Next Article in Journal
Predictive Markers for Malignant Urothelial Transformation in Balkan Endemic Nephropathy: A Case–Control Study
Next Article in Special Issue
LncRNA LINC00518 Acts as an Oncogene in Uveal Melanoma by Regulating an RNA-Based Network
Previous Article in Journal
No Impact of Smoking Status on Breast Cancer Tumor Infiltrating Lymphocytes, Response to Neoadjuvant Chemotherapy and Prognosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Non-Coding RNAs in Uveal Melanoma

by
Manuel Bande
1,2,*,
Daniel Fernandez-Diaz
1,2,
Beatriz Fernandez-Marta
1,
Cristina Rodriguez-Vidal
3,
Nerea Lago-Baameiro
4,
Paula Silva-Rodríguez
2,5,
Laura Paniagua
6,
María José Blanco-Teijeiro
1,2,
María Pardo
2,4 and
Antonio Piñeiro
1,2
1
Department of Ophthalmology, University Hospital of Santiago de Compostela, Ramon Baltar S/N, 15706 Santiago de Compostela, Spain
2
Tumores Intraoculares en el Adulto, Instituto de Investigación Sanitaria de Santiago (IDIS), 15706 Santiago de Compostela, Spain
3
Department of Ophthalmology, University Hospital of Cruces, Cruces Plaza, S/N, 48903 Barakaldo, Vizcaya, Spain
4
Grupo Obesidómica, Instituto de Investigación Sanitaria de Santiago (IDIS), 15706 Santiago de Compostela, Spain
5
Fundación Pública Galega de Medicina Xenómica, Clinical University Hospital, SERGAS, 15706 Santiago de Compostela, Spain
6
Department of Ophthalmology, University Hospital of Coruña, Praza Parrote, S/N, 15006 La Coruña, Spain
*
Author to whom correspondence should be addressed.
Submission received: 13 September 2020 / Revised: 1 October 2020 / Accepted: 9 October 2020 / Published: 12 October 2020
(This article belongs to the Special Issue Noncoding Landscapes of Uveal Melanoma)

Abstract

:

Simple Summary

The development of uveal melanoma is a multifactorial and multi-step process, in which abnormal gene expression plays a key role. Recently, several studies have highlighted the role of non-coding RNAs in the progression of uveal melanoma by affecting different signaling pathways. As important agents in the regulation of genes, non-coding RNAs have enormous potential to open up therapeutic pathways, predict response to treatment, and anticipate patient outcome for uveal melanoma. This review aims to provide a comprehensive view of what we know about ncRNAs in uveal melanoma currently.

Abstract

Uveal melanoma (UM) is the most common primary intraocular tumor in adulthood. Approximately 50% of patients develop metastatic disease, which typically affects the liver and is usually fatal within one year. This type of cancer is heterogeneous in nature and is divided into two broad groups of tumors according to their susceptibility to develop metastasis. In the last decade, chromosomal abnormalities and the aberrant expression of several signaling pathways and oncogenes in uveal melanomas have been described. Recently, importance has been given to the association of the mentioned deregulation with the expression of non-coding RNAs (ncRNAs). Here, we review the different classes of ncRNAs—such as long non-coding RNAs (lncRNAs) and microRNAs (miRNAs)—and their contribution to the development of UM. Special attention is given to miRNAs and their regulatory role in physiopathology and their potential as biomarkers. As important agents in gene regulation, ncRNAs have a huge potential for opening up therapeutic pathways, predicting response to treatment, and anticipating patient outcome for UM.

1. Introduction

The latest research on the human genome indicates that only 2% of human DNA is encodes proteins [1]. However, it is known that over 80% of the human genome contains elements linked to biochemical functions [2]. What was formerly considered junk DNA has more recently been demonstrated to be ncRNAs that perform multiple biologic processes and whose deregulation may be associated with multiple diseases, including cancer [3,4]. In addition to the well-known types of ncRNA, such as tRNA and rRNA that are involved in protein synthesis, ncRNA also comprises many new transcripts that have been identified in the last decade. They are divided into two groups according to their size: small ncRNA and long ncRNA. The size of small ncRNAs is usually less than 200 nucleotides and includes miRNA, PIWI-interacting RNA (piRNA), small nucleolar RNA (snoRNA), and promoter-associated small RNA (PASR) [5], as summarized in Figure 1.
Uveal melanoma (UM) is the most common primary intraocular malignancy in adults. At present, using any of the conservative treatment options or combinations results in a degree of local control of greater than 90% after 5 years. Despite this success, metastatic disease appears in more than 50% of patients after 15 years of local treatment. This poor prognosis is associated with clinical and molecular factors of primary UM, such as tumor height, presence of monosomy 3, and chromosome 8 gain, while 6p gain provides a protective effect [6]. Genetic prognostic markers have been found that can be used to identify patients at risk of developing metastatic disease. Gene expression profiles (GEPs) are used to classify UMs for disease-specific mortality risk, with class 1A being very low risk (2% risk at 5 years), class 1B being low risk (21% risk at 5 years), and class 2 being high risk (72% at 5 years) [7]. Likewise, tumor-specific mutations have been found in the genes GNAQ, GNA11, EIF1AX, SF3B1, and BAP1 [8,9].
In recent years, many studies have shown that ncRNAs play a key role in multiple biological processes in UM, such as tumorigenesis, proliferation, and metastasis. Some ncRNAs have also been identified as biomarkers for the clinical diagnosis and prognosis of UM. In this review, we summarize the role and mechanisms of different ncRNAs in the physiopathology of UM, including those that may have a clinical application in the diagnosis and prognosis of this type of cancer.

2. MicroRNAs in UM

Of all the types of ncRNAs, miRNAs are the best-studied and understood within the context of UM initiation and progression. MicroRNAs—RNA strands of about 17–22 nucleotides—are estimated able to regulate 60% of all protein-encoding genes, and a single miRNA can regulate up to 400 different mRNAs [10]. They regulate the expression of genes both at the transcriptional and post-transcriptional level and are implicated in several physiological and pathological processes in which they exert their regulatory effects through binding of the 3′-UTRs of target mRNAs [11,12].

2.1. Dysregulated Expression of MicroRNAs

Depending on their role in the development of UM, miRNAs can be classified as tumor suppressor miRNAs or oncogenic miRNAs. Oncogenic miRNAs downregulate the transcription of tumor suppressor genes; conversely, downregulation of tumor suppressor miRNAs results in UM progression by allowing their target oncogene to be subsequently upregulated. The MAPK/ERK pathway and PI3K/AKT pathway with its natural inhibitor PTEN have been described as direct targets of many miRNAs in the development of UM [13]. The relationships established between these types are shown schematically in Figure 2. The complete list of miRNAs that have suppressive and oncogenic actions in UM is shown in Table 1; Table 2, respectively, along with information regarding their expression profiles, function, and direct genetic targets. A summary indicating the main characteristics of the UM cell lines mentioned in these tables are shown in Table S1 [14,15,16,17,18,19].

2.1.1. miRNA Tumor Suppressors

In UM, most identified miRNAs play a role in inhibiting tumors and are defined by their characteristic of downregulating oncogenes. Yan et al. identified miR-34a as a transcriptional target of p53, thus implying an antitumor effect. UM cell cultures transfected with miR-34a had decreased cell growth and migration. Moreover, miR-34a suppressed the expression of c-Met in addition to decreasing the levels of phosphorylated Akt and other cell cycle-related proteins (Rb, cdc2, and E2F3) [23]. The c-Met tyrosine kinase receptor is implicated in the upstream pathways of PI3K/AKT and Ras/MAPK; it is overexpressed in 60–86% of solid tumors and associated with tumor aggressiveness in UM [30,45]. In addition, the suppressive role of miR-34a was reinforced when LGR4 (leucine-replication-rich G-protein coupled receptor 4) was identified as another of its targets. This receptor has been implicated in processes of embryological development, cell mobility, and metastasis. LGR4 is overexpressed in UM cells, and transfection of miR-34a has been reported to decrease its expression [24]. LGR4 can inhibit the expression of MMP2, which has been shown to be responsible for the migration and invasion of many kinds of tumors, including UM [24]. In parallel, two other miRNAs from the same miR-34 family (miR-34b and miR-34c) have also been described as tumor suppressors in UM. Their levels were found to be decreased in UM cell lines and primary samples [25]. Therefore, expression of miR-34b/c results in reduced cell growth and migration by stopping the cell cycle at G1 [25]. Similar to miR-34a, C-Met has been identified as a target of miR-34b/c; other agents involved in the c-Met signaling pathway, such as p-Akt, CDK4, and CDK6, were also detected as underexpressed upon introduction of miR-34b/c in UM [25]. In a similar way, overexpression of miR-122 and miR-144 can also reduce the expression of c-Met and ADAM10 proteins, resulting in reduced proliferation, migration, and cell cycle progression in UM [26].
Other authors have described miR-137 in UM, finding that its expression was lower in tumoral cells than in normal melanocytes [28]. Transfection of miR-137 into melanocytes induced cell cycle arrest in G1, leading to decreased cell growth. miR-137 transfection also led to decreased levels of MITF, c-Met, and other cell cycle-related proteins such as CDK2 and CDK6 [28]. Similar to the function of miR-137, miR-182 can suppresses the expression of MITF and cell cycle-related genes [32]. In another study, it was observed that miR-137 could downregulate the expression of steroid receptor coactivators (SRCs), thus decreasing the tumor’s proliferation and ability to metastasize [46].
Interleukin 10 (IL-10) suppresses the inflammatory response and immune reactions against a variety of tumors [47]. Transfection of miR-15a, miR-185, and miR-211 independently inhibited cell proliferation through specific targeting of the IL-10Rα receptor gene [35]. Accordingly, the individual or combined expression of these three miRNAs led to a significant reduction in UM cell proliferation [35].
On the other hand, Liu and coauthors revealed the previously unknown role of miR-9 in UM [20]. Its levels were reduced in highly invasive UM, and evidence suggests that suppression of cell migration and invasion of UM cells in culture occurred through modulation of the expression of NF-KB1, which is involved in proliferation, apoptosis, angiogenesis, and tumor metastasis. Other proteins related to the NF-KB1 signaling pathway are also underexpressed, such as VEGFA, MMP-2, and MMP-9. These proteins participate in a signaling pathway that leads to tumor initiation and progression through the action of MMP (matrix metallopeptidase), which degrades basement membranes and the extracellular matrix, facilitating the spread of tumor cells. This inhibitory role on MMP2/MMP-9 in UM has also been associated with miR-296-3p [48]. Likewise, miR-23a decreases the cell migration capacity by increasing E-cadherin and reversing the epithelial–mesenchymal transition (EMT) process of UM cells through zinc finger protein Zeb1 [22].
The same tumor suppressor role was also found for miR-124a, which targets CDK4, CDK6, Cyclin D2, and EZH2 proteins, with its expression being diminished in UM cells [27]. Interestingly, it was described that treatment with hypomethylating factors successfully restores miR-124a expression, suggesting that this miRNA may undergo reversible epigenetic regulation [27]. Analogously, Liu et al. [33] demonstrated the tumor-suppressive action of miR-216a-5p in UM cell lines by inhibiting HK2 (hexokinase 2), an aerobic glycolysis-limiting enzyme that is overexpressed in numerous human cancers.
It has recently been described that miRNA-145 and miRNA-205 can reduce the proliferation, migration, and invasion of UM cells by targeting the mRNA of their upstream protein NRP1 to regulate the level of expression of CDC42 [31]. CDC42 are small GTPases that have been linked to multiple human cancers and are implicated in the epithelial to mesenchymal transition, cell cycle progression, migration/invasion, tumor growth, angiogenesis, and oncogenic transformation [49]. Other miRNAs that have been described as tumor suppressors in UM capable of decreasing levels of cell growth and invasion are miR-142-3p [29] and miR-224-5p [34]. A complete list of UM-related suppressive miRNAs is summarized in Table 1.

2.1.2. Oncogenic miRNAs

miRNAs that promote the initiation and progression of UM are considered oncogenic miRNAs. Notably, miR-21—as one of the most studied and described oncogenic miRNA—also appears to be overexpressed in other types of tumors [50]. In UM, it causes suppression of p53 gene expression and increases the expression of the cell-adhesion protein LASP1 and GST-pi protein—an enzyme involved in cell detoxification—protecting tumor cells from cytotoxic drugs. Zhang et al. analyzed the action of the miR-181 members family (miR-181a, miR-181b, miR-181c), finding that their expression was increased in UM cell lines [40]. In particular, miR-181b was reported to inhibit the expression of the tumor-suppressing phosphatase enzyme (CTDSPL) that dephosphorylates tumor suppressor Rb; thus, this miRNA inhibits the role of Rb in increasing cell proliferation [40].
Cheng et al. identified miR-222 as a mediator of the oncological effect of HMBA1 (high-mobility Group A1), a transcription factor already involved in the progression of various tumors [51,52]. The introduction of miR-222 into UM cell lines resulted in an increase in PI3K, p-Akt, and MMP9 [34]. These proteins appear to participate in a signaling pathway that leads to tumor initiation and progression through the action of MMP-9 (matrix metallopeptidase 9), which degrades basement membranes and the extracellular matrix, facilitating the spread of tumor cells [41].
PTEN is a tumor suppressor involved in the PI3K/AKT pathway; its loss of expression is associated with a more aggressive UM, which augurs for future metastasis [53]. PTEN has been identified as a target of miR-367 or miR-454; its levels decrease as they increase. Inhibition of these miRNAs resulted in suppressed cell proliferation, cell cycle progression, and cell migration, while their introduction had the opposite effects, presumably via inhibition of PTEN [42,43]. Furthermore, miR-155 introduction into cultures promoted cell growth and invasion while its inhibition produced contrary effects. Protein I related to the Nedd4 family (NDF1P1) was identified as the target of miR-155. This protein is involved in the ubiquitination and nuclear translocation of PTEN [39].
On the other hand, miR-652 directly represses the expression of HOXA9 and leads to the activation of the HIF-1 signaling pathway promoting tumor proliferation and migration in UM through HK2, a mechanism that was mentioned in the previous section [44]. High expression of miR-92a-3p appears to be associated with proliferation in other types of non-UM cancers [38,54]. Treatment with the histone deacetylase inhibitor MS-275 may reduce the expression of miR-92a-3p and increase the levels of MYC 2-binding protein (MYCBP2), which is a target gene for miR-92a-3p. This would eventually lead to apoptosis of the UM cells [38]. Finally, the transfection of miR-20a into UM cell lines contributed to increased cell invasion and migration, although, in this case, its cellular target has not yet been elucidated [36].

2.2. miRNAs in Clinical Applications

2.2.1. miRNAs as Potential Biomarkers for UM

It has been shown that the miRNA expression profiles can be used to distinguish cancerous from normal samples and even allow the classification into different subtypes and clinical stages [55,56]. miRNAs have been found in multiple fluids that allow easy accessibility as well as presenting a high degree of stability because miRNAs are usually released from cells through vesicles and cannot be easily degraded by RNases, allowing them to have long half-lives of up to 24 h [57,58]. These characteristics make miRNAs excellent potential biomarkers that could be combined with other possible serum biomarkers of a protein nature [59,60].

Studies on Tissue Samples/Cell Lines

Worley et al. were the first to determine the expression of miRNAs in 24 primary frozen UM tissues through microarray miRNA analysis, correlating the results with prognosis through determining the risk of metastasis based on chromosome 3 abnormalities [61]. They further established that let-7b and miR-199a were the miRNAs that demonstrated the greatest discriminatory potential between Class 1 and Class 2 UM, respectively corresponding to the lowest and highest metastatic risk. Moreover, they describe that through the use of a combination of the six main miRNAs with discriminatory potential (let-7b, miR-143, miR-199a, miR-199*, miR-193b, and miR-652), 100% sensitivity and specificity was achieved in distinguishing Class 1 from Class 2 tumors [61]. Similarly, in another study, five miRNAs (miR-214, miR-146b, miR-143, miR-199a, and miR-134) were found to be differentially expressed in monosomy 3 and disomy 3 UM bearing tumors [62]. The miRNAs that were found to better correlate with development of liver metastases were miR-149 and miR-134 [62]. An extensive TCGA (The Cancer Genome Atlas) UM study analyzing 80 primary UM samples was conducted by Robertson et al., in which they identified four clusters of miRNAs that were associated with chromosome 3 status, risk of metastasis, and DNA methylation profile [63]. Consistent with Worley et al., miR-199a-3p/5p, miR-199b-3p, and let-7b-5p were more highly expressed in UM characterized by chromosome 3 monosomy [61].
In another study, Smit et al. classified 26 UM tissue samples into three groups based on previously described genetic mutations linked to oncological progression (GNAQ, GNA11, EIF1AX, SF3B1, BAP1, chromosome 3 monosomy). They identified 13 miRNAs whose expression varied in a differential manner between UM and were linked to its classification as high risk, intermediate, or low risk of metastasis. Five different miRNAs were overexpressed in the high-risk group (miR-132-5p, miR-151a-3p, miR-17-5p, miR-16-5p, and miR-21-5p) and eight were underexpressed (miR-181b-5p, miR-101-3p, miR-378, miR-181a-2-3p, miR-99a-5p, let-7c-5p, miR-1537-3p, and miR-99a-3p). The 106 genes possibly regulated by these miRNAs were also found to be involved in cell cycle regulation pathways and EGF and EIF2 signaling pathways [64].
Wróblewska et al. validated 6 miRNAs (of the 15 previously selected by TCGA) in 46 UM tissue samples for the distinction between primary and metastatic UM. Results corresponded to either elevated (miR-592, miR-346, and miR-1247) or decreased (miR-506 and miR-513c) expression in the metastatic stage. miR-196b did not show quantitative differences in its expression, but was correlated with the expression of BAP1 [65]. In histological samples, Radhakrishnan and coworkers related chromosomal alterations of pairs 1, 3, and 8 with miRNA expression according to whether the samples belonged to tumors that had developed liver metastases or not [66]. They also characterized two different miRNA expression profiles in tumors with or without liver metastasis, finding that a total of 30 miRNAs were differentially expressed [66].
In recent years, different computational techniques have been used to analyze miRNA expression data in the TCGA UM database to identify miRNAs as potential prognostic biomarkers. Falzone et al. determined the prognostic importance of deregulated miRNA based on TCGA data. They determined seven miRNAs associated with tumor stage, vital status, and overall survival, including six that were downregulated (miR-514a-3p, miR-508-3p, miR-509-3-5p, miR-513c-5p, miR-513a-5p, and miR-211-5p) and another six that were upregulated (let-7b-5p, let-7b-3p, miR-224, miR-452, miR-592, and miR-199a-5p) [67]. Additionally, also using bioinformatic analysis, Xin et al. established two groups of UMs according to their prognosis risk based on miRNA analysis. In their study, they inferred that this classification can be established by testing the expression of nine miRNAs (miR-195, miR-224, miR-365a, miR-365b, miR-452, miR-4709, miR-7702, miR-513c, and miR-873), which exert action on target genes related to relevant carcinogenic processes [68]. Finally, in contrast to the above, Larsen and others in their study of 26 UM samples found no association between miRNA expression and the ability to predict metastasis and survival in UM [69]. Table 3 shows the complete list of potential biomarkers determined from studies of miRNA.

miRNAs in Serum or Plasma

Stark and coworkers proposed a panel of 17 miRNAs to differentiate UM from other benign melanocytic lesions such as nevi. Thus, by means of PCR, they analyzed the blood of 10 patients with nevus, 50 with localized UM and 5 with metastatic UM. Six of the 17 studied miRNAs showed significant differences between patients with nevus and UM, namely miR-16, miR-145, miR-146a, miR-204, miR-211, and miR-363-3p. Interestingly, the proposed panel of six miRNAs identified UM with 93% sensitivity and 100% specificity when at least four of the miRNAs were used [70].
Russo et al. studied the expression of miRNAs in serum in 14 patients with UM [71]. They found significant differences compared to healthy controls in eight miRNAs. Two miRNAs were overexpressed (miR-146a and miR-523) and six were underexpressed (miR-19a, miR-30d, miR-127, miR-451, miR-518f, and miR-1274b) in UM. After validation, only miR-146a showed significant differences in upregulation between histological UM samples [71]. On the other hand, Achberger et al. studied the blood of six patients diagnosed with UM and identified miRNAs with immune regulation capacity. The plasma levels of miR-20a, miR-125b, miR-146a, miR-155, miR-181a, and miR-223 were found to be increased in patients with UM as compared to healthy controls. Of these, levels of miR-20a, miR-125b, miR-146a, miR-155, and miR-223 were found to be increased in metastatic stages, while miR-181 was decreased [72].
A study by Ragusa et al. attempted to characterize the miRNA expression profiles in vitreous humor, vitreous humor exosomes, and serum of six patients with UM as compared to six healthy subjects as controls [73]. Ninety percent of the miRNAs were shared between the miRNA profiles of vitreous humor and vitreous exosomes. This suggests the possibility that alterations in vitreous humor exosomes are responsible for those found in the vitreous humor miRNA profile. miR-618 was underexpressed in vitreous humor but overexpressed in vitreous exosomes and serum, showing the unique characteristics of the miRNA profile in vitreous humor. The levels of miR-21, miR-34a, and miR-146a were increased in vitreous humor, vitreous exosomes, and histological samples. Interestingly, miR-146a was upregulated in the serum of UM patients, as well as in serum exosomes [73].
Finally, Triozzi et al. monitored antiangiogenic treatments in UM by detecting miRNAs in the blood. They reported the plasma levels of these miRNAs and those of miR-216 and miR-16 were shown to change following treatment with interferon-alfa-2b but not dacarbazine [74]. In a later report, they detected overexpression of miR-92b, miR-223, and miR-199-5p in plasma in samples of UM with chromosome 3 monosomy [74].
A summary of the different studies is presented in Table 4 in a schematic way. It can be seen how the miR-146a is reflected in several studies with increased values in body fluids of patients with primary and metastatic UM. Although uveal melanoma is barely studied, miR-146a has been described to increase proliferation of skin melanoma cells in culture and able to form tumors in mice [75]. In addition, miR-146a is a MITF target that is responsible for regulating the development and survival of melanocytes and may also be involved in the pigmentation and proliferation processes [76]. This may make miR-146a a serious candidate as serum biomarker in UM, although further studies would be needed to statistically validate its diagnostic power in larger cohorts of patients.

2.2.2. miRNAs as Immune Regulators

miRNAs are becoming increasingly recognized in the regulation of immune responses. Natural killer cells (NK cells) inhibit oncological progress in UM by mediating cellular cytolysis and apoptosis of cancer cells [77,78]. There is evidence that cancer stem cells (CSC) are a key factor in UM metastasis [79]. Joshi et al. reported that UM CSCs could synthesize miRNA regulators of NK cells, including miR-181a, miR-146a, miR-20a, miR-223, and miR-155. The transfer of antimiR-155 to MUM2B cells (a highly invasive metastatic UM cell line) increased the sensitivity of NK cells. In addition, high levels of miR-155 expression and secretion are detected in CSC UMs, which suggests that suppression of NK cells by miR-155 contributes to metastatic progression [80]. On the other hand, miR-199a-5p participates in regulatory T-cell regulation [81] and miR-223 regulates myeloid-derived suppressor cells (MDSCs) [82].

3. Long Non-Coding RNAs in UM

Long non-coding RNAs (lncRNAs) are a type of RNA with a length exceeding 200 nucleotides and without protein coding potential [5,83,84,85]. Modifications of the epigenetic sphere include the regulation of imprinting, conformational changes at the chromosomal level and chromatin remodeling, modifications in the methylation of histone proteins and DNA, and regulation through ncRNAs [86,87].
It is well known that lncRNAs are capable of carrying out their functions through two mechanisms, either by acting directly on neighboring genes generally found on the same chromosome or by regulating genes at a distance, including those on other chromosomes [86,88]. In addition, it has been shown that lncRNAs can interact with microRNAs [12] and may even act as specific competing endogenous RNAs (ceRNAs) to suppress the function of the miRNAs themselves [89].
Several studies have shown that lncRNAs are able to perform their functions through interaction with the Enhancer of zeste homolog 2 (EZH2), a member of the polycomb family (PcG) capable of regulating the cell cycle through interaction with transcriptional factors, chromatin remodeling, and modification of nucleosomes [90,91,92]. Numerous studies have previously shown that upregulation of EZH2 is correlated with cell proliferation and poor tumor prognosis [93]; thus, in UM, a high level of EZH2 expression has been demonstrated, suggesting that it is an oncogenic protein in this type of neoplasm [90,94,95]. However, despite this carcinogenic effect, other studies have suggested that EZH2 may also have an antitumor effect in some neoplasms since its inhibition favors tumor progression [90].

3.1. Plasmacytoma Variant Translocation 1 (PVT1)

The levels of lncRNA PVT1 (Chr 8q24.21) were found to be considerably higher in UM compared to normal tissues [21,90]. It has an oncogenic function in the initiation and development of metastasis of various neoplasms, including UM [96,97,98,99]. The silencing of PVT1 impedes the biological functions of UM cells, resulting in the inhibition of proliferation, S-phase sequestration, and migration in addition to inducing apoptosis [21].
The mechanisms by which PVT1 exerts its carcinogenic function in UM are through the positive regulation of EZH2 expression [90] as well as its function in regulating the miR-17-3p/MDM2/p53 axis. Wu et al. [21] showed that there is specific binding between lncRNA PVT1 and miR-17-3p, which implies inhibition of the expression of this microRNA. In addition, they demonstrated low expression of miR-17-3p and a high level of MDM2 in UM tissues, with a targeting relationship between both. Thus, the miR-17-3p is able to inhibit the expression of MDM2, ultimately facilitating the transcriptional activation of the tumor suppressor protein p53 and reduction in the tumorigenic capacity. In conclusion, lncRNA PVT1 favors tumor emergence and development by binding to and inhibiting the expression of miR-17-3p, promoting the expression of MDM2 and the inactivation of p53.
Xu et al. demonstrated the association between a high expression of PVT1 and a number of clinicopathological parameters of poor prognosis in UM, such as advanced age, histological epithelioid-type, presence of extra-scleral extension, and distant metastasis [96]. However, it was not correlated with the thickness or size of the tumor mass. Overall, PVT1 overexpression was correlated with a higher mortality rate and was able to independently predict poor overall survival of such patients with UM (Figure 2).

3.2. CASC15-New-Transcript 1

Nuclear lncRNA CASC15-new-transcript 1 (CANT1 or CASC15-NT1), present in human uveal tissue, is an isoform of non-coding lncRNA CASC15 (Chr 6p22.3). Both CASC15 and CANT1 are downregulated in UM while an inverse expression patten is observed in normal cells [100].
CANT1 acts as a tumor suppressor, inhibiting both cell migration and tumorigenesis of UM, for which it is able to trigger and modulate a non-coding cascade (CANT1–JPX/FTX–XIST) [100]. As with CANT1, these three potential regulatory targets—JPX, FTX, and XIST—have low levels of expression in UM tissues. Inhibition of JPX/FTX transmits the signal to XIST, decreasing its expression levels and completing the sequence that ultimately leads to both activation of tumor growth and stimulation of metastatic capacity.

3.3. Metastasis-Associated Lung Adenocarcinoma Transcript 1 (MALAT1)

MALAT1 (Chr 11q13.1) has been reported to be able to function as a ceRNA since it regulates the expression of miRNAs such as miR-34 or miR-183 in cutaneous melanoma [101,102]. As with other types of tumors, MALAT1 acts as an oncoRNA in UM cells, and its abundance is upregulated [103,104]. Two main axes have been described by which lncRNA MALAT1 performs its oncogenic function:
(a) Sun et al. demonstrated that MALAT1 inhibits miR-140 and, thus, promotes the protein expression of its target genes, Slug and ADAM10, in order to induce proliferation, migration, colony formation, and invasion of UM cell lines [103]. (b) Wu et al. deciphered the MALAT1/miR-608/HOXC4 axis in UM, showing that MALAT1 is capable of enhancing HOXC4 expression through the molecular interaction and inhibition of miR-608, ultimately promoting in vitro UM cell proliferation, invasion, and migration and in vivo tumor growth [104] (Figure 2).

3.4. Homeobox A11 Antisense (HOXA11-AS)

Lu et al. described a high level of expression of HOXA11-AS (Chr 7p15.2) in UM tumor samples. This is an oncogenic lncRNA, which promotes tumorigenesis by inducing cell proliferation and invasion while inhibiting apoptotic events [94]. The pathways by which HOXA11-AS performs its functions are the following: (a) through direct interaction with EZH2, to silence the transcription of the cell cycle regulation gene p21; (b) binding directly to miR-124 and inhibiting its expression; and (c) acting as a ceRNA of miR-124, since degradation of HOXA11-AS enzyme causes a decrease in EZH2 levels (Figure 2).

3.5. Small Nucleolar RNA Host Gene 7 (SNHG7)

The expression of SNHG7 (Chr 9q34.3) is reduced in UM tissues, and especially in cases where disease is already metastatic [95]. Thus, low levels of SNHG7 expression are associated with worse prognosis in UM as well as a higher proportion of epithelioid cells, poor clinical staging, lower tumor-free survival, and a higher mortality rate. SNHG7 is inversely associated with EZH2, and high levels of EZH2 expression were found to be correlated with a poor histological type and higher clinical TNM stage [95] (Figure 2).

3.6. Ferrtin Heavy-Chain 1 Pseudogene 3 (FTH1P3)

It was found that FTH1P3 (Chr 2p23.3) is upregulated in UM tissues, and that such elevated levels of expression favor proliferation and migration while promoting cell cycle progression. The lncRNA FTH1P3 therefore functions as an oncoRNA that induces UM progression through target interaction inversely correlated with microRNA-224-5p (miR-224-5p) levels [105]. Ultimately, the increase in FTH1P3 levels implies secondary elevation of the protein expression of Rac1 and Fizzled 5 as they are direct target genes of miR-224-5p [97] (Figure 2).

3.7. Purinergic Receiver P2X7 Variant3 (P2RX7-V3)

P2RX7-V3 (Chr 12q24.31) is specifically expressed in UM and functions as an oncogene, present at high levels and promoting proliferation, migratory capacity, cell invasion, tumor growth, and metastasis [106]. Furthermore, by silencing P2RX7-V3, it was demonstrated that a wide variety of genes are deregulated and that the main target signaling pathway by which it performs its carcinogenic functions is the PI3K/AKT pathway [106].

3.8. Rhophilin RHO Gtpase Binding Protein 1 Antisense 1 (RHPN1-AS1)

Lu et al. demonstrated that RHPN1-AS1 (Chr 8q24.3) is an oncoRNA as it is overexpressed in UM cell lines and induces the proliferation, migration, invasion, and growth of tumor mass by deregulation of various biologic functions and pathways [107]. Their study was carried out by means of gene expression analysis after knockdown of RHPN1-AS1, observing that the most affected biological processes are angiogenesis, cell adhesion, and organization of the extracellular matrix, and that the most regulated signaling pathways are, among others, the metabolism of nicotinate and nicotinamide as well as the transforming growth factor-beta (TGF-β) pathway, whose constituent genes are stimulated, suggesting that RHPN1-AS1 favors UM progression by participating in this signaling pathway.

3.9. Pax6 Upstream Antisense RNA (PAUPAR)

At the cellular level, PAUPAR (Chr 11p13) is mainly located in the nucleus and is considered a neoplastic suppressor that inhibits tumor formation, both in vitro and in vivo, in addition to suppressing metastatic progression and dissemination in UM [108]. PAUPAR induces silencing of HES1 [108], a mediator of the Notch signaling pathway which plays a key role in the survival of immature melanoblasts and melanocytic stem cells by inhibiting the onset of apoptotic phenomena [109]. PAUPAR expression is reduced in UM cell lines and tissues while HES1 is highly expressed compared to normal healthy tissues. Similarly, the Notch pathway is activated [110], resulting in increased migration capacity and cell colony formation [108].

3.10. Other lncRNAs

lncRNA-numb (Chr 14q24.3) acts as a tumor suppressor, and its low levels of expression may be associated with the progression of UM by favoring cell proliferation and invasion [111].
On the other hand, recently, the high expression of the lncRNA SNHG15 (Chr 7p13) has been associated to clinical-pathological variables of UM related to the bad prognosis, including age, diameter of the tumor, pathological type, and extra scleral extension. At present, the impact of SNHG15 in UM is still not clear. Therefore, it is necessary to elucidate the biological mechanism of the SNHG15 in UM [112].

4. Conclusions

In this review, we have attempted to establish a schematic approach to the new topic of ncRNAs in UM. In the last decade, these small molecules have become very important in cancer research: in particular, the number of published works related to UM has multiplied exponentially. The number of ncRNAs being newly described new each year makes it impossible to compile all the available information in a simple way. It seems clear that ncRNAs will play an important role in the diagnosis, prognosis, and treatment of UM in the near future.
ncRNAs along with their regular target genes participate in multiple biological processes in tumors that makes them essential for the proliferation and continued metastasis of tumors. This makes them potential targets and regulators of the therapeutic response in UM. miRNA let-7b appears to suppress the expression of cyclin D1 in UM cells and enhance the radiosensitivity of UM through cell cycle arrest. This characteristic of let-7b could allow it to act as a radiosensitivity enhancer and improve the therapeutic effects of local radiation in the treatment of UM [113]. Anti-PD-1/PD-L1 antibodies (such as pembrolizumab, nivolumab) are used as drugs in the treatment of metastatic UM. PD-L1 is constitutively expressed in tumor cells and interacts with PD-1 to deactivate T cells, leading to immune escape [114]. Audrito et al. found an inverse correlation between miR-17-5p and the expression of PD-L1 in uveal melanoma cell lines [115]. A decrease in the expression of PD-1 was found in mice with miR-28-transferred melanoma, meaning that the miR-28 could silence PD-1 and reverse the exhaustive state of T cells [116].
In an interesting article, Sharma et al. described how BAP1 mutations can change the expression of a tumor-specific miRNA network, resulting in alterations to UM behavior [117]. BAP1 is, conceptually, a driver gene in UM and its alteration has been linked to an increased risk of metastasis. All this reinforces the concept that ncRNAs could serve as excellent prognostic markers, with the added bonus that they can be found in multiple body fluids along with the advantage that they can be obtained noninvasively. Unfortunately, at present, multiple studies have produced very few miRNAs that appear to be common biomarkers—in some cases, even producing conflicting results. Reasons for this could be the use of small samples and heterogeneity in the inclusion processes or in the quantification methods [118].
Though a wide variety of ncRNAs have been studied that are correlated with the pathogenesis of UM, their function and expression remain unknown in many cases. Research into the different types of ncRNAs as well as their underlying mechanisms of action may have vitally important diagnostic, therapeutic, and prognostic implications in UM. This is further complicated by the recent discovery of miRNA isoforms (isomiRs) that were initially ruled out as aberrant sequence artifacts. Some miRNA arms can produce more than 30 different isomiRs, and each isomiR of the same miRNA can target different mRNAs. Therefore, isomiR expression could greatly increase the number of miRNA-regulated target mRNAs [119,120]. Londin et al. found examples of previously reported miRNA loci that produce abundant isomiRs in UM and that their expression was correlated with clinical attributes [121].
To conclude, we hope that future studies will allow us to have a better understanding of the participation of ncRNAs in UM. We believe that these studies should include a greater number of samples and adopt bioinformatic tools for predictive analyses to generate preliminary findings that will require subsequent clinical validation.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2072-6694/12/10/2944/s1, Table S1: Characteristics of the UM cell lines used in the studies in Table 1 and Table 2.

Author Contributions

Conceptualization, M.B.; methodology, M.B.; validation, M.P., D.F.-D., M.B., and A.P.; writing—original draft preparation, C.R.-V., M.P., D.F.-D., L.P., B.F.-M., N.L.-B., and M.B.; writing—review and editing, M.P., M.J.B.-T., N.L.-B., and P.S.-R.; supervision, M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Grant PI16/00143 from Instituto de Salud Carlos III.

Acknowledgments

This work was financially supported by Instituto de Salud Carlos III. Figure 2 was created with BioRender.com

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hattori, M. Finishing the euchromatic sequence of the human genome. Tanpakushitsu Kakusan Koso 2005, 50, 162–168. [Google Scholar] [PubMed]
  2. Consortium, E.P. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef] [PubMed]
  3. Kapranov, P.; Drenkow, J.; Cheng, J.; Long, J.; Helt, G.; Dike, S.; Gingeras, T.R. Examples of the complex architecture of the human transcriptome revealed by RACE and high-density tiling arrays. Genome Res. 2005, 15, 987–997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Taft, R.J.; Pang, K.C.; Mercer, T.R.; Dinger, M.; Mattick, J.S. Non-coding RNAs: Regulators of disease. J. Pathol. 2010, 220, 126–139. [Google Scholar] [CrossRef] [PubMed]
  5. Guzel, E.; Okyay, T.M.; Yalcinkaya, B.; Karacaoglu, S.; Gocmen, M.; Akcakuyu, M.H. Tumor suppressor and oncogenic role of long non-coding RNAs in cancer. North. Clin. Istanb. 2020, 7, 81–86. [Google Scholar] [CrossRef] [PubMed]
  6. Staby, K.M.; Gravdal, K.; Mork, S.J.; Heegaard, S.; Vintermyr, O.K.; Krohn, J. Prognostic impact of chromosomal aberrations and GNAQ, GNA11 and BAP1 mutations in uveal melanoma. Acta Ophthalmol. 2018, 96, 31–38. [Google Scholar] [CrossRef]
  7. Damato, B.; Duke, C.; Coupland, S.E.; Hiscott, P.; Smith, P.A.; Campbell, I.; Douglas, A.; Howard, P. Cytogenetics of uveal melanoma: A 7-year clinical experience. Ophthalmology 2007, 114, 1925–1931. [Google Scholar] [CrossRef]
  8. Shain, A.H.; Bagger, M.M.; Yu, R.; Chang, D.; Liu, S.; Vemula, S.; Weier, J.F.; Wadt, K.; Heegaard, S.; Bastian, B.C.; et al. The genetic evolution of metastatic uveal melanoma. Nat. Genet. 2019, 51, 1123–1130. [Google Scholar] [CrossRef]
  9. Rodrigues, M.J.; Stern, M.H. Genetic landscape of uveal melanoma. J. Fr. D’ophtalmologie 2015, 38, 522–525. [Google Scholar] [CrossRef]
  10. Friedman, R.C.; Farh, K.K.; Burge, C.B.; Bartel, D.P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009, 19, 92–105. [Google Scholar] [CrossRef] [Green Version]
  11. Peng, Y.; Croce, C.M. The role of MicroRNAs in human cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Liz, J.; Esteller, M. lncRNAs and microRNAs with a role in cancer development. Biochimica et Biophysica Acta 2016, 1859, 169–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhao, D.D.; Zhao, X.; Li, W.T. Identification of differentially expressed metastatic genes and their signatures to predict the overall survival of uveal melanoma patients by bioinformatics analysis. Int. J. Ophthalmol. 2020, 13, 1046–1053. [Google Scholar] [CrossRef] [PubMed]
  14. Griewank, K.G.; Yu, X.; Khalili, J.; Sozen, M.M.; Stempke-Hale, K.; Bernatchez, C.; Wardell, S.; Bastian, B.C.; Woodman, S.E. Genetic and molecular characterization of uveal melanoma cell lines. Pigment Cell Melanoma Res. 2012, 25, 182–187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Nareyeck, G.; Zeschnigk, M.; Prescher, G.; Lohmann, D.R.; Anastassiou, G. Establishment and characterization of two uveal melanoma cell lines derived from tumors with loss of one chromosome 3. Exp. Eye Res. 2006, 83, 858–864. [Google Scholar] [CrossRef] [PubMed]
  16. Jager, M.J.; Magner, J.A.; Ksander, B.R.; Dubovy, S.R. Uveal Melanoma Cell Lines: Where do they come from? (An American Ophthalmological Society Thesis). Trans. Am. Ophthalmol. Soc. 2016, 114, T5. [Google Scholar]
  17. Yu, X.; Ambrosini, G.; Roszik, J.; Eterovic, A.K.; Stempke-Hale, K.; Seftor, E.A.; Chattopadhyay, C.; Grimm, E.; Carvajal, R.D.; Hendrix, M.J.; et al. Genetic analysis of the ‘uveal melanoma’ C918 cell line reveals atypical BRAF and common KRAS mutations and single tandem repeat profile identical to the cutaneous melanoma C8161 cell line. Pigment Cell Melanoma Res. 2015, 28, 357–359. [Google Scholar] [CrossRef]
  18. Amirouchene-Angelozzi, N.; Nemati, F.; Gentien, D.; Nicolas, A.; Dumont, A.; Carita, G.; Camonis, J.; Desjardins, L.; Cassoux, N.; Piperno-Neumann, S.; et al. Establishment of novel cell lines recapitulating the genetic landscape of uveal melanoma and preclinical validation of mTOR as a therapeutic target. Mol. Oncol. 2014, 8, 1508–1520. [Google Scholar] [CrossRef]
  19. Luyten, G.P.; Naus, N.C.; Mooy, C.M.; Hagemeijer, A.; Kan-Mitchell, J.; Van Drunen, E.; Vuzevski, V.; De Jong, P.T.; Luider, T.M. Establishment and characterization of primary and metastatic uveal melanoma cell lines. Int. J. Cancer 1996, 66, 380–387. [Google Scholar] [CrossRef]
  20. Liu, N.; Sun, Q.; Chen, J.; Li, J.; Zeng, Y.; Zhai, S.; Li, P.; Wang, B.; Wang, X. MicroRNA-9 suppresses uveal melanoma cell migration and invasion through the NF-kappaB1 pathway. Oncol. Rep. 2012, 28, 961–968. [Google Scholar] [CrossRef]
  21. Wu, S.; Chen, H.; Han, N.; Zhang, C.; Yan, H. Long Noncoding RNA PVT1 Silencing Prevents the Development of Uveal Melanoma by Impairing MicroRNA-17-3p-Dependent MDM2 Upregulation. Investig. Ophthalmol. Vis. Sci. 2019, 60, 4904–4914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Wang, Y.; Luo, Y.; Guan, W.; Zhao, H. Role of miR-23a/Zeb1 negative feedback loop in regulating epithelial-mesenchymal transition and tumorigenicity of intraocular tumors. Oncol. Lett. 2018, 16, 2462–2470. [Google Scholar] [CrossRef] [PubMed]
  23. Yan, D.; Zhou, X.; Chen, X.; Hu, D.N.; Dong, X.D.; Wang, J.; Lu, F.; Tu, L.; Qu, J. MicroRNA-34a inhibits uveal melanoma cell proliferation and migration through downregulation of c-Met. Investig. Ophthalmol. Vis. Sci. 2009, 50, 1559–1565. [Google Scholar] [CrossRef] [PubMed]
  24. Hou, Q.; Han, S.; Yang, L.; Chen, S.; Chen, J.; Ma, N.; Wang, C.; Tang, J.; Chen, X.; Chen, F.; et al. The Interplay of MicroRNA-34a, LGR4, EMT-Associated Factors, and MMP2 in Regulating Uveal Melanoma Cells. Investig. Ophthalmol. Vis. Sci. 2019, 60, 4503–4510. [Google Scholar] [CrossRef] [Green Version]
  25. Dong, F.; Lou, D. MicroRNA-34b/c suppresses uveal melanoma cell proliferation and migration through multiple targets. Mol. Vis. 2012, 18, 537–546. [Google Scholar]
  26. Amaro, A.; Croce, M.; Ferrini, S.; Barisione, G.; Gualco, M.; Perri, P.; Pfeffer, U.; Jager, M.J.; Coupland, S.E.; Mosci, C.; et al. Potential Onco-Suppressive Role of miR122 and miR144 in Uveal Melanoma through ADAM10 and C-Met Inhibition. Cancers 2020, 12, 1468. [Google Scholar] [CrossRef]
  27. Chen, X.; He, D.; Dong, X.D.; Dong, F.; Wang, J.; Wang, L.; Tang, J.; Hu, D.N.; Yan, D.; Tu, L. MicroRNA-124a is epigenetically regulated and acts as a tumor suppressor by controlling multiple targets in uveal melanoma. Investig. Ophthalmol. Vis. Sci. 2013, 54, 2248–2256. [Google Scholar] [CrossRef] [Green Version]
  28. Chen, X.; Wang, J.; Shen, H.; Lu, J.; Li, C.; Hu, D.N.; Dong, X.D.; Yan, D.; Tu, L. Epigenetics, microRNAs, and carcinogenesis: Functional role of microRNA-137 in uveal melanoma. Investig. Ophthalmol. Vis. Sci. 2011, 52, 1193–1199. [Google Scholar] [CrossRef]
  29. Peng, D.; Dong, J.; Zhao, Y.; Peng, X.; Tang, J.; Chen, X.; Wang, L.; Hu, D.N.; Reinach, P.S.; Qu, J.; et al. miR-142-3p suppresses uveal melanoma by targeting CDC25C, TGFbetaR1, GNAQ, WASL, and RAC1. Cancer Manag. Res. 2019, 11, 4729–4742. [Google Scholar] [CrossRef] [Green Version]
  30. Sun, L.; Bian, G.; Meng, Z.; Dang, G.; Shi, D.; Mi, S. MiR-144 Inhibits Uveal Melanoma Cell Proliferation and Invasion by Regulating c-Met Expression. PLoS ONE 2015, 10, e0124428. [Google Scholar] [CrossRef] [Green Version]
  31. Li, Y.; Luo, J.T.; Liu, Y.M.; Wei, W.B. miRNA-145/miRNA-205 inhibits proliferation and invasion of uveal melanoma cells by targeting NPR1/CDC42. Int. J. Ophthalmol. 2020, 13, 718–724. [Google Scholar] [CrossRef] [PubMed]
  32. Yan, D.; Dong, X.D.; Chen, X.; Yao, S.; Wang, L.; Wang, J.; Wang, C.; Hu, D.N.; Qu, J.; Tu, L. Role of microRNA-182 in posterior uveal melanoma: Regulation of tumor development through MITF, BCL2 and cyclin D2. PLoS ONE 2012, 7, e40967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Liu, Y.; Huo, Y.; Wang, D.; Tai, Y.; Li, J.; Pang, D.; Zhang, Y.; Zhao, W.; Du, N.; Huang, Y. MiR-216a-5p/Hexokinase 2 axis regulates uveal melanoma growth through modulation of Warburg effect. Biochem. Biophys. Res. Commun. 2018, 501, 885–892. [Google Scholar] [CrossRef] [PubMed]
  34. Li, J.; Liu, X.; Li, C.; Wang, W. miR-224-5p inhibits proliferation, migration, and invasion by targeting PIK3R3/AKT3 in uveal melanoma. J. Cell. Biochem. 2019, 120, 12412–12421. [Google Scholar] [CrossRef] [PubMed]
  35. Venza, I.; Visalli, M.; Beninati, C.; Benfatto, S.; Teti, D.; Venza, M. IL-10Ralpha expression is post-transcriptionally regulated by miR-15a, miR-185, and miR-211 in melanoma. BMC Med. Genom. 2015, 8, 81. [Google Scholar] [CrossRef] [Green Version]
  36. Zhou, J.; Jiang, J.; Wang, S.; Xia, X. Oncogenic role of microRNA20a in human uveal melanoma. Mol. Med. Rep. 2016, 14, 1560–1566. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, Y.C.; Yang, X.; Wei, W.B.; Xu, X.L. Role of microRNA-21 in uveal melanoma cell invasion and metastasis by regulating p53 and its downstream protein. Int. J. Ophthalmol. 2018, 11, 1258–1268. [Google Scholar] [CrossRef]
  38. Venza, M.; Visalli, M.; Beninati, C.; Benfatto, S.; Teti, D.; Venza, I. miR-92a-3p and MYCBP2 are involved in MS-275-induced and c-myc-mediated TRAIL-sensitivity in melanoma cells. Int. Immunopharmacol. 2016, 40, 235–243. [Google Scholar] [CrossRef]
  39. Peng, J.; Liu, H.; Liu, C. MiR-155 Promotes Uveal Melanoma Cell Proliferation and Invasion by Regulating NDFIP1 Expression. Technol. Cancer Res. Treat. 2017, 16, 1160–1167. [Google Scholar] [CrossRef] [Green Version]
  40. Zhang, L.; He, X.; Li, F.; Pan, H.; Huang, X.; Wen, X.; Zhang, H.; Li, B.; Ge, S.; Xu, X.; et al. The miR-181 family promotes cell cycle by targeting CTDSPL, a phosphatase-like tumor suppressor in uveal melanoma. J. Exp. Clin. Cancer Res. CR 2018, 37, 15. [Google Scholar] [CrossRef]
  41. Cheng, Y.; Cheng, T.; Zhao, Y.; Qu, Y. HMGA1 exacerbates tumor progression by activating miR-222 through PI3K/Akt/MMP-9 signaling pathway in uveal melanoma. Cell. Signal. 2019, 63, 109386. [Google Scholar] [CrossRef] [PubMed]
  42. Ling, J.W.; Lu, P.R.; Zhang, Y.B.; Jiang, S.; Zhang, Z.C. miR-367 promotes uveal melanoma cell proliferation and migration by regulating PTEN. Genet. Mol. Res. 2017, 16. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, L.; Wang, Q.; Gao, X.; Shi, D.; Mi, S.; Han, Q. MicroRNA-454 functions as an oncogene by regulating PTEN in uveal melanoma. FEBS Lett. 2015, 589, 2791–2796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Xia, Z.; Yang, C.; Yang, X.; Wu, S.; Feng, Z.; Qu, L.; Chen, X.; Liu, L.; Ma, Y. miR-652 Promotes Proliferation and Migration of Uveal Melanoma Cells by Targeting HOXA9. Med Sci. Monit. 2019, 25, 8722–8732. [Google Scholar] [CrossRef]
  45. Abdel-Rahman, M.H.; Boru, G.; Massengill, J.; Salem, M.M.; Davidorf, F.H. MET oncogene inhibition as a potential target of therapy for uveal melanomas. Investig. Ophthalmol. Vis. Sci. 2010, 51, 3333–3339. [Google Scholar] [CrossRef] [Green Version]
  46. Eedunuri, V.K.; Rajapakshe, K.; Fiskus, W.; Geng, C.; Chew, S.A.; Foley, C.; Shah, S.S.; Shou, J.; Mohamed, J.S.; Coarfa, C.; et al. miR-137 Targets p160 Steroid Receptor Coactivators SRC1, SRC2, and SRC3 and Inhibits Cell Proliferation. Mol. Endocrinol. 2015, 29, 1170–1183. [Google Scholar] [CrossRef] [Green Version]
  47. Mocellin, S.; Marincola, F.M.; Young, H.A. Interleukin-10 and the immune response against cancer: A counterpoint. J. Leukoc. Biol. 2005, 78, 1043–1051. [Google Scholar] [CrossRef]
  48. Wang, X.; Hu, Y.; Cui, J.; Zhou, Y.; Chen, L. Coordinated targeting of MMP-2/MMP-9 by miR-296-3p/FOXCUT exerts tumor-suppressing effects in choroidal malignant melanoma. Mol. Cell. Biochem. 2018, 445, 25–33. [Google Scholar] [CrossRef]
  49. Maldonado, M.D.M.; Dharmawardhane, S. Targeting Rac and Cdc42 GTPases in Cancer. Cancer Res. 2018, 78, 3101–3111. [Google Scholar] [CrossRef] [Green Version]
  50. Okayama, H.; Schetter, A.J.; Harris, C.C. MicroRNAs and inflammation in the pathogenesis and progression of colon cancer. Dig. Dis. 2012, 30 (Suppl. S2), 9–15. [Google Scholar] [CrossRef]
  51. Williams, M.D.; Zhang, X.; Belton, A.S.; Xian, L.; Huso, T.; Park, J.J.; Siems, W.F.; Gang, D.R.; Resar, L.M.; Reeves, R.; et al. HMGA1 drives metabolic reprogramming of intestinal epithelium during hyperproliferation, polyposis, and colorectal carcinogenesis. J. Proteome Res. 2015, 14, 1420–1431. [Google Scholar] [CrossRef] [PubMed]
  52. Zhou, J.; Xie, M.; He, H.; Shi, Y.; Luo, B.; Gong, G.; Li, J.; Wang, J.; Wu, X.; Wen, J. Increases urinary HMGA1 in serous epithelial ovarian cancer patients. Cancer Biomark. Sect. A Dis. Markers 2015, 15, 325–331. [Google Scholar] [CrossRef] [PubMed]
  53. Abdel-Rahman, M.H.; Yang, Y.; Zhou, X.P.; Craig, E.L.; Davidorf, F.H.; Eng, C. High frequency of submicroscopic hemizygous deletion is a major mechanism of loss of expression of PTEN in uveal melanoma. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2006, 24, 288–295. [Google Scholar] [CrossRef] [PubMed]
  54. Li, M.; Guan, X.; Sun, Y.; Mi, J.; Shu, X.; Liu, F.; Li, C. miR-92a family and their target genes in tumorigenesis and metastasis. Exp. Cell Res. 2014, 323, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lin, S.; Gregory, R.I. MicroRNA biogenesis pathways in cancer. Nat. Rev. Cancer 2015, 15, 321–333. [Google Scholar] [CrossRef]
  56. Calin, G.A.; Croce, C.M. MicroRNA signatures in human cancers. Nat. Rev. Cancer 2006, 6, 857–866. [Google Scholar] [CrossRef] [PubMed]
  57. Gilad, S.; Meiri, E.; Yogev, Y.; Benjamin, S.; Lebanony, D.; Yerushalmi, N.; Benjamin, H.; Kushnir, M.; Cholakh, H.; Melamed, N.; et al. Serum microRNAs are promising novel biomarkers. PLoS ONE 2008, 3, e3148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Kosaka, N.; Iguchi, H.; Ochiya, T. Circulating microRNA in body fluid: A new potential biomarker for cancer diagnosis and prognosis. Cancer Sci. 2010, 101, 2087–2092. [Google Scholar] [CrossRef]
  59. Reiniger, I.W.; Schaller, U.C.; Haritoglou, C.; Hein, R.; Bosserhoff, A.K.; Kampik, A.; Mueller, A.J. “Melanoma inhibitory activity” (MIA): A promising serological tumour marker in metastatic uveal melanoma. Graefes Arch. Clin. Exp. Ophthalmol. 2005, 243, 1161–1166. [Google Scholar] [CrossRef]
  60. Bande, M.F.; Santiago, M.; Mera, P.; Piulats, J.M.; Blanco, M.J.; Rodriguez-Alvarez, M.X.; Capeans, C.; Pineiro, A.; Pardo, M. ME20-S as a Potential Biomarker for the Evaluation of Uveal Melanoma. Investig. Ophthalmol. Vis. Sci. 2015, 56, 7007–7011. [Google Scholar] [CrossRef] [Green Version]
  61. Worley, L.A.; Long, M.D.; Onken, M.D.; Harbour, J.W. Micro-RNAs associated with metastasis in uveal melanoma identified by multiplexed microarray profiling. Melanoma Res. 2008, 18, 184–190. [Google Scholar] [CrossRef] [PubMed]
  62. Venkatesan, N.; Kanwar, J.; Deepa, P.R.; Khetan, V.; Crowley, T.M.; Raguraman, R.; Sugneswari, G.; Rishi, P.; Natarajan, V.; Biswas, J.; et al. Clinico-Pathological Association of Delineated miRNAs in Uveal Melanoma with Monosomy 3/Disomy 3 Chromosomal Aberrations. PLoS ONE 2016, 11, e0146128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Robertson, A.G.; Shih, J.; Yau, C.; Gibb, E.A.; Oba, J.; Mungall, K.L.; Hess, J.M.; Uzunangelov, V.; Walter, V.; Danilova, L.; et al. Integrative Analysis Identifies Four Molecular and Clinical Subsets in Uveal Melanoma. Cancer Cell 2018, 33, 151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Smit, K.N.; Chang, J.; Derks, K.; Vaarwater, J.; Brands, T.; Verdijk, R.M.; Wiemer, E.A.C.; Mensink, H.W.; Pothof, J.; de Klein, A.; et al. Aberrant MicroRNA Expression and Its Implications for Uveal Melanoma Metastasis. Cancers 2019, 11, 815. [Google Scholar] [CrossRef] [Green Version]
  65. Wroblewska, J.P.; Lach, M.S.; Ustaszewski, A.; Kulcenty, K.; Ibbs, M.; Jagiello, I.; Suchorska, W.M.; Marszalek, A. The Potential Role of Selected miRNA in Uveal Melanoma Primary Tumors as Early Biomarkers of Disease Progression. Genes 2020, 11, 271. [Google Scholar] [CrossRef] [Green Version]
  66. Radhakrishnan, A.; Badhrinarayanan, N.; Biswas, J.; Krishnakumar, S. Analysis of chromosomal aberration (1, 3, and 8) and association of microRNAs in uveal melanoma. Mol. Vis. 2009, 15, 2146–2154. [Google Scholar]
  67. Falzone, L.; Romano, G.L.; Salemi, R.; Bucolo, C.; Tomasello, B.; Lupo, G.; Anfuso, C.D.; Spandidos, D.A.; Libra, M.; Candido, S. Prognostic significance of deregulated microRNAs in uveal melanomas. Mol. Med. Rep. 2019, 19, 2599–2610. [Google Scholar] [CrossRef] [Green Version]
  68. Xin, X.; Zhang, Y.; Ling, F.; Wang, L.; Sheng, X.; Qin, L.; Zhao, X. Identification of a nine-miRNA signature for the prognosis of Uveal Melanoma. Exp. Eye Res. 2019, 180, 242–249. [Google Scholar] [CrossRef]
  69. Larsen, A.C.; Holst, L.; Kaczkowski, B.; Andersen, M.T.; Manfe, V.; Siersma, V.D.; Kolko, M.; Kiilgaard, J.F.; Winther, O.; Prause, J.U.; et al. MicroRNA expression analysis and Multiplex ligation-dependent probe amplification in metastatic and non-metastatic uveal melanoma. Acta Ophthalmol. 2014, 92, 541–549. [Google Scholar] [CrossRef]
  70. Stark, M.S.; Gray, E.S.; Isaacs, T.; Chen, F.K.; Millward, M.; McEvoy, A.; Zaenker, P.; Ziman, M.; Soyer, H.P.; Glasson, W.J.; et al. A Panel of Circulating MicroRNAs Detects Uveal Melanoma with High Precision. Transl. Vis. Sci. Technol. 2019, 8, 12. [Google Scholar] [CrossRef]
  71. Russo, A.; Caltabiano, R.; Longo, A.; Avitabile, T.; Franco, L.M.; Bonfiglio, V.; Puzzo, L.; Reibaldi, M. Increased Levels of miRNA-146a in Serum and Histologic Samples of Patients with Uveal Melanoma. Front. Pharmacol. 2016, 7, 424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Achberger, S.; Aldrich, W.; Tubbs, R.; Crabb, J.W.; Singh, A.D.; Triozzi, P.L. Circulating immune cell and microRNA in patients with uveal melanoma developing metastatic disease. Mol. Immunol. 2014, 58, 182–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ragusa, M.; Barbagallo, C.; Statello, L.; Caltabiano, R.; Russo, A.; Puzzo, L.; Avitabile, T.; Longo, A.; Toro, M.D.; Barbagallo, D.; et al. miRNA profiling in vitreous humor, vitreal exosomes and serum from uveal melanoma patients: Pathological and diagnostic implications. Cancer Biol. Ther. 2015, 16, 1387–1396. [Google Scholar] [CrossRef] [PubMed]
  74. Triozzi, P.L.; Achberger, S.; Aldrich, W.; Crabb, J.W.; Saunthararajah, Y.; Singh, A.D. Association of tumor and plasma microRNA expression with tumor monosomy-3 in patients with uveal melanoma. Clin. Epigenetics 2016, 8, 80. [Google Scholar] [CrossRef] [Green Version]
  75. Forloni, M.; Dogra, S.K.; Dong, Y.; Conte, D., Jr.; Ou, J.; Zhu, L.J.; Deng, A.; Mahalingam, M.; Green, M.R.; Wajapeyee, N. miR-146a promotes the initiation and progression of melanoma by activating Notch signaling. eLife 2014, 3, e01460. [Google Scholar] [CrossRef]
  76. Yajima, I.; Kumasaka, M.Y.; Thang, N.D.; Goto, Y.; Takeda, K.; Iida, M.; Ohgami, N.; Tamura, H.; Yamanoshita, O.; Kawamoto, Y.; et al. Molecular Network Associated with MITF in Skin Melanoma Development and Progression. J. Ski. Cancer 2011, 2011, 730170. [Google Scholar] [CrossRef] [Green Version]
  77. Trotta, R.; Chen, L.; Ciarlariello, D.; Josyula, S.; Mao, C.; Costinean, S.; Yu, L.; Butchar, J.P.; Tridandapani, S.; Croce, C.M.; et al. miR-155 regulates IFN-gamma production in natural killer cells. Blood 2012, 119, 3478–3485. [Google Scholar] [CrossRef] [Green Version]
  78. Cichocki, F.; Felices, M.; McCullar, V.; Presnell, S.R.; Al-Attar, A.; Lutz, C.T.; Miller, J.S. Cutting edge: MicroRNA-181 promotes human NK cell development by regulating Notch signaling. J. Immunol. 2011, 187, 6171–6175. [Google Scholar] [CrossRef] [Green Version]
  79. Djirackor, L.; Kalirai, H.; Coupland, S.E.; Petrovski, G. CD166high Uveal Melanoma Cells Represent a Subpopulation With Enhanced Migratory Capacity. Investig. Ophthalmol. Vis. Sci. 2019, 60, 2696–2704. [Google Scholar] [CrossRef] [Green Version]
  80. Joshi, P.; Kooshki, M.; Aldrich, W.; Varghai, D.; Zborowski, M.; Singh, A.D.; Triozzi, P.L. Expression of natural killer cell regulatory microRNA by uveal melanoma cancer stem cells. Clin. Exp. Metastasis 2016, 33, 829–838. [Google Scholar] [CrossRef]
  81. Chatila, W.M.; Criner, G.J.; Hancock, W.W.; Akimova, T.; Moldover, B.; Chang, J.K.; Cornwell, W.; Santerre, M.; Rogers, T.J. Blunted expression of miR-199a-5p in regulatory T cells of patients with chronic obstructive pulmonary disease compared to unaffected smokers. Clin. Exp. Immunol. 2014, 177, 341–352. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, Q.; Zhang, M.; Jiang, X.; Zhang, Z.; Dai, L.; Min, S.; Wu, X.; He, Q.; Liu, J.; Zhang, Y.; et al. miR-223 suppresses differentiation of tumor-induced CD11b(+) Gr1(+) myeloid-derived suppressor cells from bone marrow cells. Int. J. Cancer 2011, 129, 2662–2673. [Google Scholar] [CrossRef]
  83. Chu, C.; Qu, K.; Zhong, F.L.; Artandi, S.E.; Chang, H.Y. Genomic maps of long noncoding RNA occupancy reveal principles of RNA-chromatin interactions. Mol. Cell 2011, 44, 667–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Prensner, J.R.; Chinnaiyan, A.M. The emergence of lncRNAs in cancer biology. Cancer Discov. 2011, 1, 391–407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Batista, P.J.; Chang, H.Y. Long noncoding RNAs: Cellular address codes in development and disease. Cell 2013, 152, 1298–1307. [Google Scholar] [CrossRef] [Green Version]
  86. Lee, J.T. Epigenetic regulation by long noncoding RNAs. Science 2012, 338, 1435–1439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Schmitz, S.U.; Grote, P.; Herrmann, B.G. Mechanisms of long noncoding RNA function in development and disease. Cell. Mol. Life Sci. 2016, 73, 2491–2509. [Google Scholar] [CrossRef] [Green Version]
  88. Moran, V.A.; Perera, R.J.; Khalil, A.M. Emerging functional and mechanistic paradigms of mammalian long non-coding RNAs. Nucleic Acids Res. 2012, 40, 6391–6400. [Google Scholar] [CrossRef]
  89. Salmena, L.; Poliseno, L.; Tay, Y.; Kats, L.; Pandolfi, P.P. A ceRNA hypothesis: The Rosetta Stone of a hidden RNA language? Cell 2011, 146, 353–358. [Google Scholar] [CrossRef] [Green Version]
  90. Huang, X.M.; Shi, S.S.; Jian, T.M.; Tang, D.R.; Wu, T.; Sun, F.Y. LncRNA PVT1 knockdown affects proliferation and apoptosis of uveal melanoma cells by inhibiting EZH2. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 2880–2887. [Google Scholar] [CrossRef]
  91. Villanueva, M.T. Anticancer drugs: All roads lead to EZH2 inhibition. Nat. Rev. Drug Discov. 2017, 16, 239. [Google Scholar] [CrossRef] [PubMed]
  92. Su, M.; Xiao, Y.; Tang, J.; Wu, J.; Ma, J.; Tian, B.; Zhou, Y.; Wang, H.; Yang, D.; Liao, Q.J.; et al. Role of lncRNA and EZH2 Interaction/Regulatory Network in Lung Cancer. J. Cancer 2018, 9, 4156–4165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Jiang, T.; Wang, Y.; Zhou, F.; Gao, G.; Ren, S.; Zhou, C. Prognostic value of high EZH2 expression in patients with different types of cancer: A systematic review with meta-analysis. Oncotarget 2016, 7, 4584–4597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Lu, Q.; Zhao, N.; Zha, G.; Wang, H.; Tong, Q.; Xin, S. LncRNA HOXA11-AS Exerts Oncogenic Functions by Repressing p21 and miR-124 in Uveal Melanoma. DNA Cell Biol. 2017, 36, 837–844. [Google Scholar] [CrossRef] [PubMed]
  95. Wu, X.; Yuan, Y.; Ma, R.; Xu, B.; Zhang, R. lncRNA SNHG7 affects malignant tumor behaviors through downregulation of EZH2 in uveal melanoma cell lines. Oncol. Lett. 2020, 19, 1505–1515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Xu, H.; Gong, J.; Liu, H. High expression of lncRNA PVT1 independently predicts poor overall survival in patients with primary uveal melanoma. PLoS ONE 2017, 12, e0189675. [Google Scholar] [CrossRef] [Green Version]
  97. Zheng, X.; Tang, H.; Zhao, X.; Sun, Y.; Jiang, Y.; Liu, Y. Long non-coding RNA FTH1P3 facilitates uveal melanoma cell growth and invasion through miR-224-5p. PLoS ONE 2017, 12, e0184746. [Google Scholar] [CrossRef] [Green Version]
  98. Guan, Y.; Kuo, W.L.; Stilwell, J.L.; Takano, H.; Lapuk, A.V.; Fridlyand, J.; Mao, J.H.; Yu, M.; Miller, M.A.; Santos, J.L.; et al. Amplification of PVT1 contributes to the pathophysiology of ovarian and breast cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2007, 13, 5745–5755. [Google Scholar] [CrossRef] [Green Version]
  99. Iden, M.; Fye, S.; Li, K.; Chowdhury, T.; Ramchandran, R.; Rader, J.S. The lncRNA PVT1 Contributes to the Cervical Cancer Phenotype and Associates with Poor Patient Prognosis. PLoS ONE 2016, 11, e0156274. [Google Scholar] [CrossRef] [Green Version]
  100. Xing, Y.; Wen, X.; Ding, X.; Fan, J.; Chai, P.; Jia, R.; Ge, S.; Qian, G.; Zhang, H.; Fan, X. CANT1 lncRNA Triggers Efficient Therapeutic Efficacy by Correcting Aberrant lncing Cascade in Malignant Uveal Melanoma. Mol. Ther. J. Am. Soc. Gene Ther. 2017, 25, 1209–1221. [Google Scholar] [CrossRef]
  101. Li, F.; Li, X.; Qiao, L.; Liu, W.; Xu, C.; Wang, X. MALAT1 regulates miR-34a expression in melanoma cells. Cell Death Dis. 2019, 10, 389. [Google Scholar] [CrossRef] [PubMed]
  102. Sun, Y.; Cheng, H.; Wang, G.; Yu, G.; Zhang, D.; Wang, Y.; Fan, W.; Yang, W. Deregulation of miR-183 promotes melanoma development via lncRNA MALAT1 regulation and ITGB1 signal activation. Oncotarget 2017, 8, 3509–3518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Sun, L.; Sun, P.; Zhou, Q.Y.; Gao, X.; Han, Q. Long noncoding RNA MALAT1 promotes uveal melanoma cell growth and invasion by silencing of miR-140. Am. J. Transl. Res. 2016, 8, 3939–3946. [Google Scholar] [PubMed]
  104. Wu, S.; Chen, H.; Zuo, L.; Jiang, H.; Yan, H. Suppression of long noncoding RNA MALAT1 inhibits the development of uveal melanoma via microRNA-608-mediated inhibition of HOXC4. Am. J. Physiol. Cell Physiol. 2020, 318, C903–C912. [Google Scholar] [CrossRef] [PubMed]
  105. Di Sanzo, M.; Aversa, I.; Santamaria, G.; Gagliardi, M.; Panebianco, M.; Biamonte, F.; Zolea, F.; Faniello, M.C.; Cuda, G.; Costanzo, F. FTH1P3, a Novel H-Ferritin Pseudogene Transcriptionally Active, Is Ubiquitously Expressed and Regulated during Cell Differentiation. PLoS ONE 2016, 11, e0151359. [Google Scholar] [CrossRef]
  106. Pan, H.; Ni, H.; Zhang, L.; Xing, Y.; Fan, J.; Li, P.; Li, T.; Jia, R.; Ge, S.; Zhang, H.; et al. P2RX7-V3 is a novel oncogene that promotes tumorigenesis in uveal melanoma. Tumour Biol. J. Int. Soc. Oncodevelopmental Biol. Med. 2016, 37, 13533–13543. [Google Scholar] [CrossRef]
  107. Lu, L.; Yu, X.; Zhang, L.; Ding, X.; Pan, H.; Wen, X.; Xu, S.; Xing, Y.; Fan, J.; Ge, S.; et al. The Long Non-Coding RNA RHPN1-AS1 Promotes Uveal Melanoma Progression. Int. J. Mol. Sci. 2017, 18, 226. [Google Scholar] [CrossRef] [Green Version]
  108. Ding, X.; Wang, X.; Lin, M.; Xing, Y.; Ge, S.; Jia, R.; Zhang, H.; Fan, X.; Li, J. PAUPAR lncRNA suppresses tumourigenesis by H3K4 demethylation in uveal melanoma. FEBS Lett. 2016, 590, 1729–1738. [Google Scholar] [CrossRef]
  109. Moriyama, M.; Osawa, M.; Mak, S.S.; Ohtsuka, T.; Yamamoto, N.; Han, H.; Delmas, V.; Kageyama, R.; Beermann, F.; Larue, L.; et al. Notch signaling via Hes1 transcription factor maintains survival of melanoblasts and melanocyte stem cells. J. Cell Biol. 2006, 173, 333–339. [Google Scholar] [CrossRef]
  110. Asnaghi, L.; Ebrahimi, K.B.; Schreck, K.C.; Bar, E.E.; Coonfield, M.L.; Bell, W.R.; Handa, J.; Merbs, S.L.; Harbour, J.W.; Eberhart, C.G. Notch signaling promotes growth and invasion in uveal melanoma. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2012, 18, 654–665. [Google Scholar] [CrossRef] [Green Version]
  111. Cheng, G.; He, J.; Zhang, L.; Ge, S.; Zhang, H.; Fan, X. HIC1 modulates uveal melanoma progression by activating lncRNA-numb. Tumour Biol. J. Int. Soc. Oncodevelopmental Biol. Med. 2016, 37, 12779–12789. [Google Scholar] [CrossRef] [PubMed]
  112. Wu, X.; Li, X.F.; Wu, Q.; Ma, R.Q.; Qian, J.; Zhang, R. LncRNA SNHG15 predicts poor prognosis in uveal melanoma and its potential pathways. Int. J. Ophthalmol. 2020, 13, 1195–1201. [Google Scholar] [CrossRef] [PubMed]
  113. Zhou, Y.; Zhang, L.; Fan, J.; Jia, R.; Song, X.; Xu, X.; Dai, L.; Zhuang, A.; Ge, S.; Fan, X. Let-7b overexpression leads to increased radiosensitivity of uveal melanoma cells. Melanoma Res. 2015, 25, 119–126. [Google Scholar] [CrossRef] [PubMed]
  114. Rossi, E.; Pagliara, M.M.; Orteschi, D.; Dosa, T.; Sammarco, M.G.; Caputo, C.G.; Petrone, G.; Rindi, G.; Zollino, M.; Blasi, M.A.; et al. Pembrolizumab as first-line treatment for metastatic uveal melanoma. Cancer Immunol. Immunother. 2019, 68, 1179–1185. [Google Scholar] [CrossRef] [Green Version]
  115. Audrito, V.; Serra, S.; Stingi, A.; Orso, F.; Gaudino, F.; Bologna, C.; Neri, F.; Garaffo, G.; Nassini, R.; Baroni, G.; et al. PD-L1 up-regulation in melanoma increases disease aggressiveness and is mediated through miR-17-5p. Oncotarget 2017, 8, 15894–15911. [Google Scholar] [CrossRef] [Green Version]
  116. Li, Q.; Johnston, N.; Zheng, X.; Wang, H.; Zhang, X.; Gao, D.; Min, W. miR-28 modulates exhaustive differentiation of T cells through silencing programmed cell death-1 and regulating cytokine secretion. Oncotarget 2016, 7, 53735–53750. [Google Scholar] [CrossRef] [Green Version]
  117. Sharma, A.; Biswas, A.; Liu, H.; Sen, S.; Paruchuri, A.; Katsonis, P.; Lichtarge, O.; Chand Dakal, T.; Maulik, U.; Gromiha, M.M.; et al. Mutational Landscape of the BAP1 Locus Reveals an Intrinsic Control to Regulate the miRNA Network and the Binding of Protein Complexes in Uveal Melanoma. Cancers 2019, 11, 1600. [Google Scholar] [CrossRef] [Green Version]
  118. Bande Rodriguez, M.F.; Fernandez Marta, B.; Lago Baameiro, N.; Santiago-Varela, M.; Silva-Rodriguez, P.; Blanco-Teijeiro, M.J.; Pardo Perez, M.; Pineiro Ces, A. Blood Biomarkers of Uveal Melanoma: Current Perspectives. Clin. Ophthalmol. 2020, 14, 157–169. [Google Scholar] [CrossRef] [Green Version]
  119. Telonis, A.G.; Rigoutsos, I. Race Disparities in the Contribution of miRNA Isoforms and tRNA-Derived Fragments to Triple-Negative Breast Cancer. Cancer Res. 2018, 78, 1140–1154. [Google Scholar] [CrossRef] [Green Version]
  120. Telonis, A.G.; Loher, P.; Jing, Y.; Londin, E.; Rigoutsos, I. Beyond the one-locus-one-miRNA paradigm: MicroRNA isoforms enable deeper insights into breast cancer heterogeneity. Nucleic Acids Res. 2015, 43, 9158–9175. [Google Scholar] [CrossRef] [Green Version]
  121. Londin, E.; Magee, R.; Shields, C.L.; Lally, S.E.; Sato, T.; Rigoutsos, I. IsomiRs and tRNA-derived fragments are associated with metastasis and patient survival in uveal melanoma. Pigment Cell Melanoma Res. 2020, 33, 52–62. [Google Scholar] [CrossRef] [PubMed]
Figure 1. ncRNA categories. RNAs are divided into two major classes: messenger RNA (mRNA) and non-coding RNA (ncRNA).
Figure 1. ncRNA categories. RNAs are divided into two major classes: messenger RNA (mRNA) and non-coding RNA (ncRNA).
Cancers 12 02944 g001
Figure 2. The impact of ncRNAs-mediated PI3K/AKT and MAPK/ERK pathway in uveal melanoma (UM). The ncRNAs that have a suppressive role are colored green, while the oncogenic ncRNAs are colored red.
Figure 2. The impact of ncRNAs-mediated PI3K/AKT and MAPK/ERK pathway in uveal melanoma (UM). The ncRNAs that have a suppressive role are colored green, while the oncogenic ncRNAs are colored red.
Cancers 12 02944 g002
Table 1. miRNAs with UM tumor suppressor action. A list of known miRNAs with tumor suppression role in UM is shown indicating their chromosomal location, cell/tissue where it was described, their inhibitory action, gene target, existence of experimental validation, and reference. A summary indicating the main characteristics of the UM cell lines mentioned in this table are shown in Table S1 [14,15,16,17,18,19].
Table 1. miRNAs with UM tumor suppressor action. A list of known miRNAs with tumor suppression role in UM is shown indicating their chromosomal location, cell/tissue where it was described, their inhibitory action, gene target, existence of experimental validation, and reference. A summary indicating the main characteristics of the UM cell lines mentioned in this table are shown in Table S1 [14,15,16,17,18,19].
miRNALocation Sample TypeInhibitory EventsValidationTargetReference
miR-91q22Highly invasive cell lines: MUM2B and C918Cell migration and invasionMA, IA, PA, RT-PCR, WB, FRANF-kB1[20]
miR-17-3p13q31.3UM patient tumor tissues; UM cell lines: OCM-1A, MUM-2C, C918, and MUM-2BCell proliferation and metastasisRT-PCR, PA, MA, IA, Dual LUCA, WB, ATMMDM2[21]
miR-23a19p13.12UM cell lines: OCM-1; UM patient tumor tissuesCell migrationRT-PCR, MA
IF, WB
Zeb1[22]
miR-34a1p36.22UM cell lines: M17, M21, M23, and SP6.5; UM patient tumor tissuesCell proliferation and migrationPA, MA, IA, Dual LUCA,
IF, WB
C-MET
LGR4
[23,24]
miR-34b/c11q23.1UM cell lines: SP6.5; UM patient tumor tissues (upregulated in doxorubicin- and epigenetic drug-treated SP6.5 cells)Cell proliferation and migrationPA, AA, MA, LUCA, WBC-MET[25]
miR-12218q21.31UM cell lines: 92.1, MEL270, OMM2.5, UPMM2, and UPMM3; TCGA UM dataset, UM patient tumor tissuesCell proliferation and migrationRT-PCR, LUCA, WBC-MET, ADAM10[26]
miR-124a3 loci:
8p23.1
8q12.3
20q13.33
UM cell lines: M17, M21, M23, and SP6.5; UM patient tumor tissuesCell proliferation, migration, and invasionPA, MA, IA, RT-PCR, ATM, WBCDK4, CDK6,
cyclin D2, EZH2
[27]
miR-1371p21.3UM cell lines: OMM1.3, Mel202, 92.1, OMM1, OCM1, and OCM3Cell proliferationRT-PCR, PA, LUCA, WBMITF
CDK6
[28]
miR-142-3p17q22UM cell lines: M17 and SP6.5; UM patient tumor tissuesCell proliferation, migration, and invasionRT-PCR, PA, MA, IA, AA, LUCA, WBCDC25C, TGFβR1, GNAQ, WASL, RAC1[29]
miR-14417q11.2UM cell lines: MUM-2B, C918, MUM-2C, and OCM-1A; UM patient tumor tissuesCell proliferation and migrationLUCA, WB, PA, IAC-MET[30]
miR-1455q32UM cell lines: MUM-2B and OCM-1; UM patient tumor tissuesCell proliferation and promotes apoptosisRT-PCR, PA, IA, Dual LUCA, WBIRS-1
NPR1
[31]
miR-1827q32.2UM cell lines: M23 and SP6.5; UM patient tumor tissues;(upregulated by p53 activation)Cell proliferation, migration, and invasionRT-PCR, PA, MA, LUCA, ATM, WBMITF, BCL2, cyclin D2[32]
miR-2051q32.2UM patient tumor tissues (both high- and low invasive UM)Cell proliferation and promotes apoptosisPA, IA, Dual LUCA, WBNPR1[31]
miR-216a-5p2p16.1UM cell line: MUM-2B; UM patient tumor tissuesCell proliferationPA, RT-PCR,
LUCA, GA, WB, ATM, IHC
HK2[33]
miR-224-5pXq28UM cell line: OCM-1A; UM patient tumor tissuesCell proliferation, migration, and invasionRT-PCR, PA, MA, IA, LUCA, WBPIK3R3/AKT3,
FTH1P3, RAC1, Fizzled 5
[34]
miR-15a, miR-185, miR-21113q14.2 22q11.21 15q13.3Cell line: OCM-1; UM patient tumor tissuesCell proliferationRT-PCR, PA, WBIL-10Rα[35]
Abbreviations: RT-PCR: real time polymerase chain reaction; PA: proliferation in vitro assay; MA: migration in vitro assay; IA: invasion in vitro assay, AA: apoptosis assay; WB: western blot immunodetection; LUCA: luciferase expression assays; IF: Immunofluorescence; IHC: Immunohistochemistry; ATM: animal tumor models, GA: glycolysis assay. Information obtained from GeneCards database (www.genecards.org).
Table 2. miRNAs with oncogenic action. A description of known miRNAs with oncogenic role in UM is shown indicating their chromosomal location, cell/tissue where it was described, their tumor promoting effect, gene target, existence of experimental validation, and reference. A summary indicating the main characteristics of the UM cell lines mentioned in this table are shown in Table S1 [14,15,16,17,18,19].
Table 2. miRNAs with oncogenic action. A description of known miRNAs with oncogenic role in UM is shown indicating their chromosomal location, cell/tissue where it was described, their tumor promoting effect, gene target, existence of experimental validation, and reference. A summary indicating the main characteristics of the UM cell lines mentioned in this table are shown in Table S1 [14,15,16,17,18,19].
miRNALocation Sample TypePro-Tumorigenic EventsValidationTargetReference
miR-20a13q31.3MUM-2B and MUM-2C metastatic UM cells; UM patient tumor tissuesCell proliferation, migration, and invasionPA, IA, MA, RT-PCRNot validated[36]
miR-2117q23.1OCM-1, M619 and MUM-2B UM cellsCell proliferation, migration, and invasionPA, AA, MA, Dual LUCA, WB p53[37]
miR-92a-3p2 loci:
13q31.3
Xq26.2
OCM-1 UM cellsInhibition of apoptosisLUCA, WB MYCBP2[38]
miR-15521q21.3OCM-1A, MUM-2C, C918, and MUM-2B UM cells; UM patient tumor tissuesCell proliferation and invasion PA, IA, LUCA, WBNDFIP1[39]
miR-181 (family)Chr 1, 9, and 19SP6.5, VUP, OCM1, MUM2b, and 92-1; UM patient tumor tissuesCell proliferationDual LUCA, WB, RT-PCRCTDSPL[40]
miR-222Xp11.3C918 and MUM-2B UM cellsCell proliferation and migrationAA, PA, MA, WBPI3K/Akt/MMP-9[41]
miR-3674q25UM cell lines: M17, M23, MUM-2B, and C918; UM patient tumor tissuesCell proliferation and migrationLUCA, PA, MA, WBPTEN[42]
miR-45417q22UM cell lines: OCM-1A, MUM-2B, MUM-2C, and C918; UM patients tumor tissuesCell proliferation, colony formation, and invasionRT-PCR, PA, IA, LUCA, WBPTEN[43]
miR-652Xq23MUM-2B and MEL270 UM cell lines; UM patient tumor tissuesCell proliferation and migration PA, MA, Dual LUCA, RT-PCR, WBHOXA9[44]
Abbreviations: RT-PCR: real time polymerase chain reaction; MA: in vitro cell migration assay; IA: in vitro invasion assay; PA: proliferation in vitro assay; WB: western blot immunodetection; LUCA: luciferase expression assay; AA: apoptosis assay. Information obtained from GeneCards database (www.genecards.org).
Table 3. miRNAs in association with an increased risk of uveal melanoma metastasis.
Table 3. miRNAs in association with an increased risk of uveal melanoma metastasis.
Sample TypeOver ExpressedUnder ExpressedReference
High Risk UM MetastaticLow Risk UM MetastaticHigh Risk UM Metastatic
24 UM patient frozen tumor tissueslet-7b, miR-143, miR-193b, miR-199a, miR-199 *, miR-652 [61]
60 UM patient tumor tissues and 11 control tissuesmiR-33a, miR-99a, miR-135b, miR-196a, miR-325, miR-497, miR-512-5p, miR-549, miR-556-5p, miR-585, miR-640, miR-885-5p let-7e, miR-1, miR-10a, miR-18a, miR-19b-1, miR- 26a-2, miR-34c-5p, miR-129, miR-133a, miR-154, miR-181a, miR-218, miR-369-3p, miR-377, miR-376c, miR-493, miR-495, miR-586 †† [66]
26 UM patient tumor tissuesmiR-16-5p, miR-17-5p, miR-21-5p, miR-132-5p, miR-151a-3p let-7c-5p, miR-99a-3p, miR-99a-5p, miR-101-3p, miR-181a-2-3p, miR-181b-5p, miR-378d , miR-1537-3p[64]
50 UM patient tumor tissuesmiR-346, miR-592, miR-1247 miR-506 and miR-513c[65]
86 UM patient tumor tissuesmiR-134, miR-143, miR-146b, miR-149, miR-199a, miR-214 [62]
TCGA UM databaseLet-7b, miR-21, miR-29b, miR-142, miR-143, miR-144, miR-145, miR-150, miR-155, miR-199a, miR-199b, miR-330, miR-486 miR-30c, miR-181a-5p, miR-335, miR-140,
Xq27.3 group
[63]
TCGA UM databasemiR-195, miR-224, miR-365a, miR-365b, miR-452, miR-4709, miR-7702 miR-513c, miR-873 [68]
TCGA UM databaselet-7b(5p/3p), miR-199a-5p, miR-224-5p, miR-452-5p, miR-592 miR-211-5p,
miR-508-3p, miR-509–3-5p, miR-513a-5p, miR-513c-5p, miR-514a-3p
[67]
The repeated miRNAs appear in bold. Genomic Location (Chromosome 3); †† Genomic Location (Chromosome 6); Genomic Location (Chromosome 8). * Information obtained from GeneCards database (www.genecards.org).
Table 4. Circulating MicroRNAs as potential biomarkers in UM.
Table 4. Circulating MicroRNAs as potential biomarkers in UM.
Sample TypemiRNALevelsDifferentiate (Ratio)
Vitreous humor and serum (± exosomes)Vitreous humor (+ exosomes): miR-21, miR-34a, miR-146a
Serum (+ exosomes): miR-146a
Primary/Control[73]
Serum (14 patients with UM vs. 14 controls)miR-146aPrimary/Control[71]
miR-19a, miR-30d , miR-127, miR-451, miR-518f, miR-1274b
Plasma (33 UM M3/22 UM D3/26 controls)miR-92b, miR-199a-5p, miR-223M3/D3
M3/Control
[74]
Plasma (6 patients with UM vs. 26 controls)miR-20a, miR-125b, miR-146a, miR-155, miR-181a, miR-223Primary/Control[72]
miR-20a, miR-125b, miR-146a, miR-155, miR-223 Metastatic/Primary
miR-181a
Blood (10 uveal nevus, 50 localized UM, 5 metastatic UM)miR-16, miR-145, miR-146a, miR-204, miR-211, miR-363-3pPrimary/Metastatic
Primary/Nevus
[70]
miR-211Metastatic/Primary
miR-204
The repeated miRNAs appear in bold. Genomic Location (Chromosome 8). No miRNA in this table is located in Chr 3 or Chr 6. ↑ Overexpressed; ↓ Underexpressed. Information obtained from GeneCards database (www.genecards.org).

Share and Cite

MDPI and ACS Style

Bande, M.; Fernandez-Diaz, D.; Fernandez-Marta, B.; Rodriguez-Vidal, C.; Lago-Baameiro, N.; Silva-Rodríguez, P.; Paniagua, L.; Blanco-Teijeiro, M.J.; Pardo, M.; Piñeiro, A. The Role of Non-Coding RNAs in Uveal Melanoma. Cancers 2020, 12, 2944. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers12102944

AMA Style

Bande M, Fernandez-Diaz D, Fernandez-Marta B, Rodriguez-Vidal C, Lago-Baameiro N, Silva-Rodríguez P, Paniagua L, Blanco-Teijeiro MJ, Pardo M, Piñeiro A. The Role of Non-Coding RNAs in Uveal Melanoma. Cancers. 2020; 12(10):2944. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers12102944

Chicago/Turabian Style

Bande, Manuel, Daniel Fernandez-Diaz, Beatriz Fernandez-Marta, Cristina Rodriguez-Vidal, Nerea Lago-Baameiro, Paula Silva-Rodríguez, Laura Paniagua, María José Blanco-Teijeiro, María Pardo, and Antonio Piñeiro. 2020. "The Role of Non-Coding RNAs in Uveal Melanoma" Cancers 12, no. 10: 2944. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers12102944

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