressive phenotype of breast cancer cells and up-regulates H19/IGF2 expression through epigenetic modifications. Cancer Lett. 2017, 385, 198–206. [Google Scholar] [CrossRef]
  40. Tran, V.; Court, F.; Duputié, A.; Antoine, E.; Aptel, N.; Milligan, L.; Carbonell, F.; Lelay-Taha, M.; Piette, J.; Weber, M.; et al. H19 antisense RNA can up-regulate Igf2 transcription by activation of a novel promoter in mouse myoblasts. PLoS ONE 2012, 7, e37923. [Google Scholar] [CrossRef]
  41. Sørensen, K.; Thomassen, M.; Tan, Q.; Bak, M.; Cold, S.; Burton, M.; Larsen, M.; Kruse, T. Long non-coding RNA HOTAIR is an independent prognostic marker of metastasis in estrogen receptor-positive primary breast cancer. Breast Cancer Res. Treat. 2013, 142, 529–536. [Google Scholar] [CrossRef]
  42. Zhuang, Y.; Nguyen, H.; Burow, M.; Zhuo, Y.; El-Dahr, S.; Yao, X.; Cao, S.; Flemington, E.; Nephew, K.; Fang, F.; et al. Elevated expression of long intergenic non-coding RNA HOTAIR in a basal-like variant of MCF-7 breast cancer cells. Mol. Carcinog. 2015, 54, 1656–1667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kim, H.J.; Lee, D.W.; Yim, G.W.; Nam, E.J.; Kim, S.; Kim, S.W.; Kim, Y.T. Long non-coding RNA HOTAIR is associated with human cervical cancer progression. Int. J. Oncol. 2015, 46, 521–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Yu, Y.; Lv, F.; Liang, D.; Yang, Q.; Zhang, B.; Lin, H.; Wang, X.; Qian, G.; Xu, J.; You, W. HOTAIR may regulate proliferation, apoptosis, migration and invasion of MCF-7 cells through regulating the P53/Akt/JNK signaling pathway. Biomed. Pharm. 2017, 90, 555–561. [Google Scholar] [CrossRef] [PubMed]
  45. Huang, Y.; Du, J.; Mi, Y.; Li, T.; Gong, Y.; Ouyang, H.; Hou, Y. Long Non-coding RNAs Contribute to the Inhibition of Proliferation and EMT by Pterostilbene in Human Breast Cancer. Front. Oncol. 2018, 8, 629. [Google Scholar] [CrossRef] [PubMed]
  46. Zhao, W.; Geng, D.; Li, S.; Chen, Z.; Sun, M. LncRNA HOTAIR influences cell growth, migration, invasion, and apoptosis via the miR-20a-5p/HMGA2 axis in breast cancer. Cancer Med. 2018, 7, 842–855. [Google Scholar] [CrossRef]
  47. Cogill, S.B.; Wang, L. Co-expression Network Analysis of Human lncRNAs and Cancer Genes. Cancer Inf. 2014, 13, 49–59. [Google Scholar] [CrossRef] [Green Version]
  48. Wu, J.; Shuang, Z.; Zhao, J.; Tang, H.; Liu, P.; Zhang, L.; Xie, X.; Xiao, X. Linc00152 promotes tumorigenesis by regulating DNMTs in triple-negative breast cancer. Biomed. Pharm. 2018, 97, 1275–1281. [Google Scholar] [CrossRef]
  49. Hu, X.L.; Wang, J.; He, W.; Zhao, P.; Wu, W.Q. Down-regulation of lncRNA Linc00152 suppressed cell viability, invasion, migration, and epithelial to mesenchymal transition, and reversed chemo-resistance in breast cancer cells. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 3074–3084. [Google Scholar]
  50. Xu, S.; Wan, L.; Yin, H.; Xu, H.; Zheng, W.; Shen, M.; Zhang, Z.; Pang, D. Long Noncoding RNA Linc00152 Functions as a Tumor Propellant in Pan-Cancer. Cell. Physiol. Biochem. 2017, 44, 2476–2490. [Google Scholar] [CrossRef] [Green Version]
  51. Yang, Y.; Ren, M.; Song, C.; Li, D.; Soomro, S.H.; Xiong, Y.; Zhang, H.; Fu, H. LINC00461, a long non-coding RNA, is important for the proliferation and migration of glioma cells. Oncotarget 2017, 8, 84123–84139. [Google Scholar] [CrossRef] [Green Version]
  52. Shim, H.S.; Yoon, B.S.; Cho, N.H. Prognostic significance of paired epithelial cell adhesion molecule and E-cadherin in ovarian serous carcinoma. Hum. Pathol. 2009, 40, 693–698. [Google Scholar] [CrossRef] [PubMed]
  53. Dong, L.; Qian, J.; Chen, F.; Fan, Y.; Long, J. LINC00461 promotes cell migration and invasion in breast cancer through miR-30a-5p/integrin beta3 axis. J. Cell. Biochem. 2019, 120, 4851–4862. [Google Scholar] [CrossRef] [PubMed]
  54. Murakami, A.; Nakagawa, T.; Fukushima, C.; Torii, M.; Sueoka, K.; Nawata, S.; Takeda, O.; Ishikawa, H.; Sugino, N. Relationship between decreased expression of squamous cell carcinoma antigen 2 and E-cadherin in primary cervical cancer lesions and lymph node metastasis. Oncol. Rep. 2008, 19, 99–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Shi, Y.; Ye, P.; Long, X. Differential Expression Profiles of the Transcriptome in Breast Cancer Cell Lines Revealed by Next Generation Sequencing. Cell. Physiol. Biochem. 2017, 44, 804–816. [Google Scholar] [CrossRef]
  56. Jia, Y.; Shi, L.; Yun, F.; Liu, X.; Chen, Y.; Wang, M.; Chen, C.; Ren, Y.; Bao, Y.; Wang, L. Transcriptome sequencing profiles reveal lncRNAs may involve in breast cancer (ER/PR positive type) by interaction with RAS associated genes. Pathol. Res. Pract. 2019, 215, 152405. [Google Scholar] [CrossRef]
  57. Zhang, M.; Wu, W.B.; Wang, Z.W.; Wang, X.H. lncRNA NEAT1 is closely related with progression of breast cancer via promoting proliferation and EMT. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 1020–1026. [Google Scholar]
  58. 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]
  59. Shin, V.Y.; Chen, J.; Cheuk, I.W.; Siu, M.T.; Ho, C.W.; Wang, X.; Jin, H.; Kwong, A. Long non-coding RNA NEAT1 confers oncogenic role in triple-negative breast cancer through modulating chemoresistance and cancer stemness. Cell Death Dis. 2019, 10, 270. [Google Scholar] [CrossRef] [Green Version]
  60. Xiong, Y.; Gu, Y.; Wang, F.; Li, L.; Zhu, M.; Wang, N.; Mi, H.; Qiu, X. LINC01857 as an oncogene regulates CREB1 activation by interacting with CREBBP in breast cancer. J. Cell. Physiol. 2019. [Google Scholar] [CrossRef]
  61. Jia, D.; Augert, A.; Kim, D.; Eastwood, E.; Wu, N.; Ibrahim, A.; Kim, K.; Dunn, C.; Pillai, S.; Gazdar, A.; et al. Loss Drives Small Cell Lung Cancer and Increases Sensitivity to HDAC Inhibition. Cancer Discov. 2018, 8, 1422–1437. [Google Scholar] [CrossRef] [Green Version]
  62. Garcia-Carpizo, V.; Ruiz-Llorente, S.; Sarmentero, J.; González-Corpas, A.; Barrero, M. CREBBP/EP300 Bromodomain Inhibition Affects the Proliferation of AR-Positive Breast Cancer Cell Lines. Mol. Cancer Res. MCR 2019, 17, 720–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zhou, W.; Ye, X.L.; Xu, J.; Cao, M.G.; Fang, Z.Y.; Li, L.Y.; Guan, G.H.; Liu, Q.; Qian, Y.H.; Xie, D. The lncRNA H19 mediates breast cancer cell plasticity during EMT and MET plasticity by differentially sponging miR-200b/c and let-7b. Sci. Signal. 2017, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Ocaña, O.; Córcoles, R.; Fabra, A.; Moreno-Bueno, G.; Acloque, H.; Vega, S.; Barrallo-Gimeno, A.; Cano, A.; Nieto, M. Metastatic colonization requires the repression of the epithelial-mesenchymal transition inducer Prrx1. Cancer Cell 2012, 22, 709–724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Beerling, E.; Seinstra, D.; de Wit, E.; Kester, L.; van der Velden, D.; Maynard, C.; Schäfer, R.; van Diest, P.; Voest, E.; van Oudenaarden, A.; et al. Plasticity between Epithelial and Mesenchymal States Unlinks EMT from Metastasis-Enhancing Stem Cell Capacity. Cell Rep. 2016, 14, 2281–2288. [Google Scholar] [CrossRef] [Green Version]
  66. Matouk, I.J.; Raveh, E.; Abu-lail, R.; Mezan, S.; Gilon, M.; Gershtain, E.; Birman, T.; Gallula, J.; Schneider, T.; Barkali, M.; et al. Oncofetal H19 RNA promotes tumor metastasis. Biochim. Biophys. Acta 2014, 1843, 1414–1426. [Google Scholar] [CrossRef] [Green Version]
  67. Lu, R.; Zhang, J.; Zhang, W.; Huang, Y.; Wang, N.; Zhang, Q.; Qu, S. Circulating HOTAIR expression predicts the clinical response to neoadjuvant chemotherapy in patients with breast cancer. Cancer Biomark. 2018, 22, 249–256. [Google Scholar] [CrossRef]
  68. Barkan, D.; Kleinman, H.; Simmons, J.; Asmussen, H.; Kamaraju, A.; Hoenorhoff, M.; Liu, Z.; Costes, S.; Cho, E.; Lockett, S.; et al. Inhibition of metastatic outgrowth from single dormant tumor cells by targeting the cytoskeleton. Cancer Res. 2008, 68, 6241–6250. [Google Scholar] [CrossRef] [Green Version]
  69. Ren, Y.; Jia, H.H.; Xu, Y.Q.; Zhou, X.; Zhao, X.H.; Wang, Y.F.; Song, X.; Zhu, Z.Y.; Sun, T.; Dou, Y.; et al. Paracrine and epigenetic control of CAF-induced metastasis: The role of HOTAIR stimulated by TGF-ss1 secretion. Mol. Cancer 2018, 17, 5. [Google Scholar] [CrossRef]
  70. Teoh-Fitzgerald, M.L.; Fitzgerald, M.P.; Zhong, W.; Askeland, R.W.; Domann, F.E. Epigenetic reprogramming governs EcSOD expression during human mammary epithelial cell differentiation, tumorigenesis and metastasis. Oncogene 2014, 33, 358–368. [Google Scholar] [CrossRef] [Green Version]
  71. Lelièvre, S.A. Tissue polarity-dependent control of mammary epithelial homeostasis and cancer development: An epigenetic perspective. J. Mammary Gland Biol. Neoplasia 2010, 15, 49–63. [Google Scholar] [CrossRef] [Green Version]
  72. Li, M.; Li, X.; Zhuang, Y.; Flemington, E.K.; Lin, Z.; Shan, B. Induction of a novel isoform of the lncRNA HOTAIR in Claudin-low breast cancer cells attached to extracellular matrix. Mol. Oncol. 2017, 11, 1698–1710. [Google Scholar] [CrossRef] [PubMed]
  73. Gupta, R.; Shah, N.; Wang, K.; Kim, J.; Horlings, H.; Wong, D.; Tsai, M.; Hung, T.; Argani, P.; Rinn, J.; et al. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature 2010, 464, 1071–1076. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, Y.; Zhang, G.; Han, J. HIF1A-AS2 predicts poor prognosis and regulates cell migration and invasion in triple-negative breast cancer. J. Cell. Biochem. 2019. [Google Scholar] [CrossRef] [PubMed]
  75. Liu, M.; Xing, L.; Liu, Y. A three-long noncoding RNA signature as a diagnostic biomarker for differentiating between triple-negative and non-triple-negative breast cancers. Medicine 2017, 96, e6222. [Google Scholar] [CrossRef] [PubMed]
  76. Jia, X.; Shi, L.; Wang, X.; Luo, L.; Ling, L.; Yin, J.; Song, Y.; Zhang, Z.; Qiu, N.; Liu, H.; et al. KLF5 regulated lncRNA RP1 promotes the growth and metastasis of breast cancer via repressing p27kip1 translation. Cell Death Dis. 2019, 10, 373. [Google Scholar] [CrossRef] [Green Version]
  77. Zhang, B.; Arun, G.; Mao, Y.S.; Lazar, Z.; Hung, G.; Bhattacharjee, G.; Xiao, X.; Booth, C.J.; Wu, J.; Zhang, C.; et al. The lncRNA Malat1 is dispensable for mouse development but its transcription plays a cis-regulatory role in the adult. Cell Rep. 2012, 2, 111–123. [Google Scholar] [CrossRef] [Green Version]
  78. Miao, Y.; Fan, R.; Chen, L.; Qian, H. Clinical Significance of Long Non-coding RNA MALAT1 Expression in Tissue and Serum of Breast Cancer. Ann. Clin. Lab. Sci. 2016, 46, 418–424. [Google Scholar]
  79. Jadaliha, M.; Zong, X.; Malakar, P.; Ray, T.; Singh, D.K.; Freier, S.M.; Jensen, T.; Prasanth, S.G.; Karni, R.; Ray, P.S.; et al. Functional and prognostic significance of long non-coding RNA MALAT1 as a metastasis driver in ER negative lymph node negative breast cancer. Oncotarget 2016, 7, 40418–40436. [Google Scholar] [CrossRef]
  80. Jin, C.; Yan, B.; Lu, Q.; Lin, Y.; Ma, L. Reciprocal regulation of Hsa-miR-1 and long noncoding RNA MALAT1 promotes triple-negative breast cancer development. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2016, 37, 7383–7394. [Google Scholar] [CrossRef]
  81. Li, X.; Chen, N.; Zhou, L.; Wang, C.; Wen, X.; Jia, L.; Cui, J.; Hoffman, A.; Hu, J.; Li, W. Genome-wide target interactome profiling reveals a novel epigenetic pathway for oncogenic lncRNA in breast cancer. Am. J. Cancer Res. 2019, 9, 714–729. [Google Scholar]
  82. Li, Z.; Xu, L.; Liu, Y.; Fu, S.; Tu, J.; Hu, Y.; Xiong, Q. LncRNA MALAT1 promotes relapse of breast cancer patients with postoperative fever. Am. J. Transl. Res. 2018, 10, 3186–3197. [Google Scholar] [PubMed]
  83. Stone, J.; Kim, J.; Vukadin, L.; Richard, A.; Giannini, H.; Lim, S.; Tan, M.; Ahn, E. Hypoxia induces cancer cell-specific chromatin interactions and increases MALAT1 expression in breast cancer cells. J. Biol. Chem. 2019, 294, 11213–11224. [Google Scholar] [CrossRef] [PubMed]
  84. Arun, G.; Diermeier, S.; Akerman, M.; Chang, K.; Wilkinson, J.; Hearn, S.; Kim, Y.; MacLeod, A.; Krainer, A.; Norton, L.; et al. Differentiation of mammary tumors and reduction in metastasis upon Malat1 lncRNA loss. Genes Dev. 2016, 30, 34–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Zuo, Y.; Li, Y.; Zhou, Z.; Ma, M.; Fu, K. Long non-coding RNA MALAT1 promotes proliferation and invasion via targeting miR-129-5p in triple-negative breast cancer. Biomed. Pharmacother. 2017, 95, 922–928. [Google Scholar] [CrossRef] [PubMed]
  86. Schneider, C.; King, R.M.; Philipson, L. Genes specifically expressed at growth arrest of mammalian cells. Cell 1988, 54, 787–793. [Google Scholar] [CrossRef]
  87. Arshi, A.; Sharifi, F.S.; Khorramian Ghahfarokhi, M.; Faghih, Z.; Doosti, A.; Ostovari, S.; Mahmoudi Maymand, E.; Ghahramani Seno, M.M. Expression Analysis of MALAT1, GAS5, SRA, and NEAT1 lncRNAs in Breast Cancer Tissues from Young Women and Women over 45 Years of Age. Mol. Nucleic Acids 2018, 12, 751–757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Zhang, Z.; Zhu, Z.; Watabe, K.; Zhang, X.; Bai, C.; Xu, M.; Wu, F.; Mo, Y.Y. Negative regulation of lncRNA GAS5 by miR-21. Cell Death Differ. 2013, 20, 1558–1568. [Google Scholar] [CrossRef] [Green Version]
  89. Li, S.; Zhou, J.; Wang, Z.; Wang, P.; Gao, X.; Wang, Y. Long noncoding RNA GAS5 suppresses triple negative breast cancer progression through inhibition of proliferation and invasion by competitively binding miR-196a-5p. Biomed. Pharm. 2018, 104, 451–457. [Google Scholar] [CrossRef]
  90. Liu, L.; Yue, H.; Liu, Q.; Yuan, J.; Li, J.; Wei, G.; Chen, X.; Lu, Y.; Guo, M.; Luo, J.; et al. LncRNA MT1JP functions as a tumor suppressor by interacting with TIAR to modulate the p53 pathway. Oncotarget 2016, 7, 15787–15800. [Google Scholar] [CrossRef] [Green Version]
  91. Xu, Y.; Zhang, G.; Zou, C.; Zhang, H.; Gong, Z.; Wang, W.; Ma, G.; Jiang, P.; Zhang, W. LncRNA MT1JP Suppresses Gastric Cancer Cell Proliferation and Migration Through MT1JP/MiR-214-3p/RUNX3 Axis. Cell. Physiol. Biochem. Int. J. Exp. Cell. Physiol. Biochem. Pharmacol. 2018, 46, 2445–2459. [Google Scholar] [CrossRef]
  92. Zhang, G.; Li, S.; Lu, J.; Ge, Y.; Wang, Q.; Ma, G.; Zhao, Q.; Wu, D.; Gong, W.; Du, M.; et al. LncRNA MT1JP functions as a ceRNA in regulating FBXW7 through competitively binding to miR-92a-3p in gastric cancer. Mol. Cancer 2018, 17, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Zhu, D.; Zhang, X.; Lin, Y.; Liang, S.; Song, Z.; Dong, C. MT1JP inhibits tumorigenesis and enhances cisplatin sensitivity of breast cancer cells through competitively binding to miR-24-3p. Am. J. Transl. Res. 2019, 11, 245–256. [Google Scholar] [PubMed]
  94. Liang, Z.; Zhu, B.; Meng, D.; Shen, X.; Li, X.; Wang, Z.; Li, L. Downregulation of lncRNA-NEF indicates poor prognosis in intrahepatic cholangiocarcinoma. Biosci. Rep. 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Chang, L.; Xu, W.; Zhang, Y.; Gong, F. Long non-coding RNA-NEF targets glucose transportation to inhibit the proliferation of non-small-cell lung cancer cells. Oncol. Lett. 2019, 17, 2795–2801. [Google Scholar] [CrossRef] [Green Version]
  96. Liang, W.C.; Ren, J.L.; Wong, C.W.; Chan, S.O.; Waye, M.M.; Fu, W.M.; Zhang, J.F. LncRNA-NEF antagonized epithelial to mesenchymal transition and cancer metastasis via cis-regulating FOXA2 and inactivating Wnt/beta-catenin signaling. Oncogene 2018, 37, 1445–1456. [Google Scholar] [CrossRef]
  97. Song, X.; Liu, Z.; Yu, Z. LncRNA NEF is downregulated in triple negative breast cancer and correlated with poor prognosis. Acta Biochim. Biophys. Sin. 2019, 51, 386–392. [Google Scholar] [CrossRef]
  98. Ke, S.; Li, R.C.; Meng, F.K.; Fang, M.H. NKILA inhibits NF-kappaB signaling and suppresses tumor metastasis. Aging 2018, 10, 56–71. [Google Scholar] [CrossRef] [Green Version]
  99. Wu, W.; Chen, F.; Cui, X.; Yang, L.; Chen, J.; Zhao, J.; Huang, D.; Liu, J.; Yang, L.; Zeng, J.; et al. LncRNA NKILA suppresses TGF-beta-induced epithelial-mesenchymal transition by blocking NF-kappaB signaling in breast cancer. Int. J. Cancer 2018, 143, 2213–2224. [Google Scholar] [CrossRef]
  100. Liu, B.; Sun, L.; Liu, Q.; Gong, C.; Yao, Y.; Lv, X.; Lin, L.; Yao, H.; Su, F.; Li, D.; et al. A cytoplasmic NF-κB interacting long noncoding RNA blocks IκB phosphorylation and suppresses breast cancer metastasis. Cancer Cell 2015, 27, 370–381. [Google Scholar] [CrossRef] [Green Version]
  101. Zhou, C.X.; Wang, X.; Yang, N.; Xue, S.K.; Li, W.C.; Xie, P.P. LncRNA LET function as a tumor suppressor in breast cancer development. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 6002–6007. [Google Scholar]
  102. Zhou, B.; Guo, H.; Tang, J. Long Non-Coding RNA TFAP2A-AS1 Inhibits Cell Proliferation and Invasion in Breast Cancer via miR-933/SMAD2. Med. Sci. Monit. 2019, 25, 1242–1253. [Google Scholar] [CrossRef] [PubMed]
  103. Beltran-Anaya, F.O.; Romero-Cordoba, S.; Rebollar-Vega, R.; Arrieta, O.; Bautista-Pina, V.; Dominguez-Reyes, C.; Villegas-Carlos, F.; Tenorio-Torres, A.; Alfaro-Riuz, L.; Jimenez-Morales, S.; et al. Expression of long non-coding RNA ENSG00000226738 (LncKLHDC7B) is enriched in the immunomodulatory triple-negative breast cancer subtype and its alteration promotes cell migration, invasion, and resistance to cell death. Mol. Oncol. 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Curtis, C.; Shah, S.; Chin, S.; Turashvili, G.; Rueda, O.; Dunning, M.; Speed, D.; Lynch, A.; Samarajiwa, S.; Yuan, Y.; et al. The genomic and transcriptomic architecture of 2,000 breast tumours reveals novel subgroups. Nature 2012, 486, 346–352. [Google Scholar] [CrossRef] [PubMed]
  105. Lánczky, A.; Nagy, Á.; Bottai, G.; Munkácsy, G.; Szabó, A.; Santarpia, L.; Győrffy, B. miRpower: A web-tool to validate survival-associated miRNAs utilizing expression data from 2178 breast cancer patients. Breast Cancer Res. Treat. 2016, 160, 439–446. [Google Scholar] [CrossRef]
  106. Anaya, J. OncoLnc: Linking TCGA survival data to mRNAs, miRNAs, and lncRNAs. Peerj Comput. Sci. 2016, 2, e67. [Google Scholar] [CrossRef] [Green Version]
  107. Kim, J.; Piao, H.; Kim, B.; Yao, F.; Han, Z.; Wang, Y.; Xiao, Z.; Siverly, A.; Lawhon, S.; Ton, B.; et al. Long noncoding RNA MALAT1 suppresses breast cancer metastasis. Nat. Genet. 2018, 50, 1705–1715. [Google Scholar] [CrossRef]
  108. Xu, S.; Sui, S.; Zhang, J.; Bai, N.; Shi, Q.; Zhang, G.; Gao, S.; You, Z.; Zhan, C.; Liu, F.; et al. Downregulation of long noncoding RNA MALAT1 induces epithelial-to-mesenchymal transition via the PI3K-AKT pathway in breast cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 4881–4891. [Google Scholar]
  109. Kwok, Z.H.; Roche, V.; Chew, X.H.; Fadieieva, A.; Tay, Y. A non-canonical tumor suppressive role for the long non-coding RNA MALAT1 in colon and breast cancers. Int. J. Cancer 2018, 143, 668–678. [Google Scholar] [CrossRef] [Green Version]
  110. Chen, Z.; Liu, S.; Tian, L.; Wu, M.; Ai, F.; Tang, W.; Zhao, L.; Ding, J.; Zhang, L.; Tang, A. miR-124 and miR-506 inhibit colorectal cancer progression by targeting DNMT3B and DNMT1. Oncotarget 2015, 6, 38139–38150. [Google Scholar] [CrossRef] [Green Version]
  111. Wang, X.X.; Guo, G.C.; Qian, X.K.; Dou, D.W.; Zhang, Z.; Xu, X.D.; Duan, X.; Pei, X.H. miR-506 attenuates methylation of lncRNA MEG3 to inhibit migration and invasion of breast cancer cell lines via targeting SP1 and SP3. Cancer Cell Int. 2018, 18, 171. [Google Scholar] [CrossRef] [Green Version]
  112. Wen, S.Y.; Lin, Y.; Yu, Y.Q.; Cao, S.J.; Zhang, R.; Yang, X.M.; Li, J.; Zhang, Y.L.; Wang, Y.H.; Ma, M.Z.; et al. miR-506 acts as a tumor suppressor by directly targeting the hedgehog pathway transcription factor Gli3 in human cervical cancer. Oncogene 2015, 34, 717–725. [Google Scholar] [CrossRef] [PubMed]
  113. Peng, T.; Zhou, L.; Zuo, L.; Luan, Y. MiR-506 functions as a tumor suppressor in glioma by targeting STAT3. Oncol. Rep. 2016, 35, 1057–1064. [Google Scholar] [CrossRef] [PubMed]
  114. Li, J.; Bian, E.B.; He, X.J.; Ma, C.C.; Zong, G.; Wang, H.L.; Zhao, B. Epigenetic repression of long non-coding RNA MEG3 mediated by DNMT1 represses the p53 pathway in gliomas. Int. J. Oncol. 2016, 48, 723–733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Jiang, Y.; Lin, L.; Zhong, S.; Cai, Y.; Zhang, F.; Wang, X.; Miao, R.; Zhang, B.; Gao, S.; Hu, X. Overexpression of novel lncRNA NLIPMT inhibits metastasis by reducing phosphorylated glycogen synthase kinase 3beta in breast cancer. J. Cell. Physiol. 2018. [Google Scholar] [CrossRef]
  116. Armanious, H.; Deschenes, J.; Gelebart, P.; Ghosh, S.; Mackey, J.; Lai, R. Clinical and biological significance of GSK-3β inactivation in breast cancer-an immunohistochemical study. Hum. Pathol. 2010, 41, 1657–1663. [Google Scholar] [CrossRef]
  117. Dembowy, J.; Adissu, H.; Liu, J.; Zacksenhaus, E.; Woodgett, J. Effect of glycogen synthase kinase-3 inactivation on mouse mammary gland development and oncogenesis. Oncogene 2015, 34, 3514–3526. [Google Scholar] [CrossRef] [Green Version]
  118. Farago, M.; Dominguez, I.; Landesman-Bollag, E.; Xu, X.; Rosner, A.; Cardiff, R.; Seldin, D. Kinase-inactive glycogen synthase kinase 3beta promotes Wnt signaling and mammary tumorigenesis. Cancer Res. 2005, 65, 5792–5801. [Google Scholar] [CrossRef] [Green Version]
  119. Shakoori, A.; Mai, W.; Miyashita, K.; Yasumoto, K.; Takahashi, Y.; Ooi, A.; Kawakami, K.; Minamoto, T. Inhibition of GSK-3 beta activity attenuates proliferation of human colon cancer cells in rodents. Cancer Sci. 2007, 98, 1388–1393. [Google Scholar] [CrossRef] [Green Version]
  120. Bilim, V.; Ougolkov, A.; Yuuki, K.; Naito, S.; Kawazoe, H.; Muto, A.; Oya, M.; Billadeau, D.; Motoyama, T.; Tomita, Y. Glycogen synthase kinase-3: A new therapeutic target in renal cell carcinoma. Br. J. Cancer 2009, 101, 2005–2014. [Google Scholar] [CrossRef] [Green Version]
  121. Xu, R.; Zhu, X.; Chen, F.; Huang, C.; Ai, K.; Wu, H.; Zhang, L.; Zhao, X. LncRNA XIST/miR-200c regulates the stemness properties and tumourigenicity of human bladder cancer stem cell-like cells. Cancer Cell Int. 2018, 18, 41. [Google Scholar] [CrossRef]
  122. Tang, Y.; He, R.; An, J.; Deng, P.; Huang, L.; Yang, W. lncRNA XIST interacts with miR-140 to modulate lung cancer growth by targeting iASPP. Oncol. Rep. 2017, 38, 941–948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Zheng, R.; Lin, S.; Guan, L.; Yuan, H.; Liu, K.; Liu, C.; Ye, W.; Liao, Y.; Jia, J.; Zhang, R. Long non-coding RNA XIST inhibited breast cancer cell growth, migration, and invasion via miR-155/CDX1 axis. Biochem. Biophys. Res. Commun. 2018, 498, 1002–1008. [Google Scholar] [CrossRef] [PubMed]
  124. Hashimoto, S.; Amaya, F.; Matsuyama, H.; Ueno, H.; Kikuchi, S.; Tanaka, M.; Watanabe, Y.; Ebina, M.; Ishizaka, A.; Tsukita, S.; et al. Dysregulation of lung injury and repair in moesin-deficient mice treated with intratracheal bleomycin. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 295, L566–L574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Xing, F.; Liu, Y.; Wu, S.Y.; Wu, K.; Sharma, S.; Mo, Y.Y.; Feng, J.; Sanders, S.; Jin, G.; Singh, R.; et al. Loss of XIST in Breast Cancer Activates MSN-c-Met and Reprograms Microglia via Exosomal miRNA to Promote Brain Metastasis. Cancer Res. 2018, 78, 4316–4330. [Google Scholar] [CrossRef] [Green Version]
  126. Salter, M.W.; Stevens, B. Microglia emerge as central players in brain disease. Nat. Med. 2017, 23, 1018–1027. [Google Scholar] [CrossRef]
  127. Reya, T.; Morrison, S.J.; Clarke, M.F.; Weissman, I.L. Stem cells, cancer, and cancer stem cells. Nature 2001, 414, 105–111. [Google Scholar] [CrossRef] [Green Version]
  128. Adorno-Cruz, V.; Kibria, G.; Liu, X.; Doherty, M.; Junk, D.J.; Guan, D.; Hubert, C.; Venere, M.; Mulkearns-Hubert, E.; Sinyuk, M.; et al. Cancer stem cells: Targeting the roots of cancer, seeds of metastasis, and sources of therapy resistance. Cancer Res. 2015, 75, 924–929. [Google Scholar] [CrossRef] [Green Version]
  129. Baccelli, I.; Trumpp, A. The evolving concept of cancer and metastasis stem cells. J. Cell Biol. 2012, 198, 281–293. [Google Scholar] [CrossRef]
  130. Antoniou, A.; Hebrant, A.; Dom, G.; Dumont, J.E.; Maenhaut, C. Cancer stem cells, a fuzzy evolving concept: A cell population or a cell property? Cell Cycle 2013, 12, 3743–3748. [Google Scholar] [CrossRef] [Green Version]
  131. Castro-Oropeza, R.; Melendez-Zajgla, J.; Maldonado, V.; Vazquez-Santillan, K. The emerging role of lncRNAs in the regulation of cancer stem cells. Cell. Oncol. 2018, 41, 585–603. [Google Scholar] [CrossRef]
  132. Mani, S.; Guo, W.; Liao, M.; Eaton, E.; Ayyanan, A.; Zhou, A.; Brooks, M.; Reinhard, F.; Zhang, C.; Shipitsin, M.; et al. The epithelial-mesenchymal transition generates cells with properties of stem cells. Cell 2008, 133, 704–715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Hardin, H.; Zhang, R.; Helein, H.; Buehler, D.; Guo, Z.; Lloyd, R.V. The evolving concept of cancer stem-like cells in thyroid cancer and other solid tumors. Lab. Investig. 2017, 97, 1142–1151. [Google Scholar] [CrossRef] [PubMed]
  134. Li, H.; Zhu, L.; Xu, L.; Qin, K.; Liu, C.; Yu, Y.; Su, D.; Wu, K.; Sheng, Y. Long noncoding RNA linc00617 exhibits oncogenic activity in breast cancer. Mol. Carcinog. 2017, 56, 3–17. [Google Scholar] [CrossRef] [PubMed]
  135. Keshavarz, M.; Asadi, M.H. Long non-coding RNA ES1 controls the proliferation of breast cancer cells by regulating the Oct4/Sox2/miR-302 axis. FEBS J. 2019, 286, 2611–2623. [Google Scholar] [CrossRef]
  136. Zhou, M.; Hou, Y.; Yang, G.; Zhang, H.; Tu, G.; Du, Y.E.; Wen, S.; Xu, L.; Tang, X.; Tang, S.; et al. LncRNA-Hh Strengthen Cancer Stem Cells Generation in Twist-Positive Breast Cancer via Activation of Hedgehog Signaling Pathway. Stem Cells 2016, 34, 55–66. [Google Scholar] [CrossRef] [Green Version]
  137. Deng, J.; Yang, M.; Jiang, R.; An, N.; Wang, X.; Liu, B. Long Non-Coding RNA HOTAIR Regulates the Proliferation, Self-Renewal Capacity, Tumor Formation and Migration of the Cancer Stem-Like Cell (CSC) Subpopulation Enriched from Breast Cancer Cells. PLoS ONE 2017, 12, e0170860. [Google Scholar] [CrossRef] [Green Version]
  138. Zhang, Z.; Sun, L.; Zhang, Y.; Lu, G.; Li, Y.; Wei, Z. Long non-coding RNA FEZF1-AS1 promotes breast cancer stemness and tumorigenesis via targeting miR-30a/Nanog axis. J. Cell. Physiol. 2018, 233, 8630–8638. [Google Scholar] [CrossRef]
  139. Lu, G.; Li, Y.; Ma, Y.; Lu, J.; Chen, Y.; Jiang, Q.; Qin, Q.; Zhao, L.; Huang, Q.; Luo, Z.; et al. Long noncoding RNA LINC00511 contributes to breast cancer tumourigenesis and stemness by inducing the miR-185-3p/E2F1/Nanog axis. J. Exp. Clin. Cancer Res. 2018, 37, 289. [Google Scholar] [CrossRef]
  140. Huang, H.; Weng, H.; Sun, W.; Qin, X.; Shi, H.; Wu, H.; Zhao, B.S.; Mesquita, A.; Liu, C.; Yuan, C.L.; et al. Recognition of RNA N(6)-methyladenosine by IGF2BP proteins enhances mRNA stability and translation. Nat. Cell Biol. 2018, 20, 285–295. [Google Scholar] [CrossRef]
  141. Ma, F.; Liu, X.; Zhou, S.; Li, W.; Liu, C.; Chadwick, M.; Qian, C. Long non-coding RNA FGF13-AS1 inhibits glycolysis and stemness properties of breast cancer cells through FGF13-AS1/IGF2BPs/Myc feedback loop. Cancer Lett. 2019, 450, 63–75. [Google Scholar] [CrossRef]
  142. Zhang, S.; Hu, X.; Zhang, W.; Tam, K. Unexpected Discovery of Dichloroacetate Derived Adenosine Triphosphate Competitors Targeting Pyruvate Dehydrogenase Kinase To Inhibit Cancer Proliferation. J. Med. Chem. 2016, 59, 3562–3568. [Google Scholar] [CrossRef] [PubMed]
  143. Peng, F.; Wang, J.H.; Fan, W.J.; Meng, Y.T.; Li, M.M.; Li, T.T.; Cui, B.; Wang, H.F.; Zhao, Y.; An, F.; et al. Glycolysis gatekeeper PDK1 reprograms breast cancer stem cells under hypoxia. Oncogene 2018, 37, 1062–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Dupuy, F.; Tabariès, S.; Andrzejewski, S.; Dong, Z.; Blagih, J.; Annis, M.; Omeroglu, A.; Gao, D.; Leung, S.; Amir, E.; et al. PDK1-Dependent Metabolic Reprogramming Dictates Metastatic Potential in Breast Cancer. Cell Metab. 2015, 22, 577–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Velpula, K.; Bhasin, A.; Asuthkar, S.; Tsung, A. Combined targeting of PDK1 and EGFR triggers regression of glioblastoma by reversing the Warburg effect. Cancer Res. 2013, 73, 7277–7289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Li, C.; Shan, S.; Huang, Q.; Braun, R.; Lanzen, J.; Hu, K.; Lin, P.; Dewhirst, M. Initial stages of tumor cell-induced angiogenesis: Evaluation via skin window chambers in rodent models. J. Natl. Cancer Inst. 2000, 92, 143–147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Johnson, K.E.; Wilgus, T.A. Vascular Endothelial Growth Factor and Angiogenesis in the Regulation of Cutaneous Wound Repair. Adv. Wound Care 2014, 3, 647–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Huang, X.J.; Xia, Y.; He, G.F.; Zheng, L.L.; Cai, Y.P.; Yin, Y.; Wu, Q. MALAT1 promotes angiogenesis of breast cancer. Oncol. Rep. 2018, 40, 2683–2689. [Google Scholar] [CrossRef] [Green Version]
  149. Al-Rugeebah, A.; Alanazi, M.; Parine, N.R. MEG3: An Oncogenic Long Non-coding RNA in Different Cancers. Pathol. Oncol. Res. 2019, 25, 859–874. [Google Scholar] [CrossRef]
  150. Zhang, C.Y.; Yu, M.S.; Li, X.; Zhang, Z.; Han, C.R.; Yan, B. Overexpression of long non-coding RNA MEG3 suppresses breast cancer cell proliferation, invasion, and angiogenesis through AKT pathway. Tumour Biol. 2017, 39. [Google Scholar] [CrossRef] [Green Version]
  151. Senkus, E.; Kyriakides, S.; Ohno, S.; Penault-Llorca, F.; Poortmans, P.; Rutgers, E.; Zackrisson, S.; Cardoso, F. Primary breast cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2015, 26 (Suppl. 5), v8–v30. [Google Scholar] [CrossRef]
  152. Slamon, D.; Neven, P.; Chia, S.; Fasching, P.; De Laurentiis, M.; Im, S.; Petrakova, K.; Bianchi, G.; Esteva, F.; Martín, M.; et al. Phase III Randomized Study of Ribociclib and Fulvestrant in Hormone Receptor-Positive, Human Epidermal Growth Factor Receptor 2-Negative Advanced Breast Cancer: MONALEESA-3. J. Clin. Oncol. 2018, 36, 2465–2472. [Google Scholar] [CrossRef] [PubMed]
  153. Vogel, C.; Cobleigh, M.; Tripathy, D.; Gutheil, J.; Harris, L.; Fehrenbacher, L.; Slamon, D.; Murphy, M.; Novotny, W.; Burchmore, M.; et al. Efficacy and safety of trastuzumab as a single agent in first-line treatment of HER2-overexpressing metastatic breast cancer. J. Clin. Oncol. 2002, 20, 719–726. [Google Scholar] [CrossRef] [PubMed]
  154. Collins, D.; Conlon, N.; Kannan, S.; Verma, C.; Eli, L.; Lalani, A.; Crown, J. Preclinical Characteristics of the Irreversible Pan-HER Kinase Inhibitor Neratinib Compared with Lapatinib: Implications for the Treatment of HER2-Positive and -Mutated Breast Cancer. Cancers 2019, 11, 737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Kim, S.; Dent, R.; Im, S.; Espié, M.; Blau, S.; Tan, A.; Isakoff, S.; Oliveira, M.; Saura, C.; Wongchenko, M.; et al. Ipatasertib plus paclitaxel versus placebo plus paclitaxel as first-line therapy for metastatic triple-negative breast cancer (LOTUS): A multicentre, randomised, double-blind, placebo-controlled, phase 2 trial. Lancet Oncol. 2017, 18, 1360–1372. [Google Scholar] [CrossRef]
  156. Pan, H.; Gray, R.; Braybrooke, J.; Davies, C.; Taylor, C.; McGale, P.; Peto, R.; Pritchard, K.; Bergh, J.; Dowsett, M.; et al. 20-Year Risks of Breast-Cancer Recurrence after Stopping Endocrine Therapy at 5 Years. N. Engl. J. Med. 2017, 377, 1836–1846. [Google Scholar] [CrossRef] [Green Version]
  157. Dong, H.; Hu, J.; Zou, K.; Ye, M.; Chen, Y.; Wu, C.; Chen, X.; Han, M. Activation of LncRNA TINCR by H3K27 acetylation promotes Trastuzumab resistance and epithelial-mesenchymal transition by targeting MicroRNA-125b in breast Cancer. Mol. Cancer 2019, 18, 3. [Google Scholar] [CrossRef]
  158. Wu, Y.; Sarkissyan, M.; Vadgama, J.V. Epithelial-Mesenchymal Transition and Breast Cancer. J. Clin. Med. 2016, 5, 13. [Google Scholar] [CrossRef] [Green Version]
  159. Li, X.; Wang, S.; Li, Z.; Long, X.; Guo, Z.; Zhang, G.; Zu, J.; Chen, Y.; Wen, L. Retracted: NEAT1 induces epithelial-mesenchymal transition and 5-FU resistance through the miR-129/ZEB2 axis in breast cancer. FEBS Lett. 2016. [Google Scholar] [CrossRef]
  160. Li, X.; Wang, S.; Li, Z.; Long, X.; Guo, Z.; Zhang, G.; Zu, J.; Chen, Y.; Wen, L. The lncRNA NEAT1 facilitates cell growth and invasion via the miR-211/HMGA2 axis in breast cancer. Int. J. Biol. Macromol. 2017, 105, 346–353. [Google Scholar] [CrossRef]
  161. Hu, X.; Hou, P.; Li, T.; Quan, H.; Li, M.; Lin, T.; Liu, J.; Bai, J.; Zheng, J. The roles of Wnt/β-catenin signaling pathway related lncRNAs in cancer. Int. J. Biol. Sci. 2018, 14, 2003–2011. [Google Scholar] [CrossRef]
  162. Sun, X.; Huang, T.; Zhang, C.; Zhang, S.; Wang, Y.; Zhang, Q.; Liu, Z. Long non-coding RNA LINC00968 reduces cell proliferation and migration and angiogenesis in breast cancer through up-regulation of PROX1 by reducing hsa-miR-423-5p. Cell Cycle 2019, 18, 1908–1924. [Google Scholar] [CrossRef] [PubMed]
  163. Xiu, D.; Liu, G.; Yu, S.; Li, L.; Zhao, G.; Liu, L.; Li, X. Long non-coding RNA LINC00968 attenuates drug resistance of breast cancer cells through inhibiting the Wnt2/β-catenin signaling pathway by regulating WNT2. J. Exp. Clin. Cancer Res. CR 2019, 38, 94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Liang, Y.; Li, Y.; Song, X.; Zhang, N.; Sang, Y.; Zhang, H.; Liu, Y.; Chen, B.; Zhao, W.; Wang, L.; et al. Long noncoding RNA LINP1 acts as an oncogene and promotes chemoresistance in breast cancer. Cancer Biol. Ther. 2018, 19, 120–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Longley, D.B.; Harkin, D.P.; Johnston, P.G. 5-fluorouracil: Mechanisms of action and clinical strategies. Nat. Rev. Cancer 2003, 3, 330–338. [Google Scholar] [CrossRef]
  166. Tacar, O.; Sriamornsak, P.; Dass, C.R. Doxorubicin: An update on anticancer molecular action, toxicity and novel drug delivery systems. J. Pharm. Pharmacol. 2013, 65, 157–170. [Google Scholar] [CrossRef]
  167. The Impact of lncRNA Dysregulation on Clinicopathology and Survival of Breast Cancer: A Systematic Review and Meta-analysis. Mol. Ther. Nucleic Acids 2018, 12, 359–369. [CrossRef] [Green Version]
  168. Sun, M.; Wu, D.; Zhou, K.; Li, H.; Gong, X.; Wei, Q.; Du, M.; Lei, P.; Zha, J.; Zhu, H. An eight-lncRNA signature predicts survival of breast cancer patients: A comprehensive study based on weighted gene co-expression network analysis and competing endogenous RNA network. Breast Cancer Res. Treat. 2019, 175, 59–75. [Google Scholar] [CrossRef]
  169. Zhang, K.; Luo, Z.; Zhang, Y.; Zhang, L.; Wu, L.; Liu, L.; Yang, J.; Song, X.; Liu, J. Circulating lncRNA H19 in plasma as a novel biomarker for breast cancer. Cancer Biomark. 2016, 17, 187–194. [Google Scholar] [CrossRef]
  170. Han, L.; Ma, P.; Liu, S.M.; Zhou, X. Circulating long noncoding RNA GAS5 as a potential biomarker in breast cancer for assessing the surgical effects. Tumour Biol. 2016, 37, 6847–6854. [Google Scholar] [CrossRef]
  171. Bai, J.; Zhao, W.Y.; Li, W.J.; Ying, Z.W.; Jiang, D.Q. Long noncoding RNA LINC00473 indicates a poor prognosis of breast cancer and accelerates tumor carcinogenesis by competing endogenous sponging miR-497. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 3410–3420. [Google Scholar]
  172. Liu, X.M.; Yang, B.; Han, J. Increased long noncoding RNA LINP1 expression and its prognostic significance in human breast cancer. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 8749–8754. [Google Scholar] [PubMed]
  173. Meng, Y.; Liu, Y.L.; Li, K.; Fu, T. Prognostic value of long non-coding RNA breast cancer anti-estrogen resistance 4 in human cancers: A meta-analysis. Medicine 2019, 98, e15793. [Google Scholar] [CrossRef] [PubMed]
  174. Zhao, W.; Wang, Z.; Fang, X.; Li, N.; Fang, J. Long noncoding RNA Breast cancer antiestrogen resistance 4 is associated with cancer progression and its significant prognostic value. J. Cell. Physiol. 2019, 234, 12956–12963. [Google Scholar] [CrossRef] [PubMed]
  175. Lv, M.; Xu, P.; Wu, Y.; Huang, L.; Li, W.; Lv, S.; Wu, X.; Zeng, X.; Shen, R.; Jia, X.; et al. LncRNAs as new biomarkers to differentiate triple negative breast cancer from non-triple negative breast cancer. Oncotarget 2016, 7, 13047–13059. [Google Scholar] [CrossRef] [Green Version]
  176. Zhang, H.; Zhang, N.; Liu, Y.; Su, P.; Liang, Y.; Li, Y.; Wang, X.; Chen, T.; Song, X.; Sang, Y.; et al. Epigenetic Regulation of NAMPT by NAMPT-AS Drives Metastatic Progression in Triple-Negative Breast Cancer. Cancer Res. 2019, 79, 3347–3359. [Google Scholar] [CrossRef] [Green Version]
  177. Li, R.H.; Chen, M.; Liu, J.; Shao, C.C.; Guo, C.P.; Wei, X.L.; Li, Y.C.; Huang, W.H.; Zhang, G.J. Long noncoding RNA ATB promotes the epithelial-mesenchymal transition by upregulating the miR-200c/Twist1 axe and predicts poor prognosis in breast cancer. Cell Death Dis. 2018, 9, 1171. [Google Scholar] [CrossRef]
  178. Fu, J.; Dong, G.; Shi, H.; Zhang, J.; Ning, Z.; Bao, X.; Liu, C.; Hu, J.; Liu, M.; Xiong, B. LncRNA MIR503HG inhibits cell migration and invasion via miR-103/OLFM4 axis in triple negative breast cancer. J. Cell. Mol. Med. 2019, 23, 4738–4745. [Google Scholar] [CrossRef] [Green Version]
  179. Shore, A.N.; Rosen, J.M. Regulation of mammary epithelial cell homeostasis by lncRNAs. Int. J. Biochem. Cell Biol. 2014, 54, 318–330. [Google Scholar] [CrossRef] [Green Version]
  180. Chen, M.; Zhuang, C.; Liu, Y.; Li, J.; Dai, F.; Xia, M.; Zhan, Y.; Lin, J.; Chen, Z.; He, A.; et al. Tetracycline-inducible shRNA targeting antisense long non-coding RNA HIF1A-AS2 represses the malignant phenotypes of bladder cancer. Cancer Lett. 2016, 376, 155–164. [Google Scholar] [CrossRef] [Green Version]
  181. Campbell, P.; Yachida, S.; Mudie, L.; Stephens, P.; Pleasance, E.; Stebbings, L.; Morsberger, L.; Latimer, C.; McLaren, S.; Lin, M.; et al. The patterns and dynamics of genomic instability in metastatic pancreatic cancer. Nature 2010, 467, 1109–1113. [Google Scholar] [CrossRef] [Green Version]
  182. Wu, X.; Northcott, P.; Dubuc, A.; Dupuy, A.; Shih, D.; Witt, H.; Croul, S.; Bouffet, E.; Fults, D.; Eberhart, C.; et al. Clonal selection drives genetic divergence of metastatic medulloblastoma. Nature 2012, 482, 529–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Mizushima, N.; Yoshimori, T.; Ohsumi, Y. The role of Atg proteins in autophagosome formation. Annu. Rev. Cell. Dev. Biol. 2011, 27, 107–132. [Google Scholar] [CrossRef] [PubMed]
  184. Kenific, C.M.; Thorburn, A.; Debnath, J. Autophagy and metastasis: Another double-edged sword. Curr. Opin. Cell Biol. 2010, 22, 241–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Moraes, L.A.; Kar, S.; Foo, S.L.; Gu, T.; Toh, Y.Q.; Ampomah, P.B.; Sachaphibulkij, K.; Yap, G.; Zharkova, O.; Lukman, H.M.; et al. Annexin-A1 enhances breast cancer growth and migration by promoting alternative macrophage polarization in the tumour microenvironment. Sci. Rep. 2017, 7, 17925. [Google Scholar] [CrossRef] [PubMed]
  186. Gu, J.; Wang, Y.; Wang, X.; Zhou, D.; Wang, X.; Zhou, M.; He, Z. Effect of the LncRNA GAS5-MiR-23a-ATG3 Axis in Regulating Autophagy in Patients with Breast Cancer. Cell. Physiol. Biochem. 2018, 48, 194–207. [Google Scholar] [CrossRef]
  187. Ruiz, A.; Rockfield, S.; Taran, N.; Haller, E.; Engelman, R.; Flores, I.; Panina-Bordignon, P.; Nanjundan, M. Effect of hydroxychloroquine and characterization of autophagy in a mouse model of endometriosis. Cell Death Dis. 2016, 7, e2059. [Google Scholar] [CrossRef]
  188. Cao, Z.; Livas, T.; Kyprianou, N. Anoikis and EMT: Lethal “Liaisons” during Cancer Progression. Crit. Rev. Oncog. 2016, 21, 155–168. [Google Scholar] [CrossRef] [Green Version]
  189. Lu, J.; Tan, M.; Cai, Q. The Warburg effect in tumor progression: Mitochondrial oxidative metabolism as an anti-metastasis mechanism. Cancer Lett. 2015, 356, 156–164. [Google Scholar] [CrossRef] [Green Version]
  190. Liu, X.; Fu, Q.; Li, S.; Liang, N.; Li, F.; Li, C.; Sui, C.; Dionigi, G.; Sun, H. LncRNA FOXD2-AS1 Functions as a Competing Endogenous RNA to Regulate TERT Expression by Sponging miR-7-5p in Thyroid Cancer. Front. Endocrinol. 2019, 10, 207. [Google Scholar] [CrossRef]
  191. Gordon, M.; Babbs, B.; Cochrane, D.; Bitler, B.; Richer, J. The long non-coding RNA MALAT1 promotes ovarian cancer progression by regulating RBFOX2-mediated alternative splicing. Mol. Carcinog. 2019, 58, 196–205. [Google Scholar] [CrossRef]
  192. Seitz, A.; Christensen, L.; Christensen, E.; Faarkrog, K.; Ostenfeld, M.; Hedegaard, J.; Nordentoft, I.; Nielsen, M.; Palmfeldt, J.; Thomson, M.; et al. Profiling of long non-coding RNAs identifies LINC00958 and LINC01296 as candidate oncogenes in bladder cancer. Sci. Rep. 2017, 7, 395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Dai, X.; Li, T.; Bai, Z.; Yang, Y.; Liu, X.; Zhan, J.; Shi, B. Breast cancer intrinsic subtype classification, clinical use and future trends. Am. J. Cancer Res. 2015, 5, 2929–2943. [Google Scholar] [PubMed]
  194. Heimes, A.S.; Madjar, K.; Edlund, K.; Battista, M.J.; Almstedt, K.; Elger, T.; Krajnak, S.; Rahnenführer, J.; Brenner, W.; Hasenburg, A. Subtype-specific prognostic impact of different immune signatures in node-negative breast cancer. Breast Cancer Res. Treat. 2017, 14, 1–8. [Google Scholar] [CrossRef] [PubMed]
  195. Mathias, C.; Zambalde, E.; Rask, P.; Gradia, D.; de Oliveira, J. Long non-coding RNAs differential expression in breast cancer subtypes: What do we know? Clin. Genet. 2019, 95, 558–568. [Google Scholar] [CrossRef]
  196. Bhan, A.; Hussain, I.; Ansari, K.I.; Kasiri, S.; Bashyal, A.; Mandal, S.S. Antisense Transcript Long Noncoding RNA (lncRNA) HOTAIR is Transcriptionally Induced by Estradiol. J. Mol. Biol. 2013, 425, 3707–3722. [Google Scholar] [CrossRef] [Green Version]
  197. Yang, F.; Lyu, S.; Dong, S.; Liu, Y.; Zhang, X.; Wang, O. Expression profile analysis of long noncoding RNA in HER-2-enriched subtype breast cancer by next-generation sequencing and bioinformatics. Onco Targets Ther. 2016, 9, 761–772. [Google Scholar] [CrossRef] [Green Version]
  198. Lehmann, B.; Bauer, J.; Chen, X.; Sanders, M.; Chakravarthy, A.; Shyr, Y.; Pietenpol, J. Identification of human triple-negative breast cancer subtypes and preclinical models for selection of targeted therapies. J. Clin. Investig. 2011, 121, 2750–2767. [Google Scholar] [CrossRef] [Green Version]
  199. Liu, E.; Liu, Z.; Zhou, Y. Carboplatin-docetaxel-induced activity against ovarian cancer is dependent on up-regulated lncRNA PVT1. Int. J. Clin. Exp. Pathol. 2014, 8, 3803. [Google Scholar]
  200. Martin, D.; Smith, C.L. Ligand-independent interactions of p160/steroid receptor coactivators and CREB-binding protein (CBP) with estrogen receptor-alpha: Regulation by phosphorylation sites in the A/B region depends on other receptor domains. Mol. Endocrinol. 2003, 17, 1296–1314. [Google Scholar]
  201. Zhao, Z.; Chen, C.; Liu, Y.; Wu, C. 17β-Estradiol treatment inhibits breast cell proliferation, migration and invasion by decreasing MALAT-1 RNA level. Biochem. Biophys. Res. Commun. 2014, 445, 388–393. [Google Scholar] [CrossRef]
  202. Gupta, P.B.; Pastushenko, I.; Skibinski, A.; Blanpain, C.; Kuperwasser, C. Phenotypic Plasticity: Driver of Cancer Initiation, Progression, and Therapy Resistance. Cell Stem Cell 2019, 24, 65–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Pastushenko, I.; Blanpain, C. EMT Transition States during Tumor Progression and Metastasis. Trends Cell Biol. 2019, 29, 212–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Norouzi, S.; Gorgi Valokala, M.; Mosaffa, F.; Zirak, M.; Zamani, P.; Behravan, J. Crosstalk in cancer resistance and metastasis. Crit. Rev. Oncol. Hematol. 2018, 132, 145–153. [Google Scholar] [CrossRef] [PubMed]
  205. Lawson, D.; Kessenbrock, K.; Davis, R.; Pervolarakis, N.; Werb, Z. Tumour heterogeneity and metastasis at single-cell resolution. Nat. Cell Biol. 2018, 20, 1349–1360. [Google Scholar] [CrossRef]
  206. Wang, Q.; Fan, H.; Liu, Y.; Yin, Z.; Cai, H.; Liu, J.; Wang, Z.; Shao, M.; Sun, X.; Diao, J.; et al. Curcumin enhances the radiosensitivity in nasopharyngeal carcinoma cells involving the reversal of differentially expressed long non-coding RNAs. Int. J. Oncol. 2014, 44, 858–864. [Google Scholar] [CrossRef]
  207. Hu, X.; Ding, D.; Zhang, J.; Cui, J. Knockdown of lncRNA HOTAIR sensitizes breast cancer cells to ionizing radiation through activating miR-218. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef] [Green Version]
  208. Zhang, Y.; He, Q.; Hu, Z.; Feng, Y.; Fan, L.; Tang, Z.; Yuan, J.; Shan, W.; Li, C.; Hu, X.; et al. Long noncoding RNA LINP1 regulates repair of DNA double-strand breaks in triple-negative breast cancer. Nat. Struct. Mol. Biol. 2016, 23, 522–530. [Google Scholar] [CrossRef]
Figure 1. LncRNAs and prognosis of BC patients. Here both up-regulated lncRNA in BC tissue (HOTAIR, TINCR, LIP1, MALAT1, and LINC000473) and circulating lncRNA in plasma (HOTAIR, H19, and GAS5) predict poor prognosis of BC patients. The red or white up arrows respectively refer to up-regulation of the relevant lncRNA in BC patients’ tissue or plasma.
Figure 1. LncRNAs and prognosis of BC patients. Here both up-regulated lncRNA in BC tissue (HOTAIR, TINCR, LIP1, MALAT1, and LINC000473) and circulating lncRNA in plasma (HOTAIR, H19, and GAS5) predict poor prognosis of BC patients. The red or white up arrows respectively refer to up-regulation of the relevant lncRNA in BC patients’ tissue or plasma.
Cancers 11 01802 g001
Figure 2. LncRNAs as biomarkers for diagnosis and prognosis of TNBC patients. Up-regulated expression of DANCR, NAMPT-AS, ATB, and HIF1A-AS2 predict poor prognosis of TNBC patients. Down-regulated expression of MIR503HG predicts poor prognosis of TNBC patients. Red up arrows refer to up-regulation of the relevant lncRNA. Blue down arrows refer to down-regulation of the relevant lncRNA.
Figure 2. LncRNAs as biomarkers for diagnosis and prognosis of TNBC patients. Up-regulated expression of DANCR, NAMPT-AS, ATB, and HIF1A-AS2 predict poor prognosis of TNBC patients. Down-regulated expression of MIR503HG predicts poor prognosis of TNBC patients. Red up arrows refer to up-regulation of the relevant lncRNA. Blue down arrows refer to down-regulation of the relevant lncRNA.
Cancers 11 01802 g002
Table 1. Dual effects of lncRNAs in various stage of BC progression.
Table 1. Dual effects of lncRNAs in various stage of BC progression.
FunctionLncRNAEnsembl IDSubcellular LocalizationSignaling Axis and PathwayReferences
Promoteinvasion and migration of BC cellsDANCRENSG00000226950CytoplasmDANCR/EZH2/ABCG2[27]
DANCR/miRNA-216a-5p[30]
H19ENSG00000130600CytoplasmH19/miR-152/DNMT[36]
HOTAIRENSG00000228630Cytoplasm and nucleusHOTAIR/P53/AKT/JNK/MMPs[44]
HOTAIR/AKT/mTOR[45]
HOTAIR/miR-20a-5p/HMGA2[46]
LINC00152ENSG00000222041CytoplasmLINC00152/BRCA1/PTEN[48]
LINC00461ENSG00000245526NucleusLINC00461/miR-30a-5p/integrin β3[53]
NEAT1ENSG00000245532NucleusNEAT1/miR-448/ZEB1[58]
LINC01857ENSG00000224137NucleusLINC01857/CREBBP/H3K27Ac and CREB1[60]
distant metastasis of BC cellsH19ENSG00000130600CytoplasmH19/let-7b/CYTH3/ARF[63]
HOTAIRENSG00000228630Cytoplasm and nucleusTGF-β1/CDK5/HOTAIR/H3K27[69]
HIF1A-AS2ENSG00000258667Cytoplasmunclear[180]
RP1ENST00000420172CytoplasmKLF5/RP1/p27kip1[76]
MALAT1ENSG00000251562NucleusMALAT1/miR-1/slug[80]
MALAT1/miR-129-5p[85]
stemness of BC cellsTUNARENSG00000250366CytoplasmTUNAR/SOX2[134]
LINC00511ENSG00000227036CytoplasmLINC00511/miR-185-3p/E2F1/Nanog[139]
lncRNA-Hh lncRNA-Hh/GAS1/SOX2/OCT4[136]
FEZF1-AS1ENSG00000230316NucleusFEZF1-AS1/miR-30a/Nanog[138]
FGF13-AS1ENSG00000226031NucleusFGF13-AS1/IGF2BPs/Myc[140]
ES1ENSG00000226673NucleusES1/Oct4/Sox2/miR-302[135]
RP1ENST00000420172CytoplasmRP1/p-4E-BP1/eIF4E/p27kip1/Snail1[76]
H19ENSG00000130600CytoplasmH19/let-7/HIF-1/PDK1[143]
HOTAIRENSG00000228630Cytoplasm and nucleusHOTAIR/miR-34a/SOX2[137]
BC angiogenesisMALAT1ENSG00000251562NucleusMALAT1/miR-145/VEGF[148]
Inhibitinvasion and migration of BC cellsGAS5ENSG00000234741Cytoplasm and nucleusGAS5/miR-21/RISC[88]
GAS5/miR-196a-5p/FOXO1/PI3K/AKT[89]
MT1JPENSG00000255986CytoplasmMT1JP/miR-24-3p/Wnt/β-catenin[161]
NKILAENSG00000278709CytoplasmNKILA/NF-ĸB/MMPs[100]
NEF NEF/miRNA-155[97]
LET unclear[101]
TFAP2A-AS1 TFAP2A-AS1/miR-933/SMAD2[102]
LncKLHDC7B unclear[103]
distant metastasis of BC cellsMALAT1ENSG00000251562NucleusPTEN/microRNA/MALAT1[109]
MALAT1/YAP-TEAD[107]
MEG3ENSG00000214548NucleusmiR-506/SP3/SP1/DNMT1/MEG3[111]
NLIPMT CytoplasmNLIPMT/GSK3β and EMT proteins[115]
XISTENSG00000229807NucleusXIST/MSN/c-Met[125]
XIST/miR-155/CDX1[123]
stemness of BC cellsXISTENSG00000229807NucleusXIST/MSN/EMT proteins[125]
BC angiogenesisMEG3ENSG00000214548NucleusMEG3/AKT signaling[150]

Share and Cite

MDPI and ACS Style

Huang, Q.-Y.; Liu, G.-F.; Qian, X.-L.; Tang, L.-B.; Huang, Q.-Y.; Xiong, L.-X. Long Non-Coding RNA: Dual Effects on Breast Cancer Metastasis and Clinical Applications. Cancers 2019, 11, 1802. https://doi.org/10.3390/cancers11111802

AMA Style

Huang Q-Y, Liu G-F, Qian X-L, Tang L-B, Huang Q-Y, Xiong L-X. Long Non-Coding RNA: Dual Effects on Breast Cancer Metastasis and Clinical Applications. Cancers. 2019; 11(11):1802. https://doi.org/10.3390/cancers11111802

Chicago/Turabian Style

Huang, Qi-Yuan, Guo-Feng Liu, Xian-Ling Qian, Li-Bo Tang, Qing-Yun Huang, and Li-Xia Xiong. 2019. "Long Non-Coding RNA: Dual Effects on Breast Cancer Metastasis and Clinical Applications" Cancers 11, no. 11: 1802. https://doi.org/10.3390/cancers11111802

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