Next Article in Journal
Adipokine-Cytokine Profile in Patients with Unstable Atherosclerotic Plaques and Abdominal Obesity
Next Article in Special Issue
Determination of Common microRNA Biomarker Candidates in Stage IV Melanoma Patients and a Human Melanoma Cell Line: A Potential Anti-Melanoma Agent Screening Model
Previous Article in Journal
Early Clinical Outcomes of the First Commercialized Human Autologous Ex Vivo Cultivated Oral Mucosal Epithelial Cell Transplantation for Limbal Stem Cell Deficiency: Two Case Reports and Literature Review
Previous Article in Special Issue
Interplay between LncRNAs and microRNAs in Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of the Cell Cycle by ncRNAs Affects the Efficiency of CDK4/6 Inhibition

Department of Breast and Thyroid Surgery, Union Hospital, Tongji Medical College, Huazhong University of Science and Technology, Wuhan 430030, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(10), 8939; https://doi.org/10.3390/ijms24108939
Submission received: 20 March 2023 / Revised: 12 May 2023 / Accepted: 15 May 2023 / Published: 18 May 2023

Abstract

:
Cyclin-dependent kinases (CDKs) regulate cell division at multiple levels. Aberrant proliferation induced by abnormal cell cycle is a hallmark of cancer. Over the past few decades, several drugs that inhibit CDK activity have been created to stop the development of cancer cells. The third generation of selective CDK4/6 inhibition has proceeded into clinical trials for a range of cancers and is quickly becoming the backbone of contemporary cancer therapy. Non-coding RNAs, or ncRNAs, do not encode proteins. Many studies have demonstrated the involvement of ncRNAs in the regulation of the cell cycle and their abnormal expression in cancer. By interacting with important cell cycle regulators, preclinical studies have demonstrated that ncRNAs may decrease or increase the treatment outcome of CDK4/6 inhibition. As a result, cell cycle-associated ncRNAs may act as predictors of CDK4/6 inhibition efficacy and perhaps present novel candidates for tumor therapy and diagnosis.

1. Introduction

In human cancers, abnormal cell cycles are often identified. The beginning and progression of the cell cycle are controlled by cyclin-dependent kinases (CDKs), cyclin, and cyclin-dependent kinase inhibitors (CKIs). The activated CDK4/6-cyclin D’s complex phosphorylates RB (pRB) upregulates the E2F’s transcription factor and promotes the expression of cell division-related target genes [1]. The entire cell cycle is monitored by checkpoints [2,3]. Checkpoints halt or delay the cell cycle when there is DNA damage, replication defects, or aberrant mitosis, ensuring the structural integrity of chromosomes. The G1/S checkpoint, which is the primary regulator of the cell cycle and essential for the onset and advancement of malignant tumors, is regulated by CDK4/6 [4]. Theoretically, it would be predicted that many human cancers will show increased CDK4/6 reliance and make ideal candidates for CDK4/6 inhibitor treatment since they often activate CDK4/6 genes or have transcriptional aberrations through a variety of mechanisms [5,6].
In some solid tumors, pharmacological inhibition of CDK4/6 has demonstrated substantial effectiveness, mostly by preventing Rb phosphorylation and causing G1 cell cycle arrest in tumor cells [7]. From laboratory investigations to clinical trials, these drugs have advanced swiftly. The Food and Drug Administration (FDA) has approved three medications (palbociclib (PD0332991) [8], ribociblib (LEE011) [9], and abemaciclib (LY2835219) [10]) for the treatment of metastatic breast cancer that specifically target CDK4/6. Although CDK4/6 inhibitions are still in the preclinical and early clinical stages, an increasing number of mechanisms of resistance to CDK4/6 inhibitions have been reported.
Non-coding RNAs (ncRNAs) are RNA molecules that do not have the function of coding proteins [11]. The human genome contains 93% DNA that may be transcribed into RNA, of which only 2% encodes proteins as mRNA and the remaining 98% are known as ncRNAs [12]. Non-coding RNA can be divided into different categories roughly based on their size [13]. Small ncRNAs important in cancer include microRNAs (miRNAs), transfer RNAs (tRNAs), and PIWI-interacting RNAs (piRNAs). At the other end of the size spectrum are long non-coding RNAs (lncRNAs), which are characterized by untranslated RNAs more than 200 nucleotides in length and include subclasses such as pseudogenes and circRNAs. MicroRNAs are short ncRNAs about 22 nucleotides long; the expression of other RNAs can be regulated by binding the 5′ end of miRNAs to complementary sequences in target RNA. The roles of tsRNA and piRNA in somatic tissues and cancers, as well as the exact mechanisms of their own biogenesis, remain to be elucidated. These small ncRNAs that are abnormally expressed in cancer may also be useful biomarkers. Compared with small ncRNA, lncRNA exhibits a wide range of mechanism diversity and more functions. CircRNAs are single-stranded, covalently closed RNA molecules. Although circRNA has been widely found in cancer, most mechanisms of action and function have yet to be further investigated. While ncRNAs lack the ability to encode proteins, they can influence the expression of other genes through a variety of mechanisms. NcRNAs regulate a variety of biological activities, including cell differentiation, proliferation, apoptosis, and migration [14]. During the development and progress of cancer, ncRNAs’ expression is frequently dysregulated [15,16]. In actuality, ncRNAs also regulate the abundance of several cell cycle regulators, which affect the cell cycle and proliferation. Moreover, cell cycle-associated mechanisms also control the production of ncRNAs.
Because ncRNAs are significant participants in the cell cycle, we hypothesize that they may also play a role in the clinical response to CDK4/6 inhibition. In this article, we systematically analyze how the two most significant ncRNA types, lncRNA and miRNA, regulate the cell cycle and their effects on CDK4/6 inhibition.

2. CDKs and Cell Cycle

There are four stages in the cell cycle [17,18,19]: G1, S (DNA synthesis), G2, and M (mitosis). CDKs are an enzyme family that catalyzes protein phosphorylation during cell cycle progression and transcriptional control. CDK catalytic activity is dependent on cyclin interaction, and only particular CDK-cyclin interactions are expected to regulate the cell cycle process. Each phase of the cell cycle must be initiated under the control of cyclin, CDKs, and related pathways (Figure 1). RB suppresses cell cycle progression and proliferation by binding to and inhibiting the activity of E2F transcription factors [20,21,22]. At particular cell cycle stages, CDKs bind to and activate particular cyclins, which leads to the phosphorylation of target proteins and maintains cell cycle progression. It is possible to divide CDKs into two groups [23,24,25]: transcription-related CDKs (CDK7, CDK8, CDK9, CDK12, and CDK13), which directly control cell cycle progression, and cell cycle-related CDKs (CDK1, CDK2, CDK4, and CDK6), which are involved in the cell cycle.
CDK4/6 is thought to be essential for initiating the G1/S phase of the cell cycle. Growth factors, for example, often stimulate the MAPK pathway in the G1 phase and subsequently transcript genes encoding cyclin Ds, which bind and activate CDK4 and CDK6 [26]. The kinase activity of CDK4/6-cyclin Ds phosphorylates RB, and phosphorylation causes RB to separate from E2Fs [21,23]. To start transcription programs, E2F activates genes needed for the S phase and G2/M phase processes. At the late G1 phase, cyclin D1-CDK4/6 activity is reduced, whereas cyclin E-CDK2 activity rises. Cyclin E-CDK2 promotes the G1/S transition by phosphorylating RB and proteins involved in DNA replication, as well as many other important molecules involved in the initiation and progression of the S phase [27]. Cyclin E degrades during the early S phase, while the cyclin A-CDK2 complex accelerates the cell cycle process from the S to the G2 phase. CDK2 activity decreases throughout the middle G2 phase. The cyclin B/CDK1 complex phosphorylates APC/cyclosome and other target proteins to encourage G2 phase maturation and participate in M-phase activities [28]. After cytoplasmic division is completed in late M, cyclin B is destroyed, and the next cycle changes or enters a resting state. The progression of the cell cycle is monitored by numerous checkpoints, which interrupt or postpone the cell cycle in the event of DNA damage, replication faults, or aberrant mitosis.
The precise and timely regulation of CDK at each phase of the cell cycle also entails the negative regulation of several anti-proliferative signals, such as the expression of cyclin-dependent protein kinase inhibitors (CKIs). The N terminal of CKIs shares a common domain known as the kinase inhibition domain (KID) that works to inhibit CDKs. CKIs can specifically bind to distinct cyclin/CDK complexes, impede their activity, and impair the cell cycle. At present, there are two main CKIs, INK4 and CIP/Kip family [29,30,31].
The INK family comprises four proteins, p16INK4a [32], p15INK4b [33], p18INK4c [34], and p19INK4d [35], which compete with cyclin Ds to bind to CDK4/6, creating inactive heterodimers of the INK4 protein and CDK4/6 [36,37,38]. For instance, p16 and p15 contain ankyrin repeats to bind to CDK4 and CDK6, respectively. The INK family regulates the molecular basis of proliferation, differentiation, starvation, cell death, aging, and the development of acute and chronic diseases [39,40,41].
The Cip/Kip family contains p21Cip1/Waf1 (encoded by CDKN1A), p27Kip (encoded by CDKN1B), and p57 Kip2 (encoded by CDKN1C) [42,43,44,45]. The Cip/Kip protein, which is related or unrelated to CDK, plays a crucial role in gene transcription, apoptosis, autophagy, senescence, cytoskeletal remodeling, and cell mobility, and hence influences cell differentiation, development, organ biogenesis, and the occurrence and progression of cancer [46,47]. P21, for example, plays a significant role in the p53-dependent G1 block [42,48]. P21 binds to and inhibits cyclin/CDK activity, resulting in cell cycle growth arrest, and it also plays an environmentally dependent role in apoptosis [49,50].
In addition to INK4 and Cip/Kip, some kinases act as cell cycle checkpoints by negatively regulating CDKs. Wee1/2, also known as the Wee1A and Wee1B, work as DNA damage and cell volume checkpoints, and can inhibit CDK1 activity and cause G2/M arresting [51,52]. During DNA damage, cell cycle arrest induced by Wee1/2 is conducive to repairing damaged DNA and maintaining chromatin integrity.

3. LncRNAs and Cell Cycle

3.1. LncRNA Profile

Long non-coding RNAs (lncRNAs), which make up 80–90% of all ncRNAs, are defined as ncRNAs length 200 nucleotides—10 kb. The primary transcriptional byproducts of the human body are lncRNAs. The human genome’s intergenic regions include the majority of the DNA regions that code for lncRNAs. LncRNAs can also originate from sense or antisense transcripts with specific sequences that overlap protein-coding genes. LncRNAs are transcribed by RNA polymerase II and resemble mRNA structurally in many ways. The sequences of lncRNAs typically have a 3′-terminal polyA tail, a 5′-terminal 7-methyluridine (m7G) cap, and a promoter, but they lack an open reading frame (ORF) and thus have no protein-coding ability [53,54]. LncRNAs can be divided into five categories according to their transcription positions and directions: sense, antisense, bidirectional, intronic, and intergenic lncRNAs [55,56].
Although they do not directly encode proteins, lncRNAs play their biological functions by participating in epigenetics, transcription, post-transcriptional regulation, and other gene expression regulation at different levels, as well as various molecular mechanisms. LncRNA expression is abnormal in a variety of diseases [57,58,59]. For example, the lncRNA expression profiles of many tumor cells show significant changes compared with those of normal cells. Different lncRNAs have different functions; some are tumor suppressor genes, and some are tumor-promoting genes. LncRNAs are abnormally expressed in cancer, and the expression level can predict patients’ response to treatment and affect patients’ recurrence, metastasis, and prognosis [60,61,62].
LncRNAs influence downstream gene expression by cis and trans epigenetic mechanisms such as DNA methylation, chromatin remodeling, and histone modification. For example, the lncRNA XIST inactivates the X chromosome, which results in the cis-accumulation-mediated silencing regulation of a single X chromosome in XX female mammals [63]. LncRNAs can interact directly or indirectly with DNA and chromatin proteins to control chromatin structure. LncRNAs can function as ligands and bind to transcription factors to create complexes that regulate the activity of genes during transcription. Some lncRNAs are transcription factors [64,65]. LncRNA transcription will interfere with surrounding coding gene transcription, or lncRNA transcripts will influence the expression of neighboring coding genes. The DNA sequences of lncRNAs can interact directly with other DNA regions, changing their structure and potentially activating or inhibiting gene expression. LncRNAs can directly participate in the post-transcriptional control of mRNA via variable splicing, RNA editing, protein translation, and transport [60,66,67,68]. Antisense lncRNAs are primarily involved in the post-transcriptional control of coding genes [69,70]. Antisense lncRNAs can interact with mRNA complementary areas, influence the recruitment of splinters by some splicing sites, regulate the mRNA splicing process, and also influence the RNA editing process. LncRNAs can interact with proteins to help or inhibit the development of organelle structures, which can impact translation efficiency [71]. Furthermore, mRNA and lncRNA interactions can influence mRNA nuclear transport and intracellular distribution [72,73]. LncRNAs can also attach to miRNAs and stop them from binding to their target genes [74].

3.2. Cell Cycle Regulated by LncRNAs

LncRNAs can influence cell cycle progression via epigenetic regulation. LINC-PINT [75,76] recruits EZH2 to the promoters of target genes (CDK1, cyclin A2, AURKA, and PCNA), inducing trimethylation of H3K27 and epigenetic silencing of target genes, finally limiting the formation of melanoma (Figure 2A). Yang et al. [77] reported that lncRNA-HEIH is overexpressed in HBV-associated hepatocellular carcinoma. It interacts with EZH2 to negatively control the expression of CKIs p15, p16, p21, and p57. Hence, it plays an important role in the G0/G1 block (Figure 2B). According to reports, ANRIL [78] is linked to senescence and proliferation, as well as to the epigenetic regulation of INK4 transcription (Figure 2C). Yap et al. [79] showed that the PRC2 and ANRIL complex was involved in the regulation of cellular senescence by mediating the methylation of histone H3K27 at INK4 via EZH2. Kotake et al. [80] reported that ANRIL depletion causes human fibroblast growth arrest and increases age-related beta-galactosidase expression.
Cyclin and CDKs are key regulators of the cell cycle and are usually regulated by lncRNAs [81]. CDK1 expression can be positively regulated by lncRNA TMPO antisense transcripts (TMPO-AS) [82]. Downregulation of TMPO-AS1 reduced lung cancer cells’ viability, increased cyclin D1 and BCL expression, and triggered apoptosis. TMPO-AS1 is expected to provide a novel therapeutic target for lung cancer patients. A CDK1-related lncRNA is LINC00261 [83]. LINC00261, CDK1, and CXCL8 have been found in studies to have a mutual regulatory connection, impacting tumor angiogenesis, G2/M phase transition, and EMT. They promote tumor growth and impair immunotherapy efficacy. LINC00630 increases E2F1 binding to the CDK2 promoter region, stimulates CDK2 transcription, and thereby accelerates the malignant development of liver cancer [84]. LINC00630 is a promising molecular target for liver cancer treatment. ARAP1-AS1 stimulated cyclin D1 expression but did not affect CDK4 and CDK6 expression [85]. ARAP1-AS1 knockdown dramatically reduced lung cancer cell proliferation and clonogenicity and caused G0/G1 cell cycle arrest via reducing cyclin D1 expression.
Some lncRNAs have the potential to influence the expression of CDKs and cyclins at the same time. ALMS1-IT1 can activate the CDK pathway, increasing the proliferative potential of lung cancer cell lines and promoting malignant development [86]. LncRNA ENST00000512916 was shown to be upregulated in ameloblastoma (AB) tissues [87]. It elevates CDK2/4/6 expression in AB cells, enhancing cell cycle progression and proliferation.
CKIs’ expression is essential for controlling the cell cycle, and certain lncRNA can control both CKIs’ and CDKs’ expression simultaneously to improve cell cycle control. Maternally expressed gene 3 (Meg3) expression levels are significantly lower in non-small cell lung cancer (NSCLC) [88]. Meg3 facilitates miR-3163’s interaction with the 3′-UTR of the Skp2 mRNA, which prevents Skp2 from being translated. Skp2, a component of the E3 ubiquitin ligase SCF, particularly encourages p27’s destruction by ubiquitination. TPT1-AS1 (tumor protein translation control 1 antisense RNA1) has been linked to the development of cervical cancer, ovarian cancer, gastric cancer, and other cancers [89,90,91,92]. According to studies, TPT1-AS1 negatively regulates the expression of p21 while positively regulating the expression of CDK4 and cyclin D1. TPT1-AS1 is crucial for encouraging cell growth and controlling the G1/S transition in the cell cycle. It is possible to create more effective molecular therapy targets for TPT1-AS1. The lncRNA ABHD11-AS1 was found to be overexpressed in gastric cancer, ovarian cancer, bladder cancer, and endometrial cancer. According to research [93], the lncRNA ABHD11-AS1 upregulates Cyclin D1, CDK1, CDK2, CDK4, and Bcl-xl, downregulates p15, and increases cell proliferation and the G1/S transition while inhibiting cell death in endometrial cancer. The relationship between variations in lncRNA expression and the prevalence of triple-negative breast cancer has been discussed in previous studies. The lncRNA ZFAS1 was dramatically downregulated in breast cancer patients’ blood samples and reduced tumor cell proliferation and metastasis by positively influencing the expression levels of CDK inhibitors p21 and p27 [94]. CADM1-AS1 (cell adhesion molecule 1-antisense RNA 1, long non-coding RNA) suppresses AKT and GSK3b phosphorylation in HCC cells [95]. By inhibiting the AKT/GSK3b signaling pathway, CADM1-AS1 increases the expression of p15, p21, and p27 while decreasing the expression of cycle-related proteins such as cyclin Ds, cyclin Es, CDK2, CDK4, and CDK6, which prevents the transition from G0/G1 to S phase both in vitro and in vivo.
DNA damage is a crucial event that triggers checkpoints to stop or slow the cell cycle. The expression of several lncRNAs changes in response to DNA damage, controlling the expression of genes involved in the cell cycle. Gadd7 is a lncRNA that controls CDK6 expression post-transcriptionally [96]. Gadd7 controls CDK6 expression negatively and serves as a translational regulatory factor. Gadd7 participates in the G1 checkpoint and downregulates the cyclin D1/CDK6 complex in cisplatin-mediated DNA damage, preventing the cell cycle from responding to DNA damage [97]. More research is necessary because this might be a brand-new G1 checkpoint cascade.
Some studies have discovered that while lncRNAs control the cell cycle, they also have impacts on the expression of particular miRNAs. XIST increased PC cell line proliferation by being positively correlated with CDK1 and negatively correlated with miR-140 and miR-124 [98]. XIST knockdown caused cell cycle stalling in the G1 phase, lowered p53 and CDK1 protein levels, and elevated p21 protein levels, limiting PC cell line proliferation. XIST can be exploited as a therapeutic target for human pancreatic cancer. In addition, lncRNA and miRNA can work together to control the expression of specific genes. The level of the lncRNA CCDC144NL-AS1 was markedly increased in liver cancer tissues [99]. Through the adsorption of miR-940 via sponging, CCDC144NL-AS1 stimulates the production of WDR5, and since CDK1, CKD2, and CDK4 are all downstream targets of WDR5, this promotes the development of liver cancer. In general, lncRNAs regulate cell cycle progression in a variety of ways (Figure 3 and Table 1).

4. MiRNAs and Cell Cycle

4.1. MiRNA Profile

MicroRNAs (miRNAs) are small, single-stranded ncRNAs that have a length of 20–24 nt. MiRNAs are abundant in eukaryotes and can make up between 1 and 4% of the genes in humans [102].
Among the small ncRNA species, miRNAs are by far the most extensively studied in cancer, compared to tsRNA and piRNA. MiRNA has been found to be altered in all cancer types studied, and the alteration of miRNA has important effects on the molecular and cellular processes of cancer [103,104,105].
MiRNAs are transcribed by RNA polymerase I and RNA polymerase II, with RNA polymerase II being the most common. Long, double-stranded primary miRNAs (pri-miRNAs) are the origins of miRNAs. Pri-miRNA is detected in the nucleus by DGCR8 (Drosha chaperone) and cleaved by RNase iii (Drosha) to form a miRNA precursor (pre-miRNA) [106]. The pre-miRNA is subsequently processed in the cytoplasm by the RNase iii endonuclease Dicer to generate mature miRNA molecules, which are generally 22 nucleotides long [107,108]. One strand of RNA is destroyed, and another is loaded into the AGO protein from the 5 or 3 ends, generating a miRNA-induced silencing complex (miRISC) [109]. The name of mature miRNA is 5p or 3p, depending on the directionality of the miRNA strand [110,111].
Through base-pairing with the 3′-UTR or CDS (coding DNA sequence) of mRNA, miRNA can silence a gene by causing the degradation of the target mRNA or preventing the target mRNA from being translated [112,113]. A single miRNA molecule can target many mRNAs, and multiple miRNAs can target the same mRNA, resulting in a complex network of miRNA–mRNA interactions [107,114]. Up to 30% of human protein-coding genes are under the control of miRNAs. Many biological processes, such as cell division, proliferation, apoptosis, and migration, are regulated by miRNAs. Moreover, they affect the body’s metabolism, growth, stem cell regeneration, and the emergence of cancer [115,116].
Compared to normal tissues, the expression of many miRNAs is dysregulated in tumors, indicating that miRNA plays a crucial role in the development and propagation of malignancies [105,117]. MiRNAs have been shown to have antitumor effects; for instance, miR-15 and miR-16 are downregulated in B-cell chronic lymphoma [118], miR-122 is downregulated in liver cancer cell lines [119,120], and miR-125 is downregulated in breast cancer [121]. MiRNAs can work as oncogenes in vivo. MiR-23a/24-2/27a clusters, on the other hand, enhance the invasion of breast cancer cells and liver metastases, whereas miR-10b promotes angiogenesis and metastasis in breast cancer [122,123]. Certain miRNA levels in the circulatory system can also reflect the malignancy of tumors, making them more useful detection markers for preoperative diagnosis and recurrence monitoring of patients [124,125]. MiRNAs can be employed as prognostic predictors in various malignancies. MiRNAs can be employed as therapeutic targets in the treatment of cancer.

4.2. Cell Cycle Regulated by MiRNAs

Some studies have revealed that miRNAs can influence the levels of a wide range of cell cycle regulators (Figure 4 and Table 2). Cell cycle regulators are precisely regulated by miRNAs at the translational and post-translational levels in addition to transcriptional levels, having an impact on a variety of signaling pathways and cellular activities. In mammalian cell cycles, miRNAs are expected to play a significant role and may even regulate cell division and proliferation.
MiRNAs may regulate G1/S transitions by targeting important cell cycle regulators such as CDK1, CDK2, CDK4, and CDK6, as well as cyclin D1, D2, and E. MiR-195 binds to the 3′-UTR of CDK6 mRNA to negatively regulate CDK6 expression in gastric cancer. MiR-195 targets and suppresses cyclin D3 in non-small cell lung cancer (NSCLC), resulting in cell cycle arrest in the G1 phase [126]. The miR-124a promoter is substantially methylated and downregulated in ALL, leading to an increase in CDK6 expression and the phosphorylation of RB [149]. Promoter methylation and p53 activity control the expression of miR-125a and miR-125b in liver cancer [152,153]. MiR-125a and miR-125b suppress cyclin D1 and cause p21-dependent cell cycle arrest in the G1 phase. In several cancers, miR-17-5p prevents cyclin D1 expression and causes G1-phase arrest [144,167,168,169]. MiR-449b inhibits the expression of cyclin D1 and E2F3 by binding to their 3′-UTRs, which reduces the growth of colon cancer stem cells and has tumor-suppressive effects [184]. CDK1 is a functional target of miR-31-5p that plays an important role in cell cycle control [185,186]. MiR-212 expression specifically targets cyclin D3 and promotes apoptosis in ATL cells, causing an increase in G0/G1 cells and a decrease in S-phase cells [215]. In melanoma cells, miR-206 causes G1 arrest by inhibiting the translation of CDK4, cyclin D1, and cyclin C [211,212]. MiR-365 acts as an intermediary, downregulating cyclin D1 and cdc25A via lowering p53 abundance [130,131,216]. MiR-15a [132] and miR-16 [217,218] are part of the mir-15a-16-1 cluster [133,134,135], which downregulates CDK1, CDK2, and CDK6 as well as cyclin D1, cyclin D3, and cyclin E1 to induce cycle stagnation in the G1 phase. MiR-26a significantly decreased the expression of EZH2 and repressed the expression of c-Myc, cyclin D3, cyclin E2, and CDK4/6 in an EZH2-dependent manner [156,157,158]. Furthermore, miR-26a can directly upregulate the expression of p14 and p21 and downregulate the expression of cyclin D2 without the need for EZH2. It is essential for halting the cell cycle’s development and reducing cell proliferation. MiR-7 [136,140,141], miR-19a [142,219], miR-20a [144,145,170], and miR-34 [147,220] all downregulate cyclin D1 protein levels. Moreover, miR-34 can upregulate p21, which inhibits CDK and blocks DNA synthesis, causing cell cycle arrest [50,148]. MiR-34a causes G1-phase arrest, reduces cell development, and promotes cell death in HepG2 cells by suppressing the production of cyclin D1, CDK4, and CDK6 [221]. Prostate [222], breast [223], ovarian [223], bladder [224], stomach [159], and laryngeal squamous cell carcinomas all show low expression of miR-101 (LSCC) [160]. MiR-101 increases the expression of the tumor suppressors p14, p16, p21, and p27 while suppressing the expression of c-Myc, CDK2, CDK4, CDK6, CDK8, cyclin D2, cyclin D3, and cyclin E2. MiR-9 can suppress the proliferation of malignant tumor cells by downregulating CDK6 and cyclin D1, causing cell cycle arrest in the G0/G1 phase [161]. By specifically suppressing the CDK6/cyclin D1/p21/Waf1 pathway, miR-218 can play a significant role in limiting the growth of glioma cells [164,165]. These miRNAs exhibit anti-proliferative properties, function as cancer tumor suppressors, and are frequently repressed in malignancies in a variety of ways. Nc886 [200], a regulatory ncRNA of intermediate length (101 nucleotides), is regarded as a precursor miRNA due to its length. DNA methylation in the promoter of nc886 hinders the expression of this gene in malignant tumors. In contrast to miRNAs and nuclear lncRNAs, which attach to complementary paired RNA or DNA targets. Nc886 binds to target proteins to influence their activity, which in turn regulates the expression of genes. It has been demonstrated that nc886 silencing results in decreased expression of intracellular CDK inhibitors such as CDKN2A and CDKN2C and increased expression of CDK4.
The E2F factor, which is necessary for the G1 phase to transition into the S phase, is released when the RB is phosphorylated by the CDK/cyclin complex. The E2F family is essential for cell cycle and apoptosis regulation. The RB protein family strictly regulates their activity, but transcription, post-translational modification, and protein stability also play a role in controlling how these factors are activated. MiRNAs can target E2F and control both its expression and function. First discovered by O’Donnell et al. [171], c-Myc-regulated miRNAs (miR-17-5p and miR-20a) can control the expression of E2F1 [170]. Several miRNAs, including miR-125b [154], miR-210 [166], miR-195 [127], miR-17-5p, and miR-20a [170] have been found to negatively regulate the expression of E2Fs. The miR-17-92 cluster modulates the translation of E2Fs mRNA [171]. It was discovered that proliferative cyclin D1, E2F1, and anti-proliferative p21, PTEN, RB1, RBL1 (p107), and RBL2 (p130) are common targets of miR-17, miR-20a, and miR-106b [172]. These miRNAs alter USSC cell cycle progression because they not only directly target proteins that control the cell cycle but also encourage cell division and control G1/S transitions by boosting the intracellular activity of E2F transcription factors. In prostate and stomach cancer, miR-330 and miR-331-3p inhibit E2F1 activity, causing cell cycle arrest [181,182].
Between E2F and miRNA, there is also automatic feedback control (Figure 5). E2F1 has been shown to directly transcribe miR-17-92 and miR-106b-25, which in turn induces their expression [225,226,227]. As a result, the expression of E2F1 is suppressed. MiR-17-92 and miR-106b-25 both target the 3′UTR of E2F1. The miR-17-92 cluster member miR-20a controls translation by interacting with the 3′-UTR region of the E2F2 and E2F3 mRNAs [146]. Moreover, endogenous E2F1-3 can bind directly to the miR-17-92 cluster’s promoter and trigger transcription, indicating the possibility of an auto-regulated feedback loop between the miR-17-92 cluster and E2Fs. The automated regulation of E2F1-3 and miR-20a is crucial for mediating cell proliferation and death and can effectively avoid the aberrant aggregation of E2Fs. In conjunction with E2F, these miRNAs create a negative feedback loop that gives organisms homeostatic security. Furthermore, E2F1-induced miRNAs such as miR-449a and miR-449b can negatively control RB phosphorylation by specifically targeting and inhibiting CDK6, thereby suppressing E2F1 activity and cell cycle progression [183,228]. This is an indirect way of describing how E2F1-activated miR-449 negatively regulates E2F1 activity.
Moreover, miRNAs contribute to the later phases of the cell cycle (Figure 4 and Table 2). Following the G1/S transition, complexes of CDK1 and cyclin A or cyclin B are primarily responsible for controlling the rest of the cell cycle. Let-7 is a negative regulator of the expression of several cyclins, including CDK4/6, cyclin D, cyclin A, and cyclin B [136,137,138,139]. The expression of cyclin A or cyclin B can be inhibited by miR-125b and miR-24 [154,229]. Some miRNAs may be engaged in different cell cycle phases and can control the expression of nearly all CDKs. Hydbring et al. [230] studied a wide range of cancer cells and discovered that a class of independent miRNAs (including miR-193b, miR-193a-3p, miR-195-5p, miR-214-5p, and miR890) targeted almost all cyclins and CDKs. These miRNAs significantly reduce cancer cell proliferation.
Although they report less frequently, miRNAs can also control the mitotic process. Polo-like kinase (PLK1) is a crucial mitosis regulator [231]. PLK1 phosphorylates CDC25C, which activates the CDK1/cyclin B1 complex and causes it to translocate into the nucleus to initiate mitosis in the cell cycle. PLK1 expression is negatively regulated by miR-100, which targets PLK1 mRNA [232]. Moreover, paclitaxel therapy markedly decreases miR-100 level and enhances mRNA stability and transcription of β-tubulin I, IIA, IIB, and V, all of which are attributed to G2/M block [128]. MiR-100 targets and inhibits MTMR3, further suppressing p27, which is critical for the proliferation of breast cancer cells [129]. Downregulation of MiR-100 can activate p27, causing G2/M cell cycle arrest and apoptosis as a result. In osteosarcoma, miR-223 can act as a tumor suppressor by upregulating the expression of p21 and p27, increasing RB phosphorylation, and causing G1 phase arrest [213]. In breast cancer, miR-223 elevates p27 expression. On the one hand, p27 directly improves the stability of mature miR-223 [214]. On the other hand, p27 improves E2F1’s control of miR-223’s promoter activity. In addition, p27 increases miR-223 expression in two ways. The mutual control of p27 and miR-223 is important for cell cycle arrest and the abolishment of contact inhibition. More research is still needed to determine how functionally important miRNAs are in controlling mitosis.
The INK4 and Cip/Kip families are also regulated by miRNAs, either directly or indirectly (Figure 4 and Table 2). P16 is regulated by miR-31 [187,188]. MiR-106b and the miR-17-92 cluster (including six miRNA members: miR-17, miR-18a, miR-19a, miR-20a, miR-19b, and miR-92a-1) directly target p21 [178,179,180,233,234]. MiR-24 targets and inhibits p21, p16, and p27 [155,235,236]. In small-cell lung cancer (NSCLC) cells A549 and H1298, miR-512-5p targets and suppresses p21, which causes apoptosis and reduces glycolysis [189]. MiR-370 inhibits p27 and p21 while increasing cyclin D1 to promote the G1/S phase transition [190,237]. MiR-370 plays a crucial role in the malignant progression of prostate, lung, and bladder cancers. MiR-212 enhances p21 and p27 expression by decreasing retinoblastoma binding protein 2 (RBP2), which delays the G1/S phase transition and indirectly decreases proliferation [191,192]. MiR-196a suppresses cell cycle progression by inhibiting p27 expression and cyclin E/CDK2 activity [193,238]. Skp2’s 3′-UTR mRNA binds to miR-3163, which prevents Skp2 from being translated [194]. Skp2, a part of the E3 ubiquitin ligase SCF, especially encourages the ubiquitination degradation of the p27 and supports the development of cancer cells in various cancers. By inhibiting Skp2, miR-3163 increases the stability of p27 and slows the development of cancer cells. MiR-221/222 inhibits p27 and p57 and activates CDK2, both of which are factors in tumor progression [195,196,197,198,239]. MiR-222 increases p57 expression and accelerates the G1 to S phase transition [199]. MiR-222 also increases Capan-2 pancreatic cancer cells’ ability to survive and proliferate. A novel approach to the treatment of pancreatic cancer may involve targeting miR-222.
In the aging model, miR-17, miR-19b, miR-106a, and miR-20a were also found to be related to the regulation of the transcription level of target genes, especially p21 [143]. These miRNAs may serve as new markers of human aging. In INK4/ARF locations, polycomb repressive complexes (PRC1 and PRC2) are epigenetic regulators that control aging. MiR-9 suppresses the expression of CBX7, a component of PRC1. In turn, CBX7 suppresses miR-9, resulting in a negative feedback loop [162,163]. The miR-9/CBX7 feedback loop regulates aging by inducing the production of p16. MiR-124a can bind to the 3′-UTR region of MCP1′ (monocyte chemoattractant protein 1) and CDK2′ mRNAs and specifically inhibit their activity [150,151]. MiR-124a significantly inhibits CDK2 expression in synovial cells of patients with rheumatoid arthritis (RA), thereby inhibiting cell proliferation and arresting the cell cycle in the G1 phase. The stability of miR-17, miR-20-5p, and miR-106-5p is reduced quickly throughout the process of oxidative stress, limiting DNA synthesis and blocking the G1/S transition [177]. Overexpression of miR-20b-5p/miR-106a-5p directly inhibits p21, cyclin D1, and E2F1, which can reverse growth retardation by increasing DNA synthesis and G1/S transition.
In addition, various cell cycle conditions may alter miRNA expression and stability. The first ncRNA to be reported to be regulated by cyclin is miR-17/20, which is bound to cyclin D1 [240]. Cyclin D1 controls miRNA maturation. The RNase III endoribonuclease Dicer plays a crucial regulatory role in the development of miRNA by cleaving long-stranded RNA or pre-miRNA with stem ring structure into mature miRNA. According to studies [241], cyclin D1 regulates the generation of miRNAs by promoting the expression of the Dicer both in vivo and in vitro using targeted transcription mechanisms. Cyclin D1 can control both the processing of many miRNAs and the expression of individual miRNAs by binding to regulatory areas and stimulating the transcription of the Dicer enzyme.

5. NcRNAs and CDKI Treatment

5.1. CDKIs and Cancer Treatment

CDK inhibitors were developed to impede the development of cancer cells. Currently, there are three generations. Although they have excellent anticancer efficacy in preclinical and clinical studies, generation 1 and 2 CDKIs have not received much therapeutic attention in the treatment of cancer patients due to their low selectivity and severe toxicity [242]. By selectively targeting CDK4 and CDK6, the third generation of CDKIs has achieved outstanding success in clinical applications, significantly reducing the development of several malignant tumors (Table S1). Nowadays, drugs that precisely target CDK4/6 activity are used extensively in cancer therapy. Due to their efficacy in treating breast cancer, more than ten CDK inhibitors are presently undergoing clinical trials (Table S1).
Oral CDK inhibitors with specific chemical structures that bind to CDK4 and CDK6 include palbociclib [243,244], ribociclib [245,246], and abemaciclib [247]. They cause G1-phase arrest and inhibit RB phosphorylation [7]. Palbociclib is the first CDK4/6-cyclin Ds inhibitors FDA-approved for postmenopausal HR-positive and HER2-negative metastatic breast cancer [243]. The anticancer effects of palbociclib have been reported in neuroblastoma [248], non-small cell lung cancer [249], and breast cancer [250]. Palbociclib suppresses RB phosphorylation, downregulates E2F, causes G1-phase cell cycle arrest, and inhibits tumor growth, according to in vitro and in vivo studies [251,252]. Palbociclib can also make medulloblastoma cells more responsive to chemotherapy and ionizing radiation therapy [253]. In RB-positive cancer cells, ribociclib causes the dephosphorylation of pRB, which results in G1-phase arrest [248]. In RB-positive tumors such as ER-positive breast cancer and neuroblastoma, ribociclib’s antitumor effects have been examined [248,254]. Ribociclib dramatically reduces cell proliferation and enhanced apoptosis in ER-a breast cancer cell lines [255]. In both vitro and vivo experiments, abemaciclib improves the effects of chemotherapeutic medicines on cells overexpressing the ATP Binding Cassette Subfamily B Member 1 (ABCB1) and ATP Binding Cassette Subfamily G Member 2 (ABCG2) [256]. In breast cancer and other cancers, its anticancer properties have been studied. Ribociblib and abemaciclib have also been approved for the combination endocrine therapy of advanced HR-positive and HER2-negative breast cancer in phase III clinical trials [247,257,258,259]. Currently, in the preclinical or early clinical stages, CDK4/6 inhibitors are promising medications for the treatment of malignant cancers. In comparison to endocrine therapy alone, the combination utility of CDK4/6 inhibitors and endocrine therapy in the first-line treatment of advanced breast cancer prolonged PFS in phase III clinical trial MONARCH 3 [260,261,262].
CDKIs have been known to cause acquired resistance through several mechanisms [263,264]. Approximately 10% of patients will initially develop resistance to CDK4/6 inhibition, and a growing number will eventually lose their effectiveness [264]. To pinpoint non-responders to CDK4/6 inhibition and to individualize therapy, new biomarkers are required. Cell cycle regulators, including p16, CDK6, CCNE1/2, CDK2, CDK7, and E2F, are crucial in developing resistance to CDK4/6 inhibition [255,265,266]. The CDK inhibition palbociclib becomes partially resistant when pRB is inhibited [267,268,269]. P16 gene-deficient animals show acquired resistance to CDK4/6 drugs in preclinical studies [270]. Patients with high P16 expression and decreased RB expression typically developed resistance to CDK4/6 inhibitors and were ineligible for CDK4/6 inhibitor clinical trials [271]. Acquired resistance to CDK4/6 inhibitors may be brought on by acquired RB mutations, cyclin E amplification, CDK6 amplification, or inhibition of CDK2 inhibitors (such as p27 or p21) [271,272,273,274,275].
In addition, CDKIs tend to induce dose-related neutropenia with rare neutropenic fever. Ribociclib is associated with prolonged QT and therefore requires ECG monitoring. Abemaciclib is more likely to cause gastrointestinal toxicity, which manifests as diarrhea. Abemaciclib has a higher incidence of fatigue [276]. The main adverse event in palbociclib is neutropenia, and other adverse events include fatigue, diarrhea, nausea, dyspnea, and joint pain [277].

5.2. LncRNA and CDKI Clinical Efficiency

There are few reports and only preclinical stage studies on the reduced or enhanced effect of lncRNA on CDKIs (Table 3). LncRNA TROJAN is significantly expressed in TNBC and promotes the transfer potential of TNBC, which has detrimental effects [278]. TROJAN is associated with a decreased disease-free survival rate in patients with ER+ breast cancer [279]. Moreover, TROJAN can accelerate the growth and development of ER+ breast cancer by controlling the cell cycle during G1/S conversion. Downregulation of TROJAN causes ER+ breast cancer cells to be arrested in G1, resulting in a decrease in cyclin E1/2 and an increase in p21 and p27. TROJAN can bind to NKRF, an NF-B pathway repressor, blocking it from interacting with RELA, an NF-B pathway transcriptional activator. The TROJAN-NKRF complex can control the transcriptional frequency of CDK2 expression and prevent NKRF from binding to the CDK2 promoter, making tumor cells resistant to CDKIs [255,280]. Increased effectiveness of CDK4/6 inhibitors can be achieved by inhibiting TROJAN, which can also limit CDK2 expression and prevent non-classical CDK2-mediated entry into the S phase. Combining TROJAN inhibition with CDK4/6 inhibitors has been shown to dramatically enhance the anticancer activity of CDK4/6 inhibitors [100]. When treated with CDK4/6 inhibitors, CDK2 activation is a crucial bypass mechanism for cell cycle advancement [280].
LncRNA SNHG15 performs a cancer-promoting role by encouraging proliferation and metastasis in glioblastoma (GBM), breast cancer, lung cancer, and liver cancer [286,287,288]. SNHG15 increased the overall survival of HCC and has a favorable correlation with TNM staging [288]. Increased SNHG15 is linked to a significantly higher risk of GBM and poorer overall survival in GBM patients [289,290]. SNHG15 activates CDK6 by suppressing the tumor suppressor miR-627-5p. In tumor-bearing mice, palbociclib significantly increased the expression of miR-627-5p while decreasing the expression of SNHG15 and CDK6 [101].
LncRNA maternally expressed gene 3 (MEG3, mouse homolog Gtl2) has been shown to act as a tumor suppressor in a variety of human cancer cell lines [291,292,293,294]. By directly interacting with chromatin, MGE3 inhibits the TGF-β pathway or activates p53 to inhibit cell proliferation [292,295,296]. The CDK4/6 inhibitor palbociclib activates pRB in lung cancer cells (A549 and SK-MS-1), and pRB particularly encourages MEG3 expression [281]. This demonstrates that palbociclib or other CDKIs can be used to target a range of malignancies that depend on the activation of the MEG3 regulatory pathway. The decline in cell proliferation caused by palbociclib can be partially restored by MEG3 expression suppression. Palbociclib’s therapeutic effectiveness may therefore be reduced by the tumor cells’ low MEG3 expression. Palbociclib, a CKD4/6 inhibitor, increases MEG3 expression in tumor cells, indicating that it may be effective against tumor types whose growth is regulated by MEG3. Furthermore, CDK4/6 inhibitors may be used in combination with other medications now on the market. For example, lower MEG3 expression is associated with cisplatin resistance in NSCLC [297]. MEG3 high expression caused by CDK4/6 inhibitors may be beneficial for platinum-based therapy. More studies are required to determine whether palbociclib and cisplatin are effective when combined.

5.3. MiRNA and CDKI Clinical Efficiency

MiRNAs can affect CDKIs negatively or positively because of the intricate relationships between them and cell cycle progression. MiRNAs are a valuable alternative to or combination therapy for CDK4/6 CDK4/6 inhibitors and can be used as biomarkers to predict the therapeutic efficacy of CDK4/6 CDK4/6 inhibitors. MiR-497 inhibits the growth of anaplastic large cell lymphoma (ALCL) by negatively regulating the expression of the cell cycle regulatory genes CCNE1, CDC25A, CDK6, and E2F3 [204,205]. NPM-ALK (+) ALCL cells are susceptible to palbociclib treatment in a CDK6-dependent manner. Cell cycle arrest in the G1/S phase is caused by Nc886’s downregulation of CDK4 expression and upregulation of CKIs (such as p16 and p18) [200,201]. Nc886-silenced cells have increased G1 to S phase activity, rapid cell division, and are more sensitive to palbociclib. Hence, in cancer patients, nc886 can be utilized as a predictor of medication reactivity. In prostate cancer, miR-193b suppresses cyclin D1 expression [206,207,208,209]. Overexpression of miR-193b in prostate cancer cells prevents cells from progressing from the S phase to the G2/M phase. In contrast, prostate cancer cells with high miR-193b expression become resistant to CDK4/6 inhibitors. Reduced expression of miR-193b in prostate cancer cells results in upregulation of cyclin D1 and greater susceptibility to palbociclib therapy. Oral squamous cell carcinoma (OSCC) has decreased miR-9 expression [161]. It was discovered that miR-9 negatively regulates the expression of CDK6 and cyclin D1 and that miR-9 overexpression inhibits cell proliferation, reduces the capacity of colonies to form, and causes cell cycle arrest in the G0/G1 phase. MiR-9’s ability to inhibit cell proliferation in OSCC may serve as a theoretical foundation for the creation of small compounds that target CDK4/6.
Because CDKIs have achieved positive outcomes in the treatment of breast cancer, many research studies have shown the impact of miRNA on the sensitivity of CDKIs in breast cancer. MiR-3613-3p stops TNBC cells from proliferating, which causes a cell cycle arrest in the G0/G1 phase [282]. MiR-3613-3p increases senescence, which increases the susceptibility of breast cancer cells to palbociclib. Palbociclib-sensitive breast cancer cells produce miR-29b-3p at higher levels, but if the miR-29b-3p expression is decreased, the cells may become resistant to treatment [202,203]. Particularly in the luminal subtype of malignant breast cancers, miR-223 expression is downregulated [283]. Lower levels of miR-223 are related to a worse prognosis and reduced overall survival. In luminal breast cancer cells, downregulation of miR-223 expression is associated with palbociclib resistance. In addition, palbociclib resistance is boosted by suppressing miR-223 in HER2-positive breast cancer cells. In contrast to CDK4/6 inhibitor-sensitive cancers, miR-432-5p expression is higher in ER-positive tumors with resistance [285]. MiR-432-5p increases CDK6 expression and is attributed to palbociclib or ribociclib resistance. Loss of CDK6 causes resistant cells to regain their sensitivity to palbociclib.
MiRNAs influence the sensitivity of CDK4/6 inhibitors in addition to other tumor cell processes such as metastasis and metabolism, which may suggest novel medication combinations. By inhibiting CDK6 and creating cell cycle stasis in the G1 phase, miR-29b-3p inhibits proliferation via limiting cell motility and the epithelial–mesenchymal transition (EMT) [202,203]. In metastatic melanoma, miR-200a is down-expressed, which negatively regulates CDK6 expression and slows down G1/S checkpoint cell cycle progression [210]. In metastatic melanomas, high miR-200a and low CDK6 expression are associated with palbociclib resistance, whereas low miR-200a and high CDK6 expression make metastatic melanomas more responsive to CDK4/6 inhibitors. By blocking CDK4/6, miR-126 controls glycolysis and the advancement of the M phase, and it has a clear anti-proliferative effect on breast cancer cells [284]. In luminal and HER2-positive breast cancer cells, the combination of overexpressed miR-126 and the CDK4/6 inhibition ribociclib exhibits potent antitumor effects that are far more effective than those of either treatment alone. MiR-126 dramatically increases the susceptibility of HER2-positive breast cancer cell lines to ribociclib.
The response of tumor cells to different CDK4/6 inhibitors may be influenced by miRNAs (Table 3). A single transcript from the miR17HG gene on human chromosome 13 is processed into the miR-17-92 cluster [173,298]. A total of six mature miRNAs are present in the Mir-17-92 cluster (miR-17, miR-18a, miR-19a, miR-20a, miR-19b, and miR-92). Atypical teratoid rhabdomyoma (ATRT) has been reported to respond therapeutically to palbociclib when miR-19a, miR-17, and miR-20a block cyclin D1 expression. Both ATRT cell lines and mice with ATRT tumors show greater sensitivity to CDK4/6 inhibitors. Moreover, it was discovered that the miR-17-92 family is upregulated in glioblastoma neuropathy (GBM), increasing the sensitivity of GBM stem-like cell lines to palbociclib and ribociclib. MiR-17-19b significantly shortens the time required for the emergence of mixed lymphocytic lymphoma (MLL) by suppressing the expression of p21 [174,175]. By regulating p21 expression, miRNAs influence leukemia stem cell (LSC) self-renewal. MiR-19a or miR-19b controls cell cycle progression and inhibits Wee1 in leukemia [176]. In the G2/M phase of head and neck squamous cell carcinoma (HNSCC) cells, MiR-17-92 works as a transcription factor. MiR-17-92 negatively influences the mRNA and protein expression of cyclin G2 by targeting its 3′-UTR [298].
MiRNAs are thought to play a significant role in the effectiveness of CDK4/6 inhibitors and are crucial in the control of cancer-related pathways. To help doctors make treatment decisions, miRNA expression profiles may be employed as predictive biomarkers for the effectiveness of CDKIs.

6. Perspective and Conclusions

NcRNAs participate in different phases of the cell cycle by focusing on specific signaling pathways and related elements. NcRNAs can be valuable choices for therapeutic drugs or therapeutic targets since they can target many altered mRNAs in illness circumstances. Unrestricted proliferation brought on by an unchecked cell cycle is a hallmark of cancer [299,300,301]. CDKs play a vital role in regulating the cell cycle, so inhibitors that specifically target different CDKs have been developed to stop the proliferation of cancer cells. We postulate that because ncRNAs and the cell cycle are so closely related, dysregulated ncRNAs may provide strategies for increasing CDKIs’ therapeutic benefits in cancer. However, further investigation is required to comprehend the importance of ncRNAs in the cell cycle and the therapeutic function of CDKIs as well as to outline suitable methods for enhancing the clinical use of CDKIs.
In preclinical and clinical research, a large number of CDK4/6 inhibitors have been found to limit the growth of many malignancies, particularly breast cancer. Clinical evidence from research on the condition suggests that CDK4/6 inhibitors may be effective as a single therapy for advanced breast cancer. Moreover, CDK4/6 inhibitors can be combined with chemotherapy or endocrine treatments to enhance their effectiveness [243,247,257,258]. Even though taking CDKIs has been linked to some adverse effects, including neutropenia, this is primarily due to dose restrictions. New CDKIs with lower toxicity should be investigated, or CDKIs with different adverse features should be created to lessen side effects such as myelosuppression. More importantly, alternative molecular processes continue to drive the cell cycle in the absence of the CDK mechanism, leading to tumor resistance to CDK4/6 inhibitors and long-term growth. Thus, consideration must be given to the design of inhibitors that target various cyclin/CDK domains.
It is still unclear exactly how ncRNAs control the cell cycle, since the regulation of the cell cycle by ncRNAs is more complex. Certain ncRNAs influence proliferation not just by targeting a single cell cycle regulator but also by targeting both positive and negative cell cycle regulators, such as members of the mir-17-92 cluster [219,233]. For instance, miR-369-3p inhibits the proliferation of liver cancer cells by positively regulating the production of p21 and negatively regulating the expression of cyclin D1 in liver cancer [302]. MiR-369-3p suppresses the stability of p53 in ovarian granulosa cells and decreases apoptosis in ovarian granulosa cells (OGCs) [303]. Moreover, the stability of miRNAs may fluctuate during a particular cell cycle, and miRNA regulation may be cell cycle-dependent. Depending on the stage in the cell cycle, miRNAs may be able to either up or downregulate translation. For example, members of the miR-369-3p or let-7 family may promote translation in stationary cells and suppress translation during cell cycle/proliferation [304].
Because the links between ncRNAs and mRNAs or proteins constitute a complicated network, it is vital to investigate the interaction between ncRNAs and targets in specific settings. Many variations exist in different studies because of differences in tumor types, separation procedures, detection methods, and sample processing. It is possible that different types of cells do not exhibit the same kinds of interactions between ncRNAs and their targets. As a result, it is necessary to discuss the relationship between ncRNAs and their targets in the context of certain cellular circumstances. Many ncRNA targets are still unknown due to the lack of strong evidence-based ways to prove the direct targeting and control of ncRNAs. Proteomics is a valuable method for discovering whole-cell target genes and examining the influence of miRNAs on proteins. Most existing links between ncRNAs and the cell cycle remain undiscovered [305,306].
According to some preclinical research, ncRNAs influence how cancer cells respond to CDK4/6 inhibitors [207]. By modifying the expression of downstream targets (CDK4/6 and cyclin Ds) or by controlling associated signaling pathways, some ncRNAs can make cancer cells resistant to CDK4/6. Some ncRNAs suppress the production of cyclin Ds or interfere with the expression of cell cycle regulatory genes, making cancer cells more sensitive to CDK4/6 inhibitors. Further consideration should focus on the function of ncRNAs in CDK4/6 inhibition management. In addition to changing ncRNA expression and function, DNA damage can alter the cell cycle components, and eventually slow down the cell cycle process. The DNA damage response is associated with some lncRNAs [281,307]. At the DNA damage checkpoint, lncRNAs cause CDK inactivation, cell cycle arrest, or apoptosis, mediating the non-classical pathway of the DNA damage response [96,97]. DNA damage is the primary mechanism by which most platinum-based chemotherapy medicines treat tumors. More research is required to determine if changes in intracellular ncRNAs brought on by DNA damage treatments would increase or lessen the effects of CDKI therapy. The degree of ncRNA expression also affects the susceptibility of cancer cells to chemotherapeutic treatments [308,309,310]. As a result, the combination of CDK4/6 inhibitors and other active therapeutic medications may be effective. Significant differences in ncRNA expression patterns in cancer cells are related to an enhanced or decreased response to CDK4/6 inhibitors [311,312]. However, there is currently insufficient clinical evidence to support this connection.
Direct targeting of ncRNAs or altering intracellular levels of ncRNAs may both be effective therapeutic approaches. Although several ncRNA mimics and chemically modified anti-ncRNAs have been designed, delivering these compounds to the target location and establishing their effective concentrations remain challenging problems [313,314,315,316]. Several vectors have been created, such as cell-specific absorption of oligonucleotides based on lipids, polymers, metal nanoparticles, molecular peptides, etc. The therapeutic applicability of these vectors, however, requires more study. Challenges with whole-body administration, plasma degradation, and difficult-to-deliver targets (such as cardiovascular tissues and the central nervous system) are just a few of the clinical applications that still need to be explored.
In conclusion, the use of ncRNAs in treatment helps us comprehend the link between ncRNAs and the cell cycle. This study has concentrated on the influence of ncRNAs on the cell cycle as well as the effect of the cell cycle on the composition and function of ncRNAs. Further research is required to determine how ncRNAs affect CDK4/6 inhibitors and to generate novel treatment targets and effectiveness indicators based on this connection.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24108939/s1.

Author Contributions

Conceptualization, Q.H., T.H.; writing—original draft preparation, Q.H.; writing—review and editing, Q.H., T.H.; visualization, Q.H.; supervision, T.H.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge Johnson Chen of YiShun Biology Laboratory for his help in polishing the English manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fischer, M.; Schade, A.E.; Branigan, T.B.; Müller, G.A.; DeCaprio, J.A. Coordinating gene expression during the cell cycle. Trends Biochem. Sci. 2022, 47, 1009–1022. [Google Scholar] [CrossRef]
  2. Wang, L.; Chen, H.; Wang, C.; Hu, Z.; Yan, S. Negative regulator of E2F transcription factors links cell cycle checkpoint and DNA damage repair. Proc. Natl. Acad. Sci. USA 2018, 115, E3837–E3845. [Google Scholar] [CrossRef] [PubMed]
  3. Moreno-Layseca, P.; Streuli, C.H. Signalling pathways linking integrins with cell cycle progression. Matrix Biol. 2014, 34, 144–153. [Google Scholar] [CrossRef] [PubMed]
  4. Sherr, C.J.; Beach, D.; Shapiro, G.I. Targeting CDK4 and CDK6: From Discovery to Therapy. Cancer Discov. 2016, 6, 353–367. [Google Scholar] [CrossRef]
  5. Gong, X.; Litchfield, L.M.; Webster, Y.; Chio, L.C.; Wong, S.S.; Stewart, T.R.; Dowless, M.; Dempsey, J.; Zeng, Y.; Torres, R.; et al. Genomic Aberrations that Activate D-type Cyclins Are Associated with Enhanced Sensitivity to the CDK4 and CDK6 Inhibitor Abemaciclib. Cancer Cell 2017, 32, 761–776.e6. [Google Scholar] [CrossRef]
  6. Goel, S.; DeCristo, M.J.; McAllister, S.S.; Zhao, J.J. CDK4/6 Inhibition in Cancer: Beyond Cell Cycle Arrest. Trends Cell Biol. 2018, 28, 911–925. [Google Scholar] [CrossRef] [PubMed]
  7. O’Leary, B.; Finn, R.S.; Turner, N.C. Treating cancer with selective CDK4/6 inhibitors. Nat. Rev. Clin. Oncol. 2016, 13, 417–430. [Google Scholar] [CrossRef] [PubMed]
  8. Turner, N.C.; Ro, J.; André, F.; Loi, S.; Verma, S.; Iwata, H.; Harbeck, N.; Loibl, S.; Huang Bartlett, C.; Zhang, K.; et al. Palbociclib in Hormone-Receptor-Positive Advanced Breast Cancer. N. Engl. J. Med. 2015, 373, 209–219. [Google Scholar] [CrossRef]
  9. Hortobagyi, G.N.; Stemmer, S.M.; Burris, H.A.; Yap, Y.S.; Sonke, G.S.; Hart, L.; Campone, M.; Petrakova, K.; Winer, E.P.; Janni, W.; et al. Overall Survival with Ribociclib plus Letrozole in Advanced Breast Cancer. N. Engl. J. Med. 2022, 386, 942–950. [Google Scholar] [CrossRef]
  10. Johnston, S.R.D.; Harbeck, N.; Hegg, R.; Toi, M.; Martin, M.; Shao, Z.M.; Zhang, Q.Y.; Martinez Rodriguez, J.L.; Campone, M.; Hamilton, E.; et al. Abemaciclib Combined With Endocrine Therapy for the Adjuvant Treatment of HR+, HER2-, Node-Positive, High-Risk, Early Breast Cancer (monarchE). J. Clin. Oncol. 2020, 38, 3987–3998. [Google Scholar] [CrossRef]
  11. Wang, M.Q.; Zhu, W.J.; Gao, P. New insights into long non-coding RNAs in breast cancer: Biological functions and therapeutic prospects. Exp. Mol. Pathol. 2021, 120, 104640. [Google Scholar] [CrossRef] [PubMed]
  12. Shi, J.; Zhou, T.; Chen, Q. Exploring the expanding universe of small RNAs. Nat. Cell Biol. 2022, 24, 415–423. [Google Scholar] [CrossRef] [PubMed]
  13. Slack, F.J.; Chinnaiyan, A.M. The Role of Non-coding RNAs in Oncology. Cell 2019, 179, 1033–1055. [Google Scholar] [CrossRef] [PubMed]
  14. 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]
  15. Beermann, J.; Piccoli, M.T.; Viereck, J.; Thum, T. Non-coding RNAs in Development and Disease: Background, Mechanisms, and Therapeutic Approaches. Physiol. Rev. 2016, 96, 1297–1325. [Google Scholar] [CrossRef]
  16. Saw, P.E.; Xu, X.; Chen, J.; Song, E.W. Non-coding RNAs: The new central dogma of cancer biology. Sci. China Life Sci. 2021, 64, 22–50. [Google Scholar] [CrossRef]
  17. Barr, A.R.; Heldt, F.S.; Zhang, T.; Bakal, C.; Novák, B. A Dynamical Framework for the All-or-None G1/S Transition. Cell Syst. 2016, 2, 27–37. [Google Scholar] [CrossRef]
  18. Hume, S.; Dianov, G.L.; Ramadan, K. A unified model for the G1/S cell cycle transition. Nucleic Acids Res. 2020, 48, 12483–12501. [Google Scholar] [CrossRef]
  19. Coffman, J.A. Cell cycle development. Dev. Cell 2004, 6, 321–327. [Google Scholar] [CrossRef]
  20. Engeland, K. Cell cycle regulation: p53-p21-RB signaling. Cell Death Differ. 2022, 29, 946–960. [Google Scholar] [CrossRef]
  21. Knudsen, E.S.; Pruitt, S.C.; Hershberger, P.A.; Witkiewicz, A.K.; Goodrich, D.W. Cell Cycle and Beyond: Exploiting New RB1 Controlled Mechanisms for Cancer Therapy. Trends Cancer 2019, 5, 308–324. [Google Scholar] [CrossRef] [PubMed]
  22. Dimova, D.K.; Stevaux, O.; Frolov, M.V.; Dyson, N.J. Cell cycle-dependent and cell cycle-independent control of transcription by the Drosophila E2F/RB pathway. Genes. Dev. 2003, 17, 2308–2320. [Google Scholar] [CrossRef] [PubMed]
  23. Swaffer, M.P.; Jones, A.W.; Flynn, H.R.; Snijders, A.P.; Nurse, P. CDK Substrate Phosphorylation and Ordering the Cell Cycle. Cell 2016, 167, 1750–1761.e16. [Google Scholar] [CrossRef] [PubMed]
  24. Bertoli, C.; Skotheim, J.M.; de Bruin, R.A. Control of cell cycle transcription during G1 and S phases. Nat. Rev. Mol. Cell Biol. 2013, 14, 518–528. [Google Scholar] [CrossRef]
  25. Orlando, D.A.; Lin, C.Y.; Bernard, A.; Wang, J.Y.; Socolar, J.E.; Iversen, E.S.; Hartemink, A.J.; Haase, S.B. Global control of cell-cycle transcription by coupled CDK and network oscillators. Nature 2008, 453, 944–947. [Google Scholar] [CrossRef]
  26. Lavoie, J.N.; L’Allemain, G.; Brunet, A.; Müller, R.; Pouysségur, J. Cyclin D1 expression is regulated positively by the p42/p44MAPK and negatively by the p38/HOGMAPK pathway. J. Biol. Chem. 1996, 271, 20608–20616. [Google Scholar] [CrossRef]
  27. Hwang, H.C.; Clurman, B.E. Cyclin E in normal and neoplastic cell cycles. Oncogene 2005, 24, 2776–2786. [Google Scholar] [CrossRef]
  28. Clijsters, L.; Ogink, J.; Wolthuis, R. The spindle checkpoint, APC/C(Cdc20), and APC/C(Cdh1) play distinct roles in connecting mitosis to S phase. J. Cell Biol. 2013, 201, 1013–1026. [Google Scholar] [CrossRef]
  29. Carnero, A.; Hannon, G.J. The INK4 family of CDK inhibitors. Curr. Top. Microbiol. Immunol. 1998, 227, 43–55. [Google Scholar] [CrossRef]
  30. Vidal, A.; Koff, A. Cell-cycle inhibitors: Three families united by a common cause. Gene 2000, 247, 1–15. [Google Scholar] [CrossRef]
  31. Bury, M.; Le Calvé, B.; Ferbeyre, G.; Blank, V.; Lessard, F. New Insights into CDK Regulators: Novel Opportunities for Cancer Therapy. Trends Cell Biol. 2021, 31, 331–344. [Google Scholar] [CrossRef] [PubMed]
  32. Chan, F.K.; Zhang, J.; Cheng, L.; Shapiro, D.N.; Winoto, A. Identification of human and mouse p19, a novel CDK4 and CDK6 inhibitor with homology to p16ink4. Mol. Cell Biol. 1995, 15, 2682–2688. [Google Scholar] [CrossRef] [PubMed]
  33. Hannon, G.J.; Beach, D. p15INK4B is a potential effector of TGF-beta-induced cell cycle arrest. Nature 1994, 371, 257–261. [Google Scholar] [CrossRef]
  34. Guan, K.L.; Jenkins, C.W.; Li, Y.; Nichols, M.A.; Wu, X.; O’Keefe, C.L.; Matera, A.G.; Xiong, Y. Growth suppression by p18, a p16INK4/MTS1- and p14INK4B/MTS2-related CDK6 inhibitor, correlates with wild-type pRb function. Genes. Dev. 1994, 8, 2939–2952. [Google Scholar] [CrossRef] [PubMed]
  35. Hirai, H.; Roussel, M.F.; Kato, J.Y.; Ashmun, R.A.; Sherr, C.J. Novel INK4 proteins, p19 and p18, are specific inhibitors of the cyclin D-dependent kinases CDK4 and CDK6. Mol. Cell Biol. 1995, 15, 2672–2681. [Google Scholar] [CrossRef]
  36. Jeffrey, P.D.; Tong, L.; Pavletich, N.P. Structural basis of inhibition of CDK-cyclin complexes by INK4 inhibitors. Genes. Dev. 2000, 14, 3115–3125. [Google Scholar] [CrossRef]
  37. Serrano, M.; Hannon, G.J.; Beach, D. A new regulatory motif in cell-cycle control causing specific inhibition of cyclin D/CDK4. Nature 1993, 366, 704–707. [Google Scholar] [CrossRef]
  38. Kamb, A.; Gruis, N.A.; Weaver-Feldhaus, J.; Liu, Q.; Harshman, K.; Tavtigian, S.V.; Stockert, E.; Day, R.S., 3rd; Johnson, B.E.; Skolnick, M.H. A cell cycle regulator potentially involved in genesis of many tumor types. Science 1994, 264, 436–440. [Google Scholar] [CrossRef]
  39. Roussel, M.F. The INK4 family of cell cycle inhibitors in cancer. Oncogene 1999, 18, 5311–5317. [Google Scholar] [CrossRef]
  40. Serrano, M.; Lee, H.; Chin, L.; Cordon-Cardo, C.; Beach, D.; DePinho, R.A. Role of the INK4a locus in tumor suppression and cell mortality. Cell 1996, 85, 27–37. [Google Scholar] [CrossRef]
  41. Kim, W.Y.; Sharpless, N.E. The regulation of INK4/ARF in cancer and aging. Cell 2006, 127, 265–275. [Google Scholar] [CrossRef] [PubMed]
  42. Sherr, C.J.; Roberts, J.M. CDK inhibitors: Positive and negative regulators of G1-phase progression. Genes. Dev. 1999, 13, 1501–1512. [Google Scholar] [CrossRef] [PubMed]
  43. Besson, A.; Dowdy, S.F.; Roberts, J.M. CDK inhibitors: Cell cycle regulators and beyond. Dev. Cell 2008, 14, 159–169. [Google Scholar] [CrossRef] [PubMed]
  44. Hengst, L.; Göpfert, U.; Lashuel, H.A.; Reed, S.I. Complete inhibition of Cdk/cyclin by one molecule of p21(Cip1). Genes. Dev. 1998, 12, 3882–3888. [Google Scholar] [CrossRef]
  45. Lacy, E.R.; Filippov, I.; Lewis, W.S.; Otieno, S.; Xiao, L.; Weiss, S.; Hengst, L.; Kriwacki, R.W. p27 binds cyclin-CDK complexes through a sequential mechanism involving binding-induced protein folding. Nat. Struct. Mol. Biol. 2004, 11, 358–364. [Google Scholar] [CrossRef] [PubMed]
  46. Blain, S.W. Switching cyclin D-Cdk4 kinase activity on and off. Cell Cycle 2008, 7, 892–898. [Google Scholar] [CrossRef]
  47. Bencivenga, D.; Stampone, E.; Roberti, D.; Della Ragione, F.; Borriello, A. p27(Kip1), an Intrinsically Unstructured Protein with Scaffold Properties. Cells 2021, 10, 2254. [Google Scholar] [CrossRef] [PubMed]
  48. Deng, C.; Zhang, P.; Harper, J.W.; Elledge, S.J.; Leder, P. Mice lacking p21CIP1/WAF1 undergo normal development, but are defective in G1 checkpoint control. Cell 1995, 82, 675–684. [Google Scholar] [CrossRef]
  49. Harper, J.W.; Adami, G.R.; Wei, N.; Keyomarsi, K.; Elledge, S.J. The p21 Cdk-interacting protein Cip1 is a potent inhibitor of G1 cyclin-dependent kinases. Cell 1993, 75, 805–816. [Google Scholar] [CrossRef]
  50. Abbas, T.; Dutta, A. p21 in cancer: Intricate networks and multiple activities. Nat. Rev. Cancer 2009, 9, 400–414. [Google Scholar] [CrossRef]
  51. Matheson, C.J.; Backos, D.S.; Reigan, P. Targeting WEE1 Kinase in Cancer. Trends Pharmacol. Sci. 2016, 37, 872–881. [Google Scholar] [CrossRef] [PubMed]
  52. Geenen, J.J.J.; Schellens, J.H.M. Molecular Pathways: Targeting the Protein Kinase Wee1 in Cancer. Clin. Cancer Res. 2017, 23, 4540–4544. [Google Scholar] [CrossRef] [PubMed]
  53. Nojima, T.; Proudfoot, N.J. Mechanisms of lncRNA biogenesis as revealed by nascent transcriptomics. Nat. Rev. Mol. Cell Biol. 2022, 23, 389–406. [Google Scholar] [CrossRef] [PubMed]
  54. Mercer, T.R.; Mattick, J.S. Structure and function of long noncoding RNAs in epigenetic regulation. Nat. Struct. Mol. Biol. 2013, 20, 300–307. [Google Scholar] [CrossRef] [PubMed]
  55. Ponting, C.P.; Oliver, P.L.; Reik, W. Evolution and functions of long noncoding RNAs. Cell 2009, 136, 629–641. [Google Scholar] [CrossRef]
  56. Jarroux, J.; Morillon, A.; Pinskaya, M. History, Discovery, and Classification of lncRNAs. In Advances in Experimental Medicine and Biology; Springer: Singapore, 2017; Volume 1008, pp. 1–46. [Google Scholar] [CrossRef]
  57. Ishii, N.; Ozaki, K.; Sato, H.; Mizuno, H.; Susumu, S.; Takahashi, A.; Miyamoto, Y.; Ikegawa, S.; Kamatani, N.; Hori, M.; et al. Identification of a novel non-coding RNA, MIAT, that confers risk of myocardial infarction. J. Hum. Genet. 2006, 51, 1087–1099. [Google Scholar] [CrossRef]
  58. Ali, S.A.; Peffers, M.J.; Ormseth, M.J.; Jurisica, I.; Kapoor, M. The non-coding RNA interactome in joint health and disease. Nat. Rev. Rheumatol. 2021, 17, 692–705. [Google Scholar] [CrossRef]
  59. Chen, X.; Yan, C.C.; Zhang, X.; You, Z.H. Long non-coding RNAs and complex diseases: From experimental results to computational models. Brief. Bioinform. 2017, 18, 558–576. [Google Scholar] [CrossRef]
  60. Joung, J.; Engreitz, J.M.; Konermann, S.; Abudayyeh, O.O.; Verdine, V.K.; Aguet, F.; Gootenberg, J.S.; Sanjana, N.E.; Wright, J.B.; Fulco, C.P.; et al. Genome-scale activation screen identifies a lncRNA locus regulating a gene neighbourhood. Nature 2017, 548, 343–346. [Google Scholar] [CrossRef] [PubMed]
  61. Castellanos-Rubio, A.; Ghosh, S. Disease-Associated SNPs in Inflammation-Related lncRNAs. Front. Immunol. 2019, 10, 420. [Google Scholar] [CrossRef]
  62. Gao, P.; Wei, G.H. Genomic Insight into the Role of lncRNA in Cancer Susceptibility. Int. J. Mol. Sci. 2017, 18, 1239. [Google Scholar] [CrossRef] [PubMed]
  63. Brockdorff, N. X-chromosome inactivation: Closing in on proteins that bind Xist RNA. Trends Genet. 2002, 18, 352–358. [Google Scholar] [CrossRef] [PubMed]
  64. 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]
  65. Jiang, W.; Liu, Y.; Liu, R.; Zhang, K.; Zhang, Y. The lncRNA DEANR1 facilitates human endoderm differentiation by activating FOXA2 expression. Cell Rep. 2015, 11, 137–148. [Google Scholar] [CrossRef]
  66. Latos, P.A.; Pauler, F.M.; Koerner, M.V.; Şenergin, H.B.; Hudson, Q.J.; Stocsits, R.R.; Allhoff, W.; Stricker, S.H.; Klement, R.M.; Warczok, K.E.; et al. Airn transcriptional overlap, but not its lncRNA products, induces imprinted Igf2r silencing. Science 2012, 338, 1469–1472. [Google Scholar] [CrossRef]
  67. Lefevre, P.; Witham, J.; Lacroix, C.E.; Cockerill, P.N.; Bonifer, C. The LPS-induced transcriptional upregulation of the chicken lysozyme locus involves CTCF eviction and noncoding RNA transcription. Mol. Cell 2008, 32, 129–139. [Google Scholar] [CrossRef]
  68. Khyzha, N.; Khor, M.; DiStefano, P.V.; Wang, L.; Matic, L.; Hedin, U.; Wilson, M.D.; Maegdefessel, L.; Fish, J.E. Regulation of CCL2 expression in human vascular endothelial cells by a neighboring divergently transcribed long noncoding RNA. Proc. Natl. Acad. Sci. USA 2019, 116, 16410–16419. [Google Scholar] [CrossRef]
  69. Zhu, W.; Zhou, B.L.; Rong, L.J.; Ye, L.; Xu, H.J.; Zhou, Y.; Yan, X.J.; Liu, W.D.; Zhu, B.; Wang, L.; et al. Roles of PTBP1 in alternative splicing, glycolysis, and oncogensis. J. Zhejiang Univ. Sci. B 2020, 21, 122–136. [Google Scholar] [CrossRef]
  70. Yang, M.; Lu, H.; Liu, J.; Wu, S.; Kim, P.; Zhou, X. lncRNAfunc: A knowledgebase of lncRNA function in human cancer. Nucleic Acids Res. 2022, 50, D1295–D1306. [Google Scholar] [CrossRef]
  71. Zhu, J.; Fu, H.; Wu, Y.; Zheng, X. Function of lncRNAs and approaches to lncRNA-protein interactions. Sci. China Life Sci. 2013, 56, 876–885. [Google Scholar] [CrossRef]
  72. Lee, E.S.; Wolf, E.J.; Ihn, S.S.J.; Smith, H.W.; Emili, A.; Palazzo, A.F. TPR is required for the efficient nuclear export of mRNAs and lncRNAs from short and intron-poor genes. Nucleic Acids Res. 2020, 48, 11645–11663. [Google Scholar] [CrossRef] [PubMed]
  73. Zuckerman, B.; Ulitsky, I. Predictive models of subcellular localization of long RNAs. RNA 2019, 25, 557–572. [Google Scholar] [CrossRef] [PubMed]
  74. Paraskevopoulou, M.D.; Vlachos, I.S.; Karagkouni, D.; Georgakilas, G.; Kanellos, I.; Vergoulis, T.; Zagganas, K.; Tsanakas, P.; Floros, E.; Dalamagas, T.; et al. DIANA-LncBase v2: Indexing microRNA targets on non-coding transcripts. Nucleic Acids Res. 2016, 44, D231–D238. [Google Scholar] [CrossRef] [PubMed]
  75. Xu, Y.; Wang, H.; Li, F.; Heindl, L.M.; He, X.; Yu, J.; Yang, J.; Ge, S.; Ruan, J.; Jia, R.; et al. Long Non-coding RNA LINC-PINT Suppresses Cell Proliferation and Migration of Melanoma via Recruiting EZH2. Front. Cell Dev. Biol. 2019, 7, 350. [Google Scholar] [CrossRef]
  76. Bukhari, I.; Khan, M.R.; Hussain, M.A.; Thorne, R.F.; Yu, Y.; Zhang, B.; Zheng, P.; Mi, Y. PINTology: A short history of the lncRNA LINC-PINT in different diseases. Wiley Interdiscip. Rev. RNA 2022, 13, e1705. [Google Scholar] [CrossRef]
  77. Yang, F.; Zhang, L.; Huo, X.S.; Yuan, J.H.; Xu, D.; Yuan, S.X.; Zhu, N.; Zhou, W.P.; Yang, G.S.; Wang, Y.Z.; et al. Long noncoding RNA high expression in hepatocellular carcinoma facilitates tumor growth through enhancer of zeste homolog 2 in humans. Hepatology 2011, 54, 1679–1689. [Google Scholar] [CrossRef]
  78. Pasmant, E.; Laurendeau, I.; Héron, D.; Vidaud, M.; Vidaud, D.; Bièche, I. Characterization of a germ-line deletion, including the entire INK4/ARF locus, in a melanoma-neural system tumor family: Identification of ANRIL, an antisense noncoding RNA whose expression coclusters with ARF. Cancer Res. 2007, 67, 3963–3969. [Google Scholar] [CrossRef]
  79. Yap, K.L.; Li, S.; Muñoz-Cabello, A.M.; Raguz, S.; Zeng, L.; Mujtaba, S.; Gil, J.; Walsh, M.J.; Zhou, M.M. Molecular interplay of the noncoding RNA ANRIL and methylated histone H3 lysine 27 by polycomb CBX7 in transcriptional silencing of INK4a. Mol. Cell 2010, 38, 662–674. [Google Scholar] [CrossRef]
  80. Kotake, Y.; Nakagawa, T.; Kitagawa, K.; Suzuki, S.; Liu, N.; Kitagawa, M.; Xiong, Y. Long non-coding RNA ANRIL is required for the PRC2 recruitment to and silencing of p15(INK4B) tumor suppressor gene. Oncogene 2011, 30, 1956–1962. [Google Scholar] [CrossRef]
  81. Zangouei, A.S.; Zangoue, M.; Taghehchian, N.; Zangooie, A.; Rahimi, H.R.; Saburi, E.; Alavi, M.S.; Moghbeli, M. Cell cycle related long non-coding RNAs as the critical regulators of breast cancer progression and metastasis. Biol. Res. 2023, 56, 1. [Google Scholar] [CrossRef]
  82. Li, Q.; Bian, Y.; Li, Q. Down-Regulation of TMPO-AS1 Induces Apoptosis in Lung Carcinoma Cells by Regulating miR-143-3p/CDK1 Axis. Technol. Cancer Res. Treat. 2021, 20, 1533033820948880. [Google Scholar] [CrossRef] [PubMed]
  83. Xue, J.; Song, Y.; Xu, W.; Zhu, Y. The CDK1-Related lncRNA and CXCL8 Mediated Immune Resistance in Lung Adenocarcinoma. Cells 2022, 11, 2688. [Google Scholar] [CrossRef] [PubMed]
  84. Kang, J.; Huang, X.; Dong, W.; Zhu, X.; Li, M.; Cui, N. Long non-coding RNA LINC00630 facilitates hepatocellular carcinoma progression through recruiting transcription factor E2F1 to up-regulate cyclin-dependent kinase 2 expression. Hum. Exp. Toxicol. 2021, 40, S257–S268. [Google Scholar] [CrossRef] [PubMed]
  85. Tao, X.; Zhang, Y.; Li, J.; Ni, Z.; Tao, Z.; You, Q.; He, Z.; Huang, D.; Zheng, S. Low expression of long non-coding RNA ARAP1-AS1 can inhibit lung cancer proliferation by inducing G0/G1 cell cycle organization. J. Thorac. Dis. 2020, 12, 7326–7336. [Google Scholar] [CrossRef] [PubMed]
  86. Luan, T.; Zhang, T.-Y.; Lv, Z.-H.; Guan, B.-X.; Xu, J.-Y.; Li, J.; Li, M.-X.; Hu, S.-L. The lncRNA ALMS1-IT1 may promote malignant progression of lung adenocarcinoma via AVL9-mediated activation of the cyclin-dependent kinase pathway. Febs Open Bio 2021, 11, 1504–1515. [Google Scholar] [CrossRef]
  87. Sun, Y.; Niu, X.; Wang, G.; Qiao, X.; Chen, L.; Zhong, M. A Novel lncRNA ENST00000512916 Facilitates Cell Proliferation, Migration and Cell Cycle Progression in Ameloblastoma. Oncotargets Ther. 2020, 13, 1519–1531. [Google Scholar] [CrossRef]
  88. Su, L.; Han, D.; Wu, J.; Huo, X. Skp2 regulates non-small cell lung cancer cell growth by Meg3 and miR-3163. Tumor Biol. 2016, 37, 3925–3931. [Google Scholar] [CrossRef]
  89. Tang, J.; Huang, F.; Wang, H.; Cheng, F.; Pi, Y.; Zhao, J.; Li, Z. Knockdown of TPT1-AS1 inhibits cell proliferation, cell cycle G1/S transition, and epithelial-mesenchymal transition in gastric cancer. Bosn. J. Basic Med. Sci. 2021, 21, 39–46. [Google Scholar] [CrossRef]
  90. Lin, X.; Yang, M.; Xia, T.; Guo, J. Increased expression of long noncoding RNA ABHD11-AS1 in gastric cancer and its clinical significance. Med. Oncol. 2014, 31, 42. [Google Scholar] [CrossRef]
  91. Wu, D.D.; Chen, X.; Sun, K.X.; Wang, L.L.; Chen, S.; Zhao, Y. Role of the lncRNA ABHD11-AS(1) in the tumorigenesis and progression of epithelial ovarian cancer through targeted regulation of RhoC. Mol. Cancer 2017, 16, 138. [Google Scholar] [CrossRef]
  92. Chen, M.; Li, J.; Zhuang, C.; Cai, Z. Increased lncRNA ABHD11-AS1 represses the malignant phenotypes of bladder cancer. Oncotarget 2017, 8, 28176–28186. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, Y.; Wang, L.-L.; Chen, S.; Zong, Z.-H.; Guan, X.; Zhao, Y. LncRNA ABHD11-AS1 promotes the development of endometrial carcinoma by targeting cyclin D1. J. Cell. Mol. Med. 2018, 22, 3955–3964. [Google Scholar] [CrossRef] [PubMed]
  94. Sharma, U.; Barwal, T.S.; Khandelwal, A.; Malhotra, A.; Rana, M.K.; Rana, A.P.S.; Imyanitov, E.N.; Vasquez, K.M.; Jain, A. LncRNA ZFAS1 inhibits triple-negative breast cancer by targeting STAT3. Biochimie 2021, 182, 99–107. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, F.; Qi, X.; Li, Z.; Jin, S.; Xie, Y.; Zhong, H. lncRNA CADM1-AS1 inhibits cell-cycle progression and invasion via PTEN/AKT/GSK-3β axis in hepatocellular carcinoma. Cancer Manag. Res. 2019, 11, 3813–3828. [Google Scholar] [CrossRef]
  96. Liu, X.; Li, D.; Zhang, W.; Guo, M.; Zhan, Q. Long non-coding RNA gadd7 interacts with TDP-43 and regulates Cdk6 mRNA decay. EMBO J. 2012, 31, 4415–4427. [Google Scholar] [CrossRef]
  97. Hollander, M.C.; Alamo, I.; Fornace, A.J., Jr. A novel DNA damage-inducible transcript, gadd7, inhibits cell growth, but lacks a protein product. Nucleic Acids Res. 1996, 24, 1589–1593. [Google Scholar] [CrossRef]
  98. Liang, S.; Gong, X.; Zhang, G.; Huang, G.; Lu, Y.; Li, Y. The lncRNA XIST interacts with miR-140/miR-124/iASPP axis to promote pancreatic carcinoma growth. Oncotarget 2017, 8, 113701–113718. [Google Scholar] [CrossRef]
  99. Zhang, Y.; Zhang, H.; Wu, S. LncRNA-CCDC144NL-AS1 Promotes the Development of Hepatocellular Carcinoma by Inducing WDR5 Expression via Sponging miR-940. J. Hepatocell. Carcinoma 2021, 8, 333–348. [Google Scholar] [CrossRef]
  100. Jin, X.; Ge, L.P.; Li, D.Q.; Shao, Z.M.; Di, G.H.; Xu, X.E.; Jiang, Y.Z. LncRNA TROJAN promotes proliferation and resistance to CDK4/6 inhibitor via CDK2 transcriptional activation in ER+ breast cancer. Mol. Cancer 2020, 19, 87. [Google Scholar] [CrossRef]
  101. Li, Z.; Zhang, J.; Zheng, H.; Li, C.; Xiong, J.; Wang, W.; Bao, H.; Jin, H.; Liang, P. Modulating lncRNA SNHG15/CDK6/miR-627 circuit by palbociclib, overcomes temozolomide resistance and reduces M2-polarization of glioma associated microglia in glioblastoma multiforme. J. Exp. Clin. Cancer Res. 2019, 38, 380. [Google Scholar] [CrossRef]
  102. Meister, G.; Tuschl, T. Mechanisms of gene silencing by double-stranded RNA. Nature 2004, 431, 343–349. [Google Scholar] [CrossRef] [PubMed]
  103. Volinia, S.; Calin, G.A.; Liu, C.G.; Ambs, S.; Cimmino, A.; Petrocca, F.; Visone, R.; Iorio, M.; Roldo, C.; Ferracin, M.; et al. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc. Natl. Acad. Sci. USA 2006, 103, 2257–2261. [Google Scholar] [CrossRef] [PubMed]
  104. Nicoloso, M.S.; Spizzo, R.; Shimizu, M.; Rossi, S.; Calin, G.A. MicroRNAs--the micro steering wheel of tumour metastases. Nat. Rev. Cancer 2009, 9, 293–302. [Google Scholar] [CrossRef] [PubMed]
  105. Esquela-Kerscher, A.; Slack, F.J. Oncomirs—microRNAs with a role in cancer. Nat. Rev. Cancer 2006, 6, 259–269. [Google Scholar] [CrossRef]
  106. Lee, Y.; Jeon, K.; Lee, J.T.; Kim, S.; Kim, V.N. MicroRNA maturation: Stepwise processing and subcellular localization. EMBO J. 2002, 21, 4663–4670. [Google Scholar] [CrossRef]
  107. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef]
  108. Yi, R.; Qin, Y.; Macara, I.G.; Cullen, B.R. Exportin-5 mediates the nuclear export of pre-microRNAs and short hairpin RNAs. Genes. Dev. 2003, 17, 3011–3016. [Google Scholar] [CrossRef]
  109. Fabian, M.R.; Sonenberg, N. The mechanics of miRNA-mediated gene silencing: A look under the hood of miRISC. Nat. Struct. Mol. Biol. 2012, 19, 586–593. [Google Scholar] [CrossRef]
  110. Diederichs, S.; Haber, D.A. Dual role for argonautes in microRNA processing and posttranscriptional regulation of microRNA expression. Cell 2007, 131, 1097–1108. [Google Scholar] [CrossRef]
  111. Raisch, J.; Darfeuille-Michaud, A.; Nguyen, H.T. Role of microRNAs in the immune system, inflammation and cancer. World J. Gastroenterol. 2013, 19, 2985–2996. [Google Scholar] [CrossRef]
  112. Chen, K.; Rajewsky, N. Deep conservation of microRNA-target relationships and 3’UTR motifs in vertebrates, flies, and nematodes. Cold Spring Harb. Symp. Quant. Biol. 2006, 71, 149–156. [Google Scholar] [CrossRef] [PubMed]
  113. Tay, Y.; Zhang, J.; Thomson, A.M.; Lim, B.; Rigoutsos, I. MicroRNAs to Nanog, Oct4 and Sox2 coding regions modulate embryonic stem cell differentiation. Nature 2008, 455, 1124–1128. [Google Scholar] [CrossRef] [PubMed]
  114. 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] [PubMed]
  115. Romero-Cordoba, S.L.; Salido-Guadarrama, I.; Rodriguez-Dorantes, M.; Hidalgo-Miranda, A. miRNA biogenesis: Biological impact in the development of cancer. Cancer Biol. Ther. 2014, 15, 1444–1455. [Google Scholar] [CrossRef] [PubMed]
  116. Gangaraju, V.K.; Lin, H. MicroRNAs: Key regulators of stem cells. Nat. Rev. Mol. Cell Biol. 2009, 10, 116–125. [Google Scholar] [CrossRef]
  117. Chen, X.; Guo, X.; Zhang, H.; Xiang, Y.; Chen, J.; Yin, Y.; Cai, X.; Wang, K.; Wang, G.; Ba, Y.; et al. Role of miR-143 targeting KRAS in colorectal tumorigenesis. Oncogene 2009, 28, 1385–1392. [Google Scholar] [CrossRef]
  118. Calin, G.A.; Dumitru, C.D.; Shimizu, M.; Bichi, R.; Zupo, S.; Noch, E.; Aldler, H.; Rattan, S.; Keating, M.; Rai, K.; et al. Frequent deletions and down-regulation of micro- RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia. Proc. Natl. Acad. Sci. USA 2002, 99, 15524–15529. [Google Scholar] [CrossRef]
  119. Xu, H.; He, J.H.; Xiao, Z.D.; Zhang, Q.Q.; Chen, Y.Q.; Zhou, H.; Qu, L.H. Liver-enriched transcription factors regulate microRNA-122 that targets CUTL1 during liver development. Hepatology 2010, 52, 1431–1442. [Google Scholar] [CrossRef]
  120. Young, D.D.; Connelly, C.M.; Grohmann, C.; Deiters, A. Small molecule modifiers of microRNA miR-122 function for the treatment of hepatitis C virus infection and hepatocellular carcinoma. J. Am. Chem. Soc. 2010, 132, 7976–7981. [Google Scholar] [CrossRef]
  121. Scott, G.K.; Goga, A.; Bhaumik, D.; Berger, C.E.; Sullivan, C.S.; Benz, C.C. Coordinate suppression of ERBB2 and ERBB3 by enforced expression of micro-RNA miR-125a or miR-125b. J. Biol. Chem. 2007, 282, 1479–1486. [Google Scholar] [CrossRef]
  122. Ma, L.; Reinhardt, F.; Pan, E.; Soutschek, J.; Bhat, B.; Marcusson, E.G.; Teruya-Feldstein, J.; Bell, G.W.; Weinberg, R.A. Therapeutic silencing of miR-10b inhibits metastasis in a mouse mammary tumor model. Nat. Biotechnol. 2010, 28, 341–347. [Google Scholar] [CrossRef]
  123. Li, X.; Liu, X.; Xu, W.; Zhou, P.; Gao, P.; Jiang, S.; Lobie, P.E.; Zhu, T. c-MYC-regulated miR-23a/24-2/27a cluster promotes mammary carcinoma cell invasion and hepatic metastasis by targeting Sprouty2. J. Biol. Chem. 2013, 288, 18121–18133. [Google Scholar] [CrossRef] [PubMed]
  124. Shell, S.; Park, S.M.; Radjabi, A.R.; Schickel, R.; Kistner, E.O.; Jewell, D.A.; Feig, C.; Lengyel, E.; Peter, M.E. Let-7 expression defines two differentiation stages of cancer. Proc. Natl. Acad. Sci. USA 2007, 104, 11400–11405. [Google Scholar] [CrossRef] [PubMed]
  125. Lu, Z.; Ye, Y.; Jiao, D.; Qiao, J.; Cui, S.; Liu, Z. miR-155 and miR-31 are differentially expressed in breast cancer patients and are correlated with the estrogen receptor and progesterone receptor status. Oncol. Lett. 2012, 4, 1027–1032. [Google Scholar] [CrossRef] [PubMed]
  126. Deng, H.; Guo, Y.; Song, H.; Xiao, B.; Sun, W.; Liu, Z.; Yu, X.; Xia, T.; Cui, L.; Guo, J. MicroRNA-195 and microRNA-378 mediate tumor growth suppression by epigenetical regulation in gastric cancer. Gene 2013, 518, 351–359. [Google Scholar] [CrossRef] [PubMed]
  127. Xu, T.; Zhu, Y.; Xiong, Y.; Ge, Y.Y.; Yun, J.P.; Zhuang, S.M. MicroRNA-195 suppresses tumorigenicity and regulates G1/S transition of human hepatocellular carcinoma cells. Hepatology 2009, 50, 113–121. [Google Scholar] [CrossRef]
  128. Lobert, S.; Jefferson, B.; Morris, K. Regulation of β-tubulin isotypes by micro-RNA 100 in MCF7 breast cancer cells. Cytoskeleton 2011, 68, 355–362. [Google Scholar] [CrossRef]
  129. Gong, Y.; He, T.; Yang, L.; Yang, G.; Chen, Y.; Zhang, X. The role of miR-100 in regulating apoptosis of breast cancer cells. Sci. Rep. 2015, 5, 11650. [Google Scholar] [CrossRef]
  130. Li, H.; Jiang, M.; Cui, M.; Feng, G.; Dong, J.; Li, Y.; Xiao, H.; Fan, S. MiR-365 enhances the radiosensitivity of non-small cell lung cancer cells through targeting CDC25A. Biochem. Biophys. Res. Commun. 2019, 512, 392–398. [Google Scholar] [CrossRef]
  131. Kim, M.H.; Ham, O.; Lee, S.Y.; Choi, E.; Lee, C.Y.; Park, J.H.; Lee, J.; Seo, H.H.; Seung, M.; Choi, E.; et al. MicroRNA-365 inhibits the proliferation of vascular smooth muscle cells by targeting cyclin D1. J. Cell. Biochem. 2014, 115, 1752–1761. [Google Scholar] [CrossRef]
  132. Luo, Q.; Li, X.; Li, J.; Kong, X.; Zhang, J.; Chen, L.; Huang, Y.; Fang, L. MiR-15a is underexpressed and inhibits the cell cycle by targeting CCNE1 in breast cancer. Int. J. Oncol. 2013, 43, 1212–1218. [Google Scholar] [CrossRef] [PubMed]
  133. Bonci, D.; Coppola, V.; Musumeci, M.; Addario, A.; Giuffrida, R.; Memeo, L.; D’Urso, L.; Pagliuca, A.; Biffoni, M.; Labbaye, C.; et al. The miR-15a-miR-16-1 cluster controls prostate cancer by targeting multiple oncogenic activities. Nat. Med. 2008, 14, 1271–1277. [Google Scholar] [CrossRef] [PubMed]
  134. Cai, C.K.; Zhao, G.Y.; Tian, L.Y.; Liu, L.; Yan, K.; Ma, Y.L.; Ji, Z.W.; Li, X.X.; Han, K.; Gao, J.; et al. miR-15a and miR-16-1 downregulate CCND1 and induce apoptosis and cell cycle arrest in osteosarcoma. Oncol. Rep. 2012, 28, 1764–1770. [Google Scholar] [CrossRef] [PubMed]
  135. Bandi, N.; Vassella, E. miR-34a and miR-15a/16 are co-regulated in non-small cell lung cancer and control cell cycle progression in a synergistic and Rb-dependent manner. Mol. Cancer 2011, 10, 55. [Google Scholar] [CrossRef] [PubMed]
  136. Schultz, J.; Lorenz, P.; Gross, G.; Ibrahim, S.; Kunz, M. MicroRNA let-7b targets important cell cycle molecules in malignant melanoma cells and interferes with anchorage-independent growth. Cell Res. 2008, 18, 549–557. [Google Scholar] [CrossRef] [PubMed]
  137. Johnson, C.D.; Esquela-Kerscher, A.; Stefani, G.; Byrom, M.; Kelnar, K.; Ovcharenko, D.; Wilson, M.; Wang, X.; Shelton, J.; Shingara, J.; et al. The let-7 microRNA represses cell proliferation pathways in human cells. Cancer Res. 2007, 67, 7713–7722. [Google Scholar] [CrossRef]
  138. Han, Z.; Li, Q.; Wang, Y.; Wang, L.; Li, X.; Ge, N.; Wang, Y.; Guo, C. Niclosamide Induces Cell Cycle Arrest in G1 Phase in Head and Neck Squamous Cell Carcinoma Through Let-7d/CDC34 Axis. Front. Pharmacol. 2018, 9, 1544. [Google Scholar] [CrossRef]
  139. Yin, J.; Hu, W.; Pan, L.; Fu, W.; Dai, L.; Jiang, Z.; Zhang, F.; Zhao, J. let-7 and miR-17 promote self-renewal and drive gefitinib resistance in non-small cell lung cancer. Oncol. Rep. 2019, 42, 495–508. [Google Scholar] [CrossRef]
  140. Zhang, X.; Hu, S.; Zhang, X.; Wang, L.; Zhang, X.; Yan, B.; Zhao, J.; Yang, A.; Zhang, R. MicroRNA-7 arrests cell cycle in G1 phase by directly targeting CCNE1 in human hepatocellular carcinoma cells. Biochem. Biophys. Res. Commun. 2014, 443, 1078–1084. [Google Scholar] [CrossRef]
  141. Sanchez, N.; Gallagher, M.; Lao, N.; Gallagher, C.; Clarke, C.; Doolan, P.; Aherne, S.; Blanco, A.; Meleady, P.; Clynes, M.; et al. MiR-7 triggers cell cycle arrest at the G1/S transition by targeting multiple genes including Skp2 and Psme3. PLoS ONE 2013, 8, e65671. [Google Scholar] [CrossRef]
  142. Wu, J.; Jiang, Y.; Cao, W.; Li, X.; Xie, C.; Geng, S.; Zhu, M.; Liang, Z.; Zhu, J.; Zhu, W.; et al. miR-19 targeting of PTEN mediates butyl benzyl phthalate-induced proliferation in both ER(+) and ER(-) breast cancer cells. Toxicol. Lett. 2018, 295, 124–133. [Google Scholar] [CrossRef]
  143. Hackl, M.; Brunner, S.; Fortschegger, K.; Schreiner, C.; Micutkova, L.; Mück, C.; Laschober, G.T.; Lepperdinger, G.; Sampson, N.; Berger, P.; et al. miR-17, miR-19b, miR-20a, and miR-106a are down-regulated in human aging. Aging Cell 2010, 9, 291–296. [Google Scholar] [CrossRef] [PubMed]
  144. Yu, Z.; Wang, C.; Wang, M.; Li, Z.; Casimiro, M.C.; Liu, M.; Wu, K.; Whittle, J.; Ju, X.; Hyslop, T.; et al. A cyclin D1/microRNA 17/20 regulatory feedback loop in control of breast cancer cell proliferation. J. Cell Biol. 2008, 182, 509–517. [Google Scholar] [CrossRef] [PubMed]
  145. Sokolova, V.; Fiorino, A.; Zoni, E.; Crippa, E.; Reid, J.F.; Gariboldi, M.; Pierotti, M.A. The Effects of miR-20a on p21: Two Mechanisms Blocking Growth Arrest in TGF-β-Responsive Colon Carcinoma. J. Cell. Physiol. 2015, 230, 3105–3114. [Google Scholar] [CrossRef] [PubMed]
  146. Sylvestre, Y.; De Guire, V.; Querido, E.; Mukhopadhyay, U.K.; Bourdeau, V.; Major, F.; Ferbeyre, G.; Chartrand, P. An E2F/miR-20a autoregulatory feedback loop. J. Biol. Chem. 2007, 282, 2135–2143. [Google Scholar] [CrossRef]
  147. Qin, Y.E.; Duan, L.; He, Y.; Yuan, C.; Wang, T.; Yuan, D.; Zhang, C.; Liu, C. Saturated Fatty Acids Promote Hepatocytic Senecence through Regulation of miR-34a/Cyclin-Dependent Kinase 6. Mol. Nutr. Food Res. 2020, 64, e2000383. [Google Scholar] [CrossRef]
  148. Zhao, J.; Lammers, P.; Torrance, C.J.; Bader, A.G. TP53-independent function of miR-34a via HDAC1 and p21(CIP1/WAF1.). Mol. Ther. 2013, 21, 1678–1686. [Google Scholar] [CrossRef]
  149. Agirre, X.; Vilas-Zornoza, A.; Jiménez-Velasco, A.; Martin-Subero, J.I.; Cordeu, L.; Gárate, L.; San José-Eneriz, E.; Abizanda, G.; Rodríguez-Otero, P.; Fortes, P.; et al. Epigenetic silencing of the tumor suppressor microRNA Hsa-miR-124a regulates CDK6 expression and confers a poor prognosis in acute lymphoblastic leukemia. Cancer Res. 2009, 69, 4443–4453. [Google Scholar] [CrossRef]
  150. Nakamachi, Y.; Kawano, S.; Takenokuchi, M.; Nishimura, K.; Sakai, Y.; Chin, T.; Saura, R.; Kurosaka, M.; Kumagai, S. MicroRNA-124a is a key regulator of proliferation and monocyte chemoattractant protein 1 secretion in fibroblast-like synoviocytes from patients with rheumatoid arthritis. Arthritis Rheum. 2009, 60, 1294–1304. [Google Scholar] [CrossRef]
  151. Kawano, S.; Nakamachi, Y. miR-124a as a key regulator of proliferation and MCP-1 secretion in synoviocytes from patients with rheumatoid arthritis. Ann. Rheum. Dis. 2011, 70 (Suppl. 1), i88–i91. [Google Scholar] [CrossRef]
  152. Kim, J.K.; Noh, J.H.; Jung, K.H.; Eun, J.W.; Bae, H.J.; Kim, M.G.; Chang, Y.G.; Shen, Q.; Park, W.S.; Lee, J.Y.; et al. Sirtuin7 oncogenic potential in human hepatocellular carcinoma and its regulation by the tumor suppressors MiR-125a-5p and MiR-125b. Hepatology 2013, 57, 1055–1067. [Google Scholar] [CrossRef] [PubMed]
  153. Shi, L.; Zhang, J.; Pan, T.; Zhou, J.; Gong, W.; Liu, N.; Fu, Z.; You, Y. MiR-125b is critical for the suppression of human U251 glioma stem cell proliferation. Brain Res. 2010, 1312, 120–126. [Google Scholar] [CrossRef] [PubMed]
  154. Huang, L.; Luo, J.; Cai, Q.; Pan, Q.; Zeng, H.; Guo, Z.; Dong, W.; Huang, J.; Lin, T. MicroRNA-125b suppresses the development of bladder cancer by targeting E2F3. Int. J. Cancer 2011, 128, 1758–1769. [Google Scholar] [CrossRef] [PubMed]
  155. Lynch, S.M.; McKenna, M.M.; Walsh, C.P.; McKenna, D.J. miR-24 regulates CDKN1B/p27 expression in prostate cancer. Prostate 2016, 76, 637–648. [Google Scholar] [CrossRef]
  156. Lu, J.; He, M.L.; Wang, L.; Chen, Y.; Liu, X.; Dong, Q.; Chen, Y.C.; Peng, Y.; Yao, K.T.; Kung, H.F.; et al. MiR-26a inhibits cell growth and tumorigenesis of nasopharyngeal carcinoma through repression of EZH2. Cancer Res. 2011, 71, 225–233. [Google Scholar] [CrossRef]
  157. Trohatou, O.; Zagoura, D.; Orfanos, N.K.; Pappa, K.I.; Marinos, E.; Anagnou, N.P.; Roubelakis, M.G. miR-26a Mediates Adipogenesis of Amniotic Fluid Mesenchymal Stem/Stromal Cells via PTEN, Cyclin E1, and CDK6. Stem Cells Dev. 2017, 26, 482–494. [Google Scholar] [CrossRef]
  158. Chen, L.; Zheng, J.; Zhang, Y.; Yang, L.; Wang, J.; Ni, J.; Cui, D.; Yu, C.; Cai, Z. Tumor-specific expression of microRNA-26a suppresses human hepatocellular carcinoma growth via cyclin-dependent and -independent pathways. Mol. Ther. 2011, 19, 1521–1528. [Google Scholar] [CrossRef]
  159. Zhou, X.; Xia, Y.; Li, L.; Zhang, G. MiR-101 inhibits cell growth and tumorigenesis of Helicobacter pylori related gastric cancer by repression of SOCS2. Cancer Biol. Ther. 2015, 16, 160–169. [Google Scholar] [CrossRef]
  160. Li, M.; Tian, L.; Ren, H.; Chen, X.; Wang, Y.; Ge, J.; Wu, S.; Sun, Y.; Liu, M.; Xiao, H. MicroRNA-101 is a potential prognostic indicator of laryngeal squamous cell carcinoma and modulates CDK8. J. Transl. Med. 2015, 13, 271. [Google Scholar] [CrossRef]
  161. Shang, A.; Lu, W.-Y.; Yang, M.; Zhou, C.; Zhang, H.; Cai, Z.-X.; Wang, W.-W.; Wang, W.-X.; Wu, G.-Q. miR-9 induces cell arrest and apoptosis of oral squamous cell carcinoma via CDK 4/6 pathway. Artif. Cells Nanomed. Biotechnol. 2018, 46, 1754–1762. [Google Scholar] [CrossRef]
  162. O’Loghlen, A.; Brookes, S.; Martin, N.; Rapisarda, V.; Peters, G.; Gil, J. CBX7 and miR-9 are part of an autoregulatory loop controlling p16(INK) (4a). Aging Cell 2015, 14, 1113–1121. [Google Scholar] [CrossRef] [PubMed]
  163. Gao, S.; Wang, J.; Tian, S.; Luo, J. miR-9 depletion suppresses the proliferation of osteosarcoma cells by targeting p16. Int. J. Oncol. 2019, 54, 1921–1932. [Google Scholar] [CrossRef] [PubMed]
  164. Jun, G.J.; Zhong, G.G.; Ming, Z.S. miR-218 inhibits the proliferation of glioma U87 cells through the inactivation of the CDK6/cyclin D1/p21(Cip1/Waf1) pathway. Oncol. Lett. 2015, 9, 2743–2749. [Google Scholar] [CrossRef] [PubMed]
  165. Deng, M.; Zeng, C.; Lu, X.; He, X.; Zhang, R.; Qiu, Q.; Zheng, G.; Jia, X.; Liu, H.; He, Z. miR-218 suppresses gastric cancer cell cycle progression through the CDK6/Cyclin D1/E2F1 axis in a feedback loop. Cancer Lett. 2017, 403, 175–185. [Google Scholar] [CrossRef]
  166. Giannakakis, A.; Sandaltzopoulos, R.; Greshock, J.; Liang, S.; Huang, J.; Hasegawa, K.; Li, C.; O’Brien-Jenkins, A.; Katsaros, D.; Weber, B.L.; et al. miR-210 links hypoxia with cell cycle regulation and is deleted in human epithelial ovarian cancer. Cancer Biol. Ther. 2008, 7, 255–264. [Google Scholar] [CrossRef]
  167. Kang, H.; Yu, H.; Zeng, L.; Ma, H.; Cao, G. LncRNA Rian reduces cardiomyocyte pyroptosis and alleviates myocardial ischemia-reperfusion injury by regulating by the miR-17-5p/CCND1 axis. Hypertens. Res. 2022, 45, 976–989. [Google Scholar] [CrossRef]
  168. Wang, Q.; Han, J.; Xu, P.; Jian, X.; Huang, X.; Liu, D. Silencing of LncRNA SNHG16 Downregulates Cyclin D1 (CCND1) to Abrogate Malignant Phenotypes in Oral Squamous Cell Carcinoma (OSCC) Through Upregulating miR-17-5p. Cancer Manag. Res. 2021, 13, 1831–1841. [Google Scholar] [CrossRef]
  169. Cloonan, N.; Brown, M.K.; Steptoe, A.L.; Wani, S.; Chan, W.L.; Forrest, A.R.; Kolle, G.; Gabrielli, B.; Grimmond, S.M. The miR-17-5p microRNA is a key regulator of the G1/S phase cell cycle transition. Genome Biol. 2008, 9, R127. [Google Scholar] [CrossRef]
  170. Pickering, M.T.; Stadler, B.M.; Kowalik, T.F. miR-17 and miR-20a temper an E2F1-induced G1 checkpoint to regulate cell cycle progression. Oncogene 2009, 28, 140–145. [Google Scholar] [CrossRef]
  171. O’Donnell, K.A.; Wentzel, E.A.; Zeller, K.I.; Dang, C.V.; Mendell, J.T. c-Myc-regulated microRNAs modulate E2F1 expression. Nature 2005, 435, 839–843. [Google Scholar] [CrossRef]
  172. Trompeter, H.-I.; Abbad, H.; Iwaniuk, K.M.; Hafner, M.; Renwick, N.; Tuschl, T.; Schira, J.; Mueller, H.W.; Wernet, P. MicroRNAs MiR-17, MiR-20a, and MiR-106b Act in Concert to Modulate E2F Activity on Cell Cycle Arrest during Neuronal Lineage Differentiation of USSC. PLoS ONE 2011, 6, e16138. [Google Scholar] [CrossRef] [PubMed]
  173. Xue, Y.; Zhu, X.; Meehan, B.; Venneti, S.; Martinez, D.; Morin, G.; Maïga, R.I.; Chen, H.; Papadakis, A.I.; Johnson, R.M.; et al. SMARCB1 loss induces druggable cyclin D1 deficiency via upregulation of MIR17HG in atypical teratoid rhabdoid tumors. J. Pathol. 2020, 252, 77–87. [Google Scholar] [CrossRef] [PubMed]
  174. Wong, P.; Iwasaki, M.; Somervaille, T.C.; Ficara, F.; Carico, C.; Arnold, C.; Chen, C.Z.; Cleary, M.L. The miR-17-92 microRNA polycistron regulates MLL leukemia stem cell potential by modulating p21 expression. Cancer Res. 2010, 70, 3833–3842. [Google Scholar] [CrossRef] [PubMed]
  175. Mi, S.; Li, Z.; Chen, P.; He, C.; Cao, D.; Elkahloun, A.; Lu, J.; Pelloso, L.A.; Wunderlich, M.; Huang, H.; et al. Aberrant overexpression and function of the miR-17-92 cluster in MLL-rearranged acute leukemia. Proc. Natl. Acad. Sci. USA 2010, 107, 3710–3715. [Google Scholar] [CrossRef]
  176. Brockway, S.; Zeleznik-Le, N.J. WEE1 is a validated target of the microRNA miR-17-92 cluster in leukemia. Cancer Genet. 2015, 208, 279–287. [Google Scholar] [CrossRef]
  177. Tai, L.; Huang, C.J.; Choo, K.B.; Cheong, S.K.; Kamarul, T. Oxidative Stress Down-Regulates MiR-20b-5p, MiR-106a-5p and E2F1 Expression to Suppress the G1/S Transition of the Cell Cycle in Multipotent Stromal Cells. Int. J. Med. Sci. 2020, 17, 457–470. [Google Scholar] [CrossRef]
  178. Zhang, A.; Hao, J.; Wang, K.; Huang, Q.; Yu, K.; Kang, C.; Wang, G.; Jia, Z.; Han, L.; Pu, P. Down-regulation of miR-106b suppresses the growth of human glioma cells. J. Neuro-Oncol. 2013, 112, 179–189. [Google Scholar] [CrossRef]
  179. Kan, T.; Sato, F.; Ito, T.; Matsumura, N.; David, S.; Cheng, Y.; Agarwal, R.; Paun, B.C.; Jin, Z.; Olaru, A.V.; et al. The miR-106b-25 polycistron, activated by genomic amplification, functions as an oncogene by suppressing p21 and Bim. Gastroenterology 2009, 136, 1689–1700. [Google Scholar] [CrossRef]
  180. Gibcus, J.H.; Kroesen, B.J.; Koster, R.; Halsema, N.; de Jong, D.; de Jong, S.; Poppema, S.; Kluiver, J.; Diepstra, A.; van den Berg, A. MiR-17/106b seed family regulates p21 in Hodgkin’s lymphoma. J. Pathol. 2011, 225, 609–617. [Google Scholar] [CrossRef]
  181. Lee, K.H.; Chen, Y.L.; Yeh, S.D.; Hsiao, M.; Lin, J.T.; Goan, Y.G.; Lu, P.J. MicroRNA-330 acts as tumor suppressor and induces apoptosis of prostate cancer cells through E2F1-mediated suppression of Akt phosphorylation. Oncogene 2009, 28, 3360–3370. [Google Scholar] [CrossRef]
  182. Guo, X.; Guo, L.; Ji, J.; Zhang, J.; Zhang, J.; Chen, X.; Cai, Q.; Li, J.; Gu, Q.; Liu, B.; et al. miRNA-331-3p directly targets E2F1 and induces growth arrest in human gastric cancer. Biochem. Biophys. Res. Commun. 2010, 398, 1–6. [Google Scholar] [CrossRef] [PubMed]
  183. Lizé, M.; Pilarski, S.; Dobbelstein, M. E2F1-inducible microRNA 449a/b suppresses cell proliferation and promotes apoptosis. Cell Death Differ. 2010, 17, 452–458. [Google Scholar] [CrossRef] [PubMed]
  184. Fang, Y.; Gu, X.; Li, Z.; Xiang, J.; Chen, Z. miR-449b inhibits the proliferation of SW1116 colon cancer stem cells through downregulation of CCND1 and E2F3 expression. Oncol. Rep. 2013, 30, 399–406. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, J.; Li, B.; Wang, C.; Luo, Y.; Zhao, M.; Chen, P. Long noncoding RNA FOXD2-AS1 promotes glioma cell cycle progression and proliferation through the FOXD2-AS1/miR-31/CDK1 pathway. J. Cell. Biochem. 2019, 120, 19784–19795. [Google Scholar] [CrossRef] [PubMed]
  186. Li, Y.; Quan, J.; Chen, F.; Pan, X.; Zhuang, C.; Xiong, T.; Zhuang, C.; Li, J.; Huang, X.; Ye, J.; et al. MiR-31-5p acts as a tumor suppressor in renal cell carcinoma by targeting cyclin-dependent kinase 1 (CDK1). Biomed. Pharmacother. 2019, 111, 517–526. [Google Scholar] [CrossRef]
  187. Malhas, A.; Saunders, N.J.; Vaux, D.J. The nuclear envelope can control gene expression and cell cycle progression via miRNA regulation. Cell Cycle 2010, 9, 531–539. [Google Scholar] [CrossRef]
  188. Zhu, W.; Yu, Y.; Fang, K.; Xiao, S.; Ni, L.; Yin, C.; Huang, X.; Wang, X.; Zhang, Y.; Le, H.B.; et al. miR-31/QKI-5 axis facilitates cell cycle progression of non-small-cell lung cancer cells by interacting and regulating p21 and CDK4/6 expressions. Cancer Med. 2023, 12, 4590–4604. [Google Scholar] [CrossRef]
  189. Chu, K.; Gao, G.; Yang, X.; Ren, S.; Li, Y.; Wu, H.; Huang, Y.; Zhou, C. miR-512-5p induces apoptosis and inhibits glycolysis by targeting p21 in non-small cell lung cancer cells. Int. J. Oncol. 2016, 48, 577–586. [Google Scholar] [CrossRef]
  190. Wu, Z.; Sun, H.; Zeng, W.; He, J.; Mao, X. Upregulation of MircoRNA-370 Induces Proliferation in Human Prostate Cancer Cells by Downregulating the Transcription Factor FOXO1. PLoS ONE 2012, 7, e45825. [Google Scholar] [CrossRef]
  191. Jiping, Z.; Ming, F.; Lixiang, W.; Xiuming, L.; Yuqun, S.; Han, Y.; Zhifang, L.; Yundong, S.; Shili, L.; Chunyan, C.; et al. MicroRNA-212 inhibits proliferation of gastric cancer by directly repressing retinoblastoma binding protein 2. J. Cell Biochem. 2013, 114, 2666–2672. [Google Scholar] [CrossRef]
  192. Zhao, J.L.; Zhang, L.; Guo, X.; Wang, J.H.; Zhou, W.; Liu, M.; Li, X.; Tang, H. miR-212/132 downregulates SMAD2 expression to suppress the G1/S phase transition of the cell cycle and the epithelial to mesenchymal transition in cervical cancer cells. IUBMB Life 2015, 67, 380–394. [Google Scholar] [CrossRef] [PubMed]
  193. Sun, M.; Jin, F.Y.; Xia, R.; Kong, R.; Li, J.H.; Xu, T.P.; Liu, Y.W.; Zhang, E.B.; Liu, X.H.; De, W. Decreased expression of long noncoding RNA GAS5 indicates a poor prognosis and promotes cell proliferation in gastric cancer. BMC Cancer 2014, 14, 319. [Google Scholar] [CrossRef]
  194. Delgir, S.; Ilkhani, K.; Safi, A.; Rahmati, Y.; Montazari, V.; Zaynali-Khasraghi, Z.; Seif, F.; Bastami, M.; Alivand, M.R. The expression of miR-513c and miR-3163 was downregulated in tumor tissues compared with normal adjacent tissue of patients with breast cancer. BMC Med. Genom. 2021, 14, 180. [Google Scholar] [CrossRef]
  195. Miller, T.E.; Ghoshal, K.; Ramaswamy, B.; Roy, S.; Datta, J.; Shapiro, C.L.; Jacob, S.; Majumder, S. MicroRNA-221/222 confers tamoxifen resistance in breast cancer by targeting p27Kip1. J. Biol. Chem. 2008, 283, 29897–29903. [Google Scholar] [CrossRef]
  196. Yamashita, R.; Sato, M.; Kakumu, T.; Hase, T.; Yogo, N.; Maruyama, E.; Sekido, Y.; Kondo, M.; Hasegawa, Y. Growth inhibitory effects of miR-221 and miR-222 in non-small cell lung cancer cells. Cancer Med. 2015, 4, 551–564. [Google Scholar] [CrossRef] [PubMed]
  197. Chen, Q.; Wang, W.; Wang, Y. MiR-222 regulates the progression of oral squamous cell carcinoma by targeting CDKN1B. Am. J. Transl. Res. 2022, 14, 5215–5227. [Google Scholar] [PubMed]
  198. Liao, L.; Chen, J.; Zhang, C.; Guo, Y.; Liu, W.; Liu, W.; Duan, L.; Liu, Z.; Hu, J.; Lu, J. LncRNA NEAT1 Promotes High Glucose-Induced Mesangial Cell Hypertrophy by Targeting miR-222-3p/CDKN1B Axis. Front. Mol. Biosci. 2020, 7, 627827. [Google Scholar] [CrossRef]
  199. Zhao, Y.; Wang, Y.; Yang, Y.; Liu, J.; Song, Y.; Cao, Y.; Chen, X.; Yang, W.; Wang, F.; Gao, J.; et al. MicroRNA-222 Controls Human Pancreatic Cancer Cell Line Capan-2 Proliferation by P57 Targeting. J. Cancer 2015, 6, 1230–1235. [Google Scholar] [CrossRef]
  200. Im, W.R.; Lee, H.-S.; Lee, Y.-S.; Lee, J.-S.; Jang, H.-J.; Kim, S.-Y.; Park, J.-L.; Lee, Y.; Kim, M.S.; Lee, J.M.; et al. A Regulatory Noncoding RNA, nc886, Suppresses Esophageal Cancer by Inhibiting the AKT Pathway and Cell Cycle Progression. Cells 2020, 9, 801. [Google Scholar] [CrossRef]
  201. Fort, R.S.; Mathó, C.; Geraldo, M.V.; Ottati, M.C.; Yamashita, A.S.; Saito, K.C.; Leite, K.R.M.; Méndez, M.; Maedo, N.; Méndez, L.; et al. Nc886 is epigenetically repressed in prostate cancer and acts as a tumor suppressor through the inhibition of cell growth. BMC Cancer 2018, 18, 127. [Google Scholar] [CrossRef]
  202. Zhao, J.J.; Lin, J.; Lwin, T.; Yang, H.; Guo, J.; Kong, W.; Dessureault, S.; Moscinski, L.C.; Rezania, D.; Dalton, W.S.; et al. microRNA expression profile and identification of miR-29 as a prognostic marker and pathogenetic factor by targeting CDK6 in mantle cell lymphoma. Blood 2010, 115, 2630–2639. [Google Scholar] [CrossRef] [PubMed]
  203. Ji, W.; Zhang, W.; Wang, X.; Shi, Y.; Yang, F.; Xie, H.; Zhou, W.; Wang, S.; Guan, X. c-myc regulates the sensitivity of breast cancer cells to palbociclib via c-myc/miR-29b-3p/CDK6 axis. Cell Death Dis. 2020, 11, 760. [Google Scholar] [CrossRef] [PubMed]
  204. Ji, L.Y.; Wei, M.; Liu, Y.Y.; Di, Z.L.; Li, S.Z. miR-497/MIR497HG inhibits glioma cell proliferation by targeting CCNE1 and the miR-588/TUSC1 axis. Oncol. Rep. 2021, 46, 255. [Google Scholar] [CrossRef] [PubMed]
  205. Hoareau-Aveilla, C.; Quelen, C.; Congras, A.; Caillet, N.; Labourdette, D.; Dozier, C.; Brousset, P.; Lamant, L.; Meggetto, F. miR-497 suppresses cycle progression through an axis involving CDK6 in ALK-positive cells. Haematologica 2019, 104, 347–359. [Google Scholar] [CrossRef]
  206. Chen, J.; Feilotter, H.E.; Paré, G.C.; Zhang, X.; Pemberton, J.G.; Garady, C.; Lai, D.; Yang, X.; Tron, V.A. MicroRNA-193b represses cell proliferation and regulates cyclin D1 in melanoma. Am. J. Pathol. 2010, 176, 2520–2529. [Google Scholar] [CrossRef]
  207. Andrikopoulou, A.; Shalit, A.; Zografos, E.; Koutsoukos, K.; Korakiti, A.-M.; Liontos, M.; Dimopoulos, M.-A.; Zagouri, F. MicroRNAs as Potential Predictors of Response to CDK4/6 Inhibitor Treatment. Cancers 2021, 13, 4114. [Google Scholar] [CrossRef]
  208. Gao, J.; Ma, S.; Yang, F.; Chen, X.; Wang, W.; Zhang, J.; Li, Y.; Wang, T.; Shan, L. miR-193b exhibits mutual interaction with MYC, and suppresses growth and metastasis of osteosarcoma. Oncol. Rep. 2020, 44, 139–155. [Google Scholar] [CrossRef]
  209. Kaukoniemi, K.M.; Rauhala, H.E.; Scaravilli, M.; Latonen, L.; Annala, M.; Vessella, R.L.; Nykter, M.; Tammela, T.L.; Visakorpi, T. Epigenetically altered miR-193b targets cyclin D1 in prostate cancer. Cancer Med. 2015, 4, 1417–1425. [Google Scholar] [CrossRef]
  210. Bustos, M.A.; Ono, S.; Marzese, D.M.; Oyama, T.; Iida, Y.; Cheung, G.; Nelson, N.; Hsu, S.C.; Yu, Q.; Hoon, D.S.B. MiR-200a Regulates CDK4/6 Inhibitor Effect by Targeting CDK6 in Metastatic Melanoma. J. Invest. Dermatol. 2017, 137, 1955–1964. [Google Scholar] [CrossRef]
  211. Georgantas, R.W., 3rd; Streicher, K.; Luo, X.; Greenlees, L.; Zhu, W.; Liu, Z.; Brohawn, P.; Morehouse, C.; Higgs, B.W.; Richman, L.; et al. MicroRNA-206 induces G1 arrest in melanoma by inhibition of CDK4 and Cyclin D. Pigment. Cell Melanoma Res. 2014, 27, 275–286. [Google Scholar] [CrossRef]
  212. Dahiya, N.; Sherman-Baust, C.A.; Wang, T.L.; Davidson, B.; Shih Ie, M.; Zhang, Y.; Wood, W., 3rd; Becker, K.G.; Morin, P.J. MicroRNA expression and identification of putative miRNA targets in ovarian cancer. PLoS ONE 2008, 3, e2436. [Google Scholar] [CrossRef] [PubMed]
  213. Xu, J.; Yao, Q.; Hou, Y.; Xu, M.; Liu, S.; Yang, L.; Zhang, L.; Xu, H. MiR-223/Ect2/p21 signaling regulates osteosarcoma cell cycle progression and proliferation. Biomed. Pharmacother. 2013, 67, 381–386. [Google Scholar] [CrossRef] [PubMed]
  214. Armenia, J.; Fabris, L.; Lovat, F.; Berton, S.; Segatto, I.; D’Andrea, S.; Ivan, C.; Cascione, L.; Calin, G.A.; Croce, C.M.; et al. Contact inhibition modulates intracellular levels of miR-223 in a p27kip1-dependent manner. Oncotarget 2014, 5, 1185–1197. [Google Scholar] [CrossRef]
  215. Wang, H.; Guo, Q.; Yang, P.; Long, G. Restoration of microRNA-212 causes a G0/G1 cell cycle arrest and apoptosis in adult T-cell leukemia/lymphoma cells by repressing CCND3 expression. J. Investig. Med. 2017, 65, 82–87. [Google Scholar] [CrossRef]
  216. Guo, S.L.; Ye, H.; Teng, Y.; Wang, Y.L.; Yang, G.; Li, X.B.; Zhang, C.; Yang, X.; Yang, Z.Z.; Yang, X. Akt-p53-miR-365-cyclin D1/cdc25A axis contributes to gastric tumorigenesis induced by PTEN deficiency. Nat. Commun. 2013, 4, 2544. [Google Scholar] [CrossRef] [PubMed]
  217. Liu, Q.; Fu, H.; Sun, F.; Zhang, H.; Tie, Y.; Zhu, J.; Xing, R.; Sun, Z.; Zheng, X. miR-16 family induces cell cycle arrest by regulating multiple cell cycle genes. Nucleic Acids Res. 2008, 36, 5391–5404. [Google Scholar] [CrossRef] [PubMed]
  218. Linsley, P.S.; Schelter, J.; Burchard, J.; Kibukawa, M.; Martin, M.M.; Bartz, S.R.; Johnson, J.M.; Cummins, J.M.; Raymond, C.K.; Dai, H.; et al. Transcripts targeted by the microRNA-16 family cooperatively regulate cell cycle progression. Mol. Cell Biol. 2007, 27, 2240–2252. [Google Scholar] [CrossRef]
  219. Qin, X.; Wang, X.; Wang, Y.; Tang, Z.; Cui, Q.; Xi, J.; Li, Y.S.; Chien, S.; Wang, N. MicroRNA-19a mediates the suppressive effect of laminar flow on cyclin D1 expression in human umbilical vein endothelial cells. Proc. Natl. Acad. Sci. USA 2010, 107, 3240–3244. [Google Scholar] [CrossRef]
  220. Sun, F.; Fu, H.; Liu, Q.; Tie, Y.; Zhu, J.; Xing, R.; Sun, Z.; Zheng, X. Downregulation of CCND1 and CDK6 by miR-34a induces cell cycle arrest. FEBS Lett. 2008, 582, 1564–1568. [Google Scholar] [CrossRef]
  221. Zheng, S.Z.; Sun, P.; Wang, J.P.; Liu, Y.; Gong, W.; Liu, J. MiR-34a overexpression enhances the inhibitory effect of doxorubicin on HepG2 cells. World J. Gastroenterol. 2019, 25, 2752–2762. [Google Scholar] [CrossRef]
  222. Yang, J.; Song, Q.; Cai, Y.; Wang, P.; Wang, M.; Zhang, D. RLIP76-dependent suppression of PI3K/AKT/Bcl-2 pathway by miR-101 induces apoptosis in prostate cancer. Biochem. Biophys. Res. Commun. 2015, 463, 900–906. [Google Scholar] [CrossRef] [PubMed]
  223. Guan, H.; Dai, Z.; Ma, Y.; Wang, Z.; Liu, X.; Wang, X. MicroRNA-101 inhibits cell proliferation and induces apoptosis by targeting EYA1 in breast cancer. Int. J. Mol. Med. 2016, 37, 1643–1651. [Google Scholar] [CrossRef] [PubMed]
  224. Li, B.; Xie, D.; Zhang, H. MicroRNA-101-3p advances cisplatin sensitivity in bladder urothelial carcinoma through targeted silencing EZH2. J. Cancer 2019, 10, 2628–2634. [Google Scholar] [CrossRef] [PubMed]
  225. Aguda, B.D.; Kim, Y.; Piper-Hunter, M.G.; Friedman, A.; Marsh, C.B. MicroRNA regulation of a cancer network: Consequences of the feedback loops involving miR-17-92, E2F, and Myc. Proc. Natl. Acad. Sci. USA 2008, 105, 19678–19683. [Google Scholar] [CrossRef]
  226. Bueno, M.J.; Gómez de Cedrón, M.; Laresgoiti, U.; Fernández-Piqueras, J.; Zubiaga, A.M.; Malumbres, M. Multiple E2F-induced microRNAs prevent replicative stress in response to mitogenic signaling. Mol. Cell Biol. 2010, 30, 2983–2995. [Google Scholar] [CrossRef]
  227. Gruszka, R.; Zakrzewski, K.; Liberski, P.P.; Zakrzewska, M. mRNA and miRNA Expression Analyses of the MYC/E2F/miR-17-92 Network in the Most Common Pediatric Brain Tumors. Int. J. Mol. Sci. 2021, 22, 543. [Google Scholar] [CrossRef]
  228. Yang, X.; Feng, M.; Jiang, X.; Wu, Z.; Li, Z.; Aau, M.; Yu, Q. miR-449a and miR-449b are direct transcriptional targets of E2F1 and negatively regulate pRb-E2F1 activity through a feedback loop by targeting CDK6 and CDC25A. Genes. Dev. 2009, 23, 2388–2393. [Google Scholar] [CrossRef]
  229. Lal, A.; Navarro, F.; Maher, C.A.; Maliszewski, L.E.; Yan, N.; O’Day, E.; Chowdhury, D.; Dykxhoorn, D.M.; Tsai, P.; Hofmann, O.; et al. miR-24 Inhibits cell proliferation by targeting E2F2, MYC, and other cell-cycle genes via binding to “seedless” 3’UTR microRNA recognition elements. Mol. Cell 2009, 35, 610–625. [Google Scholar] [CrossRef]
  230. Hydbring, P.; Wang, Y.; Fassl, A.; Li, X.; Matia, V.; Otto, T.; Choi, Y.J.; Sweeney, K.E.; Suski, J.M.; Yin, H.; et al. Cell-Cycle-Targeting MicroRNAs as Therapeutic Tools against Refractory Cancers. Cancer Cell 2017, 31, 576–590.e578. [Google Scholar] [CrossRef]
  231. Glover, D.M.; Hagan, I.M.; Tavares, A.A. Polo-like kinases: A team that plays throughout mitosis. Genes. Dev. 1998, 12, 3777–3787. [Google Scholar] [CrossRef]
  232. Shi, W.; Alajez, N.M.; Bastianutto, C.; Hui, A.B.; Mocanu, J.D.; Ito, E.; Busson, P.; Lo, K.W.; Ng, R.; Waldron, J.; et al. Significance of Plk1 regulation by miR-100 in human nasopharyngeal cancer. Int. J. Cancer 2010, 126, 2036–2048. [Google Scholar] [CrossRef] [PubMed]
  233. Kim, Y.K.; Yu, J.; Han, T.S.; Park, S.Y.; Namkoong, B.; Kim, D.H.; Hur, K.; Yoo, M.W.; Lee, H.J.; Yang, H.K.; et al. Functional links between clustered microRNAs: Suppression of cell-cycle inhibitors by microRNA clusters in gastric cancer. Nucleic Acids Res. 2009, 37, 1672–1681. [Google Scholar] [CrossRef] [PubMed]
  234. Ivanovska, I.; Ball, A.S.; Diaz, R.L.; Magnus, J.F.; Kibukawa, M.; Schelter, J.M.; Kobayashi, S.V.; Lim, L.; Burchard, J.; Jackson, A.L.; et al. MicroRNAs in the miR-106b family regulate p21/CDKN1A and promote cell cycle progression. Mol. Cell Biol. 2008, 28, 2167–2174. [Google Scholar] [CrossRef] [PubMed]
  235. Lal, A.; Kim, H.H.; Abdelmohsen, K.; Kuwano, Y.; Pullmann, R., Jr.; Srikantan, S.; Subrahmanyam, R.; Martindale, J.L.; Yang, X.; Ahmed, F.; et al. p16(INK4a) translation suppressed by miR-24. PLoS ONE 2008, 3, e1864. [Google Scholar] [CrossRef] [PubMed]
  236. Giglio, S.; Cirombella, R.; Amodeo, R.; Portaro, L.; Lavra, L.; Vecchione, A. MicroRNA miR-24 promotes cell proliferation by targeting the CDKs inhibitors p27Kip1 and p16INK4a. J. Cell Physiol. 2013, 228, 2015–2023. [Google Scholar] [CrossRef]
  237. Wang, C.; Chen, Z.; Ge, Q.; Hu, J.; Li, F.; Hu, J.; Xu, H.; Ye, Z.; Li, L.C. Up-regulation of p21(WAF1/CIP1) by miRNAs and its implications in bladder cancer cells. FEBS Lett. 2014, 588, 4654–4664. [Google Scholar] [CrossRef] [PubMed]
  238. Jin, C.; Zhang, Y.; Li, J. Upregulation of MiR-196a promotes cell proliferation by downregulating p27(kip1) in laryngeal cancer. Biol. Res. 2016, 49, 40. [Google Scholar] [CrossRef] [PubMed]
  239. Visone, R.; Russo, L.; Pallante, P.; De Martino, I.; Ferraro, A.; Leone, V.; Borbone, E.; Petrocca, F.; Alder, H.; Croce, C.M.; et al. MicroRNAs (miR)-221 and miR-222, both overexpressed in human thyroid papillary carcinomas, regulate p27Kip1 protein levels and cell cycle. Endocr. Relat. Cancer 2007, 14, 791–798. [Google Scholar] [CrossRef]
  240. Yu, Z.; Willmarth, N.E.; Zhou, J.; Katiyar, S.; Wang, M.; Liu, Y.; McCue, P.A.; Quong, A.A.; Lisanti, M.P.; Pestell, R.G. microRNA 17/20 inhibits cellular invasion and tumor metastasis in breast cancer by heterotypic signaling. Proc. Natl. Acad. Sci. USA 2010, 107, 8231–8236. [Google Scholar] [CrossRef]
  241. Yu, Z.; Wang, L.; Wang, C.; Ju, X.; Wang, M.; Chen, K.; Loro, E.; Li, Z.; Zhang, Y.; Wu, K.; et al. Cyclin D1 induction of Dicer governs microRNA processing and expression in breast cancer. Nat. Commun. 2013, 4, 2812. [Google Scholar] [CrossRef]
  242. Mughal, M.J.; Bhadresha, K.; Kwok, H.F. CDK inhibitors from past to present: A new wave of cancer therapy. Semin. Cancer Biol. 2023, 88, 106–122. [Google Scholar] [CrossRef]
  243. Finn, R.S.; Martin, M.; Rugo, H.S.; Jones, S.; Im, S.A.; Gelmon, K.; Harbeck, N.; Lipatov, O.N.; Walshe, J.M.; Moulder, S.; et al. Palbociclib and Letrozole in Advanced Breast Cancer. N. Engl. J. Med. 2016, 375, 1925–1936. [Google Scholar] [CrossRef] [PubMed]
  244. Cristofanilli, M.; Turner, N.C.; Bondarenko, I.; Ro, J.; Im, S.A.; Masuda, N.; Colleoni, M.; DeMichele, A.; Loi, S.; Verma, S.; et al. Fulvestrant plus palbociclib versus fulvestrant plus placebo for treatment of hormone-receptor-positive, HER2-negative metastatic breast cancer that progressed on previous endocrine therapy (PALOMA-3): Final analysis of the multicentre, double-blind, phase 3 randomised controlled trial. Lancet Oncol. 2016, 17, 425–439. [Google Scholar] [CrossRef] [PubMed]
  245. Hortobagyi, G.N.; Stemmer, S.M.; Burris, H.A.; Yap, Y.S.; Sonke, G.S.; Paluch-Shimon, S.; Campone, M.; Blackwell, K.L.; André, F.; Winer, E.P.; et al. Ribociclib as First-Line Therapy for HR-Positive, Advanced Breast Cancer. N. Engl. J. Med. 2016, 375, 1738–1748. [Google Scholar] [CrossRef] [PubMed]
  246. Duso, B.A.; Trapani, D.; Viale, G.; Criscitiello, C.; D’Amico, P.; Belli, C.; Mazzarella, L.; Locatelli, M.; Minchella, I.; Curigliano, G. Clinical efficacy of ribociclib as a first-line therapy for HR-positive, advanced breast cancer. Expert. Opin. Pharmacother. 2018, 19, 299–305. [Google Scholar] [CrossRef] [PubMed]
  247. Sledge, G.W., Jr.; Toi, M.; Neven, P.; Sohn, J.; Inoue, K.; Pivot, X.; Burdaeva, O.; Okera, M.; Masuda, N.; Kaufman, P.A.; et al. MONARCH 2: Abemaciclib in Combination With Fulvestrant in Women With HR+/HER2- Advanced Breast Cancer Who Had Progressed While Receiving Endocrine Therapy. J. Clin. Oncol. 2017, 35, 2875–2884. [Google Scholar] [CrossRef]
  248. Rader, J.; Russell, M.R.; Hart, L.S.; Nakazawa, M.S.; Belcastro, L.T.; Martinez, D.; Li, Y.; Carpenter, E.L.; Attiyeh, E.F.; Diskin, S.J.; et al. Dual CDK4/CDK6 inhibition induces cell-cycle arrest and senescence in neuroblastoma. Clin. Cancer Res. 2013, 19, 6173–6182. [Google Scholar] [CrossRef]
  249. Puyol, M.; Martín, A.; Dubus, P.; Mulero, F.; Pizcueta, P.; Khan, G.; Guerra, C.; Santamaría, D.; Barbacid, M. A synthetic lethal interaction between K-Ras oncogenes and Cdk4 unveils a therapeutic strategy for non-small cell lung carcinoma. Cancer Cell 2010, 18, 63–73. [Google Scholar] [CrossRef]
  250. Finn, R.S.; Dering, J.; Conklin, D.; Kalous, O.; Cohen, D.J.; Desai, A.J.; Ginther, C.; Atefi, M.; Chen, I.; Fowst, C.; et al. PD 0332991, a selective cyclin D kinase 4/6 inhibitor, preferentially inhibits proliferation of luminal estrogen receptor-positive human breast cancer cell lines in vitro. Breast Cancer Res. 2009, 11, R77. [Google Scholar] [CrossRef]
  251. Lu, J. Palbociclib: A first-in-class CDK4/CDK6 inhibitor for the treatment of hormone-receptor positive advanced breast cancer. J. Hematol. Oncol. 2015, 8, 98. [Google Scholar] [CrossRef]
  252. Fry, D.W.; Harvey, P.J.; Keller, P.R.; Elliott, W.L.; Meade, M.; Trachet, E.; Albassam, M.; Zheng, X.; Leopold, W.R.; Pryer, N.K.; et al. Specific inhibition of cyclin-dependent kinase 4/6 by PD 0332991 and associated antitumor activity in human tumor xenografts. Mol. Cancer Ther. 2004, 3, 1427–1438. [Google Scholar] [CrossRef] [PubMed]
  253. Whiteway, S.L.; Harris, P.S.; Venkataraman, S.; Alimova, I.; Birks, D.K.; Donson, A.M.; Foreman, N.K.; Vibhakar, R. Inhibition of cyclin-dependent kinase 6 suppresses cell proliferation and enhances radiation sensitivity in medulloblastoma cells. J. Neurooncol. 2013, 111, 113–121. [Google Scholar] [CrossRef] [PubMed]
  254. Rascon, K.; Flajc, G.; De Angelis, C.; Liu, X.; Trivedi, M.V.; Ekinci, E. Ribociclib in HR+/HER2− Advanced or Metastatic Breast Cancer Patients. Ann. Pharmacother. 2019, 53, 501–509. [Google Scholar] [CrossRef]
  255. Jansen, V.M.; Bhola, N.E.; Bauer, J.A.; Formisano, L.; Lee, K.M.; Hutchinson, K.E.; Witkiewicz, A.K.; Moore, P.D.; Estrada, M.V.; Sánchez, V.; et al. Kinome-Wide RNA Interference Screen Reveals a Role for PDK1 in Acquired Resistance to CDK4/6 Inhibition in ER-Positive Breast Cancer. Cancer Res. 2017, 77, 2488–2499. [Google Scholar] [CrossRef]
  256. Wu, T.; Chen, Z.; To, K.K.W.; Fang, X.; Wang, F.; Cheng, B.; Fu, L. Effect of abemaciclib (LY2835219) on enhancement of chemotherapeutic agents in ABCB1 and ABCG2 overexpressing cells in vitro and in vivo. Biochem. Pharmacol. 2017, 124, 29–42. [Google Scholar] [CrossRef]
  257. Hortobagyi, G.N.; Stemmer, S.M.; Burris, H.A.; Yap, Y.S.; Sonke, G.S.; Paluch-Shimon, S.; Campone, M.; Petrakova, K.; Blackwell, K.L.; Winer, E.P.; et al. Updated results from MONALEESA-2, a phase III trial of first-line ribociclib plus letrozole versus placebo plus letrozole in hormone receptor-positive, HER2-negative advanced breast cancer. Ann. Oncol. 2018, 29, 1541–1547. [Google Scholar] [CrossRef] [PubMed]
  258. Goetz, M.P.; Toi, M.; Campone, M.; Sohn, J.; Paluch-Shimon, S.; Huober, J.; Park, I.H.; Trédan, O.; Chen, S.C.; Manso, L.; et al. MONARCH 3: Abemaciclib As Initial Therapy for Advanced Breast Cancer. J. Clin. Oncol. 2017, 35, 3638–3646. [Google Scholar] [CrossRef] [PubMed]
  259. Johnston, S.; Martin, M.; Di Leo, A.; Im, S.A.; Awada, A.; Forrester, T.; Frenzel, M.; Hardebeck, M.C.; Cox, J.; Barriga, S.; et al. MONARCH 3 final PFS: A randomized study of abemaciclib as initial therapy for advanced breast cancer. NPJ Breast Cancer 2019, 5, 5. [Google Scholar] [CrossRef]
  260. Gao, J.J.; Cheng, J.; Bloomquist, E.; Sanchez, J.; Wedam, S.B.; Singh, H.; Amiri-Kordestani, L.; Ibrahim, A.; Sridhara, R.; Goldberg, K.B.; et al. CDK4/6 inhibitor treatment for patients with hormone receptor-positive, HER2-negative, advanced or metastatic breast cancer: A US Food and Drug Administration pooled analysis. Lancet Oncol. 2020, 21, 250–260. [Google Scholar] [CrossRef]
  261. Gao, J.J.; Cheng, J.; Prowell, T.M.; Bloomquist, E.; Tang, S.; Wedam, S.B.; Royce, M.; Krol, D.; Osgood, C.; Ison, G.; et al. Overall survival in patients with hormone receptor-positive, HER2-negative, advanced or metastatic breast cancer treated with a cyclin-dependent kinase 4/6 inhibitor plus fulvestrant: A US Food and Drug Administration pooled analysis. Lancet Oncol. 2021, 22, 1573–1581. [Google Scholar] [CrossRef]
  262. Bolzacchini, E.; Pomero, F.; Fazio, M.; Civitelli, C.; Fabro, G.; Pellegrino, D.; Giordano, M.; Squizzato, A. Risk of venous and arterial thromboembolic events in women with advanced breast cancer treated with CDK 4/6 inhibitors: A systematic review and meta-analysis. Thromb. Res. 2021, 208, 190–197. [Google Scholar] [CrossRef] [PubMed]
  263. Pandey, K.; An, H.J.; Kim, S.K.; Lee, S.A.; Kim, S.; Lim, S.M.; Kim, G.M.; Sohn, J.; Moon, Y.W. Molecular mechanisms of resistance to CDK4/6 inhibitors in breast cancer: A review. Int. J. Cancer 2019, 145, 1179–1188. [Google Scholar] [CrossRef] [PubMed]
  264. McCartney, A.; Migliaccio, I.; Bonechi, M.; Biagioni, C.; Romagnoli, D.; De Luca, F.; Galardi, F.; Risi, E.; De Santo, I.; Benelli, M.; et al. Mechanisms of Resistance to CDK4/6 Inhibitors: Potential Implications and Biomarkers for Clinical Practice. Front. Oncol. 2019, 9, 666. [Google Scholar] [CrossRef] [PubMed]
  265. Ge, J.Y.; Shu, S.; Kwon, M.; Jovanović, B.; Murphy, K.; Gulvady, A.; Fassl, A.; Trinh, A.; Kuang, Y.; Heavey, G.A.; et al. Acquired resistance to combined BET and CDK4/6 inhibition in triple-negative breast cancer. Nat. Commun. 2020, 11, 2350. [Google Scholar] [CrossRef]
  266. Li, Q.; Jiang, B.; Guo, J.; Shao, H.; Del Priore, I.S.; Chang, Q.; Kudo, R.; Li, Z.; Razavi, P.; Liu, B.; et al. INK4 Tumor Suppressor Proteins Mediate Resistance to CDK4/6 Kinase Inhibitors. Cancer Discov. 2022, 12, 356–371. [Google Scholar] [CrossRef]
  267. Ismail, A.; Bandla, S.; Reveiller, M.; Toia, L.; Zhou, Z.; Gooding, W.E.; Kalatskaya, I.; Stein, L.; D’Souza, M.; Litle, V.R.; et al. Early G₁ cyclin-dependent kinases as prognostic markers and potential therapeutic targets in esophageal adenocarcinoma. Clin. Cancer Res. 2011, 17, 4513–4522. [Google Scholar] [CrossRef]
  268. Dean, J.L.; McClendon, A.K.; Hickey, T.E.; Butler, L.M.; Tilley, W.D.; Witkiewicz, A.K.; Knudsen, E.S. Therapeutic response to CDK4/6 inhibition in breast cancer defined by ex vivo analyses of human tumors. Cell Cycle 2012, 11, 2756–2761. [Google Scholar] [CrossRef]
  269. Dean, J.L.; Thangavel, C.; McClendon, A.K.; Reed, C.A.; Knudsen, E.S. Therapeutic CDK4/6 inhibition in breast cancer: Key mechanisms of response and failure. Oncogene 2010, 29, 4018–4032. [Google Scholar] [CrossRef]
  270. Heilmann, A.M.; Perera, R.M.; Ecker, V.; Nicolay, B.N.; Bardeesy, N.; Benes, C.H.; Dyson, N.J. CDK4/6 and IGF1 receptor inhibitors synergize to suppress the growth of p16INK4A-deficient pancreatic cancers. Cancer Res. 2014, 74, 3947–3958. [Google Scholar] [CrossRef]
  271. Condorelli, R.; Spring, L.; O’Shaughnessy, J.; Lacroix, L.; Bailleux, C.; Scott, V.; Dubois, J.; Nagy, R.J.; Lanman, R.B.; Iafrate, A.J.; et al. Polyclonal RB1 mutations and acquired resistance to CDK 4/6 inhibitors in patients with metastatic breast cancer. Ann. Oncol. 2018, 29, 640–645. [Google Scholar] [CrossRef]
  272. Chandarlapaty, S.; Razavi, P. Cyclin E mRNA: Assessing Cyclin-Dependent Kinase (CDK) Activation State to Elucidate Breast Cancer Resistance to CDK4/6 Inhibitors. J. Clin. Oncol. 2019, 37, 1148–1150. [Google Scholar] [CrossRef] [PubMed]
  273. Yang, C.; Li, Z.; Bhatt, T.; Dickler, M.; Giri, D.; Scaltriti, M.; Baselga, J.; Rosen, N.; Chandarlapaty, S. Acquired CDK6 amplification promotes breast cancer resistance to CDK4/6 inhibitors and loss of ER signaling and dependence. Oncogene 2017, 36, 2255–2264. [Google Scholar] [CrossRef] [PubMed]
  274. Min, A.; Kim, J.E.; Kim, Y.J.; Lim, J.M.; Kim, S.; Kim, J.W.; Lee, K.H.; Kim, T.Y.; Oh, D.Y.; Bang, Y.J.; et al. Cyclin E overexpression confers resistance to the CDK4/6 specific inhibitor palbociclib in gastric cancer cells. Cancer Lett. 2018, 430, 123–132. [Google Scholar] [CrossRef] [PubMed]
  275. Al-Qasem, A.J.; Alves, C.L.; Ehmsen, S.; Tuttolomondo, M.; Terp, M.G.; Johansen, L.E.; Vever, H.; Hoeg, L.V.A.; Elias, D.; Bak, M.; et al. Co-targeting CDK2 and CDK4/6 overcomes resistance to aromatase and CDK4/6 inhibitors in ER+ breast cancer. NPJ Precis. Oncol. 2022, 6, 68. [Google Scholar] [CrossRef]
  276. Patnaik, A.; Rosen, L.S.; Tolaney, S.M.; Tolcher, A.W.; Goldman, J.W.; Gandhi, L.; Papadopoulos, K.P.; Beeram, M.; Rasco, D.W.; Hilton, J.F.; et al. Efficacy and Safety of Abemaciclib, an Inhibitor of CDK4 and CDK6, for Patients with Breast Cancer, Non-Small Cell Lung Cancer, and Other Solid Tumors. Cancer Discov. 2016, 6, 740–753. [Google Scholar] [CrossRef] [PubMed]
  277. Flaherty, K.T.; Lorusso, P.M.; Demichele, A.; Abramson, V.G.; Courtney, R.; Randolph, S.S.; Shaik, M.N.; Wilner, K.D.; O’Dwyer, P.J.; Schwartz, G.K. Phase I, dose-escalation trial of the oral cyclin-dependent kinase 4/6 inhibitor PD 0332991, administered using a 21-day schedule in patients with advanced cancer. Clin. Cancer Res. 2012, 18, 568–576. [Google Scholar] [CrossRef]
  278. Nourbakhsh, M.; Hauser, H. Constitutive silencing of IFN-beta promoter is mediated by NRF (NF-kappaB-repressing factor), a nuclear inhibitor of NF-kappaB. EMBO J. 1999, 18, 6415–6425. [Google Scholar] [CrossRef]
  279. Györffy, B.; Lanczky, A.; Eklund, A.C.; Denkert, C.; Budczies, J.; Li, Q.; Szallasi, Z. An online survival analysis tool to rapidly assess the effect of 22,277 genes on breast cancer prognosis using microarray data of 1,809 patients. Breast Cancer Res. Treat. 2010, 123, 725–731. [Google Scholar] [CrossRef]
  280. Herrera-Abreu, M.T.; Palafox, M.; Asghar, U.; Rivas, M.A.; Cutts, R.J.; Garcia-Murillas, I.; Pearson, A.; Guzman, M.; Rodriguez, O.; Grueso, J.; et al. Early Adaptation and Acquired Resistance to CDK4/6 Inhibition in Estrogen Receptor-Positive Breast Cancer. Cancer Res. 2016, 76, 2301–2313. [Google Scholar] [CrossRef]
  281. Kruer, T.L.; Dougherty, S.M.; Reynolds, L.; Long, E.; de Silva, T.; Lockwood, W.W.; Clem, B.F. Expression of the lncRNA Maternally Expressed Gene 3 (MEG3) Contributes to the Control of Lung Cancer Cell Proliferation by the Rb Pathway. PLoS ONE 2016, 11, e0166363. [Google Scholar] [CrossRef]
  282. Yu, Y.; Liao, H.; Xie, R.; Zhang, Y.; Zheng, R.; Chen, J.; Zhang, B. Overexpression of miRNA-3613-3p Enhances the Sensitivity of Triple Negative Breast Cancer to CDK4/6 Inhibitor Palbociclib. Front. Oncol. 2020, 10, 590813. [Google Scholar] [CrossRef] [PubMed]
  283. Citron, F.; Segatto, I.; Vinciguerra, G.L.R.; Musco, L.; Russo, F.; Mungo, G.; D’Andrea, S.; Mattevi, M.C.; Perin, T.; Schiappacassi, M.; et al. Downregulation of miR-223 Expression Is an Early Event during Mammary Transformation and Confers Resistance to CDK4/6 Inhibitors in Luminal Breast Cancer. Cancer Res. 2020, 80, 1064–1077. [Google Scholar] [CrossRef] [PubMed]
  284. Baldassari, F.; Zerbinati, C.; Galasso, M.; Corrà, F.; Minotti, L.; Agnoletto, C.; Previati, M.; Croce, C.M.; Volinia, S. Screen for MicroRNA and Drug Interactions in Breast Cancer Cell Lines Points to miR-126 as a Modulator of CDK4/6 and PIK3CA Inhibitors. Front. Genet. 2018, 9, 174. [Google Scholar] [CrossRef] [PubMed]
  285. Cornell, L.; Wander, S.A.; Visal, T.; Wagle, N.; Shapiro, G.I. MicroRNA-Mediated Suppression of the TGF-β Pathway Confers Transmissible and Reversible CDK4/6 Inhibitor Resistance. Cell Rep. 2019, 26, 2667–2680.e2667. [Google Scholar] [CrossRef] [PubMed]
  286. Kong, Q.; Qiu, M. Long noncoding RNA SNHG15 promotes human breast cancer proliferation, migration and invasion by sponging miR-211-3p. Biochem. Biophys. Res. Commun. 2018, 495, 1594–1600. [Google Scholar] [CrossRef] [PubMed]
  287. Cui, H.X.; Zhang, M.Y.; Liu, K.; Liu, J.; Zhang, Z.L.; Fu, L. LncRNA SNHG15 promotes proliferation and migration of lung cancer via targeting microRNA-211-3p. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 6838–6844. [Google Scholar] [CrossRef]
  288. Zhang, J.H.; Wei, H.W.; Yang, H.G. Long noncoding RNA SNHG15, a potential prognostic biomarker for hepatocellular carcinoma. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 1720–1724. [Google Scholar]
  289. Gusev, Y.; Bhuvaneshwar, K.; Song, L.; Zenklusen, J.C.; Fine, H.; Madhavan, S. The REMBRANDT study, a large collection of genomic data from brain cancer patients. Sci. Data 2018, 5, 180158. [Google Scholar] [CrossRef]
  290. Lee, Y.; Liu, J.; Patel, S.; Cloughesy, T.; Lai, A.; Farooqi, H.; Seligson, D.; Dong, J.; Liau, L.; Becker, D.; et al. Genomic landscape of meningiomas. Brain Pathol. 2010, 20, 751–762. [Google Scholar] [CrossRef]
  291. Braconi, C.; Kogure, T.; Valeri, N.; Huang, N.; Nuovo, G.; Costinean, S.; Negrini, M.; Miotto, E.; Croce, C.M.; Patel, T. microRNA-29 can regulate expression of the long non-coding RNA gene MEG3 in hepatocellular cancer. Oncogene 2011, 30, 4750–4756. [Google Scholar] [CrossRef]
  292. Lu, K.H.; Li, W.; Liu, X.H.; Sun, M.; Zhang, M.L.; Wu, W.Q.; Xie, W.P.; Hou, Y.Y. Long non-coding RNA MEG3 inhibits NSCLC cells proliferation and induces apoptosis by affecting p53 expression. BMC Cancer 2013, 13, 461. [Google Scholar] [CrossRef] [PubMed]
  293. Wang, P.; Ren, Z.; Sun, P. Overexpression of the long non-coding RNA MEG3 impairs in vitro glioma cell proliferation. J. Cell. Biochem. 2012, 113, 1868–1874. [Google Scholar] [CrossRef] [PubMed]
  294. Zhang, X.; Gejman, R.; Mahta, A.; Zhong, Y.; Rice, K.A.; Zhou, Y.; Cheunsuchon, P.; Louis, D.N.; Klibanski, A. Maternally expressed gene 3, an imprinted noncoding RNA gene, is associated with meningioma pathogenesis and progression. Cancer Res. 2010, 70, 2350–2358. [Google Scholar] [CrossRef] [PubMed]
  295. Zhou, Y.; Zhong, Y.; Wang, Y.; Zhang, X.; Batista, D.L.; Gejman, R.; Ansell, P.J.; Zhao, J.; Weng, C.; Klibanski, A. Activation of p53 by MEG3 non-coding RNA. J. Biol. Chem. 2007, 282, 24731–24742. [Google Scholar] [CrossRef]
  296. Mondal, T.; Subhash, S.; Vaid, R.; Enroth, S.; Uday, S.; Reinius, B.; Mitra, S.; Mohammed, A.; James, A.R.; Hoberg, E.; et al. MEG3 long noncoding RNA regulates the TGF-β pathway genes through formation of RNA-DNA triplex structures. Nat. Commun. 2015, 6, 7743. [Google Scholar] [CrossRef]
  297. Liu, J.; Wan, L.; Lu, K.; Sun, M.; Pan, X.; Zhang, P.; Lu, B.; Liu, G.; Wang, Z. The Long Noncoding RNA MEG3 Contributes to Cisplatin Resistance of Human Lung Adenocarcinoma. PLoS ONE 2015, 10, e0114586. [Google Scholar] [CrossRef]
  298. Huang, Q.; Shen, Y.J.; Hsueh, C.Y.; Guo, Y.; Zhang, Y.F.; Li, J.Y.; Zhou, L. miR-17-5p drives G2/M-phase accumulation by directly targeting CCNG2 and is related to recurrence of head and neck squamous cell carcinoma. BMC Cancer 2021, 21, 1074. [Google Scholar] [CrossRef]
  299. Lee, S.; Schmitt, C.A. The dynamic nature of senescence in cancer. Nat. Cell. Biol. 2019, 21, 94–101. [Google Scholar] [CrossRef]
  300. Dang, F.; Nie, L.; Wei, W. Ubiquitin signaling in cell cycle control and tumorigenesis. Cell. Death Differ. 2021, 28, 427–438. [Google Scholar] [CrossRef]
  301. Diaz-Moralli, S.; Tarrado-Castellarnau, M.; Miranda, A.; Cascante, M. Targeting cell cycle regulation in cancer therapy. Pharmacol. Ther. 2013, 138, 255–271. [Google Scholar] [CrossRef]
  302. Xu, Q.; Liu, K. MiR-369-3p inhibits tumorigenesis of hepatocellular carcinoma by binding to PAX6. J. Biol. Regul. Homeost. Agents 2020, 34, 917–926. [Google Scholar] [CrossRef]
  303. Geng, Z.; Chen, H.; Zou, G.; Yuan, L.; Liu, P.; Li, B.; Zhang, K.; Jing, F.; Nie, X.; Liu, T.; et al. Human Amniotic Fluid Mesenchymal Stem Cell-Derived Exosomes Inhibit Apoptosis in Ovarian Granulosa Cell via miR-369-3p/YAF2/PDCD5/p53 Pathway. Oxid. Med. Cell Longev. 2022, 2022, 3695848. [Google Scholar] [CrossRef] [PubMed]
  304. Vasudevan, S.; Tong, Y.; Steitz, J.A. Switching from repression to activation: microRNAs can up-regulate translation. Science 2007, 318, 1931–1934. [Google Scholar] [CrossRef] [PubMed]
  305. Jovanovic, M.; Reiter, L.; Picotti, P.; Lange, V.; Bogan, E.; Hurschler, B.A.; Blenkiron, C.; Lehrbach, N.J.; Ding, X.C.; Weiss, M.; et al. A quantitative targeted proteomics approach to validate predicted microRNA targets in C. elegans. Nat. Methods 2010, 7, 837–842. [Google Scholar] [CrossRef] [PubMed]
  306. Archambault, V.; Chang, E.J.; Drapkin, B.J.; Cross, F.R.; Chait, B.T.; Rout, M.P. Targeted proteomic study of the cyclin-Cdk module. Mol. Cell 2004, 14, 699–711. [Google Scholar] [CrossRef]
  307. Yu, R.; Hu, Y.; Zhang, S.; Li, X.; Tang, M.; Yang, M.; Wu, X.; Li, Z.; Liao, X.; Xu, Y.; et al. LncRNA CTBP1-DT-encoded microprotein DDUP sustains DNA damage response signalling to trigger dual DNA repair mechanisms. Nucleic Acids Res. 2022, 50, 8060–8079. [Google Scholar] [CrossRef]
  308. Taheri, M.; Shoorei, H.; Tondro Anamag, F.; Ghafouri-Fard, S.; Dinger, M.E. LncRNAs and miRNAs participate in determination of sensitivity of cancer cells to cisplatin. Exp. Mol. Pathol. 2021, 123, 104602. [Google Scholar] [CrossRef]
  309. Wen, D.; Peng, Y.; Lin, F.; Singh, R.K.; Mahato, R.I. Micellar Delivery of miR-34a Modulator Rubone and Paclitaxel in Resistant Prostate Cancer. Cancer Res. 2017, 77, 3244–3254. [Google Scholar] [CrossRef]
  310. Zhang, X.; Xie, K.; Zhou, H.; Wu, Y.; Li, C.; Liu, Y.; Liu, Z.; Xu, Q.; Liu, S.; Xiao, D.; et al. Role of non-coding RNAs and RNA modifiers in cancer therapy resistance. Mol. Cancer 2020, 19, 47. [Google Scholar] [CrossRef]
  311. Lee, Y.S.; Dutta, A. MicroRNAs in cancer. Annu. Rev. Pathol. 2009, 4, 199–227. [Google Scholar] [CrossRef]
  312. Bencivenga, D.; Stampone, E.; Vastante, A.; Barahmeh, M.; Della Ragione, F.; Borriello, A. An Unanticipated Modulation of Cyclin-Dependent Kinase Inhibitors: The Role of Long Non-Coding RNAs. Cells 2022, 11, 1346. [Google Scholar] [CrossRef] [PubMed]
  313. Bader, A.G. miR-34—A microRNA replacement therapy is headed to the clinic. Front. Genet. 2012, 3, 120. [Google Scholar] [CrossRef] [PubMed]
  314. Hong, D.S.; Kang, Y.K.; Borad, M.; Sachdev, J.; Ejadi, S.; Lim, H.Y.; Brenner, A.J.; Park, K.; Lee, J.L.; Kim, T.Y.; et al. Phase 1 study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br. J. Cancer 2020, 122, 1630–1637. [Google Scholar] [CrossRef] [PubMed]
  315. Rui, M.; Qu, Y.; Gao, T.; Ge, Y.; Feng, C.; Xu, X. Simultaneous delivery of anti-miR21 with doxorubicin prodrug by mimetic lipoprotein nanoparticles for synergistic effect against drug resistance in cancer cells. Int. J. Nanomed. 2017, 12, 217–237. [Google Scholar] [CrossRef] [PubMed]
  316. Krützfeldt, J.; Rajewsky, N.; Braich, R.; Rajeev, K.G.; Tuschl, T.; Manoharan, M.; Stoffel, M. Silencing of microRNAs in vivo with ‘antagomirs’. Nature 2005, 438, 685–689. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Outline of cell cycle control. The cell cycle is positively controlled by cyclin-dependent kinases (CDKs) and cyclins’ interactions and negatively controlled by cyclin-dependent protein kinase inhibitors (CKIs). S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status. Black lines end with arrow meaning promotion, black lines end with bar meaning inhibition, colorful arrows meaning cell cycle progression.
Figure 1. Outline of cell cycle control. The cell cycle is positively controlled by cyclin-dependent kinases (CDKs) and cyclins’ interactions and negatively controlled by cyclin-dependent protein kinase inhibitors (CKIs). S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status. Black lines end with arrow meaning promotion, black lines end with bar meaning inhibition, colorful arrows meaning cell cycle progression.
Ijms 24 08939 g001
Figure 2. Schema showing the mechanisms of lncRNA-mediated regulation of the cell cycle via recruiting EZH2. (A). The model of LINC-PINT recruits EZH2, which induces H3K27me3, thus inactivating the transcription of CDK1, cyclin A2, AURKA, and PCNA genes. (B). A model of lncRNA-HEIH interacts with EZH2 to inhibit p15, p16, p21, and p57, thus leading to HCC tumorigenesis. (C). The model of lncRNA-ANRIL interacts with the PRC2 complex and recruits it to the INK4 locus to methylate histone H3K27. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status. Red arrows upward meaning cell cycle activating, red arrows downward meaning cell cycle arresting; black lines end with arrow meaning promotion, black lines end with bar meaning inhibition.
Figure 2. Schema showing the mechanisms of lncRNA-mediated regulation of the cell cycle via recruiting EZH2. (A). The model of LINC-PINT recruits EZH2, which induces H3K27me3, thus inactivating the transcription of CDK1, cyclin A2, AURKA, and PCNA genes. (B). A model of lncRNA-HEIH interacts with EZH2 to inhibit p15, p16, p21, and p57, thus leading to HCC tumorigenesis. (C). The model of lncRNA-ANRIL interacts with the PRC2 complex and recruits it to the INK4 locus to methylate histone H3K27. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status. Red arrows upward meaning cell cycle activating, red arrows downward meaning cell cycle arresting; black lines end with arrow meaning promotion, black lines end with bar meaning inhibition.
Ijms 24 08939 g002
Figure 3. An overview of cell cycle control by lncRNAs. Lines exhibiting lncRNAs that accelerate proliferation are in red, whereas lines showing lncRNAs that reduce cell cycle are in blue. Lines ending with arrows indicate lncRNA promotion on specific molecules. Lines ending with a bar indicate lncRNA inhibition on downstream molecules. Dash lines indicate the regulation of CDKs, cyclins, and CKIs within the cell cycle. S, DNA synthesis; M, mitosis; G1 and G2, index transition phases; and G0, index quiescent status.
Figure 3. An overview of cell cycle control by lncRNAs. Lines exhibiting lncRNAs that accelerate proliferation are in red, whereas lines showing lncRNAs that reduce cell cycle are in blue. Lines ending with arrows indicate lncRNA promotion on specific molecules. Lines ending with a bar indicate lncRNA inhibition on downstream molecules. Dash lines indicate the regulation of CDKs, cyclins, and CKIs within the cell cycle. S, DNA synthesis; M, mitosis; G1 and G2, index transition phases; and G0, index quiescent status.
Ijms 24 08939 g003
Figure 4. An overview of cell cycle control by microRNAs. Lines displaying miRNAs that accelerate proliferation are in red, whereas miRNAs that reduce cell cycle are in blue. Lines ending with an arrow indicate miRNA promotion on downstream molecules. Lines ending with a bar indicate miRNA inhibition on downstream molecules. Dash lines indicate the regulation of CDKs, cyclins, and CKIs within the cell cycle. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status.
Figure 4. An overview of cell cycle control by microRNAs. Lines displaying miRNAs that accelerate proliferation are in red, whereas miRNAs that reduce cell cycle are in blue. Lines ending with an arrow indicate miRNA promotion on downstream molecules. Lines ending with a bar indicate miRNA inhibition on downstream molecules. Dash lines indicate the regulation of CDKs, cyclins, and CKIs within the cell cycle. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status.
Ijms 24 08939 g004
Figure 5. Schema showing E2Fs’ transcription on miRNAs and the feedback loop between miRNAs and E2Fs during the cell cycle. MiRNAs regulated by E2Fs are anti-proliferation, and they are in blue. MiR-17-92 and miR-106b-25 clusters are upregulated by c-MYC and inhibit E2F activity. MiR-449 downregulates CDK6, then reduces phosphating RB and suppresses E2Fs. MiR-330 and miR-331-3p inhibit E2F activity, causing cell cycle arrest. Blue squares show miRNAs that regulate E2Fs. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status. Black lines end with arrow meaning promotion, black lines end with bar meaning inhibition.
Figure 5. Schema showing E2Fs’ transcription on miRNAs and the feedback loop between miRNAs and E2Fs during the cell cycle. MiRNAs regulated by E2Fs are anti-proliferation, and they are in blue. MiR-17-92 and miR-106b-25 clusters are upregulated by c-MYC and inhibit E2F activity. MiR-449 downregulates CDK6, then reduces phosphating RB and suppresses E2Fs. MiR-330 and miR-331-3p inhibit E2F activity, causing cell cycle arrest. Blue squares show miRNAs that regulate E2Fs. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status. Black lines end with arrow meaning promotion, black lines end with bar meaning inhibition.
Ijms 24 08939 g005
Table 1. LncRNA associated in the cell cycle control.
Table 1. LncRNA associated in the cell cycle control.
LncRNATarget in Cell Cycle (Phase)Deregulation in CancerReferences
MEG3Promotion of p27 by miR-3163/Skp2 mRNA translation.Downregulated in BLCA, BRCA, CCA, COCA, OVCA. Upregulated in small cell lung cancer.[88]
LINC-PINTSuppression of CDK1 and cyclin A2 with PRC2 and EZH2 (G2/M).Downregulated in HCC, BLCA, BRCA, MM, T and B-ALL. Upregulated in COCA.[75,76]
HEIHSuppression of p15, p16, p21 and p57 with EZH2 (G0/G1).Upregulated in HCC. Downregulated in BRCA, CRCA, OVCA, PRCA.[77]
ANRILSuppression of p21, p14, p15, p16 upregulated cyclin G1 by targeting miR-181a-5p (G0/G1, G2/M).Upregulated in glioma, LSCC, GC, EOC, HNSC, NSCLC, ESCC, CRCA.[78,79,80]
LINC00630Promotion of CDK2 transcription (S).Upregulated in HCC, CRCA.[84]
ZFAS1Promotion of p21 and p27 expression; destabilization of p53; promotion of CDK1/cyclin B1 complex by interaction (G1/S).Upregulated in CRCA, BLCA, GC, HCC, PRCA. Downregulated in OVCA, basal-like BRCA. [94]
TMPO-AS1Promotion of CDK1. Promotion proliferation by sponging miR-326 and targeted by E2F1 (G0/G1, G2/M).Upregulated in BLCA, basal-like BRCA, CCA, CRCA, LUCA, PRCA, HCC, OVCA.[82]
ARAP1-AS1Promotion of cyclin D1 (G0/G1).Upregulated in LUCA.[85]
CADM1-AS1Promotion of p15, p21, p27. Suppression of cyclin Ds, cyclin Es, CDK2, CDK4, CDK6 (G0/G1).Downregulated in HCC.[95]
TPT1-AS1Promotion of CDK4, cyclin D1, suppression of p21 (G1/S). Upregulated in GC, HCC. Downregulated in BLCA, BRCA, LUCA, OVCA, PRCA.[89]
ABHD11-AS1Promotion of Cyclin D1, cyclin E1, CDK1, CDK2, CDK4. Suppression of p16 (G1/S).Upregulated in EC, PCA, GC, LUCA. Downregulated in CRCA.[90,91,92,93]
CCDC144NL-AS1Promotion of CDK1, CKD2 and CDK4 (G0-1/S) through miR-940/WDR5.Upregulated in CRCA, HCC. Downregulated in LUCA.[99]
LINC00261Co-regulation with CDK1 (G0/G1; G2/M).Downregulated in choriocarcinoma, HCC, BRCA, PRCA, CRCA, GC, LUCA.[83]
Gadd7Suppression of CDK6 mRNA degradation (G1/S).Induced by DNA damage and oxidative stress.[96,97]
ALMS1- IT1ALMS1-IT1/AVL9 promote CDK pathway.Upregulated in small cell lung cancer, CCA, CRCA, OVCA. Downregulated in LUCA.[86]
XISTPromotion of cyclin D1, CDK1 and p53, suppression of p21 (G1). Suppression of Wee1 (G2).Upregulated in PCA, CRCA, ESCC. Downregulated in OVCA, NSCLC, HCC, BRCA, BLCA.[63,98]
ENST00000512916Promotion of CDK2/4/6 and cyclin D1 (G1).Upregulated in AB.[87]
TROJANPromotion of CDK2.Upregulated in ER-positive BRCA.[100]
SNHG15Promotion of CDK6 by miR-627 sponging.Upregulated in BLCA, TNBC, CRCA, GC, LUCA, PRCA.[101]
Abbreviations: BRCA, breast cancer; CCA, cervical cancer; COCA, colon cancer; BLCA, bladder cancer; HCC, hepatocellular carcinoma; CRCA, colorectal cancer; OVCA, ovarian cancer; PRCA, prostate cancer; T, T cell; B, B cell; ALL, acute lymphocytic leukemia; GC, gastric cancer; TNBC, triple-negative breast cancer; LSCC, laryngeal squamous cell cancer; EOC, epithelial ovarian cancer; HNSC, head and neck squamous cell carcinoma; NSCLC, non-small cell lung cancer; ESCC, esophageal squamous cell carcinoma; LUCA, lung cancer; EC, endometrial carcinoma; PCA, pancreatic cancer; MM, malignant melanoma; AB, ameloblastoma. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status.
Table 2. MiRNAs associated with cell cycle control.
Table 2. MiRNAs associated with cell cycle control.
miRNATarget in Cell Cycle (Phase)Deregulation in CancerReferences
miR-195Suppression of CDK6, cyclin E1, CDK4, cyclin D1, cyclin D3 (G1/S).Downregulated in NSCLC, GC, HCC, BLCA. Upregulated in BRCA, CLL, ACA.[126,127]
miR-100Targeting and suppressing PLK1; suppression of reduced β-tubulin I, IIA, IIB and V mRNA; suppression of p27 (G2/M).Upregulated in BRCA, PCA. Downregulated in PRCA, OVCA, BLCA.[128,129]
miR-365Suppression cyclin D1 and CDC25A, promotion of p27 (G1/S).Downregulated in NSCLC, BRCA, OVCA.[130,131]
miR-15aSuppression of CDK1, CDK2, CDK6, cyclin D1, cyclin D2 and cyclin E1 (G1/S).Downregulated in HCC, BRCA, OS, NSCLC, PCA, PRCA. Upregulated in CLL.[132,133,134,135]
miR-16Suppression of CDK1, CDK2, CDK6, cyclin D1, cyclin D2 and cyclin E1 (G1/S).Downregulated in PRCA, GC, LUCA. Upregulated in PCA, HCC. [133,134,135]
Let-7Suppression of CDK4, Cyclin D1, cyclin D3, Cyclin A and Cyclin B translation (G1).Downregulation in OVCA, LUCA, NSCLC, BRCA, CRCA, PRCA, HNSC.[136,137,138,139]
miR-7Suppression of p53, CDC42, Cyclin D, cyclin E1 (G0-1/S).Downregulated in BRCA, CCA, MG, HCC. Upregulated in HNSC, LUCA.[136,140,141]
miR-19Suppression of Cyclin Ds, promotion of p21 (G1/S).downregulated in MG.[142,143]
miR-20aSuppression of E2F2 and E2F3 translation; Suppression cyclin D1 mRNA transcription; Suppression of p21 (G1/S).Upregulated in HCC, CRCA, PRCA, PCA, LUCA. Downregulated in BRCA, HCC.[144,145,146]
miR-34Suppression of Cyclin D1, CDK4, CDK6 translation; promotion of p21 (G1/S).Downregulated in PCA, OVCA, NSCLC, HCC. Upregulated in CRCA.[147,148]
miR-124aSuppression CDK2, CDK6, cyclin D1, cyclin D2 translation (G1).Downregulated in all cancer.[149,150,151]
miR-125bSuppression cyclin D1, CDK6, promotion of p21 (G1/S).Downregulated in PRCA, LUCA, HCC, BRCA, CCA, CRCA. Upregulated in PCA, BLCA.[152,153,154]
miR-24Suppression of p27, p16, p21(G2/M).Downregulated in PRCA, PTC. Upregulated in LUCA, PCA, CRCA.[154,155]
miR-26aSuppression of cyclin D2, cyclin D3, cyclin E2, CDK4/6; promotion of p14.Downregulated in PTC, LUCA, HCC, BRCA, BLCA.[156,157,158]
miR-101Suppression of CDK2, CDK4, CDK6, cyclin D2, cyclin D3 and cyclin E2; Promotion of p14, p16, p21 and p27 (G1).Downregulated in LUCA, PRCA, HCC, LSCC, OVCA, BLCA.[159,160]
miR-9Suppression of CDK6, cyclin D1; promotion p16 (G0/G1).Downregulated in HSCC.[161,162,163]
miR-218Suppression CDK6, cyclin D1; promotion of p14, p16.Downregulated in MG, LUCA, CCA. Upregulated in PRCA.[164,165]
miR-210Suppression of CDK4, CDK6; promotion of p27 (G1/S).Upregulated in PRCA, LUCA, PCA, HNSC, BRCA.[166]
miR-17-5pSuppression of cyclin D1; promotion of p21 (G2/M).Upregulated in bladder cancer, HCC, BRCA, CRCA, LUCA.[144,167,168,169]
miR-17-92 cluster (MIR17HG)Suppression of E2F1, cyclin D1. Promotion of p21 (G1/S).Upregulated in BL, CRCA, LUCA, BLCA, BRCA, PRCA, PCA, HCC.[170,171,172,173,174,175,176]
miR-106aSuppression of p21, cyclin D1 (G1/S).Upregulated in LUCA, HCC, PCA, PRCA, CRCA.[143,177]
miR-106bSuppression of cyclin D1, E2F1, p21 (G1/S).upregulated in HCC, GC, CRCA. Downregulated in OVCA.[172,178,179,180]
miR-330Suppression of E2F1.Downregulated in OSCC, PRCA.[181]
miR-331-3pSuppression of E2F1.Downregulated in HNSC, OSCC. Upregulated in AML.[182]
miR-449Promotion of CDK6 and CDC25A; Suppression E2F1, E2F3, cyclin D1, cyclin A2.Downregulated in PRCA.[183,184]
miR-31Suppression of CDK1, CDK4, CDK6; promotion of p21.Upregulated in HNSC, CRCA, OSCC, LUCA. Downregulated in PRCA, BRCA, BLCA. [185,186,187,188]
miR-512-5pSuppression of p21.Downregulated in GC.[189]
miR-370Promotion of p27 and p21 (G1/S).Upregulated in PRCA. Downregulated in OSCC.[190]
miR-212Promotion of p21 and p27; suppression cyclin D3 (G1/S).Upregulated in LUCA, PCA, HCC.[191,192]
miR-196aSuppression of p27.Upregulated in BRCA, PDAC, ESCC, CRCA.[193]
miR-3163Promotion of p27 (G0/G1).Downregulated in BRCA.[88,194]
miR-221Suppression of p27 and p57, promotion of CDK2 (G2/M).Downregulated in PRCA, NSCLC. Upregulated in HCC, PCA, BLCA, CRCA, OVCA, BRCA, OSCC.[195,196]
miR-222Suppression of p27 (G1/S).Downregulated in PRCA. Upregulated in PCA, NSCLC, HCC, GC.[197,198,199]
Nc886Promotion of CDKN2A and CDKN2C, suppression of CDK4 (G1/S).Downregulated in ESCC, GC, PRCA.[200,201]
miR-29b-3pSuppression of CDK6 (G1).Downregulated in PRCA, LUCA, HNSC, CLL, AML. Upregulated in BRCA, CRCA, PCA.[202,203]
miR-497Suppression of CDK6, E2F3 and cyclin E1 (G1).Downregulated in PRCA, LUCA, CRCA, BRCA.[204,205]
miR-193bSuppression of cyclin D1 by targeting 3′UTR of mRNA (G1/S).Downregulated in BRCA, PRCA, OS, MM.[206,207,208,209]
miR-200Suppression of CDK6 (G1).Downregulated in CRCA, HCC, OV, BRCA, MM (all cancer)[210]
miR-206Suppression of CDK4, cyclin D1, cyclin C translation (G1).Downregulated in BRCA, OVCA.[211,212]
miR-223Suppression of p21, p27 (G1). E2F1 binds to the miR-223 promoter and inhibits transcription.Downregulated in AML, HCC, CLL. Upregulated in BLCA, PCA, CRCA, PRCA. [213,214]
Abbreviations: BRCA, breast cancer; CCA, cervical cancer; BLCA, bladder cancer; HCC, hepatocellular carcinoma; CRCA, colorectal cancer; OVCA, ovarian cancer; PRCA, prostate cancer; GC, gastric cancer; LSCC, laryngeal squamous cell cancer; HNSC, head and neck squamous cell carcinoma; NSCLC, non-small cell lung cancer; ESCC, esophageal squamous cell carcinoma; LUCA, lung cancer; PCA, pancreatic cancer; MM, malignant melanoma; AML, acute myeloid leukemia; CLL, chronic lymphocytic leukemia; OS, osteosarcoma; OSCC, oral squamous cell carcinoma; BL, Burkitt lymphoma; MG, malignant glioma; PTC, papillary thyroid carcinoma; ACA, adrenocortical adenomas. S, DNA synthesis; M, mitosis; G1 and G2, index transition phase; G0, index quiescent status.
Table 3. NcRNAs associated with tumor sensitivity to CDK4/6 inhibitors.
Table 3. NcRNAs associated with tumor sensitivity to CDK4/6 inhibitors.
ncRNATumor TypeCDKIsSampleFunction and MechanismReferences
MEG3LCPalbociclibcell linesPalbociclib increase MEG3 expression. Silencing MEG3 expression rescue palbociclib-mediated decrease in cell proliferation.[281]
TROJANER-positive BC, TNBCPalbociclibcell linesTROJAN inhibit NKRF/RELA and upregulate CDK2 expression, which promote cell proliferation and resistance to a CDK4/6 inhibition in ER+ breast cancer.[100]
SNHG15GBM, BC, LC, HCCPalbociclibcell linesSNHG15 upregulated CDK6 by sponging inhibit miR-627-5p. palbociclib treatment increased miR-627-5p expression and reduced SNHG15 and CDK6 expression.[101]
miR-3613-3pBCPalbociclibclinical samples; PDTX; cell linesMiR-3613-3p induce G1 cell cycle arrest and enhance TNBC sensitivity to palbociclib.[282]
miR-29b-3pHER-negative BCPalbociclibclinical samples; PDTX; cell linesMiR-29b downregulate CDK6 and induce cell cycle arrest at G1 phase. Activated c-Myc downregulate miR-29b, further activate CDK6 and resistance to palbociclib.[203]
miR-497ALCLPalbociclibcell linesHypermethylation repressed miR-497 expression. miR-497 inhibited CDK6, E2F3 and cyclin E1 (G1) caused cell cycle arrest. higher miR-497 expression cell more sensitive to palbociclib[205]
miR-193bPCAPalbociclibcell linesMiR-193b downregulate cyclin D1 by targeting 3′UTR of mRNA. palbociclib has no effect on cells with high miR-193b expression.[209]
miR-200aMMPalbociclibcell linesMiR-200 reduces CDK6 expression and reduces melanoma response to palbociclib.[210]
miR-223lum BC; HER2-positive BCPalbociclibtransgenic mouse model; clinical samples; cell linesPalbociclib restrain E2F1 activity and restore miR-223 expression. miR-223 deficiency induces luminal breast cancer resistance to palbociclib.[283]
miR-126lum BC; HER2-positive BCRibociclibcell lines.MiR-126 targets PI3K/AKT/MTOR pathway, affects cell cycle and mitosis. miR-125 preserving leukemia stem cell quiescence and promoting chemotherapy resistance.[284]
MIR17HGGBM, AT/RTPalbociclib;
Ribociclib
cell lines MIR17HG reduce cyclin D1 expression and sensitizes ATRT cells to palbociclib treatment.[173]
miR-432-5pER-positive BCpalbociclib;
Ribociclib
cell linesPalbociclib resistant cells increased miR-432-5p and CDK6 expression. In palbociclib resistance breast cancers, miR-432-5p is higher expressed.[285]
miR-9HER2-positive BC; TNBCRibociclib;cell linesSuppression of CDK6 and enhance the activity of ribociclib.[284]
miR-326HER2-positive BCRibociclibcell linesSuppression of CDK6 and enhance the activity of ribociclib.[82]
miR-124aALLPalbociclib; PD-0332991cell linesMiR-124a was down-regulated in ALL by hypermethylation of the promoter and histone modifications. MiR-124 negatively regulates CDK6 and pRB. ALL patients with overexpressed miR-124a benefit from CDK6 inhibition.[149]
miR-9ALLPalbociclibcell linesEpigenetic downregulation of MIR9 induced upregulation of CDK6.[174]
Abbreviations: LC, lung cancer; BC, breast cancer; TNBC, triple-negative breast cancer; HCC, hepatocellular carcinoma; GBM, glioblastoma; ALCL, anaplastic large cell lymphoma; PCA, prostate cancer; MM, malignant melanoma; lum BC, luminal breast cancer; AT/RT, atypical teratoid/rhabdoid tumor; ALL, acute lymphoblastic leukemia; PDTX, patient-derived tumor xenograft.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hu, Q.; Huang, T. Regulation of the Cell Cycle by ncRNAs Affects the Efficiency of CDK4/6 Inhibition. Int. J. Mol. Sci. 2023, 24, 8939. https://doi.org/10.3390/ijms24108939

AMA Style

Hu Q, Huang T. Regulation of the Cell Cycle by ncRNAs Affects the Efficiency of CDK4/6 Inhibition. International Journal of Molecular Sciences. 2023; 24(10):8939. https://doi.org/10.3390/ijms24108939

Chicago/Turabian Style

Hu, Qingyi, and Tao Huang. 2023. "Regulation of the Cell Cycle by ncRNAs Affects the Efficiency of CDK4/6 Inhibition" International Journal of Molecular Sciences 24, no. 10: 8939. https://doi.org/10.3390/ijms24108939

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