Next Article in Journal
YAP1 Regulates the YAP1/AR/PSA Axis through Autophagy in Castration-Resistant Prostate Cancer and Mediates T-Cell Immune and Inflammatory Cytokine Infiltration
Previous Article in Journal
Are There Any Differences in the Healing Capacity between the Medial Collateral Ligament’s (MCL) Proximal and Distal Parts in the Human Knee? Quantitative and Immunohistochemical Analysis of CD34, α-Smooth Muscle Actin (α-SMA), and Vascular Endothelial Growth Factor (VEGF) Expression Regarding the Epiligament (EL) Theory
Previous Article in Special Issue
New Insights into Chemoresistance Mediated by Mdm2 Inhibitors: The Benefits of Targeted Therapy over Common Cytostatics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Long Noncoding RNAs (lncRNAs) in Esophageal Cancer Therapy Resistance and Metastasis

1
School of Medicine, College of Medicine, Kaohsiung Medical University, Kaohsiung 807, Taiwan
2
Department of Biotechnology, Kaohsiung Medical University, Kaohsiung 807, Taiwan
3
Graduate Institute of Biomedical Engineering, National Chung Hsing University, Taichung 402, Taiwan
4
Department of Ophthalmology, Kaohsiung Municipal Siaogang Hospital, Kaohsiung 812, Taiwan
5
Department of Ophthalmology, Kaohsiung Medical University Hospital, Kaohsiung 807, Taiwan
*
Author to whom correspondence should be addressed.
Biomedicines 2024, 12(3), 660; https://doi.org/10.3390/biomedicines12030660
Submission received: 7 February 2024 / Revised: 27 February 2024 / Accepted: 1 March 2024 / Published: 15 March 2024

Abstract

:
Esophageal cancer (EC) is one of the most aggressive gastrointestinal cancers. Despite improvements in therapies, the survival rate of patients with EC remains low. Metastasis accounts for up to 90% of cancer-related deaths, and resistance to anti-neoplastic therapeutics is also a main cause of poor survival. Thus, metastasis and drug resistance are undoubtedly the two main challenges in cancer treatment. Among the different categories of noncoding RNAs, lncRNAs have historically drawn less attention. However, lncRNAs have gradually become a research hotspot, and increasing research has demonstrated that lncRNAs participate in the tumorigenesis of multiple types of cancer, including EC. Long noncoding RNAs (lncRNAs) are RNA transcripts longer than 200 nucleotides in length that play important roles in epigenetics, transcription regulation, and posttranscriptional processing. In this review, we elucidated the role of lncRNAs in the metastasis and drug resistance of EC and discussed their potential clinical applications and related limitations. With a better understanding of the underlying mechanisms of lncRNAs, we can identify therapeutic targets for EC in the future.

1. Introduction

Esophageal cancer (EC) is one of the most aggressive malignancies. In 2020, EC ranked seventh in terms of incidence and sixth in terms of mortality among malignant tumors [1]. There are two primary forms of EC—esophageal squamous-cell carcinoma (ESCC) and esophageal adenocarcinoma (EAC)—with the former type accounting for 90% of all EC cases worldwide; however, the incidence of the latter has significantly increased in the past few decades. However, the incidence of this disease varies with geographical location [2]. The pathogenesis of EC is complex; Barret’s esophagus is the precursor for EAC, while ESCC can be caused by multiple factors (e.g., tobacco smoking, alcohol consumption, and genetic factors) [3,4]. Currently, neoadjuvant chemoradiation therapy is a key treatment for EAC patients with locally advanced and/or lymph-node-positive tumors, and postoperative adjuvant therapy has been found to increase the disease-free survival rate of ESCC patients [5]. Despite improvements in therapies, the prognosis of patients with EC remains poor, with a global 5-year survival rate of approximately 20.5% [6]. Thus, it is essential to clarify the molecular mechanism of tumor progression, which might reveal ideas for the treatment of EC. Metastasis contributes to as many as 90% of cancer-associated deaths [7], and resistance to anti-neoplastic therapeutics is another cause of poor survival because recurrence and metastasis occur after drug resistance develops [8]. Therefore, gaining a deeper understanding of the key regulators involved in these two processes might facilitate the development of better therapeutic regimens and improve the poor prognosis of patients with EC. Previous research has focused mainly on the role of protein-coding RNAs in the progression of EC, which has led to the development of numerous promising approaches for treating this disease. However, with recent advances in genomic research, accumulating studies have revealed that noncoding RNAs are also involved in various physiological and pathological processes, including tumor progression. Among the different categories of noncoding RNAs, long noncoding RNAs (lncRNAs) have drawn less attention. Nevertheless, lncRNAs have gradually become a research hotspot because they have recently been found to govern the development of multiple types of cancer, such as prostate cancer, breast cancer and EC [9]. This review highlights the roles of lncRNAs in the metastasis and drug resistance of EC.

2. Overview of lncRNAs

In addition to protein-coding mRNAs, human genomes also contain several other types of RNAs. According to the Human Genome Project, protein-coding genes account for less than 2% of the whole genome sequence, and many noncoding regulatory regions are transcribed into ncRNAs. ncRNAs can be divided into several categories, including ‘housekeeping’ ncRNAs (such as rRNA, tRNA, snRNA and snoRNA), regulatory ncRNAs and some other poorly characterized types of ncRNAs [10,11]. LncRNAs are regulatory ncRNAs, and their significant roles in complex organisms have recently received increased attention.
LncRNAs are RNA transcripts longer than 200 nucleotides that are poorly conserved and lack an open reading frame. Like coding genes, lncRNAs are transcribed via RNA polymerase II transcription, which includes 5′-capping, poly(A) tails and splicing through spliceosomes [12]. Based on their genomic origin, lncRNAs can be classified into five categories: sense, antisense, bidirectional, intronic and intergenic lncRNAs [13]. Although lncRNAs cannot be translated into functional proteins, they are associated with diverse biological processes. According to previous studies, lncRNAs can participate in the epigenetic regulation of gene expression by recruiting chromatin-modifying proteins (e.g., methyltransferases, acetyltransferases, and deacetylases) to specific genomic loci through cis-regulation or trans-regulation [14]. For instance, the lncRNA Hox transcript antisense RNA (HOTAIR) regulates the expression of HOXD genes by recruiting polycomb chromatin remodeling complex 2 (PRC2) [15]. In addition, lncRNAs can participate in transcriptional regulation by influencing the activity of specific transcription factors and polymerases that regulate gene transcription. Moreover, lncRNAs can affect posttranscriptional processing steps, such as splicing, transport, translation and degradation, by interacting with mRNAs. For instance, they can serve as ‘‘miRNA sponges’’ that modulate the competing endogenous RNA (ceRNA) network [16]. As lncRNAs participate in a variety of mechanisms that regulate gene expression, abnormal expression of lncRNAs can result in human diseases such as cardiovascular disease and multiple types of cancer development, including EC [17,18]. In the section below, we summarize the roles of lncRNAs in the development of EC.

3. LncRNAs Participate in the Process of EC Metastasis

Metastasis is a sequential and multistep process during which malignant cells spread from the primary tumor to distant organs; it can be divided into five major steps: invasion, intravasation, dissemination, extravasation and colonization [19,20]. Accumulating research has suggested that lncRNAs can mediate EC metastasis through diverse mechanisms. LncRNAs have dual roles in metastasis: (1) Some lncRNAs function as promotors of metastasis, and (2) others function as repressors of metastasis; the detailed regulatory mechanisms of these two types of lncRNAs are discussed in the following section (Figure 1 and Table 1).

3.1. LncRNAs That Function as EC Metastasis Promoters

3.1.1. H19

H19 has been proven to play crucial roles in multiple types of cancer since it was first discovered by Bartolomei and colleagues in 1991 [21,22,23,24]. H19 expression is upregulated in human EC tissues and cell lines compared to that in corresponding normal controls. In addition to promoting the proliferation of EC cells by regulating cell cycle progression at the G0/G1 phase [25,26], H19 contributes to invasion and epithelial-to-mesenchymal transition (EMT), as its knockdown in ESCC cells leads to the upregulation of epithelial markers, e.g., E-cadherin, and the downregulation of mesenchymal markers, e.g., vimentin and MMP-9 [26]. Regarding the possible mechanism of EMT, Chen and associates demonstrated that high expression of H19 results in increased expression of signal transducer and activator of transcription 3 (STAT3) and Drosophila zeste gene enhancer human homolog 2 (EZH2), and they cooperatively promote EMT and metastasis via the SRY-box transcription factor 4 (SOX4)-β-catenin pathway [27].

3.1.2. CASC9

The lncRNA cancer susceptibility candidate 9 (CASC9) is another important player that promotes EC metastasis, and its expression is significantly greater in ESCC cells and tissues than in normal controls. Pan and associates demonstrated that the knockdown of CASC9 suppressed the migration and invasion of ESCC cells [28]. Additionally, CASC9 was shown to promote EMT, which was accompanied by the upregulation of EMT-related markers [29]. Liang et al. further reported that by binding to CREB-binding protein (CBP), CASC9 can recruit CBP to the promoter of laminin subunit gamma-2 (LAMC2), subsequently activating the focal adhesion kinase (FAK)-phosphoinositide-3-kinase (PI3K)/protein kinase B (Akt) pathway to promote ESCC metastasis [30].

3.1.3. MALAT1

The lncRNA metastasis associated with lung adenocarcinoma transcript-1 (MALAT1) is also involved in the invasion and metastasis of EC. Aberrant expression of MALAT1 is associated with metastasis in multiple types of cancers, and MALAT1 can act as a metastasis promoter or metastasis suppressor depending on the cancer type. For example, MALAT1 facilitates metastasis in colon cancer [31], gastric cancer [32] and ovarian cancer [33] but suppresses metastasis in breast cancer [34]. In EC specimens, the expression of MALAT1 is upregulated, and increased viability and migratory capacities were observed in EC cells compared to control cells. The inhibition of MALAT1 upregulated the expression of miR-1-3p via a ceRNA mechanism, which subsequently suppressed tumor cell metastasis and tumor angiogenesis by inactivating the downstream effectors Coronin-1C (CORO1C) and tropomyosin 3 (TPM3). Downregulation of CORO1C/TPM3 signaling also contributed to the inhibition of EC tumor metastasis in vivo [35]. Chen et al. demonstrated that silencing MALAT1 resulted in altered expression of EMT-related markers. However, overexpression of EZH2 reversed these effects. These results indicated that the Ezh2/Notch1 pathway is involved in EMT in EC cells [36].

3.1.4. PVT1

Recently, several studies have shown that plasmacytoma variant translocation 1 (PVT1) is overexpressed in human tumors, including non-small cell lung cancer (NSCLC) [37], small cell lung cancer (SCLC) [38], gastric cancer [39] and breast cancer [40], where it functions as an oncogene and regulates cancer cell metastasis. Zheng et al. demonstrated that PVT1 expression was markedly greater in EC tissues than in adjacent normal tissues and promoted invasion by regulating EMT-related protein expression [41]. Shen and associates indicated that PVT1 promotes EC cell metastasis by specifically binding to miR-145. Fascin actin bundling protein 1 (FSCN1), which was previously found to promote NSCLC cell growth, migration, and invasion [42], is targeted by miR-145. Thus, PVT1 represses the expression of miR-145 to induce cell migration and invasion via the upregulation of FSCN1 [43]. Moreover, by acting as a molecular sponge, PVT1 has been demonstrated to compete with zinc finger E-box binding homeobox 1 (ZEB1) for binding to miR-128. ZEB1 was identified as a transcriptional repressor of E-cadherin due to its ability to bind to its promoter. PVT1 can upregulate the expression of ZEB1 to promote ESCC cell migration and invasion [44].

3.1.5. HOTAIR

Chen and associates demonstrated that the lncRNA HOX transcript antisense RNA (HOTAIR) is upregulated in EC cells, and the invasion and migration of these cells were suppressed by transfection of KYSE30 cells with HOTAIR [45]. Xu and Zhang revealed the mechanisms by which HOTAIR promotes invasion, migration and EMT in EC. The lncRNA HOTAIR not only acts as a miRNA sponge to reduce miR-148a expression, subsequently upregulating Snail 2, but also serves as an epigenetic regulator to promote ESCC metastasis [46]. It has been reported that HOTAIR causes H3K27 trimethylation in the Wnt inhibitory factor 1 (WIF-1) promoter and inhibits its expression. With the downregulation of WIF-1, the Wnt/β-catenin pathway is activated, thereby promoting tumor migration and invasion [47].

3.1.6. HOTTIP

The lncRNA HOXA transcript at the distal tip (HOTTIP) is transcribed from a genomic region at the 5′ tip of the HOXA locus containing 3764 nucleotides [48]. The expression of HOTTIP and the HOXA gene HOXA13 is upregulated in ESCC tissues and cell lines. Both of these factors induce cell migration, invasion and EMT pathway activation. Lin et al. studied the mechanism of the HOTTIP/HOXA13 interaction and showed that the lncRNA HOTTIP can recruit WD repeat domain 5 (WDR5) to the HOXA13 promoter locus, which promotes histone H3 lysine 4 trimethylation (H3K4me3) and the binding of polymerase II to the HOXA13 promoter, thus resulting in the activation of HOXA13 gene transcription. In addition, HOTTIP induces HOXA13 expression via miR-30b and Argonaute RISC catalytic component 2 (Ago2), which serve as essential components of the RNA-induced silencing complex (RISC) and as mediators of RNAi mechanisms. Hence, HOTTIP can significantly decrease miR-30b levels and subsequently downregulate the miR-30b target HOXA13 via Ago2 [49].

3.1.7. LUCAT1

Accumulating evidence has demonstrated that lung-cancer-associated transcript 1 (LUCAT1), also known as smoke-and-cancer-associated lncRNA 1 (SCL1), is involved in the progression of multiple types of cancer [50,51,52,53,54]. Yoon and colleagues demonstrated that LUCAT1 expression is also increased in ESCC cell lines and cancer tissues. KYSE-30 cells and HCE-4 cells transfected with LUCAT1 siRNAs exhibited inhibited ESCC cell migration and invasion. Furthermore, they reported that LUCAT1 knockdown suppressed DNA methyltransferase 1 (DNMT1) protein expression by inducing DNMT1 ubiquitination through ubiquitin-like proteins, which include PHD and RING finger domain protein 1 (UHRF1). UHRF1 was previously found to trigger ubiquitination and thus lead to proteasomal degradation of DNMT1 [55]. Taken together, these findings indicate that LUCAT1 can activate DNMT1 to inhibit the expression of tumor suppressor genes such as GADD45G and SFRP2, facilitating ESCC cell invasion and migration and disease progression [56].

3.1.8. CCAT1

Colon-cancer-associated transcript-1 (CCAT1), located at chromosomal region 8q24.21, is significantly upregulated in multiple types of cancer [57,58,59,60,61] and is also highly expressed in ESCC. Zhang et al. demonstrated that knockdown of CCAT1 significantly suppressed cell migration in vivo and in vitro. In contrast, overexpression of CCAT1 promoted cell migration in ESCC. Mechanistically, in the nucleus, CCAT1 can act as a scaffold for two distinct histone methylation complexes, leading to the repression of sprouty RTK signaling antagonist 4 (SPRY4), which is considered a tumor suppressor gene. Moreover, in the cytoplasm, CCAT1 competes for the binding of miR-7, thereby mediating posttranscriptional regulation of HOXB13. Both of these mechanisms contribute to ESCC cell migration [62].

3.1.9. EZR-AS1

Ezrin (EZR), a member of the ezrin-radixin-moesin (ERM) cytoskeletal protein family, has been shown to participate in many aspects of cancer metastasis [63]. Zhang et al. suggested that the lncRNA EZR antisense AS1 (EZR-AS1) can induce ESCC migration by upregulating EZR expression. The lncRNA EZR-AS1 can complex with RNA polymerase 2 and recruit SET and MYND domain containing 3 (SMYD3) to the SMYD3 binding site within the GC-rich region downstream of the EZR promoter, inducing EZR transcription to enhance ESCC cell migration [64].

3.1.10. FTH1P3

The lncRNA ferritin heavy chain 1 pseudogene 3 (FTH1P3) has been reported to promote tumor progression in colorectal cancer [65], cervical cancer [66], and NSCLC [67]. Yang et al. reported that it also functions as an oncogene in ESCC. FTH1P3 was upregulated in ESCC, and silencing FTH1P3 inhibited the migration and invasion of ESCC cells in vitro. Sp1 and NF-κB are both transcription factors involved in cancer metastasis [68], and both are highly expressed in ESCC tissues and cells. Moreover, silencing of FTH1P3 led to the downregulation of Sp1 and NF-kB expression, revealing that FTH1P3 regulates ESCC metastasis by upregulating the SP1/NF-kB pathway [69].

3.2. LncRNAs That Function as EC Metastasis Suppressors

3.2.1. MEG3

The lncRNA maternally expressed gene 3 (MEG3) is widely expressed in many normal tissues and has been indicated to be a tumor suppressor [70]. Dong and associates demonstrated that MEG3 was significantly downregulated in EC cells. Treatment of EC cells with the DNA methyltransferase inhibitor 5-AZA-2’-deoxycytidine (5-Aza-dC) induced upregulation of MEG3, which suggested that promoter hypermethylation of MEG3 represents one of the mechanisms that decreases its expression in ESCC tissues. On the other hand, MEG3 may sponge miR-9 to upregulate the expression of E-cadherin and FOXO1, thereby affecting EMT and metastasis [71]. Phosphoserine aminotransferase 1 (PSAT1) has been found to promote the phosphorylation of glycogen synthase kinase-3 beta (GSK-3β) and activate the GSK-3β/Snail pathway, leading to EMT in ESCC [72]. Li et al. suggested that the lncRNA MEG3 can negatively regulate PSAT1 expression and thus inhibit EMT. However, the mechanism by which MEG3 downregulates PSAT1 and EMT still needs further investigation [73].

3.2.2. GAS5

The lncRNA growth-arrest-specific transcript 5 (GAS5) is suppressed in diverse cancers, such as ovarian cancer [74], cervical cancer [75] and gastric cancer [76], indicating that GAS5 may play a tumor-suppressive role. KE et al. suggested that GAS5 is expressed at low levels in ESCC tissues and cells and that the overexpression of GAS5 inhibits ESCC cell invasion and migration by regulating the expression of EMT-associated proteins [77]. Wang et al. demonstrated that the expression of the lncRNA GAS5 in serum decreases as the primary tumor stage increases, and this lncRNA could be a potential diagnostic and prognostic marker. Moreover, they indicated that GAS5 inhibits ESCC migration by inactivating the PI3K/AKT/mTOR pathway [78]. However, in public datasets, Huang et al. suggested that GAS5 is upregulated in ESCC tissues compared with normal tissues but still acts as a tumor suppressor. These authors further indicated that GAS5 was upregulated by interferon (IFN) via the JAK-STAT pathway and that GAS5 can also positively regulate IFN responses. The interaction between GAS5 and the IFN pathway leads to the inhibition of ESCC cell migration and invasion and thus may be a promising therapeutic target in the clinic [79].

3.2.3. BDNF-AS

The lncRNA brain-derived neurotrophic factor antisense (BDNF-AS) was previously found to act as a tumor suppressor in multiple types of cancer including EC [80,81,82,83]. Zhao et al. demonstrated that BDNF-AS expression was significantly decreased in both EC cell lines and tissues and that overexpression of BDNF-AS can suppress the migration, invasion, and EMT of EC cells. Mechanistically, BDNF-AS acts as a ceRNA to sponge miR-214. However, the downstream targets and pathways involved need to be further studied [84].
Cancer metastasis is the main cause of cancer-related death, and studies have shown that lncRNAs serve as double-edged swords in cancer metastasis through diverse mechanisms. Summarizing the mechanism of lncRNAs in EC metastasis might help develop promising diagnostic and prognostic tools and even for developing more effective therapeutics for EC (Table 1).
Table 1. The role of lncRNAs in the modulation of EC metastasis.
Table 1. The role of lncRNAs in the modulation of EC metastasis.
LncRNA Mechanism Location Reference
Promotion
H19 Promotes the STAT3/EZH2/β-catenin axis 11p15.5 [27]
CASC9 Recruits CBP to promoter of LAMC2 and induces LAMC2 transcription, subsequently activating FAK-PI3K/Akt pathways 8q21.11 [30]
MALAT1 Sponges miR-1-3p and activates CORO1C/TPM3 pathway; regulates the Ezh2-Notch1 pathway 11q13 [35,36]
PVT1 Binds to miR-145 and upregulate FSCN1; binds to miR-128 and increase ZEB1 expression 8q24 [43,44]
HOTAIR Sponges miR-148a and promotes the expression of Snail2; downregulates WIF-1 expression by inducing H3K27 trimethylation in the promoter region and then promoting the Wnt/β-catenin pathway 12q13.13 [46,47]
HOTTIP Recruits WDR5 and then activates HOXA13 gene transcription; sponges miR-30b and is regulated by miR-30b in the nucleus via an Ago2 7p15.2 [49,85]
LUCAT1 Activates DNMT1 to inhibit the expression of tumor suppressor genes (GADD45G and SFRP2) 5q14.3 [56,86]
CCAT1 Mediates histone methylation and thus represses the expression of tumor suppressor gene SPRY4 in nucleus; competes for miR-7 and then mediates the posttranscriptional regulation of HOXB13 in cytoplasm 8q24.21 [62]
EZR-AS1 Forms a complex with RNA polymerase 2; recruits SMYD3 to the promoter of downstream EZR and promotes EZR transcription 6q25.3 [64]
FTH1P3 Promotes the SP1/NF-κB pathway 2p23.3 [69]
Suppression
MEG3 Sponges miR-9 to regulate the expression of E-cadherin and FOXO1; downregulates expression of PSAT1 14q32.2 [71,73,87]
GAS5 Inactivates the PI3K/AKT/mTOR pathway; interacts with interferon 1q25 [78,79]
BDNF-AS Sponges miR-214 11p14.1 [80,84]

4. LncRNAs Are Involved in EC Therapy Resistance

The most suitable treatments for EC differ depending on the tumor stage or features of the patient features. Endoscopic and surgical treatments are used for those with early-stage tumors, while systemic regimens, including definitive chemoradiotherapy, are used to treat patients with locally advanced tumors or those who are not eligible for surgical treatment [88]. However, in addition to metastasis, resistance to anticancer drugs is another clinical obstacle limiting the effectiveness of therapeutic approaches such as chemo-therapy and radiation therapy in cancer patients. Therefore, to improve the poor prognosis of patients with EC, solutions for overcoming drug resistance must be explored. Recently, accumulating evidence has suggested that lncRNAs play critical roles in promoting resistance and decreasing sensitivity to therapeutic drugs in multiple cancers, including EC; thus, they have drawn increasing amounts of attention in the field of cancer (Figure 2 and Table 2).

4.1. Chemoresistance

4.1.1. Cisplatin

Cisplatin, or cis-dichlorodiammine-platinum(II), is a representative platinum-based compound that has been used to treat a wide spectrum of solid neoplasms. It interacts with purine bases to inhibit DNA repair and cause DNA damage, thereby facilitating cancer cell apoptosis [89]. Once cisplatin-induced DNA adducts result in DNA lesions, nucleotide excision repair (NER) is the pathway involved mainly in repairing DNA damage [90]. Notably, inhibition of NER effectively improved chemosensitivity to cisplatin as tumor cells with enhanced DNA repair capacity exhibited increased therapeutic resistance [91]. Therefore, targeting the DNA repair machinery seems to be a potential strategy for sensitizing tumor cells to cisplatin. Nevertheless, the involvement of lncRNAs in chemoresistance through the modulation of DNA damage repair has been relatively less reported and warrants further investigation.
Cisplatin-based chemotherapy regimens are still the foremost treatment option for patients diagnosed with EC [92]. Although cisplatin improves EC prognosis, the development of chemoresistance to cisplatin is a common cause of treatment failure [93].
LncRNAs can promote EC resistance to cisplatin by functioning as ceRNAs. The lncRNA CCAT1 contributes to ESCC chemoresistance and facilitates ESCC metastasis. Hu et al. reported that CCAT1 could downregulate miR-143 and thus indirectly increase the expression of polo-like kinase 1 (PLK1) and budding uninhibited by benzimidazole-related 1 (BUBR1), which are associated with the cell cycle. Taken together, these findings indicate that high expression of CCAT-1 facilitates ESCC resistance to cisplatin through the miR-143/PLK1/BUBR1 signaling axis [94]. Noncoding RNA activated by DNA damage (NORAD) was shown to promote multiple types of cancer [95,96], including EC. Jia et al. demonstrated that high NORAD expression was related to poorer prognosis in ESCC patients treated with cisplatin-based therapy, revealing that NORAD overexpression could contribute to cisplatin resistance in ESCC cells. Regarding the underlying mechanism, NORAD can act as a ceRNA to sponge miR-224-3p, thereby upregulating metadherin (MTDH). MTDH was found to participate in multiple pathways, such as the PI3K/Akt and MAPK pathways, and to play important roles in carcinogenesis [97]. In conclusion, the NORAD/miR-224-3p/MTDH axis facilitates Cis-diamminedichloro-platinum (CDDP) resistance by promoting the nuclear accumulation of β-catenin [98]. The roles of lnc-MCEI (a lncRNA that mediates the chemosensitivity of ESCC by regulating IGF2), a lncRNA that mediates the chemosensitivity of ESCC, were first described by Liu and associates. Liu et al. reported that lnc-MCEI was upregulated in ESCC tissues and that knockdown of lnc-MCEI increased the sensitivity of ESCC cells to cisplatin. Mechanistically, the lncRNA lnc-MCEI acts as a ceRNA by sponging miR-6759-5p, thereby inducing the insulin-like growth factor 2 (IGF2)/PI3K/AKT axis. Thus, the MCEI facilitates the chemoresistance of ESCC through the miR-6759-5p/IGF2/PI3K/AKT axis [99].
Some lncRNAs can facilitate EC resistance to cisplatin by interacting with proteins. Yang et al. suggested that LINC00337 is involved in EC progression and chemoresistance. LINC00337 expression is elevated in ESCC tissues and cells compared with normal controls, and silencing LINC00337 increased chemosensitivity to cisplatin. Mechanistically, LINC00337 recruits E2F transcription factor 4 (E2F4) to binding sites within the promotor of the targeting protein for Xenopus kinesin-like protein 2 (TPX2), which was previously found to be related to cell proliferation and a poor prognosis in ESCC patients [100]. Subsequently, TPX2 transcription is promoted, thereby increasing EC cell chemoresistance to cisplatin [101]. Jiang and associates demonstrated that taurine upregulated gene 1 (TUG1) is upregulated in ESCC, especially in tumors resistant to chemotherapy [102]. Zhang et al. further clarified the underlying mechanism related to this phenotype: TUG1 increases nuclear factor erythroid 2-related factor 2 (Nrf2) protein expression by directly binding to the Nrf2 protein. Nrf2 was previously demonstrated to be a key contributor to cisplatin resistance in ovarian cancer [103] and ESCC cells [104]. LncRNA cancer susceptibility candidate 8 (CASC8) contributes to the progression of multiple types of cancer [105,106], and Wu et al. demonstrated the mechanisms of CASC8 upregulation as well as its effect on chemoresistance in ESCC. These authors indicated that AlkB homolog 5 (ALKBH5), an RNA demethylase suppressed the m6A modification of CASC8, which resulted in increased CASC8 levels in ESCC through the stabilization of the CASC8 transcript. However, CASC8 decreases heterogeneous nuclear ribonucleoprotein L (hnRNPL) polyubiquitination and proteasomal degradation, leading to the stabilization and upregulation of hnRNPL protein expression. Taken together, these results indicate that ALKBH5-mediated m6A modification increases CASC8 levels in ESCC and that overexpression of CASC8 facilitates ESCC progression and chemoresistance to cisplatin by interacting with hnRNPL [107].
Interestingly, lncRNAs can also play a role in hypoxia-mediated drug resistance in EC cells, and the lncRNA E2F1 messenger RNA stabilizing factor (EMS) is a salient example of this phenomenon. Zhu et al. suggested that EMS is expressed at greater levels in EC tissue than in adjacent normal tissues and can be upregulated by hypoxia. EMS can sponge miRNA-758-3p and downregulate its expression. Wilms tumor 1-associated protein (WTAP), which has been revealed to be a prognostic marker of several cancers (e.g., nasopharyngeal carcinoma [108] and bladder cancer [109]), serves as the downstream target of miRNA-758-3p. Its overexpression significantly contributes to chemoresistance to cisplatin [110].
On the other hand, several lncRNAs, including LINC00261, increase ESCC sensitivity to cisplatin. Wang et al. reported that LINC00261 expression was significantly lower in ESCC tissues and cells than in normal controls and that LINC00261 could negatively regulate miR-545-3p by sponging. Subsequently, metallothionein-1 M (MT1M), which was previously identified as a tumor suppressor [111], is targeted by miR-545-3p and is thus downregulated in ESCC tissues and cells, leading to resistance to cisplatin [112]. Moreover, in ESCC cells, CASC2 is expressed at low levels. Zhu et al. demonstrated that overexpression of CASC2 in ESCC cells not only promoted DNA damage but also increased sensitivity to cisplatin. CASC2 can directly interact with miR-181a and negatively regulate its expression, thereby suppressing the Akt pathway, which plays a role in multiple cellular processes, such as cell apoptosis and proliferation. [113]. Thus, downregulation of CASC2 in ESCC decreases the antitumor activity of cisplatin by activating the miR-181a/Akt axis [114].
Cancer-associated fibroblasts (CAFs) are a major component of the tumor microenvironment and have multiple functions. They can secrete several factors (e.g., chemokines, cytokines, and growth factors) that degrade their surrounding ECM to support and maintain tumor growth. They also maintain tumor-cell-specific metabolism [115]. Intriguingly, the lncRNA POU3F3 (also known as LINC01158) can promote ESCC resistance through a CAF-associated mechanism. Tong et al. demonstrated that POU3F3 facilitates cisplatin resistance in ESCC by transforming normal fibroblasts (NFs) into cancer-associated fibroblasts (CAFs). Mechanistically, the lncRNA POU3F3 can be packaged into tumor-secreted exosomes and transferred from ESCC cells to NFs, leading to the activation of NFs. Subsequently, activated fibroblasts secrete interleukin 6 (IL-6), which participates in the communication between stromal cells and cancer cells in the tumor microenvironment [116]. Through this mechanism, POU3F3 facilitates ESCC resistance to cisplatin and contributes to a poor prognosis [117].

4.1.2. Fluorouracil (5-FU)

5-FU is widely used in cancer treatment, and it exerts its anticancer effects by inhibiting thymidylate synthase (TS) or by incorporating its metabolites into RNA and DNA [118]. Cisplatin combined with a continuous infusion of 5-FU has been identified as a standard treatment for patients with ESCC or EAC [119], but the development of resistance to 5-FU is common.
Zhang et al. demonstrated that depletion of the lncRNA HOTAIR facilitates the chemosensitivity of EC cell lines to 5-FU. They also found that the methylation level of CpG islands in the promoter region of the methylenetetrahydrofolate reductase (MTHFR) gene was greater in EC tissues than in adjacent normal tissues. Taken together, these results indicate that HOTAIR can repress MTHFR expression by upregulating MTHFR methylation, thereby promoting the chemoresistance of EC cells to 5-FU [120]. In addition, Li et al. suggested that the lncRNA LINC01270 not only promotes EC progression but also contributes to EC cell chemoresistance to 5-FU. LINC01270 is overexpressed in EC tissues, while the expression of GSTP1, which was previously found to be related to drug resistance in EC [121], is low. Mechanistically, they found that LINC01270 can recruit the DNA methyltransferases DNMT1, DNMT3A and DNMT3B to increase the methylation level of glutathione S-transferase P1 (GSTP1), thereby inhibiting the expression of GSTP1. GSTP1 plays diverse roles in the progression of different cancers. For example, GSTP1 was found to serve as an oncogene that suppresses cell apoptosis in head and neck squamous cell carcinoma [122], while another study showed that inhibition of GSTP1 expression promoted EC chemoresistance to 5-FU [123]. LINC01419 also promotes 5-FU resistance through methylation of GSTP1. High expression of LINC01419 has been shown to promote the progression of hepatocellular carcinoma and lung adenocarcinoma [124,125,126]. Chen et al. demonstrated that LINC01419 is highly expressed in ESCC tissues and can bind to the promoter region of the GSTP1 gene, thereby mediating DNA methylation of GSTP1. The hypermethylation of GSTP1 leads to the promotion of ESCC progression and chemoresistance to 5-FU therapy [127].
In contrast, several lncRNAs, including tumor suppressor candidate 7 (TUSC7), can suppress 5-FU resistance. Chang et al. reported that TUSC7 expression was downregulated in ESCC tissues and cells and that TUSC7 could negatively regulate the expression of miR-224. As a downstream target of miR-224, differentially expressed in squamous cell carcinoma 1 (DESC1) is downregulated in ESCC; it attenuates the chemoresistance of ESCC cells to 5-FU or cisplatin by downregulating the EGFR/AKT pathway [128].

4.1.3. Paclitaxel

Paclitaxel is a representative member of the taxane family and exerts its antitumor effects by forming abnormal spindles, which results in chromosome missegregation and the production of aberrant aneuploid daughter cells [129]. Paclitaxel, alone or in combination with cisplatin, has been demonstrated to be an effective first-line chemotherapy for treating EC [130,131], but the development of resistance to paclitaxel compromises its treatment efficacy. Zhang et al. suggested that the lncRNAs DDX11 antisense RNA 1 (DDX11-AS1), topoisomerase II alpha (TOP2A) and TATA-box binding protein-associated factor 1 (TAF1) are upregulated in EC and that the latter two genes contribute to chemoresistance to paclitaxel. Mechanistically, DDX11-AS1 facilitates the transcription of TOP2A by binding to and upregulating the transcription factor TAF1, thereby promoting TOP2A expression. TOP2A increases the activity of β-catenin and promotes its translocation into the nucleus, resulting in EC cell chemoresistance to paclitaxel [132].

4.1.4. Oxaliplatin

Oxaliplatin is a member of the platinum complex that is used for chemotherapy. It exerts antitumor effects by binding to nucleophilic molecules (mainly DNA) and interfering with DNA replication and transcription [133]. Oxaliplatin has become a promising chemotherapeutic agent for ESCC [134], but resistance to this agent is an obstacle in treating this disease [135]. The lncRNA LINC00152 was reported to confer oxaliplatin resistance in colon cancer [136], and it was also recently indicated to be associated with oxaliplatin resistance in EC. Zhang et al. demonstrated that LINC00152 is highly expressed in EC tissues and cells and can bind to the PRC2 protein complex (EZH2). In addition to its role in metastasis, ZEB1 was also demonstrated to regulate drug resistance through multiple different mechanisms [137]. In the present study, the transcriptional inhibition of ZEB1 was attenuated by EZH2 binding, and the overexpression of ZEB1 led to EC cell resistance to oxaliplatin [138].

4.2. Radioresistance

Radiotherapy combined with chemotherapy results in a better prognosis than surgery or chemotherapy alone for EC patients [139]. However, some patients develop radiotherapy resistance [140], which is a major obstacle to improving EC treatment efficacy.
Multiple oncogenic lncRNAs can promote radioresistance in patients with EC. The major lesion induced by ionizing radiation (IR) is considered a double-strand break (DSB), which is a break in the phosphodiester backbone of both DNA strands [141]. Nonhomologous end joining (NHEJ) and homologous recombination (HR) are two core mechanisms for DSB repair [142]. Accumulating studies have revealed that enhanced DNA repair machinery can contribute to resistance to IR [143,144]. For instance, IR exposure facilitates the phosphorylation of phosphatase and tensin homolog (PTEN), which subsequently enables the recruitment of RAD51 to augment DNA repair [145]. In addition, high expression of X-ray repair cross complement 3 (XRCC3, a RAD51 paralog) indicated resistance to chemoradiotherapy and poor survival in ESCC patients [144]. Chen et al. suggested that the expression of the lncRNA family with sequence similarity 201-member A (FAM201A) was upregulated in ESCC tissues and related to radiosensitivity. ESCC patients with high FAM201A expression have a poorer short-term response to radiotherapy and a shorter survival time than those with low FAM201A expression. The authors demonstrated that, mechanistically, FAM201A directly interacts with miR-101 as a sponge and negatively regulates miR-101 expression. Ataxia telangiectasia mutated (ATM) is an important repair protein that participates in HR repair. In addition, ATM and mTOR serve as downstream targets of miR-101 and are upregulated when miR-101 is decreased, leading to ESCC radioresistance [146]. Liu et al. suggested that in ESCC cells exposed to X-ray radiation, LINC00473 was upregulated, while miR-497-5p was downregulated. These authors further elucidated that LINC00473 overexpression can downregulate the expression of miR-497-5 and subsequently increase the protein expression of cell division cycle 25A (CDC25A). CDC25A is a significant regulator of the cell cycle, and its high expression contributes to ESCC radioresistance [147].
In addition to its role in chemoresistance, the lncRNA NORAD is also correlated with radioresistance in ESCC. Sun et al. suggested that radiotherapy promotes NORAD expression by increasing H3K4me2 levels. Overexpression of NORAD in the cytoplasm delayed pri-miR-199a1 processing by competitively binding to Pumilio RNA-Binding Family Member 1 (PUM1), thereby inhibiting miR-199a-5p maturation. Conversely, miR-199a-5p is markedly upregulated in sh-NORAD exosomes, and NORAD knockdown facilitates the transfer of exosomal miR-199a-5p into radioresistant cells. Subsequently, endonuclease/exonuclease/phosphatase family domain containing 1 (EEPD1), the target protein of miR-199a-5p, was downregulated, increasing sensitivity to radiotherapy by inhibiting DNA damage repair. In conclusion, NORAD contributes to ESCC radioresistance [148].
The lncRNA colon-cancer-associated transcript 2 (CCAT2) was first identified in colorectal cancer, and its upregulated expression in various types of cancer has been reported to promote tumor progression [149,150,151]. Wang et al. revealed the role of CCAT2 in radiotherapy resistance in EC cells. They found that CCAT2 expression was significantly greater in EC cells and tissues. More intriguingly, the CCAT2 expression level was also greater in patients who had poorer overall survival after radiotherapy treatment, and CCAT2 knockdown enhanced X-ray-induced apoptosis in EC cells, indicating that CCAT2 is involved in radiotherapy resistance in EC. Mechanistically, CCAT2 inhibits miR-145 expression, thereby downregulating the expression of the ribosomal protein S6 kinase beta-1 (p70S6K1), which is a significant downstream kinase of the apoptosis-associated Akt/ERK pathway. On the other hand, they found that CCAT2 not only inhibits the phosphorylation of Akt, ERK and p70S6K1 but also negatively regulates the expression of the p53, p21 and c-Myc proteins in EC cells. Moreover, p53 was demonstrated to increase p21 expression, thereby inhibiting the cell cycle [152], and c-Myc also regulates p21 activity [153]. In summary, CCAT2 promotes the radioresistance of EC cells by inhibiting apoptosis through the repression of the miR-145/p70S6K1 pathway, activation of the Akt/ERK/p70S6K1 pathway and suppression of the p53 pathway [154].
Moreover, some lncRNAs act as tumor suppressors to increase radiosensitivity. According to Cheng and associates, the homeobox protein Hox-B7 (HOXB7) is highly expressed in EC tissues, while the lncRNA MAGI2-AS3 is expressed at low levels. They also demonstrated that overexpression of HOXB7 increased the resistance of EC cells to radiotherapy. Regarding the underlying mechanism, MAGI2-AS3 can recruit EZH2 to the promoter of HOXB7, leading to enrichment of histone H3 lysine 27 trimethylation (H3K27me3) in the HOXB7 promoter region and suppression of HOXB7 expression. Taken together, these findings indicate that MAGI2-AS3 negatively regulates HOXB7 expression, thereby increasing EC sensitivity to radiotherapy [155].
Studies have focused on the correlation between CAFs and lncRNAs related to EC radioresistance. Zhang et al. reported that the lncRNA DNM3OS is highly expressed in ESCC cells and tissues and further suggested that CAFs induce the transcription of DNM3OS by activating the platelet-derived growth factor-β (PDGF-β)/platelet-derived growth factor receptor β (PDGFR-β)/forkhead box protein O1 (FOXO1) pathway. When it is upregulated, DNM3OS can contribute to the enhancement of radioresistance in ESCC by attenuating the DNA damage response [156].
Neoadjuvant therapy, including chemotherapy plus radiotherapy, is thought to be an effective treatment for advanced EC. Currently, therapeutic resistance remains a major obstacle to achieving good clinical responses in patients with EC, and lncRNAs have been found to participate in this process. Elucidating how lncRNAs regulate therapeutic resistance could lead to substantial progress in treating EC to improve patient prognosis (Table 2).
Table 2. The role of lncRNAs in modulating chemoresistance and radioresistance in EC cells.
Table 2. The role of lncRNAs in modulating chemoresistance and radioresistance in EC cells.
LncRNA Location Therapeutic Strategy Effect on Resistance Mechanism Reference
CCAT1 8q24.21
[157]
Cisplatin Promotion Sponges miR-143 and then increases the expression of PLK1 and BUBR1 [94]
NORAD 20q11.23
[158]
Cisplatin Promotion Sponges miR-2243p and then upregulates MTDH [98]
Radiotherapy Promotion Inhibits miR-199a-5p maturation and then upregulates EEPD1 [148]
Lnc-MCEI 1q32.1 Cisplatin Promotion Sponges miR-6759-5p, thereby upregulates IGF2 and inhibits the PI3K/AKT axis [99]
LINC00337 1p36.31
[159]
Cisplatin Promotion Recruits E2F4 to the promoter of TPX2 and increases TPX2 transcription [101]
TUG1 22q12
[160]
Cisplatin Promotion Binds to Nrf2 to increase its expression [104]
CASC8 8q24.21
[161]
Cisplatin Promotion Is upregulated with ALKBH5-mediated m6A demethylation; interacts with hnRNPL and stabilizes its protein expression [107]
EMS 7q22.1
[162]
Cisplatin Promotion Sponges miR-758-3p and induces WTAP expression [110]
POU3F3 2q12.1
[163]
Cisplatin Promotion Transforms normal fibroblasts into cancer-related fibroblasts [117]
LINC00261 20
[164]
Cisplatin Attenuation Sponges miR-545-3p and upregulates MT1M expression [112]
CASC2 10q26
[165]
Cisplatin Attenuation Downregulates miR-181a expression, and thereby suppresses the Akt pathway [114]
HOTAIR 12q13.13
[166]
5-FU Promotion Represses MTHFR expression by promoting MTHFR methylation [120]
LINC01270 20q13.13
[127]
5-FU Promotion Increases the methylation level of GSTP1 promoter by recruiting DNA methyltransferases, thereby inhibiting GSTP1 expression [123]
LINC01419 8q21.13
[167]
5-FU Promotion Binds to the promoter region of the GSTP1 gene and increases the methylation level of GSTP1 [127]
TUSC7 3q13.31
[168]
5-FU Attenuation Sponges miR-224 and then upregulates DESC1 expression [128]
DDX11-AS1 12p11.21
[169]
Paclitaxel Promotion Facilitates TOP2A transcription and then increases the activity of β-catenin [132]
LINC00512 2p11.2
[170]
Oxaliplatin Promotion Binds to EZH2 and then upregulates ZEB1 expression [138]
FAM201A 9p13.1
[171]
Radiotherapy Promotion Sponges miR-101 and then upregulates ATM expression [146]
LINC00473 6q27
[172]
Radiotherapy Promotion Negatively regulates miR-497-5p and increases CDC25A expression [147]
CCAT2 8q24.21
[173]
Radiotherapy Promotion Inhibits apoptosis by suppressing the miR-145/p70S6K1 pathway; activates the Akt/ERK/p70S6K1 pathway; represses the p53 pathway [154]
MAGI2-AS3 7q21.11
[174]
Radiotherapy Attenuation Recruits EZH2 to the promoter region of HOXB7 and then downregulates HOXB7 expression [155]
DNM3OS 1q24.3
[175]
Radiotherapy Promotion Is activated by CAFs and regulates the DNA damage response [156]

5. Discussion

Initially, lncRNAs were viewed merely as transcriptional junk RNAs, but they have been proven to play roles in various organisms and cell types. Accumulating evidence has demonstrated that lncRNAs are involved in the metastasis and drug resistance of EC and can also serve as EC biomarkers and prognostic indicators. For instance, previous studies suggested that lncRNA PVT1 expression is greater in human EC tissues than in adjacent normal tissues [41], and increased expression of HOTAIR was found to be associated with poor overall survival in ESCC patients [176]. Moreover, lncRNAs packaged in exosomes are considered more functional and integral than free lncRNAs because lncRNAs inside exosomes can be protected from RNases; for these reasons, exosomal lncRNAs are viewed as crucial cancer biomarkers [177]. Indeed, growing evidence indicates that exosomal lncRNAs play an important role in the progression of tumors, including EC [117,148,178,179,180]. Exosomes are membrane-bound, nanosized endocytic vesicles containing proteins, lipids and a variety of RNAs [181]. Exosomes are secreted by most cell types, including tumor cells. Recently, tumor-secreted exosomes have been widely studied for their roles in information exchange between cancer cells, particularly during tumor development [182]. Components of exosomes, such as RNAs, miRNAs, proteins, and even metabolites, can influence recipient cells through autocrine and paracrine signaling in many types of cancer [182]. For example, the lncRNA HNF1A-AS1 carried by exosomes upregulates the expression of tuftelin 1 (TUFT1), promoting proliferation and drug resistance in cervical cancer cells [179]. Moreover, the exosomal lncRNA 91H can promote colorectal metastasis by modifying heterogeneous nuclear ribonucleoprotein K (HNRNPK) expression [180]. In addition, as previously mentioned, the lncRNA POU3F3 can also be packaged into tumor-secreted exosomes to regulate ESCC resistance. Thus, because of their significant role in intercellular communication, exosomes are recognized as biomarkers and prognosticators of cancer and can potentially be used as vehicles for clinical gene and drug delivery [183], which largely contributes to EC treatment. However, despite their potential to serve as biomarkers and drug carriers for EC therapeutics, there are still limitations regarding the clinical implications of exosomes. For example, the efficiency of ultracentrifugation, which remains the gold standard for exosome isolation, is low [184]. Moreover, after exosomes are stored at −80 °C and subjected to freeze-thaw cycles, multiple factors, such as cargo concentration, purity, and size, are affected [185]. Thus, additional investigations regarding the utility of exosomes in the clinical setting are still needed.
Metastasis and drug resistance are undoubtedly the two main obstacles to cancer treatment and the main contributors to poor patient survival rates. In this review, we summarize the lncRNAs that are aberrantly expressed in EC and their roles in EC metastasis and drug resistance. With a better understanding of the underlying mechanisms of lncRNAs, we can identify promising biomarkers for diagnosing EC and predicting prognosis. Moreover, with growing interest in the clinical application of lncRNAs, they could become key therapeutic targets soon [186].

Author Contributions

Conceptualization, Z.-P.W., C.-C.C. and K.-C.C.; data curation, Z.-P.W. and S.-K.H.; writing—original draft preparation, Z.-P.W., C.-C.C. and K.-C.C.; writing—review and editing, H.-M.D.W., K.-J.C. and P.-Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by grants KMHK-109-035 and H-110-004 from Kaohsiung Municipal Siaogang Hospital, Taiwan; grants KMUH111-1M38 and KMUH112-2R49 from Kaohsiung Medical University Hospital, Taiwan and grants NSTC 110-2314-B-037-112 and 112-2314-B-037-024 from the National Science and Technology Council, Taiwan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All reports supporting the discussion are available in this paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA A Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. Rustgi, A.K.; El-Serag, H.B. Esophageal carcinoma. N. Engl. J. Med. 2014, 371, 2499–2509. [Google Scholar] [CrossRef] [PubMed]
  3. Reid, B.J.; Li, X.; Galipeau, P.C.; Vaughan, T.L. Barrett’s oesophagus and oesophageal adenocarcinoma: Time for a new synthesis. Nat. Rev. Cancer 2010, 10, 87–101. [Google Scholar] [CrossRef]
  4. Liu, K.; Zhao, T.; Wang, J.; Chen, Y.; Zhang, R.; Lan, X.; Que, J. Etiology, cancer stem cells and potential diagnostic biomarkers for esophageal cancer. Cancer Lett. 2019, 458, 21–28. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, F.L.; Yu, S.J. Esophageal cancer: Risk factors, genetic association, and treatment. Asian J. Surg. 2018, 41, 210–215. [Google Scholar] [CrossRef]
  6. Sheikh, M.; Roshandel, G.; McCormack, V.; Malekzadeh, R. Current Status and Future Prospects for Esophageal Cancer. Cancers 2023, 15, 765. [Google Scholar] [CrossRef]
  7. Chaffer, C.L.; Weinberg, R.A. A perspective on cancer cell metastasis. Science 2011, 331, 1559–1564. [Google Scholar] [CrossRef]
  8. Liu, X.; He, M.; Li, L.; Wang, X.; Han, S.; Zhao, J.; Dong, Y.; Ahmad, M.; Li, L.; Zhang, X.; et al. EMT and Cancer Cell Stemness Associated with Chemotherapeutic Resistance in Esophageal Cancer. Front. Oncol. 2021, 11, 672222. [Google Scholar] [CrossRef]
  9. Yang, G.; Lu, X.; Yuan, L. LncRNA: A link between RNA and cancer. Biochim. Biophys. Acta 2014, 1839, 1097–1109. [Google Scholar] [CrossRef]
  10. Pauli, A.; Rinn, J.L.; Schier, A.F. Non-coding RNAs as regulators of embryogenesis. Nat. Rev. Genet. 2011, 12, 136–149. [Google Scholar] [CrossRef] [PubMed]
  11. Consortium, E.P.; Moore, J.E.; Purcaro, M.J.; Pratt, H.E.; Epstein, C.B.; Shoresh, N.; Adrian, J.; Kawli, T.; Davis, C.A.; Dobin, A.; et al. Author Correction: Expanded encyclopaedias of DNA elements in the human and mouse genomes. Nature 2022, 605, E3. [Google Scholar] [CrossRef]
  12. Wu, Z.; Liu, X.; Liu, L.; Deng, H.; Zhang, J.; Xu, Q.; Cen, B.; Ji, A. Regulation of lncRNA expression. Cell. Mol. Biol. Lett. 2014, 19, 561–575. [Google Scholar] [CrossRef]
  13. Liu, Y.; Sharma, S.; Watabe, K. Roles of lncRNA in breast cancer. Front. Biosci. 2015, 7, 94–108. [Google Scholar] [CrossRef]
  14. Engreitz, J.M.; Ollikainen, N.; Guttman, M. Long non-coding RNAs: Spatial amplifiers that control nuclear structure and gene expression. Nat. Rev. Mol. Cell Biol. 2016, 17, 756–770. [Google Scholar] [CrossRef]
  15. Rinn, J.L.; Kertesz, M.; Wang, J.K.; Squazzo, S.L.; Xu, X.; Brugmann, S.A.; Goodnough, L.H.; Helms, J.A.; Farnham, P.J.; Segal, E.; et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell 2007, 129, 1311–1323. [Google Scholar] [CrossRef] [PubMed]
  16. Shi, X.; Sun, M.; Liu, H.; Yao, Y.; Song, Y. Long non-coding RNAs: A new frontier in the study of human diseases. Cancer Lett. 2013, 339, 159–166. [Google Scholar] [CrossRef]
  17. Archer, K.; Broskova, Z.; Bayoumi, A.S.; Teoh, J.P.; Davila, A.; Tang, Y.; Su, H.; Kim, I.M. Long Non-Coding RNAs as Master Regulators in Cardiovascular Diseases. Int. J. Mol. Sci. 2015, 16, 23651–23667. [Google Scholar] [CrossRef]
  18. Wapinski, O.; Chang, H.Y. Long noncoding RNAs and human disease. Trends Cell Biol. 2011, 21, 354–361. [Google Scholar] [CrossRef]
  19. Talmadge, J.E.; Fidler, I.J. AACR centennial series: The biology of cancer metastasis: Historical perspective. Cancer Res. 2010, 70, 5649–5669. [Google Scholar] [CrossRef] [PubMed]
  20. Su, Z.; Yang, Z.; Xu, Y.; Chen, Y.; Yu, Q. Apoptosis, autophagy, necroptosis, and cancer metastasis. Mol. Cancer 2015, 14, 48. [Google Scholar] [CrossRef]
  21. Bartolomei, M.S.; Zemel, S.; Tilghman, S.M. Parental imprinting of the mouse H19 gene. Nature 1991, 351, 153–155. [Google Scholar] [CrossRef]
  22. Peng, F.; Li, T.T.; Wang, K.L.; Xiao, G.Q.; Wang, J.H.; Zhao, H.D.; Kang, Z.J.; Fan, W.J.; Zhu, L.L.; Li, M.; et al. H19/let-7/LIN28 reciprocal negative regulatory circuit promotes breast cancer stem cell maintenance. Cell Death Dis. 2017, 8, e2569. [Google Scholar] [CrossRef]
  23. Huang, Z.; Lei, W.; Hu, H.B.; Zhang, H.; Zhu, Y. H19 promotes non-small-cell lung cancer (NSCLC) development through STAT3 signaling via sponging miR-17. J. Cell. Physiol. 2018, 233, 6768–6776. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, M.; Han, D.; Yuan, Z.; Hu, H.; Zhao, Z.; Yang, R.; Jin, Y.; Zou, C.; Chen, Y.; Wang, G.; et al. Long non-coding RNA H19 confers 5-Fu resistance in colorectal cancer by promoting SIRT1-mediated autophagy. Cell Death Dis. 2018, 9, 1149. [Google Scholar] [CrossRef]
  25. Huang, C.; Cao, L.; Qiu, L.; Dai, X.; Ma, L.; Zhou, Y.; Li, H.; Gao, M.; Li, W.; Zhang, Q.; et al. Upregulation of H19 promotes invasion and induces epithelial-to-mesenchymal transition in esophageal cancer. Oncol. Lett. 2015, 10, 291–296. [Google Scholar] [CrossRef]
  26. Tan, D.; Wu, Y.; Hu, L.; He, P.; Xiong, G.; Bai, Y.; Yang, K. Long noncoding RNA H19 is up-regulated in esophageal squamous cell carcinoma and promotes cell proliferation and metastasis. Dis. Esophagus Off. J. Int. Soc. Dis. Esophagus 2017, 30, 1–9. [Google Scholar] [CrossRef]
  27. Chen, M.J.; Deng, J.; Chen, C.; Hu, W.; Yuan, Y.C.; Xia, Z.K. LncRNA H19 promotes epithelial mesenchymal transition and metastasis of esophageal cancer via STAT3/EZH2 axis. Int. J. Biochem. Cell Biol. 2019, 113, 27–36. [Google Scholar] [CrossRef] [PubMed]
  28. Pan, Z.; Mao, W.; Bao, Y.; Zhang, M.; Su, X.; Xu, X. The long noncoding RNA CASC9 regulates migration and invasion in esophageal cancer. Cancer Med. 2016, 5, 2442–2447. [Google Scholar] [CrossRef] [PubMed]
  29. Gao, G.D.; Liu, X.Y.; Lin, Y.; Liu, H.F.; Zhang, G.J. LncRNA CASC9 promotes tumorigenesis by affecting EMT and predicts poor prognosis in esophageal squamous cell cancer. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 422–429. [Google Scholar] [CrossRef]
  30. Liang, Y.; Chen, X.; Wu, Y.; Li, J.; Zhang, S.; Wang, K.; Guan, X.; Yang, K.; Bai, Y. LncRNA CASC9 promotes esophageal squamous cell carcinoma metastasis through upregulating LAMC2 expression by interacting with the CREB-binding protein. Cell Death Differ. 2018, 25, 1980–1995. [Google Scholar] [CrossRef]
  31. Luan, N.; Wang, J.; Sheng, B.; Zhou, Q.; Ye, X.; Zhu, X.; Sun, J.; Tang, Z.; Wang, J. tRF-20-M0NK5Y93-induced MALAT1 promotes colon cancer metastasis through alternative splicing of SMC1A. Am. J. Cancer Res. 2023, 13, 852–871. [Google Scholar] [PubMed]
  32. Chen, D.; Liu, L.; Wang, K.; Yu, H.; Wang, Y.; Liu, J.; Guo, Y.; Zhang, H. The role of MALAT-1 in the invasion and metastasis of gastric cancer. Scand. J. Gastroenterol. 2017, 52, 790–796. [Google Scholar] [CrossRef] [PubMed]
  33. Jin, Y.; Feng, S.J.; Qiu, S.; Shao, N.; Zheng, J.H. LncRNA MALAT1 promotes proliferation and metastasis in epithelial ovarian cancer via the PI3K-AKT pathway. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 3176–3184. [Google Scholar] [PubMed]
  34. Kim, J.; Piao, H.L.; Kim, B.J.; Yao, F.; Han, Z.; Wang, Y.; Xiao, Z.; Siverly, A.N.; Lawhon, S.E.; Ton, B.N.; et al. Long noncoding RNA MALAT1 suppresses breast cancer metastasis. Nat. Genet. 2018, 50, 1705–1715. [Google Scholar] [CrossRef] [PubMed]
  35. Li, Q.; Dai, Z.; Xia, C.; Jin, L.; Chen, X. Suppression of long non-coding RNA MALAT1 inhibits survival and metastasis of esophagus cancer cells by sponging miR-1-3p/CORO1C/TPM3 axis. Mol. Cell Biochem. 2020, 470, 165–174. [Google Scholar] [CrossRef] [PubMed]
  36. Chen, M.; Xia, Z.; Chen, C.; Hu, W.; Yuan, Y. LncRNA MALAT1 promotes epithelial-to-mesenchymal transition of esophageal cancer through Ezh2-Notch1 signaling pathway. Anticancer. Drugs 2018, 29, 767–773. [Google Scholar] [CrossRef] [PubMed]
  37. Zeng, S.H.G.; Xie, J.H.; Zeng, Q.Y.; Dai, S.H.H.; Wang, Y.; Wan, X.M.; Liu, J.C.H. lncRNA PVT1 Promotes Metastasis of Non-Small Cell Lung Cancer Through EZH2-Mediated Activation of Hippo/NOTCH1 Signaling Pathways. Cell J. 2021, 23, 21–31. [Google Scholar] [CrossRef]
  38. Huang, C.; Liu, S.; Wang, H.; Zhang, Z.; Yang, Q.; Gao, F. LncRNA PVT1 overexpression is a poor prognostic biomarker and regulates migration and invasion in small cell lung cancer. Am. J. Transl. Res. 2016, 8, 5025–5034. [Google Scholar]
  39. Niu, J.; Song, X.; Zhang, X. Regulation of lncRNA PVT1 on miR-125 in metastasis of gastric cancer cells. Oncol. Lett. 2020, 19, 1261–1266. [Google Scholar] [CrossRef]
  40. Qu, H.; Li, X.; Chen, F.; Zhang, M.; Lu, X.; Gu, Y.; Lv, M.; Lu, C. LncRNA PVT1 influences breast cancer cells glycolysis through sponging miR-145-5p. Genes Genom. 2023, 45, 581–592. [Google Scholar] [CrossRef]
  41. Zheng, X.; Hu, H.; Li, S. High expression of lncRNA PVT1 promotes invasion by inducing epithelial-to-mesenchymal transition in esophageal cancer. Oncol. Lett. 2016, 12, 2357–2362. [Google Scholar] [CrossRef]
  42. Zhang, Y.; Lin, Q. MicroRNA-145 inhibits migration and invasion by down-regulating FSCN1 in lung cancer. Int. J. Clin. Exp. Med. 2015, 8, 8794–8802. [Google Scholar]
  43. Shen, S.N.; Li, K.; Liu, Y.; Yang, C.L.; He, C.Y.; Wang, H.R. Down-regulation of long noncoding RNA PVT1 inhibits esophageal carcinoma cell migration and invasion and promotes cell apoptosis via microRNA-145-mediated inhibition of FSCN1. Mol. Oncol. 2019, 13, 2554–2573. [Google Scholar] [CrossRef] [PubMed]
  44. Hu, J.; Gao, W. Long noncoding RNA PVT1 promotes tumour progression via the miR-128/ZEB1 axis and predicts poor prognosis in esophageal cancer. Clin. Res. Hepatol. Gastroenterol. 2021, 45, 101701. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, F.J.; Sun, M.; Li, S.Q.; Wu, Q.Q.; Ji, L.; Liu, Z.L.; Zhou, G.Z.; Cao, G.; Jin, L.; Xie, H.W.; et al. Upregulation of the long non-coding RNA HOTAIR promotes esophageal squamous cell carcinoma metastasis and poor prognosis. Mol. Carcinog. 2013, 52, 908–915. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, F.; Zhang, J. Long non-coding RNA HOTAIR functions as miRNA sponge to promote the epithelial to mesenchymal transition in esophageal cancer. Biomed. Pharmacother. 2017, 90, 888–896. [Google Scholar] [CrossRef] [PubMed]
  47. Ge, X.S.; Ma, H.J.; Zheng, X.H.; Ruan, H.L.; Liao, X.Y.; Xue, W.Q.; Chen, Y.B.; Zhang, Y.; Jia, W.H. HOTAIR, a prognostic factor in esophageal squamous cell carcinoma, inhibits WIF-1 expression and activates Wnt pathway. Cancer Sci. 2013, 104, 1675–1682. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, K.C.; Yang, Y.W.; Liu, B.; Sanyal, A.; Corces-Zimmerman, R.; Chen, Y.; Lajoie, B.R.; Protacio, A.; Flynn, R.A.; Gupta, R.A.; et al. A long noncoding RNA maintains active chromatin to coordinate homeotic gene expression. Nature 2011, 472, 120–124. [Google Scholar] [CrossRef]
  49. Lin, C.; Wang, Y.; Wang, Y.; Zhang, S.; Yu, L.; Guo, C.; Xu, H. Transcriptional and posttranscriptional regulation of HOXA13 by lncRNA HOTTIP facilitates tumorigenesis and metastasis in esophageal squamous carcinoma cells. Oncogene 2017, 36, 5392–5406. [Google Scholar] [CrossRef]
  50. Fang, F.; Zhao, M.; Jin, X.; Dong, Z.; Wang, J.; Meng, J.; Xie, S.; Shi, W. Upregulation of MCL-1 by LUCAT1 through interacting with SRSF1 promotes the migration and invasion in non-small cell lung carcinoma. Mol. Cell Biochem. 2023, 1–11. [Google Scholar] [CrossRef]
  51. Mou, E.; Wang, H. LncRNA LUCAT1 facilitates tumorigenesis and metastasis of triple-negative breast cancer through modulating miR-5702. Biosci. Rep. 2019, 39, BSR20190489. [Google Scholar] [CrossRef] [PubMed]
  52. Chi, J.; Liu, T.; Shi, C.; Luo, H.; Wu, Z.; Xiong, B.; Liu, S.; Zeng, Y. Long non-coding RNA LUCAT1 promotes proliferation and invasion in gastric cancer by regulating miR-134-5p/YWHAZ axis. Biomed. Pharmacother. 2019, 118, 109201. [Google Scholar] [CrossRef] [PubMed]
  53. Yu, H.; Xu, Y.; Zhang, D.; Liu, G. Long noncoding RNA LUCAT1 promotes malignancy of ovarian cancer through regulation of miR-612/HOXA13 pathway. Biochem. Biophys. Res. Commun. 2018, 503, 2095–2100. [Google Scholar] [CrossRef] [PubMed]
  54. Wu, R.; Li, L.; Bai, Y.; Yu, B.; Xie, C.; Wu, H.; Zhang, Y.; Huang, L.; Yan, Y.; Li, X.; et al. The long noncoding RNA LUCAT1 promotes colorectal cancer cell proliferation by antagonizing Nucleolin to regulate MYC expression. Cell Death Dis. 2020, 11, 908. [Google Scholar] [CrossRef] [PubMed]
  55. Du, Z.; Song, J.; Wang, Y.; Zhao, Y.; Guda, K.; Yang, S.; Kao, H.Y.; Xu, Y.; Willis, J.; Markowitz, S.D.; et al. DNMT1 stability is regulated by proteins coordinating deubiquitination and acetylation-driven ubiquitination. Sci. Signal. 2010, 3, ra80. [Google Scholar] [CrossRef] [PubMed]
  56. Yoon, J.H.; You, B.H.; Park, C.H.; Kim, Y.J.; Nam, J.W.; Lee, S.K. The long noncoding RNA LUCAT1 promotes tumorigenesis by controlling ubiquitination and stability of DNA methyltransferase 1 in esophageal squamous cell carcinoma. Cancer Lett. 2018, 417, 47–57. [Google Scholar] [CrossRef]
  57. Zhang, C.; Wang, H.; Liu, Q.; Dai, S.; Tian, G.; Wei, X.; Li, X.; Zhao, L.; Shan, B. LncRNA CCAT1 facilitates the progression of gastric cancer via PTBP1-mediated glycolysis enhancement. J. Exp. Clin. Cancer Res. 2023, 42, 246. [Google Scholar] [CrossRef]
  58. You, Z.; Liu, C.; Wang, C.; Ling, Z.; Wang, Y.; Wang, Y.; Zhang, M.; Chen, S.; Xu, B.; Guan, H.; et al. LncRNA CCAT1 Promotes Prostate Cancer Cell Proliferation by Interacting with DDX5 and MIR-28-5P. Mol. Cancer Ther. 2019, 18, 2469–2479. [Google Scholar] [CrossRef]
  59. Lv, L.; Jia, J.Q.; Chen, J. The lncRNA CCAT1 Upregulates Proliferation and Invasion in Melanoma Cells via Suppressing miR-33a. Oncol. Res. 2018, 26, 201–208. [Google Scholar] [CrossRef]
  60. Dong, Y.; Yuan, H.; Jin, G. Identification of long non-coding RNA CCAT1 as an oncogene in nasopharyngeal carcinoma. Oncol. Lett. 2018, 16, 2750–2756. [Google Scholar] [CrossRef]
  61. Cao, Y.; Shi, H.; Ren, F.; Jia, Y.; Zhang, R. Long non-coding RNA CCAT1 promotes metastasis and poor prognosis in epithelial ovarian cancer. Exp. Cell Res. 2017, 359, 185–194. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, E.; Han, L.; Yin, D.; He, X.; Hong, L.; Si, X.; Qiu, M.; Xu, T.; De, W.; Xu, L.; et al. H3K27 acetylation activated-long non-coding RNA CCAT1 affects cell proliferation and migration by regulating SPRY4 and HOXB13 expression in esophageal squamous cell carcinoma. Nucleic Acids Res. 2017, 45, 3086–3101. [Google Scholar] [CrossRef] [PubMed]
  63. Hunter, K.W. Ezrin, a key component in tumor metastasis. Trends Mol. Med. 2004, 10, 201–204. [Google Scholar] [CrossRef]
  64. Zhang, X.D.; Huang, G.W.; Xie, Y.H.; He, J.Z.; Guo, J.C.; Xu, X.E.; Liao, L.D.; Xie, Y.M.; Song, Y.M.; Li, E.M.; et al. The interaction of lncRNA EZR-AS1 with SMYD3 maintains overexpression of EZR in ESCC cells. Nucleic Acids Res. 2018, 46, 1793–1809. [Google Scholar] [CrossRef]
  65. Deng, H.; Zhang, Q.; Zhao, Z.; Wang, M.; Xu, Q. SNAI2/FTH1P3/miR-218-5p Positive Feedback Loop Promotes Colorectal Cancer Metastasis. Biochem. Genet. 2023. [Google Scholar] [CrossRef]
  66. Lv, R.; Zhang, Q.W. The long noncoding RNA FTH1P3 promotes the proliferation and metastasis of cervical cancer through microRNA-145. Oncol. Rep. 2020, 43, 31–40. [Google Scholar] [CrossRef]
  67. Li, Z.; Wang, Y. Long non-coding RNA FTH1P3 promotes the metastasis and aggressiveness of non-small cell lung carcinoma by inducing epithelial-mesenchymal transition. Int. J. Clin. Exp. Pathol. 2019, 12, 3782–3790. [Google Scholar] [PubMed]
  68. Safe, S.; Abdelrahim, M. Sp transcription factor family and its role in cancer. Eur. J. Cancer 2005, 41, 2438–2448. [Google Scholar] [CrossRef]
  69. Yang, L.; Sun, K.; Chu, J.; Qu, Y.; Zhao, X.; Yin, H.; Ming, L.; Wan, J.; He, F. Long non-coding RNA FTH1P3 regulated metastasis and invasion of esophageal squamous cell carcinoma through SP1/NF-kB pathway. Biomed. Pharmacother. 2018, 106, 1570–1577. [Google Scholar] [CrossRef]
  70. Zhou, Y.; Zhang, X.; Klibanski, A. MEG3 noncoding RNA: A tumor suppressor. J. Mol. Endocrinol. 2012, 48, R45–R53. [Google Scholar] [CrossRef]
  71. Dong, Z.; Zhang, A.; Liu, S.; Lu, F.; Guo, Y.; Zhang, G.; Xu, F.; Shi, Y.; Shen, S.; Liang, J.; et al. Aberrant Methylation-Mediated Silencing of lncRNA MEG3 Functions as a ceRNA in Esophageal Cancer. Mol. Cancer Res. 2017, 15, 800–810. [Google Scholar] [CrossRef]
  72. Liu, B.; Jia, Y.; Cao, Y.; Wu, S.; Jiang, H.; Sun, X.; Ma, J.; Yin, X.; Mao, A.; Shang, M. Overexpression of Phosphoserine Aminotransferase 1 (PSAT1) Predicts Poor Prognosis and Associates with Tumor Progression in Human Esophageal Squamous Cell Carcinoma. Cell. Physiol. Biochem. 2016, 39, 395–406. [Google Scholar] [CrossRef]
  73. Li, M.K.; Liu, L.X.; Zhang, W.Y.; Zhan, H.L.; Chen, R.P.; Feng, J.L.; Wu, L.F. Long non-coding RNA MEG3 suppresses epithelial-to-mesenchymal transition by inhibiting the PSAT1-dependent GSK-3beta/Snail signaling pathway in esophageal squamous cell carcinoma. Oncol. Rep. 2020, 44, 2130–2142. [Google Scholar] [CrossRef]
  74. Zhang, T.; Leng, Y.; Duan, M.; Li, Z.; Ma, Y.; Huang, C.; Shi, Q.; Wang, Y.; Wang, C.; Liu, D.; et al. LncRNA GAS5-hnRNPK axis inhibited ovarian cancer progression via inhibition of AKT signaling in ovarian cancer cells. Discov. Oncol. 2023, 14, 157. [Google Scholar] [CrossRef] [PubMed]
  75. Yang, W.; Xu, X.; Hong, L.; Wang, Q.; Huang, J.; Jiang, L. Upregulation of lncRNA GAS5 inhibits the growth and metastasis of cervical cancer cells. J. Cell. Physiol. 2019, 234, 23571–23580. [Google Scholar] [CrossRef]
  76. Liu, Y.; Yin, L.; Chen, C.; Zhang, X.; Wang, S. Long non-coding RNA GAS5 inhibits migration and invasion in gastric cancer via interacting with p53 protein. Dig. Liver Dis. 2020, 52, 331–338. [Google Scholar] [CrossRef]
  77. Ke, K.; Sun, Z.; Wang, Z. Downregulation of long non-coding RNA GAS5 promotes cell proliferation, migration and invasion in esophageal squamous cell carcinoma. Oncol. Lett. 2018, 16, 1801–1808. [Google Scholar] [CrossRef]
  78. Wang, G.; Sun, J.; Zhao, H.; Li, H. Long Non-Coding RNA (lncRNA) Growth Arrest Specific 5 (GAS5) Suppresses Esophageal Squamous Cell Carcinoma Cell Proliferation and Migration by Inactivating Phosphatidylinositol 3-kinase (PI3K)/AKT/Mammalian Target of Rapamycin (mTOR) Signaling Pathway. Med. Sci. Monit. 2018, 24, 7689–7696. [Google Scholar] [CrossRef]
  79. Huang, J.; Li, Y.; Lu, Z.; Che, Y.; Sun, S.; Mao, S.; Lei, Y.; Zang, R.; Li, N.; Sun, N.; et al. Long non-coding RNA GAS5 is induced by interferons and plays an antitumor role in esophageal squamous cell carcinoma. Cancer Med. 2018, 7, 3157–3167. [Google Scholar] [CrossRef] [PubMed]
  80. Su, R.; Ma, J.; Zheng, J.; Liu, X.; Liu, Y.; Ruan, X.; Shen, S.; Yang, C.; Wang, D.; Cai, H.; et al. PABPC1-induced stabilization of BDNF-AS inhibits malignant progression of glioblastoma cells through STAU1-mediated decay. Cell Death Dis. 2020, 11, 81. [Google Scholar] [CrossRef] [PubMed]
  81. Zhi, H.; Lian, J. LncRNA BDNF-AS suppresses colorectal cancer cell proliferation and migration by epigenetically repressing GSK-3beta expression. Cell Biochem. Funct. 2019, 37, 340–347. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, H.; Liu, C.; Yan, T.; Wang, J.; Liang, W. Long noncoding RNA BDNF-AS is downregulated in cervical cancer and has anti-cancer functions by negatively associating with BDNF. Arch. Biochem. Biophys. 2018, 646, 113–119. [Google Scholar] [CrossRef] [PubMed]
  83. Ghafouri-Fard, S.; Khoshbakht, T.; Taheri, M.; Ghanbari, M. A concise review on the role of BDNF-AS in human disorders. Biomed. Pharmacother. 2021, 142, 112051. [Google Scholar] [CrossRef]
  84. Zhao, H.; Diao, C.; Wang, X.; Xie, Y.; Liu, Y.; Gao, X.; Han, J.; Li, S. LncRNA BDNF-AS inhibits proliferation, migration, invasion and EMT in oesophageal cancer cells by targeting miR-214. J. Cell Mol. Med. 2018, 22, 3729–3739. [Google Scholar] [CrossRef] [PubMed]
  85. Lv, Z.; Xu, Q.; Sun, L.; Wen, J.; Fang, X.; Xing, C.; Yuan, Y. Four novel polymorphisms in long non-coding RNA HOTTIP are associated with the risk and prognosis of colorectal cancer. Biosci. Rep. 2019, 39, BSR20180573. [Google Scholar] [CrossRef] [PubMed]
  86. Xing, C.; Sun, S.G.; Yue, Z.Q.; Bai, F. Role of lncRNA LUCAT1 in cancer. Biomed. Pharmacother. 2021, 134, 111158. [Google Scholar] [CrossRef]
  87. Balik, V.; Srovnal, J.; Sulla, I.; Kalita, O.; Foltanova, T.; Vaverka, M.; Hrabalek, L.; Hajduch, M. MEG3: A novel long noncoding potentially tumour-suppressing RNA in meningiomas. J. Neurooncol. 2013, 112, 1–8. [Google Scholar] [CrossRef]
  88. Smyth, E.C.; Lagergren, J.; Fitzgerald, R.C.; Lordick, F.; Shah, M.A.; Lagergren, P.; Cunningham, D. Oesophageal cancer. Nat Rev Dis Primers 2017, 3, 17048. [Google Scholar] [CrossRef]
  89. Dasari, S.; Tchounwou, P.B. Cisplatin in cancer therapy: Molecular mechanisms of action. Eur. J. Pharmacol. 2014, 740, 364–378. [Google Scholar] [CrossRef]
  90. Rocha, C.R.R.; Silva, M.M.; Quinet, A.; Cabral-Neto, J.B.; Menck, C.F.M. DNA repair pathways and cisplatin resistance: An intimate relationship. Clinics 2018, 73 (Suppl. S1), e478s. [Google Scholar] [CrossRef]
  91. Oliver, T.G.; Mercer, K.L.; Sayles, L.C.; Burke, J.R.; Mendus, D.; Lovejoy, K.S.; Cheng, M.H.; Subramanian, A.; Mu, D.; Powers, S.; et al. Chronic cisplatin treatment promotes enhanced damage repair and tumor progression in a mouse model of lung cancer. Genes. Dev. 2010, 24, 837–852. [Google Scholar] [CrossRef]
  92. Yang, J.; Liu, X.; Cao, S.; Dong, X.; Rao, S.; Cai, K. Understanding Esophageal Cancer: The Challenges and Opportunities for the Next Decade. Front. Oncol. 2020, 10, 1727. [Google Scholar] [CrossRef]
  93. Galluzzi, L.; Senovilla, L.; Vitale, I.; Michels, J.; Martins, I.; Kepp, O.; Castedo, M.; Kroemer, G. Molecular mechanisms of cisplatin resistance. Oncogene 2012, 31, 1869–1883. [Google Scholar] [CrossRef] [PubMed]
  94. Hu, M.; Zhang, Q.; Tian, X.H.; Wang, J.L.; Niu, Y.X.; Li, G. lncRNA CCAT1 is a biomarker for the proliferation and drug resistance of esophageal cancer via the miR-143/PLK1/BUBR1 axis. Mol. Carcinog. 2019, 58, 2207–2217. [Google Scholar] [CrossRef] [PubMed]
  95. Huang, Q.; Xing, S.; Peng, A.; Yu, Z. NORAD accelerates chemo-resistance of non-small-cell lung cancer via targeting at miR-129-1-3p/SOX4 axis. Biosci. Rep. 2020, 40, BSR20193489. [Google Scholar] [CrossRef]
  96. Wang, B.; Xu, L.; Zhang, J.; Cheng, X.; Xu, Q.; Wang, J.; Mao, F. LncRNA NORAD accelerates the progression and doxorubicin resistance of neuroblastoma through up-regulating HDAC8 via sponging miR-144-3p. Biomed. Pharmacother. 2020, 129, 110268. [Google Scholar] [CrossRef] [PubMed]
  97. Shi, X.; Wang, X. The role of MTDH/AEG-1 in the progression of cancer. Int. J. Clin. Exp. Med. 2015, 8, 4795–4807. [Google Scholar] [PubMed]
  98. Jia, Y.; Tian, C.; Wang, H.; Yu, F.; Lv, W.; Duan, Y.; Cheng, Z.; Wang, X.; Wang, Y.; Liu, T.; et al. Long non-coding RNA NORAD/miR-224-3p/MTDH axis contributes to CDDP resistance of esophageal squamous cell carcinoma by promoting nuclear accumulation of beta-catenin. Mol. Cancer 2021, 20, 162. [Google Scholar] [CrossRef]
  99. Liu, G.; Guo, W.; Chen, G.; Li, W.; Cui, Y.; Qin, J.; Peng, J. Lnc-MCEI mediated the chemosensitivity of esophageal squamous cell carcinoma via miR-6759-5p to competitively regulate IGF2. Int. J. Biol. Sci. 2020, 16, 2938–2950. [Google Scholar] [CrossRef] [PubMed]
  100. Hsu, P.K.; Chen, H.Y.; Yeh, Y.C.; Yen, C.C.; Wu, Y.C.; Hsu, C.P.; Hsu, W.H.; Chou, T.Y. TPX2 expression is associated with cell proliferation and patient outcome in esophageal squamous cell carcinoma. J. Gastroenterol. 2014, 49, 1231–1240. [Google Scholar] [CrossRef]
  101. Yang, C.; Shen, S.; Zheng, X.; Ye, K.; Ge, H.; Sun, Y.; Lu, Y. Long non-coding RNA LINC00337 induces autophagy and chemoresistance to cisplatin in esophageal squamous cell carcinoma cells via upregulation of TPX2 by recruiting E2F4. FASEB J. 2020, 34, 6055–6069. [Google Scholar] [CrossRef]
  102. Jiang, L.; Wang, W.; Li, G.; Sun, C.; Ren, Z.; Sheng, H.; Gao, H.; Wang, C.; Yu, H. High TUG1 expression is associated with chemotherapy resistance and poor prognosis in esophageal squamous cell carcinoma. Cancer Chemother. Pharmacol. 2016, 78, 333–339. [Google Scholar] [CrossRef]
  103. Wu, J.; Bao, L.; Zhang, Z.; Yi, X. Nrf2 induces cisplatin resistance via suppressing the iron export related gene SLC40A1 in ovarian cancer cells. Oncotarget 2017, 8, 93502–93515. [Google Scholar] [CrossRef]
  104. Zhang, Z.; Xiong, R.; Li, C.; Xu, M.; Guo, M. LncRNA TUG1 promotes cisplatin resistance in esophageal squamous cell carcinoma cells by regulating Nrf2. Acta Biochim. Biophys. Sin. 2019, 51, 826–833. [Google Scholar] [CrossRef]
  105. Wu, E.R.; Hsieh, M.J.; Chiang, W.L.; Hsueh, K.C.; Yang, S.F.; Su, S.C. Association of lncRNA CCAT2 and CASC8 Gene Polymorphisms with Hepatocellular Carcinoma. Int. J. Environ. Res. Public Health 2019, 16, 2833. [Google Scholar] [CrossRef]
  106. Jiang, X.; Guan, J.; Xu, Y.; Ren, H.; Jiang, J.; Wudu, M.; Wang, Q.; Su, H.; Zhang, Y.; Zhang, B.; et al. Silencing of CASC8 inhibits non-small cell lung cancer cells function and promotes sensitivity to osimertinib via FOXM1. J. Cancer 2021, 12, 387–396. [Google Scholar] [CrossRef] [PubMed]
  107. Wu, Q.; Zhang, H.; Yang, D.; Min, Q.; Wang, Y.; Zhang, W.; Zhan, Q. The m6A-induced lncRNA CASC8 promotes proliferation and chemoresistance via upregulation of hnRNPL in esophageal squamous cell carcinoma. Int. J. Biol. Sci. 2022, 18, 4824–4836. [Google Scholar] [CrossRef] [PubMed]
  108. Tao, C.J.; Zhang, P.; Chen, Y.Y.; Chen, M. High expression of WTAP is related to poor prognosis in nasopharyngeal carcinoma. Neoplasma 2023, 70, 229–239. [Google Scholar] [CrossRef] [PubMed]
  109. Chen, L.; Wang, X. Relationship between the genetic expression of WTAP and bladder cancer and patient prognosis. Oncol. Lett. 2018, 16, 6966–6970. [Google Scholar] [CrossRef] [PubMed]
  110. Zhu, Z.J.; Pang, Y.; Jin, G.; Zhang, H.Y.; Wang, W.H.; Liu, J.W.; Tuo, G.X.; Wu, P.; Yang, Y.; Wang, Z.Q.; et al. Hypoxia induces chemoresistance of esophageal cancer cells to cisplatin through regulating the lncRNA-EMS/miR-758-3p/WTAP axis. Aging 2021, 13, 17155–17176. [Google Scholar] [CrossRef]
  111. Si, M.; Lang, J. The roles of metallothioneins in carcinogenesis. J. Hematol. Oncol. 2018, 11, 107. [Google Scholar] [CrossRef]
  112. Wang, L.; Wang, X.; Yan, P.; Liu, Y.; Jiang, X. LINC00261 Suppresses Cisplatin Resistance of Esophageal Squamous Cell Carcinoma Through miR-545-3p/MT1M Axis. Front. Cell Dev. Biol. 2021, 9, 687788. [Google Scholar] [CrossRef]
  113. Mitsiades, C.S.; Mitsiades, N.; Koutsilieris, M. The Akt pathway: Molecular targets for anti-cancer drug development. Curr. Cancer Drug Targets 2004, 4, 235–256. [Google Scholar] [CrossRef]
  114. Zhu, D.; Yu, Y.; Qi, Y.; Wu, K.; Liu, D.; Yang, Y.; Zhang, C.; Zhao, S. Long Non-coding RNA CASC2 Enhances the Antitumor Activity of Cisplatin Through Suppressing the Akt Pathway by Inhibition of miR-181a in Esophageal Squamous Cell Carcinoma Cells. Front. Oncol. 2019, 9, 350. [Google Scholar] [CrossRef]
  115. Simon, T.; Salhia, B. Cancer-Associated Fibroblast Subpopulations with Diverse and Dynamic Roles in the Tumor Microenvironment. Mol. Cancer Res. 2022, 20, 183–192. [Google Scholar] [CrossRef]
  116. Kalluri, R. The biology and function of fibroblasts in cancer. Nat. Rev. Cancer 2016, 16, 582–598. [Google Scholar] [CrossRef] [PubMed]
  117. Tong, Y.; Yang, L.; Yu, C.; Zhu, W.; Zhou, X.; Xiong, Y.; Wang, W.; Ji, F.; He, D.; Cao, X. Tumor-Secreted Exosomal lncRNA POU3F3 Promotes Cisplatin Resistance in ESCC by Inducing Fibroblast Differentiation into CAFs. Mol. Ther. Oncolytics 2020, 18, 1–13. [Google Scholar] [CrossRef] [PubMed]
  118. 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] [PubMed]
  119. Ilson, D.H. Esophageal cancer chemotherapy: Recent advances. Gastrointest. Cancer Res. 2008, 2, 85–92. [Google Scholar] [PubMed]
  120. Zhang, S.; Zheng, F.; Zhang, L.; Huang, Z.; Huang, X.; Pan, Z.; Chen, S.; Xu, C.; Jiang, Y.; Gu, S.; et al. LncRNA HOTAIR-mediated MTHFR methylation inhibits 5-fluorouracil sensitivity in esophageal cancer cells. J. Exp. Clin. Cancer Res. 2020, 39, 131. [Google Scholar] [CrossRef] [PubMed]
  121. Ogino, S.; Konishi, H.; Ichikawa, D.; Matsubara, D.; Shoda, K.; Arita, T.; Kosuga, T.; Komatsu, S.; Shiozaki, A.; Okamoto, K.; et al. Glutathione S-transferase Pi 1 is a valuable predictor for cancer drug resistance in esophageal squamous cell carcinoma. Cancer Sci. 2019, 110, 795–804. [Google Scholar] [CrossRef] [PubMed]
  122. Mutallip, M.; Nohata, N.; Hanazawa, T.; Kikkawa, N.; Horiguchi, S.; Fujimura, L.; Kawakami, K.; Chiyomaru, T.; Enokida, H.; Nakagawa, M.; et al. Glutathione S-transferase P1 (GSTP1) suppresses cell apoptosis and its regulation by miR-133alpha in head and neck squamous cell carcinoma (HNSCC). Int. J. Mol. Med. 2011, 27, 345–352. [Google Scholar] [CrossRef]
  123. Li, N.; Zhao, Z.; Miao, F.; Cai, S.; Liu, P.; Yu, Y.; Wang, B. Correction: Silencing of long non-coding RNA LINC01270 inhibits esophageal cancer progression and enhances chemosensitivity to 5-fluorouracil by mediating GSTP1methylation. Cancer Gene Ther. 2022, 29, 1299–1300. [Google Scholar] [CrossRef] [PubMed]
  124. Abdelsattar, S.; Fahim, S.A.; Kamel, H.F.M.; Al-Amodi, H.; Kasemy, Z.A.; Khalil, F.O.; Abdallah, M.S.; Bedair, H.M.; Gadallah, A.A.; Sabry, A.; et al. The Potential Role of Circulating Long Miscellaneous RNAs in the Diagnosis and Prognosis of Hepatitis C Related Hepatocellular Carcinoma. Noncoding RNA 2023, 9, 62. [Google Scholar] [CrossRef]
  125. Zhang, G.; Chen, X.; Ma, L.; Ding, R.; Zhao, L.; Ma, F.; Deng, X. LINC01419 facilitates hepatocellular carcinoma growth and metastasis through targeting EZH2-regulated RECK. Aging 2020, 12, 11071–11084. [Google Scholar] [CrossRef]
  126. Cheng, Z.; Hou, S.; Wu, Y.; Wang, X.; Sun, Y.; Liu, B.; Yuan, M. LINC01419 promotes cell proliferation and metastasis in lung adenocarcinoma via sponging miR-519b-3p to up-regulate RCCD1. Biochem. Biophys. Res. Commun. 2019, 520, 107–114. [Google Scholar] [CrossRef]
  127. Chen, J.L.; Lin, Z.X.; Qin, Y.S.; She, Y.Q.; Chen, Y.; Chen, C.; Qiu, G.D.; Zheng, J.T.; Chen, Z.L.; Zhang, S.Y. Overexpression of long noncoding RNA LINC01419 in esophageal squamous cell carcinoma and its relation to the sensitivity to 5-fluorouracil by mediating GSTP1 methylation. Ther. Adv. Med. Oncol. 2019, 11, 1758835919838958. [Google Scholar] [CrossRef] [PubMed]
  128. Chang, Z.W.; Jia, Y.X.; Zhang, W.J.; Song, L.J.; Gao, M.; Li, M.J.; Zhao, R.H.; Li, J.; Zhong, Y.L.; Sun, Q.Z.; et al. LncRNA-TUSC7/miR-224 affected chemotherapy resistance of esophageal squamous cell carcinoma by competitively regulating DESC1. J. Exp. Clin. Cancer Res. 2018, 37, 56. [Google Scholar] [CrossRef]
  129. Weaver, B.A. How Taxol/paclitaxel kills cancer cells. Mol. Biol. Cell 2014, 25, 2677–2681. [Google Scholar] [CrossRef]
  130. Liu, Y.; Ren, Z.; Yuan, L.; Xu, S.; Yao, Z.; Qiao, L.; Li, K. Paclitaxel plus cisplatin vs. 5-fluorouracil plus cisplatin as first-line treatment for patients with advanced squamous cell esophageal cancer. Am. J. Cancer Res. 2016, 6, 2345–2350. [Google Scholar]
  131. Ilson, D.H.; Wadleigh, R.G.; Leichman, L.P.; Kelsen, D.P. Paclitaxel given by a weekly 1-h infusion in advanced esophageal cancer. Ann. Oncol. 2007, 18, 898–902. [Google Scholar] [CrossRef] [PubMed]
  132. Zhang, S.; Jiang, H.; Xu, Z.; Jiang, Y.; She, Y.; Huang, X.; Feng, S.; Chen, W.; Chen, S.; Chen, Y.; et al. The resistance of esophageal cancer cells to paclitaxel can be reduced by the knockdown of long noncoding RNA DDX11-AS1 through TAF1/TOP2A inhibition. Am. J. Cancer Res. 2019, 9, 2233–2248. [Google Scholar] [PubMed]
  133. Martinez-Balibrea, E.; Martinez-Cardus, A.; Gines, A.; Ruiz de Porras, V.; Moutinho, C.; Layos, L.; Manzano, J.L.; Buges, C.; Bystrup, S.; Esteller, M.; et al. Tumor-Related Molecular Mechanisms of Oxaliplatin Resistance. Mol. Cancer Ther. 2015, 14, 1767–1776. [Google Scholar] [CrossRef] [PubMed]
  134. Cunningham, D.; Starling, N.; Rao, S.; Iveson, T.; Nicolson, M.; Coxon, F.; Middleton, G.; Daniel, F.; Oates, J.; Norman, A.R.; et al. Capecitabine and oxaliplatin for advanced esophagogastric cancer. N. Engl. J. Med. 2008, 358, 36–46. [Google Scholar] [CrossRef] [PubMed]
  135. Shen, Z.; Xu, L.; Li, J.; Zhang, N. Capilliposide C Sensitizes Esophageal Squamous Carcinoma Cells to Oxaliplatin by Inducing Apoptosis Through the PI3K/Akt/mTOR Pathway. Med. Sci. Monit. 2017, 23, 2096–2103. [Google Scholar] [CrossRef]
  136. Yue, B.; Cai, D.; Liu, C.; Fang, C.; Yan, D. Linc00152 Functions as a Competing Endogenous RNA to Confer Oxaliplatin Resistance and Holds Prognostic Values in Colon Cancer. Mol. Ther. 2016, 24, 2064–2077. [Google Scholar] [CrossRef]
  137. Zhang, P.; Sun, Y.; Ma, L. ZEB1: At the crossroads of epithelial-mesenchymal transition, metastasis and therapy resistance. Cell Cycle 2015, 14, 481–487. [Google Scholar] [CrossRef]
  138. Zhang, S.; Liao, W.; Wu, Q.; Huang, X.; Pan, Z.; Chen, W.; Gu, S.; Huang, Z.; Wang, Y.; Tang, X.; et al. LINC00152 upregulates ZEB1 expression and enhances epithelial-mesenchymal transition and oxaliplatin resistance in esophageal cancer by interacting with EZH2. Cancer Cell Int. 2020, 20, 569. [Google Scholar] [CrossRef]
  139. Shridhar, R.; Almhanna, K.; Meredith, K.L.; Biagioli, M.C.; Chuong, M.D.; Cruz, A.; Hoffe, S.E. Radiation therapy and esophageal cancer. Cancer Control 2013, 20, 97–110. [Google Scholar] [CrossRef] [PubMed]
  140. Kim, M.S.; Lebron, C.; Nagpal, J.K.; Chae, Y.K.; Chang, X.; Huang, Y.; Chuang, T.; Yamashita, K.; Trink, B.; Ratovitski, E.A.; et al. Methylation of the DFNA5 increases risk of lymph node metastasis in human breast cancer. Biochem. Biophys. Res. Commun. 2008, 370, 38–43. [Google Scholar] [CrossRef] [PubMed]
  141. Lomax, M.E.; Folkes, L.K.; O’Neill, P. Biological consequences of radiation-induced DNA damage: Relevance to radiotherapy. Clin. Oncol. R. Coll. Radiol. 2013, 25, 578–585. [Google Scholar] [CrossRef]
  142. Shrivastav, M.; De Haro, L.P.; Nickoloff, J.A. Regulation of DNA double-strand break repair pathway choice. Cell Res. 2008, 18, 134–147. [Google Scholar] [CrossRef] [PubMed]
  143. Kheyrandish, M.R.; Mir, S.M.; Sheikh Arabi, M. DNA repair pathways as a novel therapeutic strategy in esophageal cancer: A review study. Cancer Rep. 2022, 5, e1716. [Google Scholar] [CrossRef] [PubMed]
  144. Cheng, J.; Liu, W.; Zeng, X.; Zhang, B.; Guo, Y.; Qiu, M.; Jiang, C.; Wang, H.; Wu, Z.; Meng, M.; et al. XRCC3 is a promising target to improve the radiotherapy effect of esophageal squamous cell carcinoma. Cancer Sci. 2015, 106, 1678–1686. [Google Scholar] [CrossRef] [PubMed]
  145. Ma, J.; Benitez, J.A.; Li, J.; Miki, S.; Ponte de Albuquerque, C.; Galatro, T.; Orellana, L.; Zanca, C.; Reed, R.; Boyer, A.; et al. Inhibition of Nuclear PTEN Tyrosine Phosphorylation Enhances Glioma Radiation Sensitivity through Attenuated DNA Repair. Cancer Cell 2019, 35, 504–518.e507. [Google Scholar] [CrossRef] [PubMed]
  146. Chen, M.; Liu, P.; Chen, Y.; Chen, Z.; Shen, M.; Liu, X.; Li, X.; Li, A.; Lin, Y.; Yang, R.; et al. Long Noncoding RNA FAM201A Mediates the Radiosensitivity of Esophageal Squamous Cell Cancer by Regulating ATM and mTOR Expression via miR-101. Front. Genet. 2018, 9, 611. [Google Scholar] [CrossRef] [PubMed]
  147. Liu, W.H.; Qiao, H.Y.; Xu, J.; Wang, W.Q.; Wu, Y.L.; Wu, X. LINC00473 contributes to the radioresistance of esophageal squamous cell carcinoma by regulating microRNA-497-5p and cell division cycle 25A. Int. J. Mol. Med. 2020, 46, 571–582. [Google Scholar] [CrossRef]
  148. Sun, Y.; Wang, J.; Ma, Y.; Li, J.; Sun, X.; Zhao, X.; Shi, X.; Hu, Y.; Qu, F.; Zhang, X. Radiation induces NORAD expression to promote ESCC radiotherapy resistance via EEPD1/ATR/Chk1 signalling and by inhibiting pri-miR-199a1 processing and the exosomal transfer of miR-199a-5p. J. Exp. Clin. Cancer Res. 2021, 40, 306. [Google Scholar] [CrossRef]
  149. Deng, H.; Gao, J.; Cao, B.; Qiu, Z.; Li, T.; Zhao, R.; Li, H.; Wei, B. LncRNA CCAT2 promotes malignant progression of metastatic gastric cancer through regulating CD44 alternative splicing. Cell Oncol. 2023, 46, 1675–1690. [Google Scholar] [CrossRef]
  150. Zhang, Z.; Wang, X.; Wang, Y.; Zhou, D.; Wu, H.; Cheng, W.; Wang, Q.; Zheng, G.; Wang, J.; Gu, J. Effect of long noncoding RNA CCAT2 on drug sensitivity to 5-fluorouracil of breast cancer cells through microRNA-145 meditated by p53. J. Biochem. Mol. Toxicol. 2022, 36, e23176. [Google Scholar] [CrossRef]
  151. Xin, Y.; Li, Z.; Zheng, H.; Chan, M.T.V.; Ka Kei Wu, W. CCAT2: A novel oncogenic long non-coding RNA in human cancers. Cell Prolif. 2017, 50, e12342. [Google Scholar] [CrossRef] [PubMed]
  152. Engeland, K. Cell cycle regulation: p53-p21-RB signaling. Cell Death Differ. 2022, 29, 946–960. [Google Scholar] [CrossRef] [PubMed]
  153. Hermeking, H.; Funk, J.O.; Reichert, M.; Ellwart, J.W.; Eick, D. Abrogation of p53-induced cell cycle arrest by c-Myc: Evidence for an inhibitor of p21WAF1/CIP1/SDI1. Oncogene 1995, 11, 1409–1415. [Google Scholar]
  154. Wang, M.; Wang, L.; He, X.; Zhang, J.; Zhu, Z.; Zhang, M.; Li, X. lncRNA CCAT2 promotes radiotherapy resistance for human esophageal carcinoma cells via the miR-145/p70S6K1 and p53 pathway. Int. J. Oncol. 2020, 56, 327–336. [Google Scholar] [CrossRef] [PubMed]
  155. Cheng, W.; Shi, X.; Lin, M.; Yao, Q.; Ma, J.; Li, J. LncRNA MAGI2-AS3 Overexpression Sensitizes Esophageal Cancer Cells to Irradiation Through Down-Regulation of HOXB7 via EZH2. Front. Cell Dev. Biol. 2020, 8, 552822. [Google Scholar] [CrossRef]
  156. Zhang, H.; Hua, Y.; Jiang, Z.; Yue, J.; Shi, M.; Zhen, X.; Zhang, X.; Yang, L.; Zhou, R.; Wu, S. Cancer-associated Fibroblast-promoted LncRNA DNM3OS Confers Radioresistance by Regulating DNA Damage Response in Esophageal Squamous Cell Carcinoma. Clin. Cancer Res. 2019, 25, 1989–2000. [Google Scholar] [CrossRef] [PubMed]
  157. Ozawa, T.; Matsuyama, T.; Toiyama, Y.; Takahashi, N.; Ishikawa, T.; Uetake, H.; Yamada, Y.; Kusunoki, M.; Calin, G.; Goel, A. CCAT1 and CCAT2 long noncoding RNAs, located within the 8q.24.21 ‘gene desert’, serve as important prognostic biomarkers in colorectal cancer. Ann. Oncol. 2017, 28, 1882–1888. [Google Scholar] [CrossRef]
  158. Huang, C.; Liu, J.; He, L.; Wang, F.; Xiong, B.; Li, Y.; Yang, X. The long noncoding RNA noncoding RNA activated by DNA damage (NORAD)-microRNA-496-Interleukin-33 axis affects carcinoma-associated fibroblasts-mediated gastric cancer development. Bioengineered 2021, 12, 11738–11755. [Google Scholar] [CrossRef]
  159. Zhou, R.T.; Ni, Y.R.; Zeng, F.J. The roles of long noncoding RNAs in the regulation of OCT4 expression. Stem Cell Res. Ther. 2022, 13, 383. [Google Scholar] [CrossRef]
  160. Li, T.; Chen, Y.; Zhang, J.; Liu, S. LncRNA TUG1 promotes cells proliferation and inhibits cells apoptosis through regulating AURKA in epithelial ovarian cancer cells. Medicine 2018, 97, e12131. [Google Scholar] [CrossRef]
  161. Hashemi, M.; Bahari, G.; Naderi, M.; Sadeghi Bojd, S.; Taheri, M. Association of lnc-LAMC2-1:1 rs2147578 and CASC8 rs10505477 Polymorphisms with Risk of Childhood Acute Lymphoblastic Leukemia. Asian Pac. J. Cancer Prev. 2016, 17, 4985–4989. [Google Scholar] [CrossRef]
  162. Priyanka, P.; Sharma, M.; Das, S.; Saxena, S. E2F1-induced lncRNA, EMSLR regulates lncRNA LncPRESS1. Sci. Rep. 2022, 12, 2548. [Google Scholar] [CrossRef]
  163. Mutai, H.; Nagashima, R.; Sugitani, Y.; Noda, T.; Fujii, M.; Matsunaga, T. Expression of Pou3f3/Brn-1 and its genomic methylation in developing auditory epithelium. Dev. Neurobiol. 2009, 69, 913–930. [Google Scholar] [CrossRef]
  164. Shi, J.; Ma, H.; Wang, H.; Zhu, W.; Jiang, S.; Dou, R.; Yan, B. Overexpression of LINC00261 inhibits non-small cell lung cancer cells progression by interacting with miR-522-3p and suppressing Wnt signaling. J. Cell Biochem. 2019, 120, 18378–18387. [Google Scholar] [CrossRef]
  165. Wang, P.; Liu, Y.H.; Yao, Y.L.; Li, Z.; Li, Z.Q.; Ma, J.; Xue, Y.X. Long non-coding RNA CASC2 suppresses malignancy in human gliomas by miR-21. Cell Signal. 2015, 27, 275–282. [Google Scholar] [CrossRef] [PubMed]
  166. Yuan, C.; Ning, Y.; Pan, Y. Emerging roles of HOTAIR in human cancer. J. Cell Biochem. 2020, 121, 3235–3247. [Google Scholar] [CrossRef] [PubMed]
  167. Liu, P.; Zhang, X.; Fu, Q.; Liu, C.; Luo, Q.; Yu, P.; Chen, S.; Zhang, H.; Qin, T. LINC01419 Promotes the Proliferation of Hepatoma Cells by Recruiting XRCC5 and Regulating Its Phosphorylation to Repair DNA Damage. Dis. Markers 2022, 2022, 9313680. [Google Scholar] [CrossRef]
  168. Ren, W.; Chen, S.; Liu, G.; Wang, X.; Ye, H.; Xi, Y. TUSC7 acts as a tumor suppressor in colorectal cancer. Am. J. Transl. Res. 2017, 9, 4026–4035. [Google Scholar] [PubMed]
  169. Feng, Y.; Wu, M.; Hu, S.; Peng, X.; Chen, F. LncRNA DDX11-AS1: A novel oncogene in human cancer. Hum. Cell 2020, 33, 946–953. [Google Scholar] [CrossRef]
  170. Li, S.; Yao, W.; Liu, R.; Gao, L.; Lu, Y.; Zhang, H.; Liang, X. Long non-coding RNA LINC00152 in cancer: Roles, mechanisms, and chemotherapy and radiotherapy resistance. Front. Oncol. 2022, 12, 960193. [Google Scholar] [CrossRef] [PubMed]
  171. He, W.; Qiao, Z.X.; Ma, B. Long noncoding RNA FAM201A mediates the metastasis of lung squamous cell cancer via regulating ABCE1 expression. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 10343–10353. [Google Scholar] [CrossRef]
  172. Mo, J.; Li, B.; Zhou, Y.; Xu, Y.; Jiang, H.; Cheng, X.; Wu, X.; Zhang, Y. LINC00473 promotes hepatocellular carcinoma progression via acting as a ceRNA for microRNA-195 and increasing HMGA2 expression. Biomed. Pharmacother. 2019, 120, 109403. [Google Scholar] [CrossRef]
  173. Pirlog, R.; Drula, R.; Nutu, A.; Calin, G.A.; Berindan-Neagoe, I. The Roles of the Colon Cancer Associated Transcript 2 (CCAT2) Long Non-Coding RNA in Cancer: A Comprehensive Characterization of the Tumorigenic and Molecular Functions. Int. J. Mol. Sci. 2021, 22, 12491. [Google Scholar] [CrossRef]
  174. Kai-Xin, L.; Cheng, C.; Rui, L.; Zheng-Wei, S.; Wen-Wen, T.; Peng, X. Roles of lncRNA MAGI2-AS3 in human cancers. Biomed. Pharmacother. 2021, 141, 111812. [Google Scholar] [CrossRef]
  175. Wang, H.; Ji, X. SMAD6, positively regulated by the DNM3OS-miR-134-5p axis, confers promoting effects to cell proliferation, migration and EMT process in retinoblastoma. Cancer Cell Int. 2020, 20, 23. [Google Scholar] [CrossRef]
  176. Song, W.; Zou, S.B. Prognostic role of lncRNA HOTAIR in esophageal squamous cell carcinoma. Clin. Chim. Acta 2016, 463, 169–173. [Google Scholar] [CrossRef] [PubMed]
  177. Dragomir, M.; Chen, B.; Calin, G.A. Exosomal lncRNAs as new players in cell-to-cell communication. Transl. Cancer Res. 2018, 7 (Suppl. S2), S243–S252. [Google Scholar] [CrossRef] [PubMed]
  178. Ren, J.; Ding, L.; Zhang, D.; Shi, G.; Xu, Q.; Shen, S.; Wang, Y.; Wang, T.; Hou, Y. Carcinoma-associated fibroblasts promote the stemness and chemoresistance of colorectal cancer by transferring exosomal lncRNA H19. Theranostics 2018, 8, 3932–3948. [Google Scholar] [CrossRef] [PubMed]
  179. Luo, X.; Wei, J.; Yang, F.L.; Pang, X.X.; Shi, F.; Wei, Y.X.; Liao, B.Y.; Wang, J.L. Exosomal lncRNA HNF1A-AS1 affects cisplatin resistance in cervical cancer cells through regulating microRNA-34b/TUFT1 axis. Cancer Cell Int. 2019, 19, 323. [Google Scholar] [CrossRef] [PubMed]
  180. Gao, T.; Liu, X.; He, B.; Nie, Z.; Zhu, C.; Zhang, P.; Wang, S. Exosomal lncRNA 91H is associated with poor development in colorectal cancer by modifying HNRNPK expression. Cancer Cell Int. 2018, 18, 11. [Google Scholar] [CrossRef] [PubMed]
  181. Kalluri, R. The biology and function of exosomes in cancer. J. Clin. Investig. 2016, 126, 1208–1215. [Google Scholar] [CrossRef] [PubMed]
  182. Zhang, L.; Yu, D. Exosomes in cancer development, metastasis, and immunity. Biochim. Biophys. Acta Rev. Cancer 2019, 1871, 455–468. [Google Scholar] [CrossRef] [PubMed]
  183. Kourembanas, S. Exosomes: Vehicles of intercellular signaling, biomarkers, and vectors of cell therapy. Annu. Rev. Physiol. 2015, 77, 13–27. [Google Scholar] [CrossRef] [PubMed]
  184. Colao, I.L.; Corteling, R.; Bracewell, D.; Wall, I. Manufacturing Exosomes: A Promising Therapeutic Platform. Trends Mol. Med. 2018, 24, 242–256. [Google Scholar] [CrossRef]
  185. Gelibter, S.; Marostica, G.; Mandelli, A.; Siciliani, S.; Podini, P.; Finardi, A.; Furlan, R. The impact of storage on extracellular vesicles: A systematic study. J. Extracell. Vesicles 2022, 11, e12162. [Google Scholar] [CrossRef]
  186. Winkle, M.; El-Daly, S.M.; Fabbri, M.; Calin, G.A. Noncoding RNA therapeutics—Challenges and potential solutions. Nat. Rev. Drug Discov. 2021, 20, 629–651. [Google Scholar] [CrossRef]
Figure 1. Long noncoding RNA (lncRNA)-controlled pathways in the metastasis of esophageal carcinoma cells (ECC). These pathways involve migration, epithelial–mesenchymal transition (EMT), and invasion, which lead to the metastasis of ECC. Two distinct types of lncRNAs are depicted: oncogenic lncRNAs (blue) promote metastasis, while tumor suppressor lncRNAs (red) attenuate metastasis of ECC.
Figure 1. Long noncoding RNA (lncRNA)-controlled pathways in the metastasis of esophageal carcinoma cells (ECC). These pathways involve migration, epithelial–mesenchymal transition (EMT), and invasion, which lead to the metastasis of ECC. Two distinct types of lncRNAs are depicted: oncogenic lncRNAs (blue) promote metastasis, while tumor suppressor lncRNAs (red) attenuate metastasis of ECC.
Biomedicines 12 00660 g001
Figure 2. Long noncoding RNA (lncRNA)-controlled pathways in chemoresistance/radiotherapy resistance in esophageal carcinoma cells (ECC). (A). These lncRNAs confer cisplatin resistance to ECC cells. (B). In addition to cisplatin, these lncRNAs confer resistance to clinical chemotherapeutics including paclitaxel, oxaliplatin, and 5-fluorouracil (5-FU), and these lncRNAs confer radioresistance to patients with ECC. Oncogenic lncRNAs (blue) promote chemoresistance and radiotherapy resistance, while tumor suppressor lncRNAs (red) attenuate chemoresistance and radiotherapy resistance in ECC.
Figure 2. Long noncoding RNA (lncRNA)-controlled pathways in chemoresistance/radiotherapy resistance in esophageal carcinoma cells (ECC). (A). These lncRNAs confer cisplatin resistance to ECC cells. (B). In addition to cisplatin, these lncRNAs confer resistance to clinical chemotherapeutics including paclitaxel, oxaliplatin, and 5-fluorouracil (5-FU), and these lncRNAs confer radioresistance to patients with ECC. Oncogenic lncRNAs (blue) promote chemoresistance and radiotherapy resistance, while tumor suppressor lncRNAs (red) attenuate chemoresistance and radiotherapy resistance in ECC.
Biomedicines 12 00660 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Weng, Z.-P.; Hsu, S.-K.; Wang, H.-M.D.; Chen, K.-J.; Lee, P.-Y.; Chiu, C.-C.; Cheng, K.-C. The Role of Long Noncoding RNAs (lncRNAs) in Esophageal Cancer Therapy Resistance and Metastasis. Biomedicines 2024, 12, 660. https://doi.org/10.3390/biomedicines12030660

AMA Style

Weng Z-P, Hsu S-K, Wang H-MD, Chen K-J, Lee P-Y, Chiu C-C, Cheng K-C. The Role of Long Noncoding RNAs (lncRNAs) in Esophageal Cancer Therapy Resistance and Metastasis. Biomedicines. 2024; 12(3):660. https://doi.org/10.3390/biomedicines12030660

Chicago/Turabian Style

Weng, Zong-Ping, Shen-Kai Hsu, Hui-Min David Wang, Kuo-Jen Chen, Po-Yen Lee, Chien-Chih Chiu, and Kai-Chun Cheng. 2024. "The Role of Long Noncoding RNAs (lncRNAs) in Esophageal Cancer Therapy Resistance and Metastasis" Biomedicines 12, no. 3: 660. https://doi.org/10.3390/biomedicines12030660

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