Next Article in Journal
Characterization of Circulating T Cell Receptor Repertoire Provides Information about Clinical Outcome after PD-1 Blockade in Advanced Non-Small Cell Lung Cancer Patients
Next Article in Special Issue
The lncRNAs in HBV-Related HCCs: Targeting Chromatin Dynamics and Beyond
Previous Article in Journal
DNA Methylation-Based Estimates of Circulating Leukocyte Composition for Predicting Colorectal Cancer Survival: A Prospective Cohort Study
Previous Article in Special Issue
MicroRNAs Deregulated in Intraductal Papillary Mucinous Neoplasm Converge on Actin Cytoskeleton-Related Pathways That Are Maintained in Pancreatic Ductal Adenocarcinoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Long Intergenic Non-Coding RNAs in HNSCC: From “Junk DNA” to Important Prognostic Factor

1
Laboratory of Cancer Genetics, Greater Poland Cancer Centre, Garbary 15, 61-866 Poznan, Poland
2
Research and Implementation Unit, Greater Poland Cancer Centre, Garbary 15, 61-866 Poznan, Poland
3
Department of Cancer Immunology, Chair of Medical Biotechnology, Poznan University of Medical Sciences, 8 Rokietnicka Street, 60-806 Poznan, Poland
4
Department of Oncologic Pathology and Prophylaxis, Poznan University of Medical Sciences, Greater Poland Cancer Centere, Garbary 15, 61-866 Poznan, Poland
5
Department of Pathology, Pomeranian Medical University, Rybacka 1, 70-204 Szczecin, Poland
6
Postgraduate School of Molecular Medicine, Medical University of Warsaw, ul. Zwirki 61 and ul. Wigury, 02-091 Warsaw, Poland
*
Authors to whom correspondence should be addressed.
Cancers 2021, 13(12), 2949; https://doi.org/10.3390/cancers13122949
Submission received: 24 May 2021 / Revised: 8 June 2021 / Accepted: 9 June 2021 / Published: 12 June 2021
(This article belongs to the Collection Regulatory and Non-Coding RNAs in Cancer Epigenetic Mechanisms)

Abstract

:

Simple Summary

Head and neck squamous cell carcinoma (HNSCC) has one of the highest incidence and mortality rates among all cancers. Diagnostic process and treatment results are far from satisfactory. These are the main reasons behind studying the microenvironment of the tumor and finding the connection between the aberrant expression levels of non-coding RNAs and patients’ outcomes. In this paper, we tried to present the function and the promising diagnostic potential of long intergenic non-coding RNAs (lincRNAs). We proved that a multitude of them play a pivotal role in the different processes involved in the progression of the disease—e.g., proliferation, migration, and the epithelial-to-mesenchymal transition. Even though there is a lot of work ahead of us, lincRNAs could become unique and valuable biomarkers or future targets for personalized medicine.

Abstract

Head and neck squamous cell carcinoma is one of the most common and fatal cancers worldwide. Even a multimodal approach consisting of standard chemo- and radiotherapy along with surgical resection is only effective in approximately 50% of the cases. The rest of the patients develop a relapse of the disease and acquire resistance to treatment. Especially this group of individuals needs novel, personalized, targeted therapy. The first step to discovering such solutions is to investigate the tumor microenvironment, thus understanding the role and mechanism of the function of coding and non-coding sequences of the human genome. In recent years, RNA molecules gained great interest when the complex character of their impact on our biology allowed them to come out of the shadows of the “junk DNA” label. Furthermore, long non-coding RNAs (lncRNA), specifically the intergenic subgroup (lincRNA), are one of the most aberrantly expressed in several malignancies, which makes them particularly promising future diagnostic biomarkers and therapeutic targets. This review contains characteristics of known and validated lincRNAs in HNSCC, such as XIST, MALAT, HOTAIR, HOTTIP, lincRNA-p21, LINC02487, LINC02195, LINC00668, LINC00519, LINC00511, LINC00460, LINC00312, and LINC00052, with a description of their prognostic abilities. Even though much work remains to be done, lincRNAs are important factors in cancer biology that will become valuable biomarkers of tumor stage, outcome prognosis, and contribution to personalized medicine.

1. Introduction

For decades, cancer has been one of the greatest medical challenges. Just head and neck squamous cell carcinoma (HNSCC) alone, the seventh most common cancer worldwide, is responsible for approximately 450,000 deaths per year [1]. This particularly fatal disease has a nearly 50% mortality rate in the first five years after diagnosis [2]. Head and neck aggressive heterogeneous malignancies form tumors in epithelial tissue of the upper aerodigestive tract. Major risk factors for developing HNSCC are alcohol and/or tobacco consumption [3,4], environmental carcinogens [5], and human papillomavirus (HPV) infection [6]. The latter one is responsible for the higher incidence in the younger group of patients, especially men [7]. Moreover, there is growing evidence that it is associated with 47% of tonsillar squamous cell carcinomas (TSCCs) and 22% of oropharyngeal squamous cell carcinomas (OPSCC) [8,9,10,11]. It is worth mentioning that HPV-positive patients (HPV(+)) have a significantly better prognosis than HPV(–) individuals, due to different clinicopathological, molecular, and even epigenetic characteristics [12,13]. Searching for such distinctive features in all types of HNSCC brings us closer to the creation of precision medicine solutions that will become alternative for the standard therapy consisting of chemo- and radiotherapy, surgical intervention, or systemic treatment, e.g., cisplatin, carboplatin, and cetuximab [14,15,16]. The main reason behind seeking personal medicine approaches is that more than 50% of the cases relapse despite administering aggressive multimodal therapy [17]. After that, high risk of complications significantly narrows down options for further treatment. Currently, cetuximab, pembrolizumab, or nivolumab is used in combination with chemotherapy or alone as a treatment for patients with recurrent or metastatic disease. However, in most cases, the response is only partial and often leads to acquired resistance followed by tumor regrowth. This demonstrates the urgent need to better understand the tumor microenvironment in order to propose targeted therapy that could significantly improve the quality and life expectancy of patients with HNSCC in the future [18].
In recent years, more and more studies indicate a disturbed expression of many coding and non-coding RNA molecules, which could soon be used as predictive or prognostic biomarkers specific for a particular type, or even stage, of a tumor [19,20,21]. This complex regulatory network composed of microRNAs (miRNAs), messenger RNAs (mRNAs), and long non-coding RNAs (lncRNAs), including long intergenic non-coding RNAs (lincRNAs), gained significant interest after being correlated with changes in genome stability, cell proliferation, differentiation, and migration, leading to metastasis caused by the epithelial-to-mesenchymal transition (EMT), and subsequent relapse of patients [22,23,24]. The aim of this review is to elucidate the role of lincRNAs in HNSCC, which are the least characterized molecules from all of the above.

2. lncRNA or lincRNA: Difference, Biogenesis, Function

Long non-coding RNAs belong to a subgroup of RNA molecules that are at least 200 nucleotides long and the majority of them do not have proteincoding ability [20,25]. It is worth emphasizing that roughly 93% of the human genome can be transcribed as RNA, but only 2% of it may be translated into protein. For many years, the rest of the genome was considered a “junk DNA” that does not bear any function [20]. However, with time, lncRNAs were associated with a great number of pivotal biological processes such as regulation of gene expression, e.g., chromatin modification, interaction with transcriptional factors, mRNA processing, cell metabolism, proliferation, apoptosis, acting as “molecular sponge,” and creating ribonucleoprotein complexes [22,23,24,25]. Genes encoding lncRNAs could be situated in intergenic and intragenic positions. The intragenic lncRNA transcripts can be found in intronic, enhancer, promoter as well as in 3′UTR regions of the specified gene [25].
More than 50% of lncRNAs are long intergenic non-coding RNAs (lincRNAs), which have been distinguished based on the lack of overlapping annotated protein-coding genes in their transcripts [26]. Creating this subgroup of lncRNAs was proposed after conducting studies using tiling arrays across the genome [27]. Extensive research allowed the division of lincRNAs into four classes depending on the distance from the protein-coding transcript and direction of transcription, sense or antisense: (i) same strand, (ii) convergent, (iii) divergent, and (iv) isolated—the only one placed further than 50kb from the nearest protein-coding gene [26] (Figure 1A). It should be noted that divergent (bidirectional) lncRNAs are transcribed often concordant with expression of the nearby protein-coding gene [28].
Fluorescent in situ hybridization and ribosome profiling provided information on nuclear enrichment of lincRNAs compared to the cytoplasm compartment where more of lncRNAs can be found [29,30]. Localization within the nucleus corroborates an important role in the process of cell differentiation, providing proper nuclear architecture and modulating states of the chromatin condensation [23,26,31]. However, it has been proven that the whole group of lncRNA also regulates the mechanism of transcription itself and post-transcriptional modifications, such as polyadenylation, 5′capping or splicing [32].
The broad range of lincRNA functions includes also acting as a protein and RNA scaffold or decoy, sequestering different intracellular molecules or improving their function, and producing micropeptides (Figure 1B). Furthermore, the low abundance of these molecules does not affect their ability to form multiple macromolecular complexes alteringthe epigenetic state of neighboring genes [26]. The vast majority of described lincRNAs are also involved in developmental pathways, including linc-RoR, TINCR, ANCR, LINC00261, PNKY, or lincRNA-EPS, which are responsible for, respectively, establishing and maintaining pluripotency, promoting epidermal differentiation, and maintaining its progenitor state, cardiac lineage specification, neurogenesis, and immunomodulation [29,33,34,35,36]. Even though there are a multitude of aforementioned functions, a description of the exact role of specific lincRNAs is still difficult to accomplish. This class of molecules remains one of the poorest understood so far.
The study by Cabili et al. indicated that lincRNA expression is remarkably more tissue specific compared to coding genes [37]. Surprisingly, the enrichment of the repressive H3K9me3 modification at lincRNA gene promoters is associated with higher tissue specificity, instead of low-expressed tissue-specific mRNAs [38]. This feature makes lincRNAs very promising diagnostic or prognostic biomarkers, not only in HNSCC but in many more diseases [23,26].

3. lincRNAs as Biomarkers and How to Find Them

Epidemiology with the mortality rate of HNSCC designates a tremendous need for finding a specific, non-invasive biomarker of the early stages of the disease. The abundance of lncRNAs, their tissue specificity, and association with different cancers’ abnormal expression landscape suggest promising results within this group of potential diagnostic markers. Furthermore, it has been proven that they can be detected not only in tissue samples but also in body fluids [23,26,39,40,41]. However, the fact that they are tissue and even cell specific makes their profile particular for each type of collected biological material [42]. Studies investigating lncRNAs as salivary, plasma, or urine biomarkers proved that their stability, half-life, and resistance to RNase digestion meet the requirements of a good biomarker [20,23,42]. Surprisingly, PCA3 became the first FDA-approved (Food and Drug Administration, USA) lncRNA-based biomarker whose predictive values exceed those characteristic of PSA serum testing [23,43].
There are two types of methods used in lincRNAs profiling: RNA-centric methods, especially ChIRP (chromatin isolation by RNA purification), CHART (capture hybridization of RNA targets) or RAP (RNA antisense purification), andprotein-centric methods, such as different variants of nRIP (native RNA immunoprecipitation), and CLIP (crosslinking immunoprecipitation), which are based on lincRNAs’ known ability to encode micropeptides [26]. These techniques are excellent for academic purposes but too expensive and complex for diagnostic use. Even though the classical method of RNA isolation with TRIzol or column-based protocols raises questions about its accuracy for this kind of molecule, it appears to not affect the quantification results [20,44]. Moreover, carefully designed collection tubes and column protocols can minimize the negative effect of background RNAs from coagulation, blood cell contamination, as well as progressive hemolysis on results [44,45]. After isolation of RNA, the most common methods to perform are microarrays, NGS (new generation sequencing), or simply qRT-PCR. The latter one, being the least expensive and the most popular, is a golden standard for lincRNA quantification, especially in the field of diagnostics [46]. The lincRNA amplification could be a challenging step due to the different evolutionary conservation patterns than in protein-coding genes. The level of conservation can depend on the function of the molecule, and linker sequence patterns will differ from functional modules [47]. However, it has been proven that thousands of lncRNAs are in fact evolutionarily conserved and the ones that are not are likely to have conserved promoter regions [47,48].
Liquid biopsy may be a good response to the aforementioned need for a non-invasive, accurate method of detecting the disease, assessing its stage, and monitoring the course of its treatment. For this type of analysis, non-solid biological materials, such as blood and its fractions, are obtained [49,50,51,52]. They contain many biomarkers that allow the assessment of changes in the cancer phenotype (modifications of the genome, epigenome, and transcriptome), immunophenotype, response to therapy, and the incidence of infection [49,50,51,52,53,54,55,56]. Such characterization cannot always be performed by traditional biopsy, because it often does not fully reflect tumor heterogeneity [57,58] or cannot be obtained at all, as in the case when neoadjuvant therapy reduces the volume of the neoplastic lesion to an undetectable size [50]. Liquid biopsy can be based on cells, e.g., circulating tumor cells (CTCs), or circulating endothelial cells (CECs) and molecules, like DNA, e.g., circulating tumor DNAs (ctDNAs), or cell-free fetal DNAs (cffDNAs) and RNA, e.g., circulating tumor RNAs (ctRNAs), or circulating free RNAs (cfRNAs) [49,59,60,61]. The study by Umu et al. indicated that some of the RNA classes are highly expressed in serum and presented the percentage distribution of their uniquely-mapped reads: miRNA (45.7%), mRNA (20.3%), miscRNA (11.8%), lncRNA (10.7%), piRNA (4.3%), tRFs (1.9%), and others (5.3%) [53]. Even though lncRNAs abundance is not very high, knowledge about their miRNA targets can provide us with a specific landscape of molecular changes resulting from tumor growth or its post-treatment remission. Carefully designed steps of liquid biopsy can help minimize its challenges, such as sample processing, extraction techniques, quality and quantity assessment, and data normalization [62]. Moreover, our team previously made recommendations on how to cope with lncRNA stability and its low copy number [63]. We have also elucidated that detection of rare cfRNAs during qRT-PCR can be accomplished with the proper use of stem-loop-specific primers or adding poly(A) tails. Additionally, if choosing the right reference gene and performing isoform specification will not improve the efficiency of qRT-PCR, we can replace it with a droplet digital PCR method (ddPCR), which is sensitive to a very low amount of material [63,64]. It is worth mentioning that two FDA-approved liquid biopsy methods are already available: one of them is based on CTCs and dedicated to various types of cancers, and the second one is Cobas EGFR Mutation Test (Roche Molecular Systems, Inc.), which uses cfDNA isolated from plasma of individuals with metastatic non-small cell lung cancer (NSCLC) [65,66].
The aforementioned features prove that lincRNAs are promising, specific, easy-to-access future biomarkers that can revolutionize the process of disease detection and monitoring the effects of the applied treatment. However, much research remains to be done to characterize these molecules in detail, understand their role in cancer biology, and validate their assay methods.

4. Known lincRNA Biomarkers in HNSCC

The growing number of studies indicates the importance of changing the landscape of lincRNAs expression levels in different malignancies. Their potential role as biomarkers was discussed considering colorectal, gastric, prostate cancer, or HNSCC [38]. The earliest discovered lincRNA molecules XIST, MALAT1, HOTAIR, and HOTTIP are rarely distinguished as an intergenic subtype of long non-coding RNAs [26]. Despite the little information on the role of lincRNAs in HNSCC, we tried to collect and describe molecules with the potential to become biomarkers of early detection or prognosis in HNSCC below.
X-inactive specific transcript (XIST) is a lincRNA molecule, whose sequence is localized within the XIST gene (Xq13.2) [67]. It has been proven that XIST is up-regulated in many tumors, including glioblastoma [68], hepatocellular carcinoma (HCC) [69], breast cancer (BC) [70], NSCLC [71], as well as nasopharyngeal carcinoma (NPC) [72], which suggests that it can become a valuable diagnostic biomarker specific for this group of diseases. Studies carried out on mouse models indicated that silencing or knocking down XIST caused decreased cell growth and metastasis, which implies an essential role in the development and progression of malignancies [68,69,70]. Song et al. proved the prognostic value of XIST in NPC. Furthermore, they showed that the XIST expression level increased with tumor size and stage, leading to poor survival in a group of patients with a high level of this molecule [72]. All of the above underline the diagnostic and therapeutic potential of XIST in HNSCC.
Metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) has one of the most conserved primary and secondary structures of all lincRNAs [73]. Its sequence is localized inside the non-coding nuclear-enriched abundant transcript 2 (NEAT2) [74]. MALAT1 is one of the modulators of pre-mRNA processing, principally by regulating splicing efficiency, with its ability to sequester splicing factors while being retained in nuclear speckles [73,74]. The study by Hu et al. showed that this molecule acts as an oncogene in esophageal squamous cell carcinoma, promoting its growth by regulating the ATM-CHK2 pathway, which is associated with G2/M transition and processes of DNA damage response [75]. Oncogenic MALAT1 was also positively correlated with clinical stage in other malignancies such as glioma, pancreas, prostate, and lung cancer [74,75,76]. The above lincRNA is overexpressed in all HNSCC localizations [74]. Hu et al. proposed that a higher level of MALAT1 can be caused by its amplification in tumor tissue [75]. Moreover, Zhou et al. implicated that patients with overexpression of this molecule are characterized by unfavorable prognosis and significantly shorter overall survival (OS) [74].
HOX transcript antisense RNA (HOTAIR) is expressed from locus HOXC, interacts with polycomb repressive complex 2 (PRC2) and plays a pivotal role in the H3K27 methylation of many genes. This complex is responsible for epigenetic silencing of different sequences during many important cellular processes, even cancerogenesis [77]. lincRNA HOTAIR is significantly overexpressed in several types of malignant tumors, including esophageal squamous cell carcinoma (ESCC) and oral squamous cell carcinoma (OSCC) [78,79,80]. In recent years, many studies have indicated that a high level of HOTAIR molecules is associated with poor prognosis and overall survival of cancer patients [78,79,80]. Ge et al. discovered that cell cultures with up-regulation of HOTAIR have abnormally activated Wnt signaling pathways due to a decrease in Wnt-inhibitory factor 1 expression, which results in progression, increased migration, and the ability to create metastasis [78]. Analysis by Li et al. described the regulatory function of HOTAIR, which may interact with a broad range of genes involved in cell differentiation, death, adhesion, and cell cycle [79]. What is more, Tang et al. documented that this lincRNA can be easily detected in saliva, particularly in more advanced, metastatic stages of the disease [42]. HOTAIR lincRNA has significant prognostic potential, which after further investigation may lead in the future to the creation of the specific diagnostic or prognostic molecular test.
The HOXA transcript at the distal tip (HOTTIP) is localized within the HOXA cluster and regulates the activation of multiple HOXA genes by controlling H3K4 methylation [41,81,82]. Over the years, different studies have implied that up-regulation of this lincRNA expression is crucial for tumor development, growth, and metastasis in many cancers, e.g., HCC [82] and tongue squamous cell carcinoma (TSCC) [41]. Zhang et al. demonstrated that overexpression of HOTTIP is characteristic of TSCC and its high level is positively associated with several clinicopathological features along with the ability to create distant lesions of metastasis. Moreover, this correlation allows suggesting that HOTTIP may become a valuable prognostic biomarker in the future [41]. A few years later, Yin et al. proved that HOTTIP has the strongest prognostic value within a group of approximately 1000 lncRNAs differently expressed in HNSCC. The researchers confirmed the association of the expression of this molecule with the grade, stage, and overall survival of patients [83]. Additionally, it was characterized as an independent prognostic factor. The above implies that a high level of HOTTIP is pivotal for cancer development, proliferation, and progression, which together with a strong correlation with survival prognosis makes it a unique, valuably diagnostic biomarker in HNSCC [41,83].
Long intergenic non-coding RNA p21 (lincRNA-p21) was first described as a p53-dependent apoptotic response repressor in studies investigating this process in mouse embryonic fibroblasts with wild-type TP53 [84]. These molecules’ impact is especially interesting due to the fact that 85% of HNSCC patients have mutated versions of the TP53 gene [85]. Recently, its pivotal role in the development and progression of multiple cancer was described. Jin et al. observed that a low level of lincRNA-p21 causes drastic progression of HNSCC due to the lack of induced G1 phase arrest and inhibition of apoptosis. Interestingly, researchers observed that this lincRNA displays their suppressor function by decreasing activity of the Janus kinase 2 (JAK2)/signal transducer and activator of transcription 3 (STAT3) pathway by binding to the latter one. It is worth mentioning that HNSCC patients with down-regulated expression of lincRNA-p21 had an unfavorable prognosis [86]. Studying the role and molecular mechanism of lincRNA-p21 interactions is especially important because of its direct association with TP53 and its targets. Future research should focus on exploring this subject because it can become a unique predicting factor for HNSCC patients.
Long non-protein-coding RNA 2487 (LINC02487) is a molecule mostly localized in the cytoplasm, retained close to the nuclear membrane, where it displays its regulatory function at the post-transcriptional or post-translational level [87]. Recently, it has been proven that this lincRNA is dysregulated in OSCC and its characteristic for oral carcinogenic tissue low expression is correlated with unfavorable clinical outcome and poor survival [88]. Feng et al. determined that LINC02487 expression is correlated with the OSCC development stage and its level increases from the amount characteristic for adjacent normal tissue, through cancer tissue to the highest expression value in samples from patients with metastasis. Part of their study based on cancerous cell cultures resulted in the observation that overexpression of LINC02487 has an inhibiting impact on OSCC proliferation, migration, and invasiveness [87]. Additionally, up-regulation of LINC02487 regulates the levels of EMT markers and causes a decrease in N-cadherin and vimentin, along with an increase in E-cadherin, through interaction with ubiquitin carboxyl-terminal hydrolase 17 (USP17), a known EMT regulator [89]. Significant differences between tumor tissue and healthy oral mucosa, taken together with its proven OSCC suppressor role [86,87], indicate that LINC02487 can become a unique diagnostic and prognostic biomarker that could lower the high mortality rate caused by distant metastasis [88,90].
Long non-protein-coding RNA 2195 (LINC02195) is closely associated with major histocompatibility complex class I (MHC I) molecules, whose lack of function leads to the mechanism of escaping immunosurveillance by tumor in HNSCC [91,92]. Analysis of LINC02195 expression patterns showed significantly higher lincRNA levels in HNSCC tumor samples than in cell lines derived from dysplastic tissue or normal mucosa. Li et al. also found out that silencing LINC02195 expression causes a decrease in the level of MHC I molecules and definitely proved that patients with high expression of this lincRNA have a better prognosis. Further research indicated that this lincRNA is immune-related due to its significant correlation with the T cell receptor pathway, chemokines (class I and II), and cytokines [91]. The discovered association with lymphocytes T suggests better infiltration of tumor tissue, which leads to the positive response to immunotherapy and better prognosis, not only in HNSCC but in many different cancers [91,92,93]. The above information implies that LINC02195 is a promising prognostic factor and therapeutic target; however, more research needs to be done in this field.
Long intergenic non-protein-coding RNA 668 (LINC00668) is one of the least characterized lincRNAs described in this paper. Its up-regulation in HNSCC—particularly in OSCC and laryngeal squamous cell carcinoma (LSCC) [94,95]—has been proven. LINC00668 promotes tumor growth in OSCC cells by interaction with miR-297 and VEGFA signaling pathways. Zhang et al. implied that this lincRNA acts as an oncogene promoting OSCC tumorigenesis [94]. Furthermore, Zhao et al. showed that this lincRNA’s expression level in LSCC is associated with age, stage, and cervical lymph node metastasis. They also suggest that these molecules can enhance the proliferation, migration, and invasion ability of the studied cell lines [95]. LINC00668 might become a valuable diagnostic or prognostic biomarker of LSCC.
Long intergenic non-protein-coding RNA 519 (LINC00519) was first described in lung squamous cell carcinoma (LUSC) as an oncogene [96]. Localization analysis indicated that the vast majority of these lincRNA molecules can be found in the cytoplasm where they act as an miRNA-sequestering sponge [97]. In TSCC, one of those short RNAs is miR-876-3p, which was earlier widely described as aberrantly expressed and tumorigenic in many cancers [98]. The study based on TSCC patients with high expression of LINC00519 demonstrated shorter overall survival and unsatisfactory prognosis [97]. Although the above lincRNA displays the potential to become an HNSCC biomarker, still a lot of work remains to be done.
Long intergenic non-protein-coding RNA 511 (LINC00511) is up-regulated in different malignancies including also HNSCC [99]. This molecule has a variety of functions, including regulation of the developmental process, apoptosis, programmed cell death, focal adhesion through hemostasis, and different carcinogenic pathways [100]. It has been proven that LINC00511 modulates TSCC progression by promoting cell proliferation and migration [101]. Moreover, a high level of LINC00511 is strongly associated with age, tumor size, clinical stage, lymph node metastasis, as well as unsatisfactory prognosis [102]. Growing evidence supports the proposition of these lincRNAs as a potential novel therapeutic target and biomarker.
Long intergenic non-protein-coding RNA 460 (LINC00460) has in recent years gained more and more interest as it appears to be an oncogene in different cancers, e.g., NPC, ESCC, and lung cancer [103,104,105,106,107]. Chaudhary et al. investigated LINC00460 expression levels in different subgroups of HNSCC patients. Their study proved this lincRNA may serve in the future as an independent prognostic biomarker of patients’ survival, especially in a subgroup of individuals who have not undergone HPV infection. Further analysis associated a high level of LINC00460 with several carcinogenic pathways, indicating its involvement in cell development, proliferation, the EMT, and adhesion [107]. Additionally, ESCC-based research showed that this lincRNA not only promotes tumor growth but also exerts its oncogenic role by apoptosis [104]. Xie et al. investigated the mechanism underlying the tumorigenic impact of LINC00460. They proved that p21, E-cadherin, N-cadherin, and cyclin D1 are above lincRNA targets, which corroborates the importance of LINC00460 overexpression in cell cycle, migration, the EMT, and invasion [105]. Interestingly, this lincRNA regulates the activity of miR-162 by acting as a molecular sponge. Inhibition of miR-612 causes up-regulation of serine/threonine kinase 2 (AKT2), which leads to progression, metastasis, and unfavorable prognosis [106,108,109]. LINC00460 exhibits the features of a good prognostic marker in HNSCC.
Long intergenic non-protein-coding RNA 312 (LINC00312) is localized on 3p.25.3 loci on a chromosome, which is a very common region of allelic loss, especially in NPC [110]. Another name for this lincRNA is a novel putative tumor suppressor, ornasopharyngeal carcinoma candidate 7 (NAG7). LINC00312 is overexpressed in NPC cell lines, which leads to an increase of adhesion, motility, and invasiveness and inhibits proliferation, by arresting progression from G1 to S phase of the cell cycle [110,111]. The study by Zhang et al. discovered a positive correlation of LINC00312 expression with lymph node metastasis and negative association with the stages and size of the tumor. Their results showed that this lincRNA could be a useful biomarker that allows distinguishing healthy individuals from NPC patients and, within the latter group, determine who developed distant metastasis [110,111]. Considering the above data, LINC00312 could serve in the future as a biomarker of NPC, specifically its metastatic stages [110].
Long intergenic non-protein-coding RNA 52 (LINC00052) displays a regulatory function by acting as a molecular sponge and sequestering different miRNAs [112,113,114]. Even though its mechanism of action is not fully known, multiple studies have indicated its aberrant expression level in several cancers and corresponding cell lines, e.g., hepatocellular carcinoma (HCC) [115,116,117], BC [113], and HNSCC [114]. We have conducted research analyzing the expression of LINC00052 in HNSCC cell lines and patients’ samples to study its impact on cancer biology and investigated its potential as a future biomarker; however, the results obtained have not been published yet. Our team managed to prove that the level of this molecule is significantly higher in tumor tissue compared to the control sample. Interestingly, top values of expression were characteristic for patients with a mutated version of the TP53 gene. We have found out that the LINC00052 level is associated with gender, cancer, T and N stage, as well as perineural invasion and HPV status. It has been proven that LINC00052 is negatively correlated with miR-27b-5p, a known modulator of the EMT process [118], whose targets are involved in several important cellular processes. Moreover, we have observed that patients with higher expression of miR-27b-5p and lower LINC00052 have significantly longer survival, higher infiltration of immune cells, and substantial down-regulation of EMT regulators such as vimentin, MMP16, MMP2, TGFBR2, TGFBI, ITGB3, PDGFRB, SOX11, ZEB2, and FOXD1. The presented results signify that LINC00052 could be a very promising prognostic biomarker, especially in combination with miR-27b; nevertheless, this correlation needs to be further explored. A summary of these well-described lincRNAs is presented in Table 1.
Previously, our team performed a liquid biopsy on HNSCC patients and healthy individuals to analyze lncRNA plasma expression differences between these two groups and to determine their diagnostic potential. In agroup of 90 lncRNA transcripts, 20 lincRNAs, including ANRIL, Dios3os, Emx2os, GAS5, H19, HAR1B, HULC, Jpx, lincRNA-RoR, MALAT1, MEG9, ncR-uPAR, NEAT1, NRON, RNCR3, SNHG1, SNHG6, Tsix, UCA1, and Zfas1, had significantly higher expression levels in metastatic and/or recurrent patients’ samples in comparison to healthy individuals [52]. We checked whether lncRNA is lincRNA in the LNCipedia database (version 5.2) [119].
The up-regulation of lincRNA ANRIL, also called CDKN2B antisense RNA 1 (CDKN2B-AS1),has been described in many different cancers and correlated with tumor progression [120,121,122]. The study by Zhang et al. indicated that this lincRNA promotes HNSCC tumorigenesis by regulating EGFR1 expression through sponging mir-125a-3p [120]. The Dio3os overexpression has been described as an oncogenic molecule in pancreatic [123] and thyroid cancer, and as a risk factor for the latter patients’ overall survival [124]. Our recent study indicated that this lincRNA level in cell cultures is sensitive to radiation and tends to decrease after irradiation [125]. The Emx2os molecule is an antisense transcript of homeobox protein Emx2os, a known transcription factor with tumor suppressor abilities, e.g., in LSCC [126]. Its up-regulation is associated with intensified tumor proliferation and migration along with a poor prognosis for ovarian cancer patients [127]. Moreover, we have also proved that high Emx2os expression levels can negatively affect progression-free survival (PFS) of individuals with HNSCC [52]. Growth arrest-specific 5 (GAS5) lincRNA was identified as a poorly conserved tumor suppressor that also acts as a decoy for the glucocorticoid receptor (GR) [47,128,129]. The study by Fayda et al. proposed that this lincRNA could become a useful biomarker of chemotherapy treatment response in head and neck cancer [130]. H19 is up-regulated in many different malignancies and promotes oncogenesis along with drug resistance by regulating DNA methylation genome-wide through interactions with S-adenosylhomocysteine hydrolase [131]. Guan et al. proved that high expression levels of this lincRNA together with up-regulation of miR-675 promote tumor growth in HNSCC patients [132]. Additionally, H19 induces EMT and promotes invasion in NPC by regulating the miR-630/EZH2 axis [133]. HAR1B (HAR reverse) is an antisense lincRNA transcribed from the opposite strand of the “human-accelerated” region 1 (HAR1) [134]. Our previous studies indicated that these molecules’ expression level decreases after radiation, which can lead to disruption of many important processes and pathways, e.g., cell cycle, cadherin, Wnt, and angiogenesis signaling pathways [125]. Interestingly, Yamada et al. proved that HAR1B could be a useful biomarker of pazopanib therapy response in patients with bone or soft-tissue sarcomas [135]. The lncRNA highly up-regulated in liver cancer (HULC) is associated with tumor progression not only in, as the name suggests, hepatic cancer but also in gastric and pancreatic cancer, liver metastasis of colorectal cancer, and OSCC [136,137,138,139,140]. Su et al. indicated that suppression of this lincRNA expression in OSCC cell lines increases their apoptosis rate and inhibits their proliferation, migration, and invasion [140]. The just proximal to XIST lncRNA (JPX) overexpression is known to promote tumorigenesis and metastatic lesion development in lung cancer by targeting miR-33a-5p, which causes up-regulation of its downstream gene Twist1, leading to activation of the Wnt/β-catenin signaling pathway [141]. The lincRNA regulator of reprogramming (lincRNA-RoR) is an oncogene involved in EMT along with drug resistance in different malignancies, e.g., NPC, BC, and HNSCC [142,143,144]. Interestingly, we have observed that its expression level increases after exposure to cisplatin [52]. One of the most poorly studied lincRNA is maternally expressed 9 (MEG9), speculated to be induced by hypoxia in a mouse model by Voellenkle et al. [145]. We have found that lincRNA upstream of the PAR-1 (ncR-uPAR) is down-regulated after cisplatin administration compared to non-treated controls [52]. The aberrant expression of lincRNA nuclear paraspeckle assembly transcript 1 (NEAT1) in different cancers has been widely described. Previously, we conducted a study regarding its still pending role in HNSCC biology [146]. The non-coding RNA repressor of NFAT (NRON) acts as a part of a scaffold that binds the nuclear factor of activated T cells (NFAT) and subsequently could affect the T cell activation and immune system response to cancer [147]. Shang et al. proved that the lincRNA retinal non-coding RNA3 (RNCR3) is significantly up-regulated in an inflammatory and tumor microenvironment, promotes myeloid-derived suppressor cells (MDSCs) differentiation, and functions as a sponge for miR-185-5p in a mouse model [148]. The lincRNA small nucleolar RNA host genes (SNHGs) such as SNHG1 and SNHG6 are also aberrantly expressed in HNSCC [52]. In LSCC, up-regulation of SNHG1 was proved to promote proliferation, EMT, and metastasis and was connected with patients’ poor survival [149]. Interestingly, we have indicated that high expression of this lincRNA is correlated with better OS in HNSCC patients [52]. The SNHG6 lincRNA has been described as an oncogene in tongue cancer and OPSCC [150,151]. In our study, we proved its aberrant expression in HNSCC patients, especially in the group with the progressive disease [52]. Additionally, we observed an increase in its level after cisplatin treatment, which could disrupt the mechanism of different molecular pathways, such as cadherin, Wnt signaling pathway, or TP53 pathway [125]. The lincRNA Tsix is a negative regulator of lincRNA XIST that inhibits its function, which in this case can lead to tumor progression [152,153]. Salama et al. suggested that lincRNAs XIST and Tsix could become stable non-invasive immune biomarkers for BC patients [154]. The lincRNA urothelial cancer-associated 1 (UCA1) is described as an oncogene associated with cancer progression [24]. It was proved that elevated levels of the above lincRNA can induce cell migration and are correlated with lymph node metastasis in TSCC [155]. The ZNFX1 antisense RNA 1 (Zfas1) was described as an oncogenic lincRNA in a multitude of different cancers, e.g., NSCLC [156], HC [157], ESCC [158], and HNSCC [159]. We have elucidated its role in HNSCC biology and proved its diagnostic potential in previous research [159].
Additionally, our lncRNA-based study indicated that lincRNAs HAR1B, Jpx, and NEAT1 levels differ between localizations of HNSSC tumors. As we mentioned above, some of the lincRNAs are correlated with treatment response and can become biomarkers of primary chemotherapy resistance in the future. We have indicated that lncRNAs can serve as diagnostic biomarkers that will help to distinguish healthy individuals from HNSCC patients [52].
According to the UALCAN database presenting results based on available TCGA data [160], additional significantly (p<0.05) changed lincRNAs, which were not previously described in HNSCC, were indicated. Among the changed lincRNAs, expression levels of 24 are up-regulated, 4 down-regulated, and 5 not changed compared to normal samples. These changed lincRNAs were connected in different ways with stages and cancer grades. Moreover, only 4 of 29 lincRNAs are associated with patients’ survival time: higher levels of LINC00115 (p = 0.049), LINC00158 (p = 0.0076), and LINC00167 (p = 0.042), as well as a lower level of LINC00460 (p = 0.00074) were connected with significantly better patients’ survival. A schematic representation of UALCAN results is shown in Figure 2A. The detailed results are freely available on the UALACAN database: http://ualcan.path.uab.edu (accessed on 15 April 2021).

5. lincRNAs in HPV-Positive HNSCC

As mentioned above, HPV infection is a well-known predictive factor for HNSCC patients [7,10]. HPV infection causes differences in the cellular program, which is manifested by changes in protein-coding and non-coding RNA transcripts [161,162]. More and more studies are focusing on the role of lncRNAs in HPV infection, e.g., TCGA analysis revealed differences in 177 lncRNAs between HPV(+) and HPV(−) HNSCC patients including 75 up- and 102 down-regulated [163]. However, the function of different lincRNAs is not fully understood. In the case of lncRNAs, our results indicated the role of lncRNA EGOT, PRINS, and CDKN2B-AS1 (ANRIL) [164,165]. Moreover, in our published work, we observed changes in lincRNAs including up-regulation of TTTY14, and TTTY15, and down-regulation of MEG3 and H19 in HPV(+) patients in comparison to HPV(−). We also observed that TTY14, TTY15, and MEG3 showed the high discrimination potential of HPV(−) and HPV(+) patients. However, no differences in the case of MALAT1 and CYTOR (LINC00152) depending on the HPV status based on TCGA data were noticed [165]. Tomar et al. study based on HNSCC samples indicated changes of TTTY14, TTTY15, XIST, and CYTOR (LINC00152) depending on the HPV infection and activity status [166]. A previous publication indicated that MALAT1, MEG3, and H19 are associated with HPV infection in the case of cervical cancer and probably it could take some function in HPV(+) HNSCC cases but there was no fully described experimental evidence based on the HNSCC model [167]. However, in the doctoral dissertation of Tomar (2013), changes associated with HPV infection in HNSCC for MALAT1 as well as other lincRNAs, such as LINC0002, LINC00028, LINC00087, LINC00152, LINC00173, LINC00174, LINC00230A, LINC00240, LINC00263, LINC00319, LINC00426, LINC00472, LINC00487, LINC00277, LINC00339, and LINC00476, were presented [168] and are summarized in Figure 2B. We postulate that the biological role of these lincRNAs needs to be verified based on the in vitro model and that it should be established whether they may be used as potential biomarkers. Song et al. identified that lincRNA lnc-IL17RA-11 had the highest correlation with HPV infection among those analyzed using TCGA data. The authors observed that lnc-IL17RA-11 expression is up-regulated by transcription factor ER-alpha, which is associated with HPV infection. High levels of lnc-IL17RA-11 and co-expressed genes are involved in the cell cycle, DNA replication, and base excision repair pathways, which influences the cellular phenotype. Moreover, patients with a higher level of lnc-IL17RA-11 displayed better survival, and the expression level of this lincRNA was higher in the group of HPV(+) than HPV(−). However, it is difficult to clearly say that only lnc-IL17RA-11 influences the obtained results because in high- and low-expression groups of patients, we could find HPV(+) and HPV(−) patients. Moreover, no evidence was presented by Song et al. about the direct or indirect regulation of lnc-IL17RA-11 on co-expressed genes and the role of lincRNA in potential sensitivity to radiotherapy [163].

6. Conclusions and Future Perspectives

HNSCC is one of the most common cancers, with a very high mortality rate at the same time. Moreover, the proposed standard treatment often leads to serious side effects, acquired resistance, subsequent relapse, and distant metastasis. Despite several clinical trials testing novel therapy solutions, there is still tremendous demand for personalized medicine and a panel of unique, specific biomarkers detecting the early stage of HNSCC or implicating its prognosis.
In recent years, RNA molecules have gained great interest due to their broad range of functions and crucial impact on every molecular pathway and process. However, their ability to create complex networks of interactions does not make it easy to understand and characterize their role, mechanism of action, and influence on human biology. lincRNAs belong to one of the least described subgroups of RNA molecules. Even though nowadays we can identify particular molecules in a large group of aberrantly expressed ones, still much work in the field of functional studies remains to be done. The lincRNAs discussed in this review are promising diagnostic and predicting biomarkers of the whole group of HNSCC tumors. Additionally, some of them act as tumor suppressors, e.g., lincRNA-p21 or LINC02487, Emx2os, and GAS5, but others play an oncogenic role, e.g., MALAT1, LINC00668, LINC00519, LINC00460, Dio3os, H19, SNHG6, UCA1, or Zfas1, and can be used as a therapeutic target in the future. Nevertheless, each one of them needs to be studied in a cohort of patients and to pass the validation process. This can be very challenging due to the carefulness required in designing sample processing, extraction techniques, quality and quantity assessment, and data normalization. It is worth mentioning that more and more publications describe the interactions between different groups of RNA molecules with proteins or transcription factors, but very few address the impact of the chemotherapy or radiation phenomenon on the function of the lincRNAs network. However, the use of TCGA data such as those in the UALCAN database is the best solution for further extension of lincRNA knowledge in HNSCC, especially for selection and validation of predicted transcripts. In our opinion, it is the best solution and brings new discoveries in the biology of HNSCC, especially in the diagnostics field as well as in personalization of therapy.
Although considered “junk DNA,” long intergenic non-protein-coding RNAs became valuable and significant players in the field of cancer research. Even though there is still much work ahead of us, we can confidently say that lincRNAs will revolutionize the diagnostics and understanding of cancer biology.

Author Contributions

Conceptualization, J.K., T.K. and K.L.; methodology, J.K. and T.K.; investigation, J.K., T.K., P.P., K.G., J.S. and M.S.; data curation, J.K., T.K., P.P., K.G., J.S. and M.S.; writing—original draft preparation, J.K. and T.K.; writing—review and editing, J.K., T.K., R.B., A.T. and K.L.; visualization, J.K.; supervision, T.K. and K.L. J.K.’s contribution to this paper is estimated to be 70% of all authors’ contribution. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Greater Poland Cancer Centre through an internal grant for the Laboratory of Cancer Genetics.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are available online with common access. The data analyzed during the current study are available from the corresponding author on reasonable request.

Acknowledgments

We would like to say thanks to the Greater Poland Cancer Centre for supporting our work.

Conflicts of Interest

The authors declare that there is no conflict of interest regarding the publication of this paper.

Abbreviations

linc-RoRlong intergenic non-protein-coding RNA, regulator of reprogramming
TINCRtissue differentiation-inducing non-protein-coding RNA
ANCRAngelman syndrome chromosome region
PNKYlong intergenic non-protein-coding RNA
lincRNA-EPSerythroid prosurvival lincRNA, also known as Ttc39aos1
miscRNAmiscellaneous RNA
piRNApiwi-interacting RNA
tRFstRNA-related RNA fragments
XISTX-inactive specific transcript
MALAT1metastasis-associated lung adenocarcinoma transcript 1
HOTAIRHOX transcript antisense RNA
HOTTIPHOXA transcript at the distal tip
NEAT2nuclear-enriched abundant transcript 2
lincRNA-p21long intergenic non-coding RNA p21
ANRILCDKN2B-AS1-CDKN2B antisense RNA 1
Dios3osDIO3 opposite strand upstream RNA
Emx2osEMX2 opposite strand/antisense RNA
GAS5growth arrest-specific 5
H19H19 imprinted maternally expressed transcript
HAR1Bhighly accelerated region 1B
HULChepatocellular carcinoma up-regulated long non-coding RNA
JpxJPX transcript, XIST activator
MEG3/9maternally expressed 3/9
ncR-uPARnon-coding RNA upstream of the PAR-1
NEAT1nuclear paraspeckle assembly transcript 1
NRONnon-coding repressor of NFAT
RNCR3retinal non-coding RNA3
SNHG1/SNHG6small nucleolar RNA host gene 1/6
TsixXIST antisense RNA
UCA1urothelial cancer-associated 1
Zfas1ZNFX1 antisense RNA 1
EGOTeosinophil granule ontogeny transcript
PRINSpsoriasis-associated non-protein-coding RNA induced by stress
TTTY14testis-specific transcript, Y-linked 14
TTTY15testis-specific transcript, Y-linked 15
CYTORcytoskeleton regulator RNA
lnc-IL17RA-11long, non-coding interleukin 17 receptor A-11

References

  1. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [Green Version]
  2. Leemans, C.R.; Braakhuis, B.J.M.; Brakenhoff, R.H. The molecular biology of head and neck cancer. Nat. Rev. Cancer 2010, 11, 9–22. [Google Scholar] [CrossRef]
  3. Maier, H.; Dietz, A.; Gewelke, U.; Heller, W.D.; Weidauer, H. Tobacco and alcohol and the risk of head and neck cancer. J. Mol. Med. 1992, 70, 320–327. [Google Scholar] [CrossRef]
  4. Blot, W.J.; McLaughlin, J.K.; Winn, D.M.; Austin, D.F.; Greenberg, R.S.; Preston-Martin, S.; Bernstein, L.; Schoenberg, J.B.; Stemhagen, A.; Fraumeni, J.F. Smoking and drinking in relation to oral and pharyngeal cancer. Cancer Res. 1988, 48, 3282–3287. [Google Scholar]
  5. Solomon, B.; Young, R.J.; Rischin, D. Head and neck squamous cell carcinoma: Genomics and emerging biomarkers for im-munomodulatory cancer treatments. Semin. Cancer Biol. 2018, 52 Pt 2, 228–240. [Google Scholar] [CrossRef]
  6. Curado, M.P.; Hashibe, M. Recent changes in the epidemiology of head and neck cancer. Curr. Opin. Oncol. 2009, 21, 194–200. [Google Scholar] [CrossRef]
  7. Gillison, M.L.; Chaturvedi, A.K.; Anderson, W.F.; Fakhry, C. Epidemiology of Human Papillomavirus–Positive Head and Neck Squamous Cell Carcinoma. J. Clin. Oncol. 2015, 33, 3235–3242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Llorente, J.L.; López, F.; Suárez, C.; Hermsen, M.A. Sinonasal carcinoma: Clinical, pathological, genetic and therapeutic ad-vances. Nat. Rev. Clin. Oncol. 2014, 11, 460–472. [Google Scholar] [CrossRef] [PubMed]
  9. Cramer, J.D.; Burtness, B.; Le, Q.T.; Ferris, R.L. The changing therapeutic landscape of head and neck cancer. Nat. Rev. Clin. Oncol. 2019, 16, 669–683. [Google Scholar] [CrossRef] [PubMed]
  10. Wang, Z.; Xia, R.-H.; Ye, D.-X.; Li, J. Human Papillomavirus 16 Infection and TP53 Mutation: Two Distinct Pathogeneses for Oropharyngeal Squamous Cell Carcinoma in an Eastern Chinese Population. PLoS ONE 2016, 11, e0164491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Chen, S.M.Y.; Krinsky, A.L.; Woolaver, R.A.; Wang, X.; Chen, Z.; Wang, J.H. Tumor immune microenvironment in head and neck cancers. Mol. Carcinog. 2020, 59, 766–774. [Google Scholar] [CrossRef]
  12. Gillison, M.L. Human papillomavirus-associated head and neck cancer is a distinct epidemiologic, clinical, and molecular entity. Semin. Oncol. 2004, 31, 744–754. [Google Scholar] [CrossRef]
  13. Fakhry, C.; Westra, W.H.; Li, S.; Cmelak, A.; Ridge, J.A.; Pinto, H.; Forastiere, A.; Gillison, M.L. Improved Survival of Patients With Human Papillomavirus-Positive Head and Neck Squamous Cell Carcinoma in a Prospective Clinical Trial. J. Natl. Cancer Inst. 2008, 100, 261–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Alsahafi, E.; Begg, K.; Amelio, I.; Raulf, N.; Lucarelli, P.; Sauter, T.; Tavassoli, M. Clinical update on head and neck cancer: Molecular biology and ongoing challenges. Cell Death Dis. 2019, 10, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Irani, S. miRNAs Signature in Head and Neck Squamous Cell Carcinoma Metastasis: A Literature Review. J. Dent. 2016, 17, 71–83. [Google Scholar]
  16. Vermorken, J.B.; Mesia, R.; Rivera, F.; Remenar, E.; Kawecki, A.; Rottey, S.; Erfan, J.; Zabolotnyy, D.; Kienzer, H.-R.; Cupissol, D.; et al. Platinum-Based Chemotherapy plus Cetuximab in Head and Neck Cancer. N. Engl. J. Med. 2008, 359, 1116–1127. [Google Scholar] [CrossRef] [Green Version]
  17. Karabajakian, A.; Toussaint, P.; Neidhardt, E.-M.; Paulus, V.; Saintigny, P.; Fayette, J. Chemotherapy for recurrent/metastatic head and neck cancers. Anti-Cancer Drugs 2017, 28, 357–361. [Google Scholar] [CrossRef]
  18. Ortiz-Cuaran, S.; Bouaoud, J.; Karabajakian, A.; Fayette, J.; Saintigny, P. Precision Medicine Approaches to Overcome Resistance to Therapy in Head and Neck Cancers. Front. Oncol. 2021, 11, 614332. [Google Scholar] [CrossRef] [PubMed]
  19. Martinez, B.V.; Dhahbi, J.M.; Lopez, Y.N.; Lamperska, K.; Golusinski, P.; Luczewski, L.; Kolenda, T.; Atamna, H.; Spindler, S.R.; Golusiński, W.; et al. Circulating small non coding RNA signature in head and neck squamous cell carcinoma. Oncotarget 2015, 6, 19246–19263. [Google Scholar] [CrossRef] [PubMed]
  20. Guglas, K.; Bogaczyńska, M.; Kolenda, T.; Ryś, M.; Teresiak, A.; Bliźniak, R.; Łasińska, I.; Mackiewicz, J.; Lamperska, K. lncRNA in HNSCC: Challenges and potential. WspółczesnaOnkologia 2017, 21, 259–266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Kolenda, T.; Przybyła, W.; Teresiak, A.; Mackiewicz, A.; Lamperska, K.M. The mystery of let-7d—A small RNA with great power. WspółczesnaOnkologia 2014, 18, 293–301. [Google Scholar] [CrossRef]
  22. Prensner, J.; Chinnaiyan, A.M. The Emergence of lncRNAs in Cancer Biology. Cancer Discov. 2011, 1, 391–407. [Google Scholar] [CrossRef] [Green Version]
  23. Schmitt, A.M.; Chang, H.Y. Long Noncoding RNAs in Cancer Pathways. Cancer Cell 2016, 29, 452–463. [Google Scholar] [CrossRef] [Green Version]
  24. Jiang, C.; Li, X.; Zhao, H.; Liu, H. Long non-coding RNAs: Potential new biomarkers for predicting tumor invasion and me-tastasis. Mol. Cancer 2016, 15, 62. [Google Scholar] [CrossRef] [Green Version]
  25. Mercer, T.R.; Dinger, M.; Mattick, J.S. Long non-coding RNAs: Insights into functions. Nat. Rev. Genet. 2009, 10, 155–159. [Google Scholar] [CrossRef] [PubMed]
  26. Ransohoff, J.D.; Wei, Y.; Khavari, P.A. The functions and unique features of long intergenic non-coding RNA. Nat. Rev. Mol. Cell Biol. 2018, 19, 143–157. [Google Scholar] [CrossRef]
  27. Djebali, S.; Davis, C.A.; Merkel, A.; Dobin, A.; Lassmann, T.; Mortazavi, A.; Tanzer, A.; Lagarde, J.; Lin, W.; Schlesinger, F.; et al. Landscape of transcription in human cells. Nature 2012, 489, 101–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Knauss, J.; Sun, T. Regulatory mechanisms of long noncoding RNAs in vertebrate central nervous system development and function. Neuroscience 2013, 235, 200–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Kretz, M.; Siprashvili, Z.; Chu, C.; Webster, D.; Zehnder, A.; Qu, K.; Lee, C.S.; Flockhart, R.J.; Groff, A.F.; Chow, J.; et al. Control of somatic tissue differentiation by the long non-coding RNA TINCR. Nat. Cell Biol. 2012, 493, 231–235. [Google Scholar] [CrossRef] [PubMed]
  30. Carlevaro-Fita, J.; Rahim, A.; Guigó, R.; Vardy, L.A.; Johnson, R. Cytoplasmic long noncoding RNAs are frequently bound to and degraded at ribosomes in human cells. RNA 2016, 22, 867–882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Koziol, M.; Rinn, J.L. RNA traffic control of chromatin complexes. Curr. Opin. Genet. Dev. 2010, 20, 142–148. [Google Scholar] [CrossRef] [Green Version]
  32. Statello, L.; Guo, C.-J.; Chen, L.-L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef]
  33. Loewer, S.; Cabili, M.N.; Guttman, M.; Loh, Y.-H.; Thomas, K.; Park, I.H.; Garber, M.; Curran, M.; Onder, T.; Agarwal, S.; et al. Large intergenic non-coding RNA-RoR modulates reprogramming of human induced pluripotent stem cells. Nat. Genet. 2010, 42, 1113–1117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Kretz, M.; Webster, D.; Flockhart, R.J.; Lee, C.S.; Zehnder, A.; Lopez-Pajares, V.; Qu, K.; Zheng, G.X.; Chow, J.; Kim, G.E.; et al. Suppression of progenitor differentiation requires the long noncoding RNA ANCR. Genes Dev. 2012, 26, 338–343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ramos, A.D.; Andersen, R.; Liu, S.J.; Nowakowski, T.J.; Hong, S.J.; Gertz, C.C.; Salinas, R.D.; Zarabi, H.; Kriegstein, A.R.; Lim, D.A. The Long Noncoding RNA Pnky Regulates Neuronal Differentiation of Embryonic and Postnatal Neural Stem Cells. Cell Stem Cell 2015, 16, 439–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Atianand, M.K.; Hu, W.; Satpathy, A.; Shen, Y.; Ricci, E.P.; Alvarez-Dominguez, J.R.; Bhatta, A.; Schattgen, S.; McGowan, J.D.; Blin, J.; et al. A Long Noncoding RNA lincRNA-EPS Acts as a Transcriptional Brake to Restrain Inflammation. Cell 2016, 165, 1672–1685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Cabili, M.N.; Trapnell, C.; Goff, L.; Koziol, M.; Tazon-Vega, B.; Regev, A.; Rinn, J.L. Integrative annotation of human large intergenic noncoding RNAs reveals global properties and specific subclasses. Genes Dev. 2011, 25, 1915–1927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Mele, M.; Mattioli, K.; Mallard, W.; Shechner, D.M.; Gerhardinger, C.; Rinn, J.L. Chromatin environment, transcriptional regulation, and splicing distinguish lincRNAs and mRNAs. Genome Res. 2017, 27, 27–37. [Google Scholar] [CrossRef] [Green Version]
  39. Yarmishyn, A.A.; Kurochkin, I.V. Long noncoding RNAs: A potential novel class of cancer biomarkers. Front. Genet. 2015, 6, 145. [Google Scholar] [CrossRef]
  40. Eissa, S.; Matboli, M.; Essawy, N.O.; Shehta, M.; Kotb, Y.M. Rapid detection of urinary long non-coding RNA urothelial car-cinoma associated one using a PCR-free nanoparticle-based assay. Biomarkers 2015, 20, 212–217. [Google Scholar] [CrossRef]
  41. Zhang, H.; Zhao, L.; Wang, Y.-X.; Xi, M.; Liu, S.-L.; Luo, L.-L. Long non-coding RNA HOTTIP is correlated with progression and prognosis in tongue squamous cell carcinoma. Tumor Biol. 2015, 36, 8805–8809. [Google Scholar] [CrossRef]
  42. Tang, H.; Wu, Z.; Zhang, J.; Su, B. Salivary lncRNA as a potential marker for oral squamous cell carcinoma diagnosis. Mol. Med. Rep. 2012, 7, 761–766. [Google Scholar] [CrossRef] [Green Version]
  43. Wei, J.T.; Feng, Z.; Partin, A.W.; Brown, E.; Thompson, I.; Sokoll, L.; Chan, D.W.; Lotan, Y.; Kibel, A.S.; Busby, J.E.; et al. Can Urinary PCA3 Supplement PSA in the Early Detection of Prostate Cancer? J. Clin. Oncol. 2014, 32, 4066–4072. [Google Scholar] [CrossRef] [Green Version]
  44. Qi, P.; Zhou, X.-Y.; Du, X. Circulating long non-coding RNAs in cancer: Current status and future perspectives. Mol. Cancer 2016, 15, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Qin, J.; Williams, T.L.; Fernando, M.R. A novel blood collection device stabilizes cell-free RNA in blood during sample shipping and storage. BMC Res. Notes 2013, 6, 380. [Google Scholar] [CrossRef] [Green Version]
  46. Murphy, J.; Bustin, S.A. Reliability of real-time reverse-transcription PCR in clinical diagnostics: Gold standard or substandard? Expert Rev. Mol. Diagn. 2009, 9, 187–197. [Google Scholar] [CrossRef] [PubMed]
  47. Johnsson, P.; Lipovich, L.; Grandér, D.; Morris, K.V. Evolutionary conservation of long non-coding RNAs; sequence, structure, function. Biochim. Biophys. Acta (BBA) Gen. Subj. 2014, 1840, 1063–1071. [Google Scholar] [CrossRef] [Green Version]
  48. Guttman, M.; Amit, I.; Garber, M.; French, C.; Lin, M.F.; Feldser, D.M.; Huarte, M.; Zuk, O.; Carey, B.W.; Cassady, J.P.; et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nat. Cell Biol. 2009, 458, 223–227. [Google Scholar] [CrossRef] [PubMed]
  49. Neumann, M.H.; Bender, S.; Krahn, T.; Schlange, T. ctDNA and CTCs in Liquid Biopsy—Current Status and Where We Need to Progress. Comput. Struct. Biotechnol. J. 2018, 16, 190–195. [Google Scholar] [CrossRef] [PubMed]
  50. Elazezy, M.; Joosse, S.A. Techniques of using circulating tumor DNA as a liquid biopsy component in cancer management. Comput. Struct. Biotechnol. J. 2018, 16, 370–378. [Google Scholar] [CrossRef] [PubMed]
  51. Palmirotta, R.; Lovero, D.; Cafforio, P.; Felici, C.; Mannavola, F.; Pellè, E.; Quaresmini, D.; Tucci, M.; Silvestris, F. Liquid biopsy of cancer: A multimodal diagnostic tool in clinical oncology. Ther. Adv. Med Oncol. 2018, 10, 1758835918794630. [Google Scholar] [CrossRef] [PubMed]
  52. Łasińska, I.; Kolenda, T.; Guglas, K.; Kopczyńska, M.; Sobocińska, J.; Teresiak, A.; Strzelecki, N.O.; Lamperska, K.; Mackiewicz, A.; Mackiewicz, J. Liquid lncRNA Biopsy for the Evaluation of Locally Advanced and Metastatic Squamous Cell Carcinomas of the Head and Neck. J. Pers. Med. 2020, 10, 131. [Google Scholar] [CrossRef]
  53. Umu, S.U.; Langseth, H.; Bucher-Johannessen, C.; Fromm, B.; Keller, A.; Meese, E.; Lauritzen, M.; Leithaug, M.; Lyle, R.; Rounge, T.B. A comprehensive profile of circulating RNAs in human serum. RNA Biol. 2018, 15, 242–250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Joosse, S.A.; Pantel, K. Genetic traits for hematogeneous tumor cell dissemination in cancer patients. Cancer Metastasis Rev. 2016, 35, 41–48. [Google Scholar] [CrossRef]
  55. Hong, D.K.; Blauwkamp, T.A.; Kertesz, M.; Bercovici, S.; Truong, C.; Banaei, N. Liquid biopsy for infectious diseases: Sequencing of cell-free plasma to detect pathogen DNA in patients with invasive fungal disease. Diagn. Microbiol. Infect. Dis. 2018, 92, 210–213. [Google Scholar] [CrossRef] [PubMed]
  56. Pezzuto, F.; Buonaguro, L.; Buonaguro, F.M.; Tornesello, M.L. The Role of Circulating Free DNA and MicroRNA in Non-Invasive Diagnosis of HBV- and HCV-Related Hepatocellular Carcinoma. Int. J. Mol. Sci. 2018, 19, 1007. [Google Scholar] [CrossRef] [Green Version]
  57. Bedard, P.L.; Hansen, A.R.; Ratain, M.J.; Siu, L.L. Tumour heterogeneity in the clinic. Nat. Cell Biol. 2013, 501, 355–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. McGranahan, N.; Swanton, C. Clonal Heterogeneity and Tumor Evolution: Past, Present, and the Future. Cell 2017, 168, 613–628. [Google Scholar] [CrossRef] [Green Version]
  59. Karachaliou, N.; Mayo-De-Las-Casas, C.; Molina-Vila, M.A.; Rosell, R. Real-time liquid biopsies become a reality in cancer treatment. Ann. Transl. Med. 2015, 3, 309. [Google Scholar] [CrossRef]
  60. Farinacci, M.; Krahn, T.; Dinh, W.; Volk, H.-D.; Düngen, H.-D.; Wagner, J.; Konen, T.; Von Ahsen, O. Circulating endothelial cells as biomarker for cardiovascular diseases. Res.Pract. Thromb. Haemost. 2018, 3, 49–58. [Google Scholar] [CrossRef] [Green Version]
  61. Nagy, B. Cell-free nucleic acids in prenatal diagnosis and pregnancy-associated diseases. EJIFCC 2019, 30, 215–223. [Google Scholar]
  62. Kolenda, T.; Guglas, K.; Baranowski, D.; Sobocińska, J.; Kopczyńska, M.; Teresiak, A.; Bliźniak, R.; Lamperska, K. cfRNAs as biomarkers in oncology—Still experimental or applied tool for personalized medicine already? Rep. Pract. Oncol. Radiother. 2020, 25, 783–792. [Google Scholar] [CrossRef]
  63. Kolenda, T.; Ryś, M.; Guglas, K.; Teresiak, A.; Bliźniak, R.; Mackiewicz, J.; Lamperska, K. Quantification of long non-coding RNAs using qRT-PCR: Comparison of different cDNA synthesis methods and RNA stability. Arch. Med. Sci. 2019, 15, 2–10. [Google Scholar] [CrossRef] [Green Version]
  64. Metzenmacher, M.; Váraljai, R.; Hegedüs, B.; Cima, I.; Forster, J.; Schramm, A.; Scheffler, B.; Horn, P.A.; Klein, C.A.; Szarvas, T.; et al. Plasma Next Generation Sequencing and Droplet Digital-qPCR-Based Quantification of Circulating Cell-Free RNA for Noninvasive Early Detection of Cancer. Cancers 2020, 12, 353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. cobas EGFR Mutation Test v2. Available online: https://www.fda.gov/Drugs/InformationOnDrugs/ApprovedDrugs/ucm504540.htm (accessed on 10 April 2021).
  66. 510(k) Substantial Equivalence Determination Decision Summary. Available online: https://www.accessdata.fda.gov/cdrh_docs/reviews/K031588.pdf (accessed on 10 April 2021).
  67. Brown, C.; Ballabio, A.; Rupert, J.L.; LaFreniere, R.G.; Grompe, M.; Tonlorenzi, R.; Willard, H.F. A gene from the region of the human X inactivation centre is expressed exclusively from the inactive X chromosome. Nat. Cell Biol. 1991, 349, 38–44. [Google Scholar] [CrossRef] [PubMed]
  68. Yao, Y.; Ma, J.; Xue, Y.; Wang, P.; Li, Z.; Liu, J.; Chen, L.; Xi, Z.; Teng, H.; Wang, Z.; et al. Knockdown of long non-coding RNA XIST exerts tumor-suppressive functions in human glioblastoma stem cells by up-regulating miR-152. Cancer Lett. 2015, 359, 75–86. [Google Scholar] [CrossRef]
  69. Zhuang, L.K.; Yang, Y.T.; Ma, X.; Han, B.; Wang, Z.S.; Zhao, Q.Y.; Wu, L.Q.; Qu, Z.Q. MicroRNA-92b promotes hepatocellular carcinoma progression by targeting Smad7 and is mediated by long non-coding RNA XIST. Cell Death Dis. 2016, 7, e2203. [Google Scholar] [CrossRef] [Green Version]
  70. Huang, Y.-S.; Chang, C.-C.; Lee, S.-S.; Jou, Y.-S.; Shih, H.-M. Xist reduction in breast cancer upregulates AKT phosphorylation via HDAC3-mediated repression of PHLPP1 expression. Oncotarget 2016, 7, 43256–43266. [Google Scholar] [CrossRef] [Green Version]
  71. Tantai, J.; Hu, D.; Yang, Y.; Geng, J. Combined identification of long non-coding RNA XIST and HIF1A-AS1 in serum as an effective screening for non-small cell lung cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 7887–7895. [Google Scholar] [PubMed]
  72. Song, P.; Ye, L.-F.; Zhang, C.; Peng, T.; Zhou, X.-H. Long non-coding RNA XIST exerts oncogenic functions in human naso-pharyngeal carcinoma by targeting miR-34a-5p. Gene 2016, 592, 8–14. [Google Scholar] [CrossRef]
  73. Tripathi, V.; Ellis, J.D.; Shen, Z.; Song, D.Y.; Pan, Q.; Watt, A.T.; Freier, S.M.; Bennett, C.F.; Sharma, A.; Bubulya, P.A.; et al. The Nuclear-Retained Noncoding RNA MALAT1 Regulates Alternative Splicing by Modulating SR Splicing Factor Phosphorylation. Mol. Cell 2010, 39, 925–938. [Google Scholar] [CrossRef] [Green Version]
  74. Zhou, X.; Liu, S.; Cai, G.; Kong, L.; Zhang, T.; Ren, Y.; Wu, Y.; Mei, M.; Zhang, L.; Wang, X. Long Non Coding RNA MALAT1 Promotes Tumor Growth and Metastasis by inducing Epithelial-Mesenchymal Transition in Oral Squamous Cell Carcinoma. Sci. Rep. 2015, 5, 15972. [Google Scholar] [CrossRef] [Green Version]
  75. Hu, L.; Wu, Y.; Tan, D.; Meng, H.; Wang, K.; Bai, Y.; Yang, K. Up-regulation of long noncoding RNA MALAT1 contributes to proliferation and metastasis in esophageal squamous cell carcinoma. J. Exp. Clin. Cancer Res. 2015, 34, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Gutschner, T.; Hämmerle, M.; Diederichs, S. MALAT1—A paradigm for long noncoding RNA function in cancer. J. Mol. Med. 2013, 91, 791–801. [Google Scholar] [CrossRef] [PubMed]
  77. Davidovich, C.; Zheng, L.; Goodrich, K.J.; Cech, T.R. Promiscuous RNA binding by Polycomb repressive complex 2. Nat. Struct. Mol. Biol. 2013, 20, 1250–1257. [Google Scholar] [CrossRef] [Green Version]
  78. 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]
  79. Li, X.; Wu, Z.; Mei, Q.; Guo, M.; Fu, X.; Han, W. Long non-coding RNA HOTAIR, a driver of malignancy, predicts negative prognosis and exhibits oncogenic activity in oesophageal squamous cell carcinoma. Br. J. Cancer 2013, 109, 2266–2278. [Google Scholar] [CrossRef] [Green Version]
  80. Lu, M.-Y.; Liao, Y.-W.; Chen, P.-Y.; Hsieh, P.-L.; Fang, C.-Y.; Wu, C.-Y.; Yen, M.-L.; Peng, B.-Y.; Wang, D.P.; Cheng, H.-C.; et al. Targeting LncRNA HOTAIR suppresses cancer stemness and metastasis in oral carcinomas stem cells through modulation of EMT. Oncotarget 2017, 8, 98542–98552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Wang, K.C.; Yang, Y.W.; Liu, B.; Sanyal, A.; Corces-Zimmerman, R.; Chen, Y.; Lajoie, B.R.; Protacio, A.; Flynn, R.; Gupta, R.A.; et al. A long noncoding RNA maintains active chromatin to coordinate homeotic gene expression. Nat. Cell Biol. 2011, 472, 120–124. [Google Scholar] [CrossRef] [Green Version]
  82. Tsang, F.H.; Au, S.L.; Wei, L.; Fan, D.N.; Lee, J.M.; Wong, C.C.; Ng, I.O.; Wong, C.-M. Long non-coding RNA HOTTIP is frequently up-regulated in hepatocellular carcinoma and is targeted by tumour suppressive miR-125b. Liver Int. 2014, 35, 1597–1606. [Google Scholar] [CrossRef] [PubMed]
  83. Yin, X.; Yang, W.; Xie, J.; Wei, Z.; Tang, C.; Song, C.; Wang, Y.; Cai, Y.; Xu, W.; Han, W. HOTTIP Functions as a Key Candidate Biomarker in Head and Neck Squamous Cell Carcinoma by Integrated Bioinformatic Analysis. BioMed Res. Int. 2019, 2019, 5450617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Huarte, M.; Guttman, M.; Feldser, D.; Garber, M.; Koziol, M.; Kenzelmann-Broz, D.; Khalil, A.M.; Zuk, O.; Amit, I.; Rabani, M.; et al. A Large Intergenic Noncoding RNA Induced by p53 Mediates Global Gene Repression in the p53 Response. Cell 2010, 142, 409–419. [Google Scholar] [CrossRef] [Green Version]
  85. Zhou, G.; Wang, J.; Zhao, M.; Xie, T.-X.; Tanaka, N.; Sano, D.; Patel, A.A.; Ward, A.M.; Sandulache, V.C.; Jasser, S.A.; et al. Gain-of-Function Mutant p53 Promotes Cell Growth and Cancer Cell Metabolism via Inhibition of AMPK Activation. Mol. Cell 2014, 54, 960–974. [Google Scholar] [CrossRef] [Green Version]
  86. Jin, S.; Yang, X.; Li, J.; Yang, W.; Ma, H.; Zhang, Z. p53-targeted lincRNA-p21 acts as a tumor suppressor by inhibiting JAK2/STAT3 signaling pathways in head and neck squamous cell carcinoma. Mol. Cancer 2019, 18, 1–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Feng, L.; Zhang, J.; Sun, M.; Qiu, F.; Chen, W.; Qiu, W. Tumor Suppressor LINC02487 Inhibits Oral Squamous Cell Carcinoma Cell Migration and Invasion Through the USP17–SNAI1 Axis. Front. Oncol. 2020, 10, 559808. [Google Scholar] [CrossRef]
  88. Xu, K.; Tian, H.; Zhao, S.; Yuan, D.; Jiang, L.; Liu, X.; Zou, B.; Zhang, J. Long Noncoding RNA LOC441178 Reduces the Invasion and Migration of Squamous Carcinoma Cells by Targeting ROCK1. BioMed Res. Int. 2018, 2018, 4357647. [Google Scholar] [CrossRef] [Green Version]
  89. Song, C.; Liu, W.; Li, J. USP17 is upregulated in osteosarcoma and promotes cell proliferation, metastasis, and epithelial–mesenchymal transition through stabilizing SMAD4. Tumor Biol. 2017, 39, 1010428317717138. [Google Scholar] [CrossRef] [Green Version]
  90. Rivera, C. Essentials of oral cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 11884–11894. [Google Scholar]
  91. Li, H.; Xiong, H.-G.; Xiao, Y.; Yang, Q.-C.; Yang, S.-C.; Tang, H.-C.; Zhang, W.-F.; Sun, Z.-J. Long Non-coding RNA LINC02195 as a Regulator of MHC I Molecules and Favorable Prognostic Marker for Head and Neck Squamous Cell Carcinoma. Front. Oncol. 2020, 10, 615. [Google Scholar] [CrossRef] [PubMed]
  92. Leone, P.; Shin, E.-C.; Perosa, F.; Vacca, A.; Dammacco, F.; Racanelli, V. MHC Class I Antigen Processing and Presenting Machinery: Organization, Function, and Defects in Tumor Cells. J. Natl. Cancer Inst. 2013, 105, 1172–1187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Matsushita, H.; Vesely, M.D.; Koboldt, D.C.; Rickert, C.G.; Uppaluri, R.; Magrini, V.J.; Arthur, C.D.; White, J.M.; Chen, Y.S.; Shea, L.K.; et al. Cancer exome analysis reveals a T-cell-dependent mechanism of cancer immunoediting. Nature 2012, 482, 400–404. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, C.-Z. Long intergenic non-coding RNA 668 regulates VEGFA signaling through inhibition of miR-297 in oral squamous cell carcinoma. Biochem. Biophys. Res. Commun. 2017, 489, 404–412. [Google Scholar] [CrossRef] [PubMed]
  95. Zhao, L.; Cao, H.; Chi, W.; Meng, W.; Cui, W.; Guo, W.; Wang, B. Expression profile analysis identifies the long non-coding RNA landscape and the potential carcinogenic functions of LINC00668 in laryngeal squamous cell carcinoma. Gene 2019, 687, 47–55. [Google Scholar] [CrossRef]
  96. Ye, P.; Lv, X.; Aizemaiti, R.; Cheng, J.; Xia, P.; Di, M. H3K27ac-activated LINC00519 promotes lung squamous cell carcinoma progression by targeting miR-450b-5p/miR-515-5p/YAP1 axis. Cell Prolif. 2020, 53, e12797. [Google Scholar] [CrossRef] [Green Version]
  97. Liu, D.; Zhao, J.; Wang, H.; Li, H.; Li, Y.; Qin, W. Long Intergenic Non-Protein Coding RNA 519 Promotes the Biological Activities of Tongue Squamous Cell Carcinoma by Sponging microRNA-876-3p and Consequently Upregulating MACC1. OncoTargetsTher. 2020, 13, 11975–11990. [Google Scholar] [CrossRef]
  98. Li, Q.; Wang, Y.; Hu, R.; Yang, G. Dysregulation of SPRR3/miR-876-3p Axis Contributes to Tumorigenesis in Non-Small-Cell Lung Cancer. OncoTargetsTher. 2020, 13, 2411–2419. [Google Scholar] [CrossRef] [Green Version]
  99. Zhang, M.; Chen, Z.; Zhang, S.; Wu, L.; Jie, Y.; Liao, Y.; Huang, Y.; Chen, J.; Shi, B. Analysis of Differentially Expressed Long Non-coding RNAs and the Associated TF-mRNA Network in Tongue Squamous Cell Carcinoma. Front. Oncol. 2020, 10, 1421. [Google Scholar] [CrossRef]
  100. Wang, P.; Li, X.; Gao, Y.; Guo, Q.; Ning, S.; Zhang, Y.; Shang, S.; Wang, J.; Wang, Y.; Zhi, H.; et al. LnCeVar: A comprehensive database of genomic variations that disturb ceRNA network regulation. Nucleic Acids Res. 2019, 48, D111–D117. [Google Scholar] [CrossRef] [PubMed]
  101. Ding, J.; Yang, C.; Yang, S. LINC00511 interacts with miR-765 and modulates tongue squamous cell carcinoma progression by targeting LAMC2. J. Oral Pathol. Med. 2018, 47, 468–476. [Google Scholar] [CrossRef] [PubMed]
  102. Wang, X.-F.; Liang, B.; Chen, C.; Zeng, D.-X.; Zhao, Y.-X.; Su, N.; Ning, W.-W.; Yang, W.; Huang, J.-A.; Gu, N.; et al. Long Intergenic Non-protein Coding RNA 511 in Cancers. Front. Genet. 2020, 11, 667. [Google Scholar] [CrossRef] [PubMed]
  103. Kong, Y.G.; Cui, M.; Chen, S.M.; Xu, Y.; Xu, Y.; Tao, Z.Z. LncRNA-LINC00460 facilitates nasopharyngeal carcinoma tumor-igenesis through sponging miR-149-5p to up-regulate IL6. Gene 2018, 639, 77–84. [Google Scholar] [CrossRef]
  104. Liang, Y.; Wu, Y.; Chen, X.; Zhang, S.; Wang, K.; Guan, X.; Yang, K.; Li, J.; Bai, Y. A novel long noncoding RNA linc00460 up-regulated by CBP/P300 promotes carcinogenesis in esophageal squamous cell carcinoma. Biosci. Rep. 2017, 37, 37. [Google Scholar] [CrossRef] [PubMed]
  105. Xie, X.; Xiong, G.; Wang, Q.; Ge, Y.; Cui, X. Long non-coding RNA LINC00460 promotes head and neck squamous cell carci-noma cell progression by sponging miR-612 to up-regulate AKT2. Am. J. Transl. Res. 2019, 11, 6326–6340. [Google Scholar] [PubMed]
  106. Li, K.; Sun, D.; Gou, Q.; Ke, X.; Gong, Y.; Zuo, Y.; Zhou, J.-K.; Guo, C.; Xia, Z.; Liu, L.; et al. Long non-coding RNA linc00460 promotes epithelial-mesenchymal transition and cell migration in lung cancer cells. Cancer Lett. 2018, 420, 80–90. [Google Scholar] [CrossRef]
  107. Chaudhary, R.; Wang, X.; Cao, B.; De La Iglesia, J.; Masannat, J.; Song, F.; Hernandez-Prera, J.C.; Gimbrone, N.T.; Slebos, R.J.; Chung, C.H. Long noncoding RNA, LINC00460, as a prognostic biomarker in head and neck squamous cell carcinoma (HNSCC). Am. J. Trans. Res. 2020, 12, 684–696. [Google Scholar]
  108. Nicholson, K.M.; Anderson, N.G. The protein kinase B/Akt signalling pathway in human malignancy. Cell. Signal. 2002, 14, 381–395. [Google Scholar] [CrossRef]
  109. Dillon, R.L.; Marcotte, R.; Hennessy, B.T.; Woodgett, J.R.; Mills, G.B.; Muller, W.J. Akt1 and akt2 play distinct roles in the initi-ation and metastatic phases of mammary tumor progression. Cancer Res. 2009, 69, 5057–5064. [Google Scholar] [CrossRef] [Green Version]
  110. Zhang, W.; Huang, C.; Gong, Z.; Zhao, Y.; Tang, K.; Li, X.; Fan, S.; Shi, L.; Li, X.; Zhang, P.; et al. Expression of LINC00312, a long intergenic non-coding RNA, is negatively correlated with tumor size but positively correlated with lymph node metastasis in nasopharyngeal carcinoma. J. Mol. Histol. 2013, 44, 545–554. [Google Scholar] [CrossRef]
  111. Huang, C.; Wu, M.; Tang, Y.; Li, X.; Ouyang, J.; Xiao, L.; Li, D.; Li, G. NAG7 promotes human nasopharyngeal carcinoma in-vasion through inhibition of estrogen receptor alpha and up-regulation of JNK2/AP-1/MMP1 pathways. J. Cell Physiol. 2009, 221, 394–401. [Google Scholar] [CrossRef]
  112. Shan, Y.; Ying, R.; Jia, Z.; Kong, W.; Wu, Y.; Zheng, S.; Jin, H. LINC00052 Promotes Gastric Cancer Cell Proliferation and Metastasis via Activating the Wnt/β-Catenin Signaling Pathway. Oncol. Res. Featur. Preclin. Clin. Cancer Ther. 2017, 25, 1589–1599. [Google Scholar] [CrossRef]
  113. Salameh, A.; Fan, X.; Choi, B.-K.; Zhang, S.; Zhang, N.; An, Z. HER3 and LINC00052 interplay promotes tumor growth in breast cancer. Oncotarget 2016, 8, 6526–6539. [Google Scholar] [CrossRef] [Green Version]
  114. Ouyang, T.; Zhang, Y.; Tang, S.; Wang, Y. Long non-coding RNA LINC00052 regulates miR-608/EGFR axis to promote pro-gression of head and neck squamous cell carcinoma. Exp. Mol.Pathol. 2019, 111, 104321. [Google Scholar] [CrossRef] [PubMed]
  115. Zhu, L.; Yang, N.; Chen, J.; Zeng, T.; Yan, S.; Liu, Y.; Yu, G.; Chen, Q.; Du, G.; Pan, W.; et al. LINC00052 upregulates EPB41L3 to inhibit migration and invasion of hepatocellular carcinoma by binding miR-452-5p. Oncotarget 2017, 8, 63724–63737. [Google Scholar] [CrossRef] [Green Version]
  116. Xiong, D.; Sheng, Y.; Ding, S.; Chen, J.; Tan, X.; Zeng, T.; Qin, D.; Zhu, L.; Huang, A.; Tang, H. LINC00052 regulates the expression of NTRK3 by miR-128 and miR-485-3p to strengthen HCC cells invasion and migration. Oncotarget 2016, 7, 47593–47608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Yan, S.; Shan, X.; Chen, K.; Liu, Y.; Yu, G.; Chen, Q.; Zeng, T.; Zhu, L.; Dang, H.; Chen, F.; et al. LINC00052/miR-101-3p axis inhibits cell proliferation and metastasis by targeting SOX9 in hepatocellular carcinoma. Gene 2018, 679, 138–149. [Google Scholar] [CrossRef] [PubMed]
  118. Ding, L.; Ni, J.; Yang, F.; Huang, L.; Deng, H.; Wu, Y.; Ding, X.; Tang, J. Promising therapeutic role of miR-27b in tumor. Tumor Biol. 2017, 39, 1010428317691657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Volders, P.-J.; Anckaert, J.; Verheggen, K.; Nuytens, J.; Martens, L.; Mestdagh, P.; Vandesompele, J. LNCipedia 5: Towards a reference set of human long non-coding RNAs. Nucleic Acids Res. 2019, 47, D135–D139. [Google Scholar] [CrossRef] [Green Version]
  120. Zhang, L.-M.; Ju, H.-Y.; Wu, Y.-T.; Guo, W.; Mao, L.; Ma, H.-L.; Xia, W.-Y.; Hu, J.-Z.; Ren, G.-X. Long non-coding RNA ANRIL promotes tumorgenesis through regulation of FGFR1 expression by sponging miR-125a-3p in head and neck squamous cell carcinoma. Am. J. Cancer Res. 2018, 8, 2296–2310. [Google Scholar]
  121. Zhao, J.-J.; Hao, S.; Wang, L.-L.; Hu, C.-Y.; Zhang, S.; Guo, L.-J.; Zhang, G.; Gao, B.; Jiang, Y.; Tian, W.-G.; et al. Long non-coding RNA ANRIL promotes the invasion and metastasis of thyroid cancer cells through TGF-β/Smad signaling pathway. Oncotarget 2016, 7, 57903–57918. [Google Scholar] [CrossRef] [Green Version]
  122. Chai, L.; Yuan, Y.; Chen, C.; Zhou, J.; Wu, Y. The role of long non-coding RNA ANRIL in the carcinogenesis of oral cancer by targeting miR-125a. Biomed. Pharmacother. 2018, 103, 38–45. [Google Scholar] [CrossRef]
  123. Cui, K.; Jin, S.; Du, Y.; Yu, J.; Feng, H.; Fan, Q.; Ma, W. Long noncoding RNA DIO3OS interacts with miR-122 to promote proliferation and invasion of pancreatic cancer cells through upregulating ALDOA. Cancer Cell Int. 2019, 19, 1–10. [Google Scholar] [CrossRef] [Green Version]
  124. Wang, M.; Li, J.; Zuo, Z.; Ren, C.; Tang, T.; Long, C.; Gong, Y.; Ye, F.; Wang, Z.; Huang, J. Long non-coding RNA DIO3OS/let-7d/NF-κB2 axis regulates cells proliferation and metastasis of thyroid cancer cells. J. Cell Commun. Signal. 2020, 15, 1–14. [Google Scholar] [CrossRef]
  125. Guglas, K.; Kolenda, T.; Teresiak, A.; Kopczyńska, M.; Łasińska, I.; Mackiewicz, J.; Mackiewicz, A.; Lamperska, K. lncRNA Expression after Irradiation and Chemoexposure of HNSCC Cell Lines. Non-Coding RNA 2018, 4, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Yue, D.; Li, H.; Che, J.; Zhang, Y.; Tolani, B.; Mo, M.; Zhang, H.; Zheng, Q.; Yang, Y.; Cheng, R.; et al. EMX2 Is a Predictive Marker for Adjuvant Chemotherapy in Lung Squamous Cell Carcinomas. PLoS ONE 2015, 10, e0132134. [Google Scholar] [CrossRef] [PubMed]
  127. Duan, M.; Fang, M.; Wang, C.; Wang, H.; Li, M. LncRNA EMX2OS Induces Proliferation, Invasion and Sphere Formation of Ovarian Cancer Cells via Regulating the miR-654-3p/AKT3/PD-L1 Axis. Cancer Manag. Res. 2020, 12, 2141–2154. [Google Scholar] [CrossRef] [Green Version]
  128. Schneider, C.; King, R.M.; Philipson, L. Genes specifically expressed at growth arrest of mammalian cells. Cell 1988, 54, 787–793. [Google Scholar] [CrossRef]
  129. Kino, T.; Hurt, D.E.; Ichijo, T.; Nader, N.; Chrousos, G.P. Noncoding RNA Gas5 Is a Growth Arrest- and Starvation-Associated Repressor of the Glucocorticoid Receptor. Sci. Signal. 2010, 3, ra8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Fayda, M.; Isin, M.; Tambas, M.; Guveli, M.; Meral, R.; Altun, M.; Sahin, D.; Ozkan, G.; Sanli, Y.; Isin, H.; et al. Do circulating long non-coding RNAs (lncRNAs) (LincRNA-p21, GAS 5, HOTAIR) predict the treatment response in patients with head and neck cancer treated with chemoradiotherapy? Tumor Biol. 2015, 37, 3969–3978. [Google Scholar] [CrossRef]
  131. Zhou, J.; Yang, L.; Zhong, T.; Mueller, M.; Men, Y.; Zhang, N.; Xie, J.; Giang, K.; Chung, H.; Sun, X.; et al. H19 lncRNA alters DNA methylation genome wide by regulating S-adenosylhomocysteine hydrolase. Nat. Commun. 2015, 6, 10221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Guan, G.-F.; Zhang, D.-J.; Wen, L.-J.; Xin, D.; Liu, Y.; Yu, D.-J.; Su, K.; Zhu, L.; Guo, Y.-Y.; Wang, K. Overexpression of lncRNA H19/miR-675 promotes tumorigenesis in head and neck squamous cell carcinoma. Int. J. Med Sci. 2016, 13, 914–922. [Google Scholar] [CrossRef] [Green Version]
  133. Li, X.; Lin, Y.; Yang, X.; Wu, X.; He, X. Long noncoding RNA H19 regulates EZH2 expression by interacting with miR-630 and promotes cell invasion in nasopharyngeal carcinoma. Biochem. Biophys. Res. Commun. 2016, 473, 913–919. [Google Scholar] [CrossRef]
  134. Pollard, K.S.; Salama, S.; Lambert, N.; Lambot, M.-A.; Coppens, S.; Pedersen, J.S.; Katzman, S.; King, B.; Onodera, C.; Siepel, A.; et al. An RNA gene expressed during cortical development evolved rapidly in humans. Nat. Cell Biol. 2006, 443, 167–172. [Google Scholar] [CrossRef]
  135. Yamada, H.; Takahashi, M.; Watanuki, M.; Watanabe, M.; Hiraide, S.; Saijo, K.; Komine, K.; Ishioka, C. lncRNA HAR1B has potential to be a predictive marker for pazopanib therapy in patients with sarcoma. Oncol. Lett. 2021, 21, 1–14. [Google Scholar] [CrossRef]
  136. Panzitt, K.; Tschernatsch, M.M.; Guelly, C.; Moustafa, T.; Stradner, M.; Strohmaier, H.M.; Buck, C.R.; Denk, H.; Schroeder, R.; Trauner, M.; et al. Characterization of HULC, a Novel Gene with Striking Up-Regulation in Hepatocellular Carcinoma, as Noncoding RNA. Gastroenterology 2007, 132, 330–342. [Google Scholar] [CrossRef] [PubMed]
  137. Zhao, Y.; Guo, Q.; Chen, J.; Hu, J.; Wang, S.; Sun, Y. Role of long non-coding RNA HULC in cell proliferation, apoptosis and tumor metastasis of gastric cancer: A clinical and in vitro investigation. Oncol. Rep. 2013, 31, 358–364. [Google Scholar] [CrossRef] [Green Version]
  138. Peng, W.; Gao, W.; Feng, J. Long noncoding RNA HULC is a novel biomarker of poor prognosis in patients with pancreatic cancer. Med. Oncol. 2014, 31, 1–7. [Google Scholar] [CrossRef] [PubMed]
  139. Matouk, I.J.; Abbasi, I.; Hochberg, A.; Galun, E.; Dweik, H.; Akkawi, M. Highly upregulated in liver cancer noncoding RNA is overexpressed in hepatic colorectal metastasis. Eur. J. Gastroenterol. Hepatol. 2009, 21, 688–692. [Google Scholar] [CrossRef] [PubMed]
  140. Su, W.; Tang, J.; Wang, Y.; Sun, S.; Shen, Y.; Yang, H. Long non-coding RNA highly up-regulated in liver cancer promotes epithelial-to-mesenchymal transition process in oral squamous cell carcinoma. J. Cell. Mol. Med. 2019, 23, 2645–2655. [Google Scholar] [CrossRef]
  141. Pan, J.; Fang, S.; Tian, H.; Zhou, C.; Zhao, X.; Tian, H.; He, J.; Shen, W.; Meng, X.; Jin, X.; et al. lncRNA JPX/miR-33a-5p/Twist1 axis regulates tumorigenesis and metastasis of lung cancer by activating Wnt/β-catenin signaling. Mol. Cancer 2020, 19, 9–17. [Google Scholar] [CrossRef]
  142. Li, L.; Gu, M.; You, B.; Shi, S.; Shan, Y.; Bao, L.; You, Y. Long non-coding RNA ROR promotes proliferation, migration and chemoresistance of nasopharyngeal carcinoma. Cancer Sci. 2016, 107, 1215–1222. [Google Scholar] [CrossRef]
  143. Zhang, H.-Y.; Liang, F.; Zhang, J.-W.; Wang, F.; Wang, L.; Kang, X.-G. Effects of long noncoding RNA-ROR on tamoxifen resistance of breast cancer cells by regulating microRNA-205. Cancer Chemother. Pharmacol. 2017, 79, 327–337. [Google Scholar] [CrossRef]
  144. Kolenda, T.; Guglas, K.; Ryś, M.; Bogaczyńska, M.; Teresiak, A.; Bliźniak, R.; Łasińska, I.; Mackiewicz, J.; Lamperska, K.M. Biological role of long non-coding RNA in head and neck cancers. Rep. Pract. Oncol. Radiother. 2017, 22, 378–388. [Google Scholar] [CrossRef]
  145. Voellenkle, C.; Garcia-Manteiga, J.M.; Pedrotti, S.; Perfetti, A.; De Toma, I.; Da Silva, D.; Maimone, B.; Greco, S.; Fasanaro, P.; Creo, P.; et al. Implication of Long noncoding RNAs in the endothelial cell response to hypoxia revealed by RNA-sequencing. Sci. Rep. 2016, 6, 24141. [Google Scholar] [CrossRef] [Green Version]
  146. Kozłowska, J.; Kozioł, K.; Stasiak, M.; Obacz, J.; Guglas, K.; Poter, P.; Mackiewicz, A.; Kolenda, T. The role of NEAT1 lncRNA in squamous cell carcinoma of the head and neck is still difficult to define. WspółczesnaOnkologia 2020, 24, 96–105. [Google Scholar] [CrossRef]
  147. Willingham, A.T.; Orth, A.P.; Batalov, S.; Peters, E.C.; Wen, B.G.; Aza-Blanc, P.; Hogenesch, J.B.; Schultz, P.G. A Strategy for Probing the Function of Noncoding RNAs Finds a Repressor of NFAT. Science 2005, 309, 1570–1573. [Google Scholar] [CrossRef]
  148. Shang, W.; Tang, Z.; Gao, Y.; Qi, H.; Su, X.; Zhang, Y.; Yang, R. LncRNA RNCR3 promotes Chop expression by sponging miR-185-5p during MDSC differentiation. Oncotarget 2017, 8, 111754–111769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Lin, S.-X.; Jiang, H.; Xiang, G.-Z.; Zhang, W.-R.; Weng, Y.-H.; Qiu, F.-D.; Wu, J.; Wang, H.-G. Up-regulation of long non-coding RNA SNHG1 contributes to proliferation and metastasis in laryngeal squamous cell carcinoma. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 1333–1341. [Google Scholar] [PubMed]
  150. Zhao, Y.; Wang, J.; Ma, K. Knockdown of lncRNA SNHG6 inhibites the proliferation and epithelial mesenchymal transition in tongue cancer cells. Chin. J. Cell. Mol. Immunol. 2018, 34, 806–811. [Google Scholar]
  151. Guo, T.; Zambo, K.D.A.; Zamuner, F.T.; Ou, T.; Hopkins, C.; Kelley, D.Z.; Wulf, H.A.; Winkler, E.; Erbe, R.; Danilova, L.; et al. Chromatin structure regulates cancer-specific alternative splicing events in primary HPV-related oropharyngeal squamous cell carcinoma. Epigenetics 2020, 15, 959–971. [Google Scholar] [CrossRef]
  152. Gayen, S.; Maclary, E.; Buttigieg, E.; Hinten, M.; Kalantry, S. A Primary Role for the Tsix lncRNA in Maintaining Random X-Chromosome Inactivation. Cell Rep. 2015, 11, 1251–1265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Zheng, R.; Lin, S.; Guan, L.; Yuan, H.; Liu, K.; Liu, C.; Ye, W.; Liao, Y.; Jia, J.; Zhang, R. Long non-coding RNA XIST inhibited breast cancer cell growth, migration, and invasion via miR-155/CDX1 axis. Biochem. Biophys. Res. Commun. 2018, 498, 1002–1008. [Google Scholar] [CrossRef] [PubMed]
  154. Salama, E.A.; Adbeltawab, R.E.; El Tayebi, H.M. XIST and TSIX: Novel Cancer Immune Biomarkers in PD-L1-Overexpressing Breast Cancer Patients. Front. Oncol. 2020, 9, 1459. [Google Scholar] [CrossRef]
  155. Fang, Z.; Wu, L.; Wang, L.; Yang, Y.; Meng, Y.; Yang, H. Increased expression of the long non-coding RNA UCA1 in tongue squamous cell carcinomas: A possible correlation with cancer metastasis. Oral Surg. Oral Med. Oral Pathol. Oral Radiol. 2014, 117, 89–95. [Google Scholar] [CrossRef] [PubMed]
  156. Tian, F.M.; Meng, F.Q.; Wang, X.B. Overexpression of long-noncoding RNA ZFAS1 decreases survival in human NSCLC pa-tients. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 5126–5131. [Google Scholar]
  157. Li, T.; Xie, J.; Shen, C.; Cheng, D.; Shi, Y.; Wu, Z.; Deng, X.; Chen, H.; Shen, B.; Peng, C.; et al. Amplification of Long Noncoding RNA ZFAS1 Promotes Metastasis in Hepatocellular Carcinoma. Cancer Res. 2015, 75, 3181–3191. [Google Scholar] [CrossRef] [Green Version]
  158. Shi, H.; Liu, Z.; Pei, D.; Jiang, Y.; Zhu, H.; Chen, B. Development and validation of nomogram based on lncRNA ZFAS1 for predicting survival in lymph node-negative esophageal squamous cell carcinoma patients. Oncotarget 2017, 8, 59048–59057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Kolenda, T.; Guglas, K.; Kopczyńska, M.; Teresiak, A.; Bliźniak, R.; Mackiewicz, A.; Lamperska, K.; Mackiewicz, J. Oncogenic Role of ZFAS1 lncRNA in Head and Neck Squamous Cell Carcinomas. Cells 2019, 8, 366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Chandrashekar, D.S.; Bashel, B.; Balasubramanya, S.A.H.; Creighton, C.J.; Ponce-Rodriguez, I.; Chakravarthi, B.V.; Varambally, S. UALCAN: A Portal for Facilitating Tumor Subgroup Gene Expression and Survival Analyses. Neoplasia 2017, 19, 649–658. [Google Scholar] [CrossRef]
  161. Aggarwal, N.; Yadav, J.; Thakur, K.; Bibban, R.; Chhokar, A.; Tripathi, T.; Bhat, A.; Singh, T.; Jadli, M.; Singh, U.; et al. Human Papillomavirus Infection in Head and Neck Squamous Cell Carcinomas: Transcriptional Triggers and Changed Disease Patterns. Front. Cell. Infect. Microbiol. 2020, 10, 746. [Google Scholar] [CrossRef] [PubMed]
  162. Sannigrahi, M.K.; Sharma, R.; Panda, N.K.; Khullar, M. Role of non-coding RNAs in head and neck squamous cell carcinoma: A narrative review. Oral Dis. 2017, 24, 1417–1427. [Google Scholar] [CrossRef] [PubMed]
  163. Song, L.; Xie, H.; Tong, F.; Yan, B.; Zhang, S.; Fu, E.; Jing, Q.; Wei, L. Association of lnc-IL17RA-11 with increased radiation sensitivity and improved prognosis of HPV-positive HNSCC. J. Cell. Biochem. 2019, 120, 17438–17448. [Google Scholar] [CrossRef]
  164. Kolenda, T.; Kopczyńska, M.; Guglas, K.; Teresiak, A.; Bliźniak, R.; Łasińska, I.; Mackiewicz, J.; Lamperska, K. EGOT lncRNA in head and neck squamous cell carcinomas. Pol. J. Pathol. 2018, 69, 356–365. [Google Scholar] [CrossRef] [PubMed]
  165. Kopczyńska, M.; Kolenda, T.; Guglas, K.; Sobocińska, J.; Teresiak, A.; Bliźniak, R.; Mackiewicz, A.; Mackiewicz, J.; Lamperska, K. PRINS lncRNA Is a New Biomarker Candidate for HPV Infection and Prognosis of Head and Neck Squamous Cell Carcinomas. Diagnostics 2020, 10, 762. [Google Scholar] [CrossRef] [PubMed]
  166. Tomar, S.; Bs, C.A.G.; Altomare, D.; Kowli, S.; Ms, S.K.; Sutkowski, N.; Gillespie, M.B.; Creek, K.E.; Pirisi, L. Human papillomavirus status and gene expression profiles of oropharyngeal and oral cancers from European American and African American patients. Head Neck 2016, 38, E694–E704. [Google Scholar] [CrossRef] [Green Version]
  167. Casarotto, M.; Fanetti, G.; Guerrieri, R.; Palazzari, E.; Lupato, V.; Steffan, A.; Polesel, J.; Boscolo-Rizzo, P.; Fratta, E. Beyond MicroRNAs: Emerging Role of Other Non-Coding RNAs in HPV-Driven Cancers. Cancers 2020, 12, 1246. [Google Scholar] [CrossRef] [PubMed]
  168. Tomar, S. Differential Gene Expression Patterns in HPV-Positive and HPV-Negative Oropharyngeal Carcinomas. Doctoral Dissertation, University of South Carolina, Columbia, SC, USA, 2013. Available online: https://scholarcommons.sc.edu/cgi/viewcontent.cgi?article=4577&context=etd (accessed on 18 April 2021).
Figure 1. Characteristics of lncRNAs: (A) the classification of transcripts depending on the localization in the genome; (B) the main biological functions of lncRNAs, including modification of RNA as well as chromatin, changes in nucleus architecture, molecular sponging, production of short peptides, and regulation of transcription.
Figure 1. Characteristics of lncRNAs: (A) the classification of transcripts depending on the localization in the genome; (B) the main biological functions of lncRNAs, including modification of RNA as well as chromatin, changes in nucleus architecture, molecular sponging, production of short peptides, and regulation of transcription.
Cancers 13 02949 g001
Figure 2. Changes in lincRNA expression levels in HNSCC patients based on (A) UALCAN database with differences in expression level between HNSCC samples and normal samples, and patients’ survival depending on the expression level of selected lincRNAs (p<0.05 were considered statistically significant in the UALCAN results); (B)HPV status based on Tomar’s results where 65 oropharyngeal samples including HPV+ and HPV− were analyzed using microarray analysis. The differences in the expression of lincRNA depending on HPV status as well as viral infection activities are listed [168].
Figure 2. Changes in lincRNA expression levels in HNSCC patients based on (A) UALCAN database with differences in expression level between HNSCC samples and normal samples, and patients’ survival depending on the expression level of selected lincRNAs (p<0.05 were considered statistically significant in the UALCAN results); (B)HPV status based on Tomar’s results where 65 oropharyngeal samples including HPV+ and HPV− were analyzed using microarray analysis. The differences in the expression of lincRNA depending on HPV status as well as viral infection activities are listed [168].
Cancers 13 02949 g002
Table 1. The summary of the well-described lincRNAs in HNSCC and other cancers.
Table 1. The summary of the well-described lincRNAs in HNSCC and other cancers.
lincRNALocationPossible TargetsRoleRef.
XISTXq13.2miR-137, miR-92b, mir-34a-5P
-
essential role in the development and progression of may cancers, e.g., glioblastoma, HCC, BC, NSCLC, NPC
-
high expression level increased with tumor size and stage; leads to poor survival of patients with NPC
[68,69,70,71,72]
MALAT111q13.1AIM1, LAYN, HMMR, SLC26A2, CCT4, ROD1, CTHRC1, FHL1
-
modulator of pre-mRNA processing, regulating splicing efficiency
-
oncogene in ESCC, promoting its growth by regulating the ATM-CHK2 pathway, which is associated with G2/M transition and processes of DNA damage response
-
correlated with the clinical stage in, e.g., glioma, pancreas, prostate, and lung cancer
-
overexpression is associated with unfavorable prognosis and significantly shorter OS
[73,74,75,76]
HOTAIR12q13.13PRC2, ZEB1, SNAIL, MMP13, MMP9
-
interacts with PRC2 and plays a pivotal role in the H3K27 methylation, causes epigenetic silencing during many cellular processes, e.g., cell differentiation, adhesion, and cell cycle
-
high level is associated with poor prognosis and patients’ OS
-
up-regulation in cell cultures leads to abnormally activated Wnt signaling pathways, which results in progression, increased migration, and the ability to create metastasis
[77,78,79,80]
HOTTIP7p15.2WDR5, UPF1, PTB, FUS, IF4AIII, DGCR8, HOXA10, HOXA11, HOXA13
-
regulates the activation of multiple HOXA genes by controlling H3K4 methylation
-
overexpression is crucial for tumor development growth, and metastasis in many cancers, e.g., HCC, TSCC
-
expression associated with the grade, stage, and overall survival of patients; independent prognostic factor
[41,81,82,83]
lincRNA-p216p21.2STAT3, CTNNB1, JUNB
-
pivotal role in the development and progression of multiple cancers
-
displays their suppressor function by decreasing the activity of the JAK2/STAT3 pathway
-
low level causes drastic progression of HNSCC due to lack of induced G1 phase arrest and inhibition of apoptosis
[86]
LINC024876q27USP17
-
displays regulatory function at the post-transcriptional or post-translational level
-
dysregulated in OSCC and correlated with development stage, unfavorable clinical outcome, poor survival
-
overexpressed in cell cultures, inhibits OSCC proliferation, migration, and invasiveness, and regulates levels of EMT markers
[87,88]
LINC0219516p12.1HLA-A, HLA-B, HLA-C
-
closely associated with MHC I molecules, whose lack of function leads to the mechanism of escaping immunosurveillance
-
silencing causes a decrease in the level of MHC I
-
correlation with the T cell receptor pathway, chemokines (class I and II), and cytokines
-
high expression correlated with better prognosis and positive response to immunotherapy in different cancers
[91,92,93]
LINC0066818p11.31ABL2, RAB3B, ENAH, HMGA2
-
oncogene; promotes tumor growth in OSCC cells by interaction with miR-297 and VEGFA signaling pathways
-
expression level in LSCC is associated with age, stage, and cervical lymph node metastasis; enhances the proliferation, migration, and invasion ability of LSCC cell lines
[94,95]
LINC0051914q22.1miR-450b-5p, miR-515-5p, YAP1, miR-876-3p
-
a known oncogene in LSCC
-
acts as an miRNA-sequestering sponge
-
binds miR-876-3p, which is aberrantly expressed and tumorigenic in many cancers
-
high expression associated with shorter OS and unsatisfactory prognosis in TSCC
[96,97,98]
LINC0051117q24.3miR-765, LAMC2
-
regulates the developmental process, apoptosis, programmed cell death, focal adhesion through hemostasis, and different carcinogenic pathways
-
modulates TSCC progression by promoting cell proliferation and migration
-
associated with age, tumor size, clinical stage, lymph node metastasis along with unsatisfactory prognosis
[100,101,102]
LINC0046013q33.2p21, E-cadherin, N-cadherin, cyclin D1, miR-162, miR-149-5p, miR-612
-
oncogene in different cancers, e.g., NPC, ESCC, and lung cancer
-
high level associated with several carcinogenic pathways; involvement in cell development, proliferation, the EMT, and adhesion
-
promotes tumor growth, affects cell cycle, migration, the EMT along with invasion
-
regulates the activity of miR-162 by acting as a molecular sponge, which leads to progression, metastasis, and unfavorable prognosis
[103,104,105,108,109]
LINC003123p25.3JNK2, c-Jun, c-Fos, H-Ras, ER-alpha
-
overexpression in NPC cell lines leads to an increase of adhesion, motility, invasiveness and inhibits proliferation, by arresting progression from G1 to S phase of the cell cycle
-
expression positively correlated with lymph node metastasis and negatively associated with stages and size of the tumor
-
allows distinguishing healthy individuals from NPC patients and, within the latter group, determine who developed distant metastasis
[110,111]
LINC0005215q25.3SMYD2, NTRK3, HER3, miR-608
-
displays regulatory function through acting as a molecular sponge and sequestering different miRNAs
[112,113,114]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kozłowska, J.; Kolenda, T.; Poter, P.; Sobocińska, J.; Guglas, K.; Stasiak, M.; Bliźniak, R.; Teresiak, A.; Lamperska, K. Long Intergenic Non-Coding RNAs in HNSCC: From “Junk DNA” to Important Prognostic Factor. Cancers 2021, 13, 2949. https://doi.org/10.3390/cancers13122949

AMA Style

Kozłowska J, Kolenda T, Poter P, Sobocińska J, Guglas K, Stasiak M, Bliźniak R, Teresiak A, Lamperska K. Long Intergenic Non-Coding RNAs in HNSCC: From “Junk DNA” to Important Prognostic Factor. Cancers. 2021; 13(12):2949. https://doi.org/10.3390/cancers13122949

Chicago/Turabian Style

Kozłowska, Joanna, Tomasz Kolenda, Paulina Poter, Joanna Sobocińska, Kacper Guglas, Maciej Stasiak, Renata Bliźniak, Anna Teresiak, and Katarzyna Lamperska. 2021. "Long Intergenic Non-Coding RNAs in HNSCC: From “Junk DNA” to Important Prognostic Factor" Cancers 13, no. 12: 2949. https://doi.org/10.3390/cancers13122949

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