Abstract
Cancer is difficult to cure due to frequent metastasis, and developing effective therapeutic approaches to treat cancer is urgently important. Long non-coding RNAs (lncRNAs) have diverse roles in regulating gene expression at both the transcriptional and translational levels and have been reported to be involved in tumorigenesis and tumor metastasis. In this article, we review the emerging roles of lncRNAs in cancer, especially in cancer immunity, cancer metabolism and cancer metastasis. We also discuss the use of novel technologies, such as antisense oligonucleotides, CRISPR-Cas9 and nanomedicines, to target lncRNAs and thus control cancers.
Keywords: LncRNA, tumorigenesis, tumor metastasis, ASO, CRISPR-Cas9, virus, nanomedicine
Introduction
Cancer is among the most life-threatening diseases, and its morbidity and mortality rank first or second among noncommunicable diseases [1]. According to global cancer data from 2018, cancer morbidity and mortality are increasing annually with the rapid growth of the population and the problem of the aging population. The number of cancer deaths worldwide reached 9.55 million, with an incidence rate of 18.08 million [2]. During the past several decades, great success in treating cancer has been achieved. However, the survival time of most cancer patients is still poor, especially that of advanced cancer patients with metastasis. There is an urgent need to understand more about the molecular mechanisms governing tumor progression and to develop more effective clinical strategies for cancer treatment.
RNA-based therapeutics against cancer has gradually changed from concept to reality [3,4]. Among these therapeutics, non-coding RNA (ncRNA), which refers to a class of RNA that does not encode protein, exerts clinical therapeutic effects against tumors by inhibiting the transcription of mRNA and binding to protein to block its function [5]. According to the molecular size of ncRNA, it can be classified as either small non-coding RNA (sncRNA), measuring under 200 nucleotides in length, or long non-coding RNA (lncRNA), measuring over 200 nucleotides in length [6].
An increasing number of studies have documented that lncRNAs play diverse roles in regulating gene transcription, post-transcription, translation, and epigenetic modification. Aberrant expression or dysfunction of lncRNA is closely associated with various diseases [7-10]. LncRNAs may regulate cell proliferation, apoptosis, migration, invasion and maintenance of stemness during cancer development [11,12]. Recent studies have shown that lncRNAs may also engage in remodeling the tumor microenvironment and tumor metastasis. In this review, we will discuss the emerging roles of lncRNAs in tumorigenesis, tumor immunity, tumor metabolism and tumor metastasis. Considering the pivotal roles of lncRNAs in cancer, lncRNA-based therapeutics may represent promising approaches in treating cancer.
LncRNAs in tumorigenesis and tumor progression
Tumor formation is not only marked by the unrestricted expansion of tumor cells but also includes many biological processes, such as genomic mutation, DNA damage, immune escape, and metabolic disorder. LncRNAs are distributed in both the nucleus and cytoplasm and serve as important post-transcriptional translational regulators in these processes (Figure 1).
Figure 1.
LncRNA in tumorigenesis and tumor metastasis. Divided into six subtypes: Immune escape, DNA damage, Metabolic disorders, EMT, Cell stemness, Chemical resistance.
DNA damage
DNA damage caused by exogenous factors and chemicals or its abnormal cellular localization could transmit dangerous signals in cells [13] (Figure 1). The accumulation of DNA damage is one of the symbols of tumor progression. Many molecules participate in facilitating or inhibiting DNA damage.
P53, which has been extensively studied as a tumor suppressor, plays important roles in DNA damage. The lncRNA MEG3 activates p53 to exert its anti-cancer effect [14]. Many other lncRNAs are associated with the downstream activities of p53 [15-17]. When DNA is damaged, the transcription of the lncRNA damage induced noncoding (DINO) is activated via p53, thereby controlling the stress response after DNA damage [18]. Additionally, specific expression of lncRNA-DINO activates the impaired signaling pathway and cell cycle arrest in the absence of DNA damage [18] (Figure 1). In addition, lncRNA CCND1-upstream intergenic DNA repair 1 and 2 (CUPID1 and CUPID2) involves in the progression of breast cancer by modulating the stress response to DNA damage [19]. These studies indicate that lncRNAs respond to DNA damage and may be involved in DNA repair, which is a fundamental issue for carcinogenesis.
Immune escape
Immune escape has long been considered as one of the hallmarks of cancer. Tumor cells may train macrophages and regulatory T cells (Treg), forming an inseparable “fortress”, to escape from threats of killer T cells. LncRNAs have been documented to involve innate and adaptive immune responses by modulating the functional status of immune cells [20-24]. For instance, lnc-EGFR promoted immune escape of hepatocellular carcinoma cells by stimulating the differentiation of Treg cells [25]. In the tumor microenvironment, tumor-associated macrophages display limited phagocytosis function and promote the progression of cancer. LncRNA lymph node metastasis associated transcript 1 (LNMAT1) was involved in the regulation of CCL2 recruiting macrophages into the tumor [26] (Figure 1).
In tumor-specific cytotoxic T lymphocytes (CTLs) and type 1 helper T (TH1) cells, an NF-κB-interacting long noncoding RNA (NKILA) may enhance T cell sensitivity to activation-induced cell death by mechanically inhibiting the NF-κB signaling pathway [27]. In addition, lncRNA CamK-A participates in the remodeling of the tumor microenvironment via activation of Ca2+-triggered signaling [28] (Figure 1).
Taken together, these evidences suggest that lncRNAs may be pivotal regulators in remodeling the tumor immune microenvironment.
Metabolic disorders
Cellular metabolic disorder is one of the most prominent characteristics of cancer. Abnormal cellular metabolic processes not only provide energy for the proliferation of cancer cells, but also maintain cellular redox homeostasis by inhibiting reactive oxygen species production. Notably, the proportion of cellular metabolites ATP/AMP is altered by various stimulations. Energy stress may increase the ratio of AMP/ATP which activates AMP-activated protein kinase (AMPK) [29] (Figure 1).
Under energy stress, the lncRNA neighbor of BRCA1 gene 2 (NBR2) activated AMPK via direct binding. Knockdown of lncRNA-NBR2 lead to cell metabolism disorders and subsequently promoted cell proliferation [30]. Mitochondria are the center of energy metabolism, and their homeostasis is also affected by lncRNA. The lncRNA-SAMMSON bound to the major mitochondrial regulator p32 protein in melanoma cells and enhanced its cancer-promoting function [31] (Figure 1).
In addition, glycolysis replacing oxidative phosphorylation is the principal mode of energy metabolism in cancer cells. Hypoxia-inducible factor 1-alpha (HIF-1α) plays an important role in this process. Recent studies have reported an interaction between HIF-1α and lncRNAs. Long intergenic non-coding RNA for kinase activation (LINK-A) regulated the phosphorylation of HIF-1α, maintained its stability, and activated the transcriptional program of HIF-1α to promote tumorigenesis of triple-negative breast cancer (TNBC) [32]. HIF-1α up-regulated the expression of lincRNA-p21 and participated in tumor formation by regulating the Warburg effect [33]. Long noncoding HIF-1α co-activating RNA (LncHIFCAR) is a co-activator of HIF-1α that drives the progression of oral cancer [34] (Figure 1).
The redox reaction is inhibited during the proliferation of tumors, and ATP is mainly formed by the decomposition of pyruvate by lactate dehydrogenase, creating a hypoxia microenvironment inside of the tumor. In a hypoxic environment, histone deacetylase 3 inhibits the expression of lncRNA low expression in tumor (LET) by reducing histone acetylation of the lncRNA-LET promoter region, and the low expression of lncRNA-LET is a key step in stabilizing the nuclear factor 90 protein, thereby promoting cancer cell invasion [35] (Figure 1).
In addition, lncRNA HOXB-AS3 participated in the metabolism of cancer by affecting the expression of a conserved 53-amino acid peptide [36], while lncRNA FoxO-induced long non-coding RNA 1 (FILNC1) functioned as a tumor suppressor. The down-regulation of FILNC1 enhanced glucose metabolism and lactic acid production by increasing the expression of c-Myc [37] (Figure 1).
These evidences suggest that lncRNA is involved in many aspects of cell metabolism, such as ATP production, the hypoxic environment, and Warburg effect regulation. Therefore, these lncRNAs may serve as potential therapeutic targets through inhibiting tumor energy production and reprogramming its growth microenvironment.
LncRNA in tumor metastasis
EMT
Epithelial-mesenchymal transition (EMT) is a complex multi-step biological process that is orchestrated by a variety of EMT-inducing transcription factors. Briefly, epithelial-like cells transdifferentiate into mesenchymal-like cells, facilitating their invasion and migration into blood vessels and lymphatic vessels, thereby participating in the metastasis of a variety of cancers [38-41]. Previous studies have also found that lncRNAs are involved in the regulation of EMT in tumors [42] (Figure 1).
Transforming growth factor β (TGF-β) acts as an initial agonist in EMT. It promoted cell migration and invasion by inducing the occurrence of EMT [43]. LncRNA activated by TGF-β (lncRNA-ATB) induced distant metastasis of liver cancer by up-regulating the levels of zinc finger E-box-binding homeobox (ZEB1 and ZEB2) to stimulate the EMT cascade [12]. In addition, TGF-β induced the production of LIN28B, which participated in the development of pancreatic ductal adenocarcinoma (PDAC) [44]. LncRNA MEG3 participated in the TGF-β signaling pathway via RNA-DNA triplex structures [45] (Figure 1).
In addition, lncRNA human ortholog RNA of Dreh (hDREH) was down-regulated by hepatitis B virus X protein (HBx), which is an inhibitor of EMT in hepatocellular carcinoma (HCC) [46]. Further studies have revealed that the transcription factor PNUTS has its corresponding lncRNA-PNUTS, which is involved in the metastasis of breast cancer by affecting the EMT process [47] (Figure 1).
HOTAIR is also known to promote the metastasis of various cancers, such as breast cancer, liver cancer, and pancreatic cancer [42,48-50]. TGF-β secreted by carcinoma-associated fibroblasts stimulated the expression of HOTAIR in cancer cells to activate the SMAD cascade signaling pathway and subsequently induced the EMT process and promoted cancer metastasis [51] (Figure 1).
Collectively, many lncRNAs have been documented to regulate the EMT process during tumor metastasis. However, EMT is an intricate multi-cascade process. The roles of lncRNAs in the trans-vascular migration process and vascular circulation require more in-depth research. On the other hand, it is known that spread tumor cells may undergo mesenchymal-epithelial transition (MET) when they arrive at the distant organ, which facilitates the growth of overt metastasis lesions. Whether lncRNAs also participate in the MET process remains an intriguing issue.
Cancer cell stemness
Stemness is an important property of tumor metastasis-initiating cells. After colonization of spread tumor cells into distant tissues, tumor metastasis-initiating cells with high stemness can survive and form micrometastases and subsequent overt macrometastases. This step is critical for tumor cell colonization and metastasis and is coordinated by various signaling pathways.
Various studies have shown that lncRNAs are involved in stemness-related signaling pathways. For example, lncRNA tumor suppressor long noncoding RNA on chromosome 8p12 (TSLNC8) exerted a tumor suppressor function by inhibiting the STAT3 signaling pathway [52], and lncRNA MST1/2-antagonizing for YAP activation (MAYA) participated in the Hippo-YAP signaling pathway [53]. These lncRNAs might be directly regulating cell stemness [52,53] (Figure 1). LncBRM stimulated YAP1 to regulate the self-renewal of liver cancer stem cells [54]. LncSox4 participated in the self-renewal of liver cancer-initiating cells through the STAT3 pathway [55]. The expression of lncTCF7 recruited the SWI/SNF complex and further activated the Wnt signaling pathway based on the activation of the TCF7 transcriptional promoter, thereby promoting the stemness of cancer cells [56] (Figure 1).
In summary, lncRNA plays a pivotal role in cancer metastasis. Emerging findings in this field have revealed more previously unknown functions of lncRNA in tumor progression and metastasis. Thus, targeting lncRNAs as a promising approach to treat cancer has attracted interest from researchers in recent years.
Therapeutic opportunities of lncRNA in controlling cancer
Non-coding RNAs play a significant role in tumorigenesis and tumor progression [57]. LncRNAs may be promising targets for controlling cancer. Some efforts have been paid to the lncRNA therapy in animal models via various methods. In this section, we list four types of approaches targeting lncRNAs in cancer treatment (Figure 2).
Figure 2.
Targeting lncRNA to treat cancer. ① ASOs-mediated knock down of cytoplasmic and nuclear lncRNA transcription levels through RNaseH-dependent degradation. ② CRISPR-cas9 based knock-out strategy of lncRNA via specific gDNA. ③ Two methods of virus therapy, encapsulated shRNA or lncRNA-mediated knock down or up-regulation of targeted lncRNA, or NATs-mediated down-regulation of lncRNA elevating the expression of neighboring tumor suppressor genes. ④ Nano-carrier-absorbed on nanodrugs detached within cytoplasm after specific stimulation.
Antisense oligonucleotides
Antisense oligonucleotides (ASOs), which may form a DNA-RNA structure with target RNA through base pairing rules, can trigger RNase-H-mediated RNA degradation. ASOs have been clinically tested for targeting mRNA in cancer [58]. As we have reviewed above, the aberrant expression of lncRNAs regulates tumorigenesis and tumor progression [59-63]. Targeting lncRNA by ASO may be a promising method for treating cancer.
Knocking down metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) by ASOs significantly inhibited tumor growth and metastasis, including breast cancer [64] and lung cancer [65]. Prostate cancer is mild in most patients, while only a small percentage of patients have a definite deterioration in symptoms [66,67]. One study has found that a small percentage of patients with high expression of the lncRNA second chromosome locus associated with prostate-1 (SChLAP1) showed significantly limited tumor formation and metastasis after ASO-mediated down-regulation of lncRNA-SChLAP1 [57], providing idea for the clinical treatment of malignant prostate cancer. In addition, targeting lnc-USMycN by LNA-ASO markedly inhibited tumor formation in mice with neuroblastoma [68].
Due to the poor membrane permeability of ASOs, ASOs are mainly constricted within the cytoplasm and it is difficult for ASOs to manipulate sub-nucleus lncRNAs [69]. Although recent studies have claimed the abundance of RNase-H inside nucleus, it is still difficult to obtain accurate therapeutic effects. Linking ASOs with nanotechnology may be a latent method and we will discuss them in detail in the following sections (Figure 2).
CRISPR/Cas9 genome editing technique
As a technology for specific DNA modification of targeted genes, CRISPR/Cas9 has also received extensive attention in the treatment of cancer. Recent studies have found that CRISPR/Cas9 can successfully silence the transcription of the lncRNA-expressing loci [70,71]. CRISPR/Cas9 targeted the transcriptional site of a gene promoter to silence transcription [72]. Studies have found that more than 16,000 lncRNA promoters in the human genome could be targeted by guide RNAs [73].
The CRISPR/Cas9 system has been used to target genomic DNA in cancer cells and animal cancer models. For example, knockout of lncRNA-NEAT1 and lncRNA-MALAT1 dramatically inhibited the metastasis of cancer cells [74,75]. LncRNA-NEAT1 is involved in the regulation of the replication stress response and chemosensitivity of cancer cells. Knockout of LncRNA-NEAT1 induced the sensitivity of precancerous cells towards DNA damage-induced cell death and promoted the lethality of chemotherapy drugs on cancer cells [74]. Previously, we reported that lncRNA-GMAN is a gastric cancer metastasis-associated long non-coding RNA. GMAN is highly expressed in gastric cancer cells and is associated with poor prognosis [76]. A well-designed proof-of-concept animal experiment found that the delivery of a CRISPR/Cas9 system targeting GMAN significantly suppressed the metastasis of gastric cancer cells and improved overall survival in mice [76].
Many lncRNAs are expressed specifically in different tissues and even different people. Therefore, it is clinically possible to make personalized treatment depending on the situation of patients. Although CRISPR/Cas9 has broad adaptability and target specificity as a genome editor theoretically, off-target cleavage events can still occur in practical applications [77,78]. Therefore, oncologists should be more cautious in designing gene-editing therapy. Now, the clinical application of the CRISPR/Cas9 system targeting lncRNA to treat cancer can be vague. Moreover, developing more specific gene-editing tools is important [79] (Figure 2).
Virus
As a superior RNA interference (RNAi) transfection method, viral vectors mainly include recombinant vectors of adenovirus, lentivirus and retrovirus. RNAi is a biological process of specific gene knockdown via neutralizing targeted RNA by exogenous double-stranded RNA, first found in Caenorhabditis elegans [80-82], which includes short interfering RNAs (siRNAs) and short hairpin RNAs (shRNAs). Despite its specificity, siRNA’s efficiency is transient due to its instability, while stem-loop shRNA may provide a durable and long-lasting effect in vivo [83-86].
In particular, the application of adenovirus vectors is far more extensive, and there have been emerging clinical trials [87-89]. Adeno-associated viruses (AAVs) are structures of uncoated, single-stranded DNA [90]. The AAV-based vector is an efficient gene delivery system, mainly due its non-pathogenicity, free from immune response and stability within live cells [91]. After large-scale screening, more ideal AAVs have been developed for human cancer cells [92]. AAVs have laid a solid foundation for the clinical treatment of tumors by targeting lncRNA.
There have been many reports on the use of shRNAs to target lncRNAs in treating cancer. In a recent study, the lncRNA-BCAR4 knockdown cell line constructed by lentiviral transfection significantly inhibited the formation of metastases in breast cancer in vivo in mice [93]. In addition, transfecting the HOTAIR shRNA with a retrovirus in a gastric cancer cell line significantly inhibited cells from spreading in the peritoneal dissemination [94]. It was also found that knockdown of lncRNA-PNUTS by an adenovirus system could reduce the formation of primary breast cancer and metastases via inhibiting the expression of metastatic Ki-67 [47] (Figure 2).
On the other hand, some lncRNAs with tumor suppressor function are expressed at low levels in tumors. It becomes feasible to upregulate the expression of these lncRNAs to achieve cancer treatment. Virus transfection, as the main method for accurately transmitting the shRNA plasmid to the target site, can also be used to transfect exogenously synthesized lncRNA plasmids into cancer cells to upregulate corresponding lncRNAs. However, solid experimental data are required to verify the feasibility and practicability of this method (Figure 2).
There is also a small class of special non-coding RNAs in the human genome, natural antisense RNAs (NATs), which belong to non-coding RNAs and are antisense to the overlapped protein-coding gene [95]. Studies have shown that inhibition of NATs expression can upregulate the expression of neighbor/overlapping coding genes [96,97]. Therefore, it is possible to apply a method such as ASO to knock down NATs that are adjacent to or overlap with certain tumor suppressor genes, such as CDKN2B (ANRIL) and CDKN1A (P21-AS) [98,99], and then upregulate their expression to treat cancer (Figure 2).
Although viral transfection has achieved excellent therapeutic results in basic research, the complexity of clinical trials is considerable. It is noteworthy that the off-target effect has species differences [100,101]. Therefore, dose control and improved accuracy of viral infections should be seriously considered in future applications.
Nanomedicine
After being proposed in the 1990s, nanotechnology gradually gained prominence due to its small size, biodegradability, ability to covalently combine with a large variety of small molecule drugs, and ability to reach sub-nucleus targets. As a consequence, its potential in treating cancer and related diseases has gradually developed.
Nanomedicine is usually composed of four main structures: drug, targeting agent, imaging agent, and linker. In the classification of third-generation nanomedicines, five types of nanomedicines have been proposed based on the specific resistance of cancer, including the following: (I) Lipid-based nanoparticles (liposomes), which are lipid-based vesicles that are capable of carrying payloads in either an aqueous compartment or embedded in the lipid bilayer and have been previously used in conjunction with paclitaxel to target human epidermal growth factor receptor 2 (HER2) for clinical trials [102]. (II) Polymer-based nanoparticles and micelles are composed of biodegradable or natural multimer covalently cross-linked therapeutic particles. Among these nanoparticles and micelles, two multimers, polylactide (PLA) and poly (lactide-co-glycolide) (PLGA), have been used to synthesize FDA-approved nanomedicines [103]. (III) Dendrimers are well-defined globular structures with a central core composed of multiple branched polymers. Clinical trials using dendrimer complex structures to transport paclitaxel for the targeted treatment of breast cancer and non-small-cell lung cancer have been conducted [104]. (IV) Carbon-based nanoparticles mainly release drug particles into the cytoplasm through pinhole-like penetration. Therefore, the constituent materials of carbon-based nanoparticles must be fine and small in size, as well as biodegradable. Previous studies have used this material in the treatment of cancer [105]. (V) Metallic and magnetic nanoparticles, such as gold nanoparticles, can be used to transport small molecules, such as proteins, and can release covalently bound drugs through photophysical properties, such as tumor necrosis factor α (TNFα) bound to colloidal gold for the treatment of solid tumors [106,107].
Although scientists have developed diverse nanomedicines, multi-drug resistance (MDR) in cancer cells is still challenging. P-glycoprotein, which is highly expressed on the surface of cancer cells, can activate the efflux of anti-cancer drugs within tumor cells, leading to a significant decrease in drug potency. Similarly, lncRNAs can also regulate cancer cell sensitivity towards different types of drugs. LncRNA urothelial cancer-associated 1 (UCA1) is involved in the chemical resistance of bladder cancer by regulating Wnt6 [108], and lncRNA MACC1-AS1 is involved in the chemical resistance of gastric cancer through fatty acid oxidation [109].
Additionally, a system has been established for the stable transportation of nanomedicine into the nucleus. Mechanically, the nanodrug binds to the nanotruck through an aptamer, which facilitates its cell membrane permeability and then detaches the nanodrug from the nanotruck through near-infrared (NIR)-range (700-900 nm) radiation in the cytoplasm [110,111], eventually entering into the nucleus [111]. Since the majority of lncRNAs are located in the cell nucleus, via this system, sub-nucleus lncRNAs can be accurately and effectively targeted to obtain the desired therapeutic effect (Figure 2).
Conclusions
Numerous studies have documented that lncRNA plays a key role in tumorigenesis and tumor progression. In particular, abnormal lncRNA expression may accompany DNA damage, immune escape as well as cellular metabolic disorders in cancer cells. The diversity and heterogeneity of lncRNAs make the complicated tumorigenesis process even more intriguing. In addition, lncRNA is also strongly associated with EMT, as well as the regulation of cell stemness. These findings together make lncRNA a solid component of tumor metastasis. Therefore, targeting lncRNA could be an opportune clinical approach in cancer treatment.
Combination therapy has made great progress in the clinical treatment of tumors. Combining surgery with chemotherapy, or the emergence of specific targeted-drugs, has further improved the survival rate of patients. Recently, with the significant clinical achievements in immunotherapy of PD1/PD-L1, ideas focusing on the re-stimulation of the suppressed cellular immunity of patients through immunotherapy have been under the limelight of cancer research. As a part of the immune escape of cancer cells, lncRNAs may provide new insight for future cancer treatment combined with immunotherapy.
The rapid development of a new generation of gene-editing tools makes it possible to target lncRNA inside tumor cells. ASOs or CRISPR/Cas9-based therapy has already shown the feasibility of gene editing in treating cancer, but their off-target event or unstable efficiency due to the spatiotemporal specificity of lncRNA should also be carefully evaluated before further application. Moreover, the combination of nanotechnology and bioinformatics accelerates new nanoparticle development as well as its optimization under clinical conditions via the deep analysis of lncRNA functions and distribution. The emergence of new cancer therapeutic strategies worldwide exhibits promise in the treatment of cancer, one of the most serious human diseases.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (31771540, 91740205 and 31301149). Natural Scientific Foundation of Zhejiang Province, China (LYY19H310011, LQ13H160013), and Fundamental Research Funds for the Central Universities (2017QNA7005).
Disclosure of conflict of interest
None.
References
- 1.GBD 2013 Mortality and Causes of Death Collaborators. Global, regional, and national age-sex specific all-cause and cause-specific mortality for 240 causes of death, 1990-2013: a systematic analysis for the global burden of disease study 2013. Lancet. 2015;385:117–171. doi: 10.1016/S0140-6736(14)61682-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Bray F, Ferlay J, Soerjomataram I, Siegel RL, Torre LA, Jemal A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2018;68:394–424. doi: 10.3322/caac.21492. [DOI] [PubMed] [Google Scholar]
- 3.Burnett JC, Rossi JJ. RNA-based therapeutics: current progress and future prospects. Chem Biol. 2012;19:60–71. doi: 10.1016/j.chembiol.2011.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Drolet DW, Green LS, Gold L, Janjic N. Fit for the eye: aptamers in ocular disorders. Nucleic Acid Ther. 2016;26:127–146. doi: 10.1089/nat.2015.0573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Sullenger BA, Nair S. From the RNA world to the clinic. Science. 2016;352:1417–1420. doi: 10.1126/science.aad8709. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Iyer MK, Niknafs YS, Malik R, Singhal U, Sahu A, Hosono Y, Barrette TR, Prensner JR, Evans JR, Zhao S, Poliakov A, Cao X, Dhanasekaran SM, Wu YM, Robinson DR, Beer DG, Feng FY, Iyer HK, Chinnaiyan AM. The landscape of long noncoding RNAs in the human transcriptome. Nat Genet. 2015;47:199–208. doi: 10.1038/ng.3192. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Wang KC, Chang HY. Molecular mechanisms of long noncoding RNAs. Mol Cell. 2011;43:904–914. doi: 10.1016/j.molcel.2011.08.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Fang Y, Fullwood MJ. Roles, functions, and mechanisms of long non-coding RNAs in cancer. Genomics Proteomics Bioinformatics. 2016;14:42–54. doi: 10.1016/j.gpb.2015.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Mercer TR, Dinger ME, Mattick JS. Long non-coding RNAs: insights into functions. Nat Rev Genet. 2009;10:155–159. doi: 10.1038/nrg2521. [DOI] [PubMed] [Google Scholar]
- 10.Cesana M, Cacchiarelli D, Legnini I, Santini T, Sthandier O, Chinappi M, Tramontano A, Bozzoni I. A long noncoding RNA controls muscle differentiation by functioning as a competing endogenous RNA. Cell. 2011;147:358–369. doi: 10.1016/j.cell.2011.09.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Batista PJ, Chang HY. Long noncoding RNAs: cellular address codes in development and disease. Cell. 2013;152:1298–1307. doi: 10.1016/j.cell.2013.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Yuan JH, Yang F, Wang F, Ma JZ, Guo YJ, Tao QF, Liu F, Pan W, Wang TT, Zhou CC, Wang SB, Wang YZ, Yang Y, Yang N, Zhou WP, Yang GS, Sun SH. A long noncoding RNA activated by TGF-beta promotes the invasion-metastasis cascade in hepatocellular carcinoma. Cancer Cell. 2014;25:666–681. doi: 10.1016/j.ccr.2014.03.010. [DOI] [PubMed] [Google Scholar]
- 13.Napirei M, Karsunky H, Zevnik B, Stephan H, Mannherz HG, Moroy T. Features of systemic lupus erythematosus in Dnase1-deficient mice. Nat Genet. 2000;25:177–181. doi: 10.1038/76032. [DOI] [PubMed] [Google Scholar]
- 14.Zhou Y, Zhong Y, Wang Y, Zhang X, Batista DL, Gejman R, Ansell PJ, Zhao J, Weng C, Klibanski A. Activation of p53 by MEG3 non-coding RNA. J Biol Chem. 2007;282:24731–24742. doi: 10.1074/jbc.M702029200. [DOI] [PubMed] [Google Scholar]
- 15.Huarte M, Guttman M, Feldser D, Garber M, Koziol MJ, Kenzelmann-Broz D, Khalil AM, Zuk O, Amit I, Rabani M, Attardi LD, Regev A, Lander ES, Jacks T, Rinn JL. A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell. 2010;142:409–419. doi: 10.1016/j.cell.2010.06.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Hung T, Wang Y, Lin MF, Koegel AK, Kotake Y, Grant GD, Horlings HM, Shah N, Umbricht C, Wang P, Wang Y, Kong B, Langerod A, Borresen-Dale AL, Kim SK, van de Vijver M, Sukumar S, Whitfield ML, Kellis M, Xiong Y, Wong DJ, Chang HY. Extensive and coordinated transcription of noncoding RNAs within cell-cycle promoters. Nat Genet. 2011;43:621–629. doi: 10.1038/ng.848. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Marin-Bejar O, Marchese FP, Athie A, Sanchez Y, Gonzalez J, Segura V, Huang L, Moreno I, Navarro A, Monzo M, Garcia-Foncillas J, Rinn JL, Guo S, Huarte M. Pint lincRNA connects the p53 pathway with epigenetic silencing by the Polycomb repressive complex 2. Genome Biol. 2013;14:R104. doi: 10.1186/gb-2013-14-9-r104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Schmitt AM, Garcia JT, Hung T, Flynn RA, Shen Y, Qu K, Payumo AY, Peres-da-Silva A, Broz DK, Baum R, Guo S, Chen JK, Attardi LD, Chang HY. An inducible long noncoding RNA amplifies DNA damage signaling. Nat Genet. 2016;48:1370–1376. doi: 10.1038/ng.3673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Betts JA, Moradi Marjaneh M, Al-Ejeh F, Lim YC, Shi W, Sivakumaran H, Tropee R, Patch AM, Clark MB, Bartonicek N, Wiegmans AP, Hillman KM, Kaufmann S, Bain AL, Gloss BS, Crawford J, Kazakoff S, Wani S, Wen SW, Day B, Moller A, Cloonan N, Pearson J, Brown MA, Mercer TR, Waddell N, Khanna KK, Dray E, Dinger ME, Edwards SL, French JD. Long noncoding RNAs CUPID1 and CUPID2 mediate breast cancer risk at 11q13 by modulating the response to DNA damage. Am J Hum Genet. 2017;101:255–266. doi: 10.1016/j.ajhg.2017.07.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Heward JA, Lindsay MA. Long non-coding RNAs in the regulation of the immune response. Trends Immunol. 2014;35:408–419. doi: 10.1016/j.it.2014.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Atianand MK, Hu W, Satpathy AT, Shen Y, Ricci EP, Alvarez-Dominguez JR, Bhatta A, Schattgen SA, McGowan JD, Blin J, Braun JE, Gandhi P, Moore MJ, Chang HY, Lodish HF, Caffrey DR, Fitzgerald KA. A long noncoding RNA lincRNA-EPS acts as a transcriptional brake to restrain inflammation. Cell. 2016;165:1672–1685. doi: 10.1016/j.cell.2016.05.075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Wang P, Xue Y, Han Y, Lin L, Wu C, Xu S, Jiang Z, Xu J, Liu Q, Cao X. The STAT3-binding long noncoding RNA lnc-DC controls human dendritic cell differentiation. Science. 2014;344:310–313. doi: 10.1126/science.1251456. [DOI] [PubMed] [Google Scholar]
- 23.Ranzani V, Rossetti G, Panzeri I, Arrigoni A, Bonnal RJ, Curti S, Gruarin P, Provasi E, Sugliano E, Marconi M, De Francesco R, Geginat J, Bodega B, Abrignani S, Pagani M. The long intergenic noncoding RNA landscape of human lymphocytes highlights the regulation of T cell differentiation by linc-MAF-4. Nat Immunol. 2015;16:318–325. doi: 10.1038/ni.3093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Hu G, Tang Q, Sharma S, Yu F, Escobar TM, Muljo SA, Zhu J, Zhao K. Expression and regulation of intergenic long noncoding RNAs during T cell development and differentiation. Nat Immunol. 2013;14:1190–1198. doi: 10.1038/ni.2712. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Jiang R, Tang J, Chen Y, Deng L, Ji J, Xie Y, Wang K, Jia W, Chu WM, Sun B. The long noncoding RNA lnc-EGFR stimulates T-regulatory cells differentiation thus promoting hepatocellular carcinoma immune evasion. Nat Commun. 2017;8:15129. doi: 10.1038/ncomms15129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Chen C, He W, Huang J, Wang B, Li H, Cai Q, Su F, Bi J, Liu H, Zhang B, Jiang N, Zhong G, Zhao Y, Dong W, Lin T. LNMAT1 promotes lymphatic metastasis of bladder cancer via CCL2 dependent macrophage recruitment. Nat Commun. 2018;9:3826. doi: 10.1038/s41467-018-06152-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Huang D, Chen J, Yang L, Ouyang Q, Li J, Lao L, Zhao J, Liu J, Lu Y, Xing Y, Chen F, Su F, Yao H, Liu Q, Su S, Song E. NKILA lncRNA promotes tumor immune evasion by sensitizing T cells to activation-induced cell death. Nat Immunol. 2018;19:1112–1125. doi: 10.1038/s41590-018-0207-y. [DOI] [PubMed] [Google Scholar]
- 28.Sang LJ, Ju HQ, Liu GP, Tian T, Ma GL, Lu YX, Liu ZX, Pan RL, Li RH, Piao HL, Marks JR, Yang LJ, Yan Q, Wang W, Shao J, Zhou Y, Zhou T, Lin A. LncRNA CamK-A regulates Ca(2+)-signaling-mediated tumor microenvironment remodeling. Mol Cell. 2018;72:71–83. e77. doi: 10.1016/j.molcel.2018.08.014. [DOI] [PubMed] [Google Scholar]
- 29.Hardie DG, Ross FA, Hawley SA. AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nat Rev Mol Cell Biol. 2012;13:251–262. doi: 10.1038/nrm3311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Liu X, Xiao ZD, Han L, Zhang J, Lee SW, Wang W, Lee H, Zhuang L, Chen J, Lin HK, Wang J, Liang H, Gan B. LncRNA NBR2 engages a metabolic checkpoint by regulating AMPK under energy stress. Nat Cell Biol. 2016;18:431–442. doi: 10.1038/ncb3328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Leucci E, Vendramin R, Spinazzi M, Laurette P, Fiers M, Wouters J, Radaelli E, Eyckerman S, Leonelli C, Vanderheyden K, Rogiers A, Hermans E, Baatsen P, Aerts S, Amant F, Van Aelst S, van den Oord J, de Strooper B, Davidson I, Lafontaine DL, Gevaert K, Vandesompele J, Mestdagh P, Marine JC. Melanoma addiction to the long non-coding RNA SAMMSON. Nature. 2016;531:518–522. doi: 10.1038/nature17161. [DOI] [PubMed] [Google Scholar]
- 32.Lin A, Li C, Xing Z, Hu Q, Liang K, Han L, Wang C, Hawke DH, Wang S, Zhang Y, Wei Y, Ma G, Park PK, Zhou J, Zhou Y, Hu Z, Zhou Y, Marks JR, Liang H, Hung MC, Lin C, Yang L. The LINK-A lncRNA activates normoxic HIF1alpha signalling in triple-negative breast cancer. Nat Cell Biol. 2016;18:213–224. doi: 10.1038/ncb3295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Yang F, Zhang H, Mei Y, Wu M. Reciprocal regulation of HIF-1alpha and lincRNA-p21 modulates the warburg effect. Mol Cell. 2014;53:88–100. doi: 10.1016/j.molcel.2013.11.004. [DOI] [PubMed] [Google Scholar]
- 34.Shih JW, Chiang WF, Wu ATH, Wu MH, Wang LY, Yu YL, Hung YW, Wang WC, Chu CY, Hung CL, Changou CA, Yen Y, Kung HJ. Long noncoding RNA LncHIFCAR/MIR31HG is a HIF-1alpha co-activator driving oral cancer progression. Nat Commun. 2017;8:15874. doi: 10.1038/ncomms15874. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Yang F, Huo XS, Yuan SX, Zhang L, Zhou WP, Wang F, Sun SH. Repression of the long noncoding RNA-LET by histone deacetylase 3 contributes to hypoxia-mediated metastasis. Mol Cell. 2013;49:1083–1096. doi: 10.1016/j.molcel.2013.01.010. [DOI] [PubMed] [Google Scholar]
- 36.Huang JZ, Chen M, Chen , Gao XC, Zhu S, Huang H, Hu M, Zhu H, Yan GR. A peptide encoded by a putative lncRNA HOXB-AS3 suppresses colon cancer growth. Mol Cell. 2017;68:171–184. e176. doi: 10.1016/j.molcel.2017.09.015. [DOI] [PubMed] [Google Scholar]
- 37.Xiao ZD, Han L, Lee H, Zhuang L, Zhang Y, Baddour J, Nagrath D, Wood CG, Gu J, Wu X, Liang H, Gan B. Energy stress-induced lncRNA FILNC1 represses c-Myc-mediated energy metabolism and inhibits renal tumor development. Nat Commun. 2017;8:783. doi: 10.1038/s41467-017-00902-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Nieto MA. Epithelial-mesenchymal transitions in development and disease: old views and new perspectives. Int J Dev Biol. 2009;53:1541–1547. doi: 10.1387/ijdb.072410mn. [DOI] [PubMed] [Google Scholar]
- 39.Nieto MA, Huang RY, Jackson RA, Thiery JP. Emt: 2016. Cell. 2016;166:21–45. doi: 10.1016/j.cell.2016.06.028. [DOI] [PubMed] [Google Scholar]
- 40.Kalluri R, Weinberg RA. The basics of epithelial-mesenchymal transition. J Clin Invest. 2009;119:1420–1428. doi: 10.1172/JCI39104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Thiery JP, Acloque H, Huang RY, Nieto MA. Epithelial-mesenchymal transitions in development and disease. Cell. 2009;139:871–890. doi: 10.1016/j.cell.2009.11.007. [DOI] [PubMed] [Google Scholar]
- 42.Gupta RA, Shah N, Wang KC, Kim J, Horlings HM, Wong DJ, Tsai MC, Hung T, Argani P, Rinn JL, Wang Y, Brzoska P, Kong B, Li R, West RB, van de Vijver MJ, Sukumar S, Chang HY. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature. 2010;464:1071–1076. doi: 10.1038/nature08975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Akhurst RJ, Hata A. Targeting the TGFbeta signalling pathway in disease. Nat Rev Drug Discov. 2012;11:790–811. doi: 10.1038/nrd3810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Ottaviani S, Stebbing J, Frampton AE, Zagorac S, Krell J, de Giorgio A, Trabulo SM, Nguyen VTM, Magnani L, Feng H, Giovannetti E, Funel N, Gress TM, Jiao LR, Lombardo Y, Lemoine NR, Heeschen C, Castellano L. TGF-beta induces miR-100 and miR-125b but blocks let-7a through LIN28B controlling PDAC progression. Nat Commun. 2018;9:1845. doi: 10.1038/s41467-018-03962-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Mondal T, Subhash S, Vaid R, Enroth S, Uday S, Reinius B, Mitra S, Mohammed A, James AR, Hoberg E, Moustakas A, Gyllensten U, Jones SJ, Gustafsson CM, Sims AH, Westerlund F, Gorab E, Kanduri C. MEG3 long noncoding RNA regulates the TGF-beta pathway genes through formation of RNA-DNA triplex structures. Nat Commun. 2015;6:7743. doi: 10.1038/ncomms8743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Huang JF, Guo YJ, Zhao CX, Yuan SX, Wang Y, Tang GN, Zhou WP, Sun SH. Hepatitis B virus X protein (HBx)-related long noncoding RNA (lncRNA) down-regulated expression by HBx (Dreh) inhibits hepatocellular carcinoma metastasis by targeting the intermediate filament protein vimentin. Hepatology. 2013;57:1882–1892. doi: 10.1002/hep.26195. [DOI] [PubMed] [Google Scholar]
- 47.Grelet S, Link LA, Howley B, Obellianne C, Palanisamy V, Gangaraju VK, Diehl JA, Howe PH. A regulated PNUTS mRNA to lncRNA splice switch mediates EMT and tumour progression. Nat Cell Biol. 2017;19:1105–1115. doi: 10.1038/ncb3595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Pádua Alves C, Fonseca AS, Muys BR, de Barros E Lima Bueno R, Bürger MC, de Souza JE, Valente V, Zago MA, Silva WA Jr. Brief report: the lincRNA hotair is required for epithelial-to-mesenchymal transition and stemness maintenance of cancer cell lines. Stem Cells. 2013;31:2827–2832. doi: 10.1002/stem.1547. [DOI] [PubMed] [Google Scholar]
- 49.Yang Z, Zhou L, Wu LM, Lai MC, Xie HY, Zhang F, Zheng SS. Overexpression of long non-coding RNA HOTAIR predicts tumor recurrence in hepatocellular carcinoma patients following liver transplantation. Ann Surg Oncol. 2011;18:1243–1250. doi: 10.1245/s10434-011-1581-y. [DOI] [PubMed] [Google Scholar]
- 50.Kim K, Jutooru I, Chadalapaka G, Johnson G, Frank J, Burghardt R, Kim S, Safe S. HOTAIR is a negative prognostic factor and exhibits pro-oncogenic activity in pancreatic cancer. Oncogene. 2013;32:1616–1625. doi: 10.1038/onc.2012.193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Ren Y, Jia HH, Xu YQ, Zhou X, Zhao XH, Wang YF, Song X, Zhu ZY, Sun T, Dou Y, Tian WP, Zhao XL, Kang CS, Mei M. Paracrine and epigenetic control of CAF-induced metastasis: the role of HOTAIR stimulated by TGF-ss1 secretion. Mol Cancer. 2018;17:5. doi: 10.1186/s12943-018-0758-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Zhang J, Li Z, Liu L, Wang Q, Li S, Chen D, Hu Z, Yu T, Ding J, Li J, Yao M, Huang S, Zhao Y, He X. Long noncoding RNA TSLNC8 is a tumor suppressor that inactivates the interleukin-6/STAT3 signaling pathway. Hepatology. 2018;67:171–187. doi: 10.1002/hep.29405. [DOI] [PubMed] [Google Scholar]
- 53.Li C, Wang S, Xing Z, Lin A, Liang K, Song J, Hu Q, Yao J, Chen Z, Park PK, Hawke DH, Zhou J, Zhou Y, Zhang S, Liang H, Hung MC, Gallick GE, Han L, Lin C, Yang L. A ROR1-HER3-lncRNA signalling axis modulates the hippo-YAP pathway to regulate bone metastasis. Nat Cell Biol. 2017;19:106–119. doi: 10.1038/ncb3464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Zhu P, Wang Y, Wu J, Huang G, Liu B, Ye B, Du Y, Gao G, Tian Y, He L, Fan Z. LncBRM initiates YAP1 signalling activation to drive self-renewal of liver cancer stem cells. Nat Commun. 2016;7:13608. doi: 10.1038/ncomms13608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Chen ZZ, Huang L, Wu YH, Zhai WJ, Zhu PP, Gao YF. LncSox4 promotes the self-renewal of liver tumour-initiating cells through Stat3-mediated Sox4 expression. Nat Commun. 2016;7:12598. doi: 10.1038/ncomms12598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Wang Y, He L, Du Y, Zhu P, Huang G, Luo J, Yan X, Ye B, Li C, Xia P, Zhang G, Tian Y, Chen R, Fan Z. The long noncoding RNA lncTCF7 promotes self-renewal of human liver cancer stem cells through activation of Wnt signaling. Cell Stem Cell. 2015;16:413–425. doi: 10.1016/j.stem.2015.03.003. [DOI] [PubMed] [Google Scholar]
- 57.Huarte M. The emerging role of lncRNAs in cancer. Nat Med. 2015;21:1253–1261. doi: 10.1038/nm.3981. [DOI] [PubMed] [Google Scholar]
- 58.Bennett CF, Baker BF, Pham N, Swayze E, Geary RS. Pharmacology of antisense drugs. Annu Rev Pharmacol Toxicol. 2017;57:81–105. doi: 10.1146/annurev-pharmtox-010716-104846. [DOI] [PubMed] [Google Scholar]
- 59.Buller HR, Bethune C, Bhanot S, Gailani D, Monia BP, Raskob GE, Segers A, Verhamme P, Weitz JI FXI-ASO TKA Investigators. Factor XI antisense oligonucleotide for prevention of venous thrombosis. N Engl J Med. 2015;372:232–240. doi: 10.1056/NEJMoa1405760. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Gaudet D, Brisson D, Tremblay K, Alexander VJ, Singleton W, Hughes SG, Geary RS, Baker BF, Graham MJ, Crooke RM, Witztum JL. Targeting APOC3 in the familial chylomicronemia syndrome. N Engl J Med. 2014;371:2200–2206. doi: 10.1056/NEJMoa1400284. [DOI] [PubMed] [Google Scholar]
- 61.Hong D, Kurzrock R, Kim Y, Woessner R, Younes A, Nemunaitis J, Fowler N, Zhou T, Schmidt J, Jo M, Lee SJ, Yamashita M, Hughes SG, Fayad L, Piha-Paul S, Nadella MV, Mohseni M, Lawson D, Reimer C, Blakey DC, Xiao X, Hsu J, Revenko A, Monia BP, MacLeod AR. AZD9150, a next-generation antisense oligonucleotide inhibitor of STAT3 with early evidence of clinical activity in lymphoma and lung cancer. Sci Transl Med. 2015;7:314ra185. doi: 10.1126/scitranslmed.aac5272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Meng L, Ward AJ, Chun S, Bennett CF, Beaudet AL, Rigo F. Towards a therapy for angelman syndrome by targeting a long non-coding RNA. Nature. 2015;518:409–412. doi: 10.1038/nature13975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Ling H, Fabbri M, Calin GA. MicroRNAs and other non-coding RNAs as targets for anticancer drug development. Nat Rev Drug Discov. 2013;12:847–865. doi: 10.1038/nrd4140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Arun G, Diermeier S, Akerman M, Chang KC, Wilkinson JE, Hearn S, Kim Y, MacLeod AR, Krainer AR, Norton L, Brogi E, Egeblad M, Spector DL. Differentiation of mammary tumors and reduction in metastasis upon Malat1 lncRNA loss. Genes Dev. 2016;30:34–51. doi: 10.1101/gad.270959.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Gutschner T, Hammerle M, Eissmann M, Hsu J, Kim Y, Hung G, Revenko A, Arun G, Stentrup M, Gross M, Zornig M, MacLeod AR, Spector DL, Diederichs S. The noncoding RNA MALAT1 is a critical regulator of the metastasis phenotype of lung cancer cells. Cancer Res. 2013;73:1180–1189. doi: 10.1158/0008-5472.CAN-12-2850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Etzioni R, Cha R, Feuer EJ, Davidov O. Asymptomatic incidence and duration of prostate cancer. Am J Epidemiol. 1998;148:775–785. doi: 10.1093/oxfordjournals.aje.a009698. [DOI] [PubMed] [Google Scholar]
- 67.Cooperberg MR, Moul JW, Carroll PR. The changing face of prostate cancer. J. Clin. Oncol. 2005;23:8146–8151. doi: 10.1200/JCO.2005.02.9751. [DOI] [PubMed] [Google Scholar]
- 68.Liu PY, Erriquez D, Marshall GM, Tee AE, Polly P, Wong M, Liu B, Bell JL, Zhang XD, Milazzo G, Cheung BB, Fox A, Swarbrick A, Huttelmaier S, Kavallaris M, Perini G, Mattick JS, Dinger ME, Liu T. Effects of a novel long noncoding RNA, lncUSMycN, on N-Myc expression and neuroblastoma progression. J Natl Cancer Inst. 2014;106 doi: 10.1093/jnci/dju113. [DOI] [PubMed] [Google Scholar]
- 69.Astriab-Fisher A, Sergueev D, Fisher M, Shaw BR, Juliano RL. Conjugates of antisense oligonucleotides with the tat and antennapedia cell-penetrating peptides: effects on cellular uptake, binding to target sequences, and biologic actions. Pharm Res. 2002;19:744–754. doi: 10.1023/a:1016136328329. [DOI] [PubMed] [Google Scholar]
- 70.Koch L. Functional genomics: screening for lncRNA function. Nat Rev Genet. 2017;18:70. doi: 10.1038/nrg.2016.168. [DOI] [PubMed] [Google Scholar]
- 71.Gilbert LA, Horlbeck MA, Adamson B, Villalta JE, Chen Y, Whitehead EH, Guimaraes C, Panning B, Ploegh HL, Bassik MC, Qi LS, Kampmann M, Weissman JS. Genome-scale CRISPR-mediated control of gene repression and activation. Cell. 2014;159:647–661. doi: 10.1016/j.cell.2014.09.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Thakore PI, D’Ippolito AM, Song L, Safi A, Shivakumar NK, Kabadi AM, Reddy TE, Crawford GE, Gersbach CA. Highly specific epigenome editing by CRISPR-Cas9 repressors for silencing of distal regulatory elements. Nat Methods. 2015;12:1143–1149. doi: 10.1038/nmeth.3630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Liu SJ, Horlbeck MA, Cho SW, Birk HS, Malatesta M, He D, Attenello FJ, Villalta JE, Cho MY, Chen Y, Mandegar MA, Olvera MP, Gilbert LA, Conklin BR, Chang HY, Weissman JS, Lim DA. CRISPRi-based genome-scale identification of functional long noncoding RNA loci in human cells. Science. 2017;355 doi: 10.1126/science.aah7111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Adriaens C, Standaert L, Barra J, Latil M, Verfaillie A, Kalev P, Boeckx B, Wijnhoven PW, Radaelli E, Vermi W, Leucci E, Lapouge G, Beck B, van den Oord J, Nakagawa S, Hirose T, Sablina AA, Lambrechts D, Aerts S, Blanpain C, Marine JC. p53 induces formation of NEAT1 lncRNA-containing paraspeckles that modulate replication stress response and chemosensitivity. Nat Med. 2016;22:861–868. doi: 10.1038/nm.4135. [DOI] [PubMed] [Google Scholar]
- 75.Mendell JT. Targeting a long noncoding RNA in breast cancer. N Engl J Med. 2016;374:2287–2289. doi: 10.1056/NEJMcibr1603785. [DOI] [PubMed] [Google Scholar]
- 76.Zhuo W, Liu Y, Li S, Guo D, Sun Q, Jin J, Rao X, Li M, Sun M, Jiang M, Xu Y, Teng L, Jin Y, Si J, Liu W, Kang Y, Zhou T. Long noncoding RNA GMAN, up-regulated in gastric cancer tissues, is associated with metastasis in patients and promotes translation of Ephrin A1 by competitively binding GMAN-AS. Gastroenterology. 2019;156:676–691. e611. doi: 10.1053/j.gastro.2018.10.054. [DOI] [PubMed] [Google Scholar]
- 77.Yang H, Wang H, Shivalila CS, Cheng AW, Shi L, Jaenisch R. One-step generation of mice carrying reporter and conditional alleles by CRISPR/Cas-mediated genome engineering. Cell. 2013;154:1370–1379. doi: 10.1016/j.cell.2013.08.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Cho SW, Kim S, Kim Y, Kweon J, Kim HS, Bae S, Kim JS. Analysis of off-target effects of CRISPR/Cas-derived RNA-guided endonucleases and nickases. Genome Res. 2014;24:132–141. doi: 10.1101/gr.162339.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Tsai SQ, Joung JK. Defining and improving the genome-wide specificities of CRISPR-Cas9 nucleases. Nat Rev Genet. 2016;17:300–312. doi: 10.1038/nrg.2016.28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC. Potent and specific genetic interference by double-stranded RNA in caenorhabditis elegans. Nature. 1998;391:806–811. doi: 10.1038/35888. [DOI] [PubMed] [Google Scholar]
- 81.Elbashir SM, Harborth J, Lendeckel W, Yalcin A, Weber K, Tuschl T. Duplexes of 21-nucleotide RNAs mediate RNA interference in cultured mammalian cells. Nature. 2001;411:494–498. doi: 10.1038/35078107. [DOI] [PubMed] [Google Scholar]
- 82.Brummelkamp TR, Bernards R, Agami R. A system for stable expression of short interfering RNAs in mammalian cells. Science. 2002;296:550–553. doi: 10.1126/science.1068999. [DOI] [PubMed] [Google Scholar]
- 83.Zuber J, McJunkin K, Fellmann C, Dow LE, Taylor MJ, Hannon GJ, Lowe SW. Toolkit for evaluating genes required for proliferation and survival using tetracycline-regulated RNAi. Nat Biotechnol. 2011;29:79–83. doi: 10.1038/nbt.1720. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Fellmann C, Hoffmann T, Sridhar V, Hopfgartner B, Muhar M, Roth M, Lai DY, Barbosa IA, Kwon JS, Guan Y, Sinha N, Zuber J. An optimized microRNA backbone for effective single-copy RNAi. Cell Rep. 2013;5:1704–1713. doi: 10.1016/j.celrep.2013.11.020. [DOI] [PubMed] [Google Scholar]
- 85.Gu S, Jin L, Zhang Y, Huang Y, Zhang F, Valdmanis PN, Kay MA. The loop position of shRNAs and pre-miRNAs is critical for the accuracy of dicer processing in vivo. Cell. 2012;151:900–911. doi: 10.1016/j.cell.2012.09.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Watanabe C, Cuellar TL, Haley B. Quantitative evaluation of first, second, and third generation hairpin systems reveals the limit of mammalian vector-based RNAi. RNA Biol. 2016;13:25–33. doi: 10.1080/15476286.2015.1128062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Bainbridge JW, Smith AJ, Barker SS, Robbie S, Henderson R, Balaggan K, Viswanathan A, Holder GE, Stockman A, Tyler N, Petersen-Jones S, Bhattacharya SS, Thrasher AJ, Fitzke FW, Carter BJ, Rubin GS, Moore AT, Ali RR. Effect of gene therapy on visual function in leber’s congenital amaurosis. N Engl J Med. 2008;358:2231–2239. doi: 10.1056/NEJMoa0802268. [DOI] [PubMed] [Google Scholar]
- 88.Gaudet D, Methot J, Dery S, Brisson D, Essiembre C, Tremblay G, Tremblay K, de Wal J, Twisk J, van den Bulk N, Sier-Ferreira V, van Deventer S. Efficacy and long-term safety of alipogene tiparvovec (AAV1-LPLS447X) gene therapy for lipoprotein lipase deficiency: an open-label trial. Gene Ther. 2013;20:361–369. doi: 10.1038/gt.2012.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Bennett J, Ashtari M, Wellman J, Marshall KA, Cyckowski LL, Chung DC, McCague S, Pierce EA, Chen Y, Bennicelli JL, Zhu X, Ying GS, Sun J, Wright JF, Auricchio A, Simonelli F, Shindler KS, Mingozzi F, High KA, Maguire AM. AAV2 gene therapy readministration in three adults with congenital blindness. Sci Transl Med. 2012;4:120ra115. doi: 10.1126/scitranslmed.3002865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Schaffer DV, Koerber JT, Lim KI. Molecular engineering of viral gene delivery vehicles. Annu Rev Biomed Eng. 2008;10:169–194. doi: 10.1146/annurev.bioeng.10.061807.160514. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Kaeppel C, Beattie SG, Fronza R, van Logtenstein R, Salmon F, Schmidt S, Wolf S, Nowrouzi A, Glimm H, von Kalle C, Petry H, Gaudet D, Schmidt M. A largely random AAV integration profile after LPLD gene therapy. Nat Med. 2013;19:889–891. doi: 10.1038/nm.3230. [DOI] [PubMed] [Google Scholar]
- 92.Lisowski L, Dane AP, Chu K, Zhang Y, Cunningham SC, Wilson EM, Nygaard S, Grompe M, Alexander IE, Kay MA. Selection and evaluation of clinically relevant AAV variants in a xenograft liver model. Nature. 2014;506:382–386. doi: 10.1038/nature12875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Xing Z, Lin A, Li C, Liang K, Wang S, Liu Y, Park PK, Qin L, Wei Y, Hawke DH, Hung MC, Lin C, Yang L. lncRNA directs cooperative epigenetic regulation downstream of chemokine signals. Cell. 2014;159:1110–1125. doi: 10.1016/j.cell.2014.10.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Endo H, Shiroki T, Nakagawa T, Yokoyama M, Tamai K, Yamanami H, Fujiya T, Sato I, Yamaguchi K, Tanaka N, Iijima K, Shimosegawa T, Sugamura K, Satoh K. Enhanced expression of long non-coding RNA HOTAIR is associated with the development of gastric cancer. PLoS One. 2013;8:e77070. doi: 10.1371/journal.pone.0077070. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Katayama S, Tomaru Y, Kasukawa T, Waki K, Nakanishi M, Nakamura M, Nishida H, Yap CC, Suzuki M, Kawai J, Suzuki H, Carninci P, Hayashizaki Y, Wells C, Frith M, Ravasi T, Pang KC, Hallinan J, Mattick J, Hume DA, Lipovich L, Batalov S, Engström PG, Mizuno Y, Faghihi MA, Sandelin A, Chalk AM, Mottagui-Tabar S, Liang Z, Lenhard B, Wahlestedt C RIKEN Genome Exploration Research Group; Genome Science Group (Genome Network Project Core Group); FANTOM Consortium. Antisense transcription in the mammalian transcriptome. Science. 2005;309:1564–1566. doi: 10.1126/science.1112009. [DOI] [PubMed] [Google Scholar]
- 96.Faghihi MA, Wahlestedt C. Regulatory roles of natural antisense transcripts. Nat Rev Mol Cell Biol. 2009;10:637–643. doi: 10.1038/nrm2738. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Modarresi F, Faghihi MA, Lopez-Toledano MA, Fatemi RP, Magistri M, Brothers SP, van der Brug MP, Wahlestedt C. Inhibition of natural antisense transcripts in vivo results in gene-specific transcriptional upregulation. Nat Biotechnol. 2012;30:453–459. doi: 10.1038/nbt.2158. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Yap KL, Li S, Munoz-Cabello AM, Raguz S, Zeng L, Mujtaba S, Gil J, Walsh MJ, Zhou MM. Molecular interplay of the noncoding RNA ANRIL and methylated histone H3 lysine 27 by polycomb CBX7 in transcriptional silencing of INK4a. Mol Cell. 2010;38:662–674. doi: 10.1016/j.molcel.2010.03.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Dimitrova N, Zamudio JR, Jong RM, Soukup D, Resnick R, Sarma K, Ward AJ, Raj A, Lee JT, Sharp PA, Jacks T. LincRNA-p21 activates p21 in cis to promote polycomb target gene expression and to enforce the G1/S checkpoint. Mol Cell. 2014;54:777–790. doi: 10.1016/j.molcel.2014.04.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Anderson EM, Birmingham A, Baskerville S, Reynolds A, Maksimova E, Leake D, Fedorov Y, Karpilow J, Khvorova A. Experimental validation of the importance of seed complement frequency to siRNA specificity. RNA. 2008;14:853–861. doi: 10.1261/rna.704708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Burchard J, Jackson AL, Malkov V, Needham RH, Tan Y, Bartz SR, Dai H, Sachs AB, Linsley PS. MicroRNA-like off-target transcript regulation by siRNAs is species specific. Rna. 2009;15:308–315. doi: 10.1261/rna.1326809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Mattheolabakis G, Rigas B, Constantinides PP. Nanodelivery strategies in cancer chemotherapy: biological rationale and pharmaceutical perspectives. Nanomedicine (Lond) 2012;7:1577–1590. doi: 10.2217/nnm.12.128. [DOI] [PubMed] [Google Scholar]
- 103.Webster DM, Sundaram P, Byrne ME. Injectable nanomaterials for drug delivery: carriers, targeting moieties, and therapeutics. Eur J Pharm Biopharm. 2013;84:1–20. doi: 10.1016/j.ejpb.2012.12.009. [DOI] [PubMed] [Google Scholar]
- 104.Fabbro C, Ali-Boucetta H, Da Ros T, Kostarelos K, Bianco A, Prato M. Targeting carbon nanotubes against cancer. Chem Commun (Camb) 2012;48:3911–3926. doi: 10.1039/c2cc17995d. [DOI] [PubMed] [Google Scholar]
- 105.Libutti SK, Paciotti GF, Byrnes AA, Alexander HR Jr, Gannon WE, Walker M, Seidel GD, Yuldasheva N, Tamarkin L. Phase I and pharmacokinetic studies of CYT-6091, a novel PEGylated colloidal gold-rhTNF nanomedicine. Clin Cancer Res. 2010;16:6139–6149. doi: 10.1158/1078-0432.CCR-10-0978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Yang T, Choi MK, Cui FD, Lee SJ, Chung SJ, Shim CK, Kim DD. Antitumor effect of paclitaxel-loaded PEGylated immunoliposomes against human breast cancer cells. Pharm Res. 2007;24:2402–2411. doi: 10.1007/s11095-007-9425-y. [DOI] [PubMed] [Google Scholar]
- 107.Markman JL, Rekechenetskiy A, Holler E, Ljubimova JY. Nanomedicine therapeutic approaches to overcome cancer drug resistance. Adv Drug Deliv Rev. 2013;65:1866–1879. doi: 10.1016/j.addr.2013.09.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Fan Y, Shen B, Tan M, Mu X, Qin Y, Zhang F, Liu Y. Long non-coding RNA UCA1 increases chemoresistance of bladder cancer cells by regulating wnt signaling. FEBS J. 2014;281:1750–1758. doi: 10.1111/febs.12737. [DOI] [PubMed] [Google Scholar]
- 109.He W, Liang B, Wang C, Li S, Zhao Y, Huang Q, Liu Z, Yao Z, Wu Q, Liao W, Zhang S, Liu Y, Xiang Y, Liu J, Shi M. MSC-regulated lncRNA MACC1-AS1 promotes stemness and chemoresistance through fatty acid oxidation in gastric cancer. Oncogene. 2019;38:4637–4654. doi: 10.1038/s41388-019-0747-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Huang X, El-Sayed IH, Qian W, El-Sayed MA. Cancer cell imaging and photothermal therapy in the near-infrared region by using gold nanorods. J Am Chem Soc. 2006;128:2115–2120. doi: 10.1021/ja057254a. [DOI] [PubMed] [Google Scholar]
- 111.Qiu L, Chen T, Ocsoy I, Yasun E, Wu C, Zhu G, You M, Han D, Jiang J, Yu R, Tan W. A cell-targeted, size-photocontrollable, nuclear-uptake nanodrug delivery system for drug-resistant cancer therapy. Nano Lett. 2015;15:457–463. doi: 10.1021/nl503777s. [DOI] [PMC free article] [PubMed] [Google Scholar]