- Review
- Open access
- Published:
Non-coding RNAs, a double-edged sword in breast cancer prognosis
Cancer Cell International volume 25, Article number: 123 (2025)
Abstract
Cancer is a rising issue worldwide, and numerous studies have focused on understanding the underlying reasons for its occurrence and finding proper ways to defeat it. By applying technological advances, researchers are continuously uncovering and updating treatments in cancer therapy. Their vast functions in the regulation of cell growth and proliferation and their significant role in the progression of diseases, including cancer. This review provides a comprehensive analysis of ncRNAs in breast cancer, focusing on long non-coding RNAs such as HOTAIR, MALAT1, and NEAT1, as well as microRNAs such as miR-21, miR-221/222, and miR-155. These ncRNAs are pivotal in regulating cell proliferation, metastasis, drug resistance, and apoptosis. Additionally, we discuss experimental approaches that are useful for studying them and highlight the advantages and challenges of each method. We then explain the results of these clinical trials and offer insights for future studies by discussing major existing gaps. On the basis of an extensive number of studies, this review provides valuable insights into the potential of ncRNAs in cancer therapy. Key findings show that even though the functions of ncRNAs are vast and undeniable in cancer, there are still complications associated with their therapeutic use. Moreover, there is an absence of sufficient experiments regarding their application in mouse models, which is an area to work on. By emphasizing the crucial role of ncRNAs, this review underscores the need for innovative approaches and further studies to explore their potential in cancer therapy.
Graphical Abstract

Introduction
Breast cancer is a global issue that affects women around the world. In 2020, the World Health Organization reported that 2.3 million women were diagnosed with it and that 685,000 of those women lost their lives to it [1]. Moreover, it is forecasted that by 2040, these rates will increase to 3 million new cases and 1 million deaths [2]. The statistics can be observed in Fig. 1. These rates highlight the undeniable importance of research into the underlying mechanisms and precursors contributing to the rise of breast cancer and the need to develop effective strategies to overcome it.
Breast cancer is classified into four main subtypes: luminal A, luminal B, human epidermal growth factor receptor 2 (HER2)-positive, and triple-negative breast cancer (TNBC). These subtypes differ in their molecular profiles, prognoses, and treatment strategies (Fig. 2). Luminal subtypes, which are characterized by hormone receptor positivity, are often treated with hormonal therapies, whereas the HER2-positive and TNBC subtypes are more aggressive and resistant to standard treatments [3,4,5,6,7,8,9,10,11,12]. Understanding the molecular characteristics of these subtypes is essential for identifying therapeutic targets that could improve outcomes.
Current breast cancer treatments, including surgery, radiation, chemotherapy, and targeted therapies, face significant challenges, such as therapy resistance and adverse side effects [13, 14]. For example, resistance to conventional therapies often arises from changes in key regulatory pathways, such as phosphatidylinositol 3-kinase/protein kinase B/mechanistic target of rapamycin (PI3K/AKT/mTOR), which lead to drug resistance [15]. Moreover, side effects such as heart disease, bone issues, and ovarian failure following treatment highlight the limitations of existing approaches [16, 17]. Therefore, there is a need to identify new and effective therapeutic targets for treating different stages of breast cancer. Understanding and investigating the factors involved in its formation and progression is crucial for designing more practical and targeted treatments.
There are different types of RNAs in cells, and they can be coding or non-coding. In higher organisms, less than 3% of the genes that are transcribed are encoded to proteins, and projects such as the Encyclopedia of DNA Elements (ENCODE) suggest that 80% of the genome is transcribed to non-coding RNA (ncRNA). Long non-coding RNAs (lncRNAs) and microRNAs (miRNAs) are two types of well-known ncRNAs [18]. These RNAs can play a significant role in breast cancer progression and drug resistance [19] and therefore have emerged as promising candidates for breast cancer treatment. Non-coding RNAs are instrumental in various cancer pathways [20] because of their wide range of interactions with materials such as DNA, mRNA proteins, and other non-coding RNAs [21,22,23]. Dysregulated miRNAs can play a role in maintaining proliferative signals, evading growth suppressors, resisting cell death, activating invasion and metastasis, and promoting the formation of new blood vessels (angiogenesis) [24,25,26,27,28]. Recent research has shown that secreted miRNAs can act as ligands to activate premetastatic inflammatory responses inside the tumor microenvironment [29]. MicroRNA-21’s (miR-21) dysregulation in various organs, such as the breast and brain, is linked to apoptosis (the process of programmed cell death), metastasis, and cell growth and differentiation pathways [30,31,32].
Another type of non-coding RNA known as PIWI-interacting RNAs (piRNAs) regulate cancer cell metastasis, tumor cell proliferation, apoptosis, and invasion [33, 34]. In addition, the levels of piR-4987 in breast cancer are positively associated with lymph node metastases [35]. piR-823 affects transcriptional activity, apoptosis, DNA methylation, metastasis, and cell proliferation [36,37,38].
Several lncRNAs have been studied and linked with cancer through their roles in angiogenesis, posttranscriptional gene regulation, metastasis, cell proliferation, and treatment response [26, 39]. For example, HOX transcript antisense intergenic RNA’s (HOTAIR) dysregulation in the thyroid, prostate, ovarian, endometrial, and lung can lead to methylation and subsequent silencing of tumor suppressor genes [30, 40, 41]. In response to stress, H19 induces survival pathways and epithelial-to-mesenchymal and mesenchymal-to-epithelial transitions, and this non-coding RNA is dysregulated in the bladder, ovarian, endometrial, breast, prostate, and colorectal regions [42]. The long non-coding RNA X-inactivation-specific transcript (XIST) is an important factor in the development of breast cancer [43]. Moreover, some studies have indicated that the overexpression of XIST can repress cancer cell growth and invasion [44, 45].
As highlighted above, the role of ncRNAs in cancer is vast and complex and is constantly being updated and identified. Given the importance of their function in various serious diseases, our aim in this review is to provide a general classification of non-coding RNAs, highlight the timeline of their discovery, and explain their role in breast cancer. Later, explore the advantages of targeting non-coding RNAs, discuss experimental approaches, and finally, review clinical trials involving non-coding RNAs were reviewed to present prospects and challenges.
Brief history of non-coding RNAs
The discovery of ribosomal RNA (rRNA) was first documented by Georges Palade in 1955 via micrographs obtained from electron microscopy [46]. Two years later, Francis Crick proposed a fundamental framework, the Central Dogma, which explains how genetic information flows from DNA to RNA to protein during gene expression [47, 48]. Later, in 1958, Paul Zamecnik and colleagues led groundbreaking research on transfer RNA (tRNA), a crucial element in the translation process [49]. In 1968, small nuclear RNAs (snRNAs) were identified in the laboratories of Harris Busch and Sheldon Penman [50, 51], increasing the complexity of our understanding of RNA regulatory activities. Sanger et al. reported the existence of circular RNA (circRNA) in 1976 [52], and the first lncRNA, H19, was discovered in 1990 by Brannan et al. [53], which was a great milestone in the investigation of ncRNA. In 1993, another revolutionary ncRNA, microRNA (miRNA), was discovered in the roundworm Caenorhabditis elegans by Lee and colleagues [54,55,56]. In 1998, Andrew Fire and Craig Mello described RNA interference (RNAi), which inactivates genes by corrupting the corresponding mRNAs in the nematode Caenorhabditis elegans [57]. The journey of ncRNAs continues, and important discoveries were made in the late 1990s and early 2000s. David Baulcombe's group published a paper in 1999 suggesting the existence of small interfering RNA (siRNA) in plants [58]. Three years later, Calin and Croce reported a connection between dysregulated miR-16–1 and miR-15a and chronic lymphocytic leukemia (CLL) [59], providing insights into the complex interactions of ncRNAs and cancer. In 2003, Lee et al. discovered that human Drosha is the core nuclease responsible for initiating miRNA processing in the nucleus [60], providing more knowledge on the molecular mechanisms involved. Two years later, a technical breakthrough occurred when researchers presented an integrated system that could efficiently handle large sequences, leading to advances in whole-genome sequencing [61]. In the coming years, scientists have performed different studies with different RNA-seq methodologies. 454 Sequencing was employed for deep sequencing, making it easier to identify uncommon and rare transcripts despite a limited sequence portion [62]. Building on these innovations, in 2006, the first comprehensive DNA methylation map of an entire genome was reported for Arabidopsis thaliana with 35 base pair resolutions [63]. In 2012, the National Human Genome Institute (NHGI) and the ENCODE project announced a significant discovery: three-quarters of the human genome can be transcribed [64], revealing a new landscape in the genome world. Our understanding became more complicated than ever, as we discovered the range, expression levels, localization, and processing fates of both known and unknown RNAs [65]. Following academic exploration, ncRNAs entered medicinal approaches as miRNA therapeutics in Austin, Texas, which launched phase 1 studies for liposomal miR-34a mimic (MRX34), the first microRNA mimic, in April 2013 [66]. The non-coding RNA timeline is shown in Table 1.
Classification of non-coding RNAs
Non-coding RNAs can be divided into two groups: housekeeping and regulatory [67]. Housekeeping ncRNAs, such as rRNA and tRNA, are involved in messenger RNA (mRNA) translation, whereas regulatory ncRNAs are created in response to an external trigger [68]. In terms of transcript size, ncRNAs can be divided into two domains: small ncRNAs (200 nucleotides) and long ncRNAs (> 200 nucleotides) [69]. RNA maturation, RNA processing, signaling, gene expression, and protein synthesis are important areas in which ncRNAs play a role [70,71,72].
lncRNAs are important non-coding RNAs that contain at least 20,000 distinct genes [73]. Some lncRNAs specifically interact with RNA, DNA, and proteins in a sequence-dependent manner (e.g., non-coding RNA activated by DNA damage (NORAD), nuclear enriched abundant transcript 1 (NEAT1), and metastasis-associated lung adenocarcinoma transcript 1 (MALAT1)) [74]. LncRNAs can overlap with or cis-regulate DNA regulatory elements such as enhancers to harness PCGs [75, 76]. In general, HOX transcript antisense RNA (HOTAIR), nuclear enriched abundant transcript 1 (NEAT1), and metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) are among the most researched long non-coding RNAs. HOTAIR is a transcript of the homeobox C (HOXC) gene and has multiple functions in cancer and metastasis [77]. NEAT plays a role as a tumor oncogene in several cancers [78,79,80]. MALAT1 has been found to increase proliferation and hinder cell death and apoptosis [81].
Among small non-coding RNAs, microRNAs (miRNAs), small nuclear RNAs (snRNAs), small nucleolar RNAs (snoRNAs), small interfering RNAs (siRNAs) and Piwi-interacting RNAs (piRNAs) are the most studied. miRNAs can activate gene expression [82]. Recent studies suggest that miRNAs shuttle between subcellular compartments to regulate translation and transcription [83]. Abnormal expression of microRNAs is linked with numerous human diseases [84, 85]. In addition, extracellular miRNAs, which serve as signaling molecules, have been widely reported as potential biomarkers for a variety of diseases [86,87,88].
A biomarker, or biological marker, is an objective indicator that reflects the biological activity of a cell or organism at a specific time. Biomarkers improve our understanding of the connections between environmental chemicals and human diseases and improve our ability to diagnose, monitor, and predict disease risk, including cancers [89]. Therefore, ncRNAs such as miRNAs can be considered reliable indicators for various pathological conditions.
Eukaryotic cells contain small, highly abundant, nucleus-localized non-coding RNAs (snRNAs), which play important roles in the splicing of introns from primary genomic transcripts [90]. snRNAs are ribonucleoprotein particles (snRNPs) along with additional proteins that create a large particulate complex (spliceosome) bound to unsliced primary RNA transcripts to facilitate the process [91]. In addition to splicing, snRNPs also function in the nuclear maturation of primary transcripts in mRNAs, the splicing of donors in noncanonical systems, the regulation of gene expression, and 3′-end processing of replication-dependent histone mRNAs [92]. Small nucleolar RNAs (snoRNAs) are a type of non-coding RNA that are ubiquitous in the nucleoli of eukaryotic cells. They play an essential role in modifying ribosomal RNA (rRNA), which is critical for the proper functioning of the ribosome and protein synthesis. Studies have shown that snoRNAs, both those that promote and suppress tumors, not only affect tumors but also have a significant effect on the prognosis of patients [93]. Piwi-interacting RNAs (piRNAs) are short non-coding RNAs that play important roles in genome integrity by regulating transposons [94]. They are also linked to major cancer hallmarks, including proliferation, invasion, and chemoresistance, and have the potential to be used as biomarkers for cancer diagnosis and prognosis [95]. Small interfering RNAs (siRNAs) are double-stranded RNA molecules that play important roles in eukaryotic gene regulation after transcription [96]. Recently, they have been known as a naturally occurring gene-silencing mechanism [97] Table 2.
Role of non-coding RNAs in breast cancer
Our review focused on the role of non-coding RNAs in breast cancer. The two most investigated ncRNAs in breast cancer are miRNAs and long non-coding RNAs (lncRNAs), which we discuss below.
Long non-coding RNAs
Data obtained from The Cancer Genome Atlas (TCGA) revealed approximately 1059 dysregulated lncRNAs in breast cancer tissues [98]. Studies over the last few decades have shown a link between patient survival rates and the recurrence of breast cancer and the lncRNAs involved in the disease [99, 100], highlighting its crucial impact and the need for further research. Below, we briefly discuss several lncRNAs and their molecular pathways in breast cancer.
HOTAIR
HOX transcript antisense intergenic RNA (HOTAIR) is involved in various signaling pathways in breast cancer. The expression of HOTAIR is significantly greater in breast cancer tissues than in normal tissues, and through its inhibition, the migration and proliferation of cancer cells can be reduced [101]. HOTAIR impacts breast cancer progression by restricting miRNAs, which can act as tumor suppressors and activate pro-oncogenic miRNAs, leading to the expression of epithelial-to-mesenchymal transition (EMT)-related proteins that are important in metastasis. EMT is a complex biological process that results in epithelial cells transitioning to mesenchymal cells; it is linked with cancer because it provides motility for cancer cells and aids in their spread [102, 103]. Wenxing He et al. reported that HOTAIR reduces the level of miR-130a-3p, leading to the regulation of the suppressor of variegation 3–9 homolog 1 (Suv39H1)/protein kinase B (AKT)/mammalian target of rapamycin (mTOR) pathway in BC cells, which promotes the growth and metastasis of BC cells [104]. HOTAIR promotes EMT by regulating the miR-129-5p/Frizzled-7 (FZD7) axis [105] and E-cadherin, N-cadherin and vimentin [106]. It also activates AKT, which promotes metastasis by activating the miR-601/zinc finger E-box-binding homeobox 1 (ZEB1) axis [107]. Furthermore, HOTAIR can activate angiogenesis via activation of the glucose-regulated protein 78 (GRP78)/angiopoietin-2 (Ang2) axis [105]. Additionally, it can be correlated with autophagy, a cellular process that degrades and recycles damaged components to maintain homeostasis under stress, which is essential for cancer cell viability [108]. HOTAIR is detectable in circulating exosomes and has been linked to erythroblastic oncogene B2 (ErbB2)/HER2-positive breast cancer, making it a potential biomarker for liquid biopsy, a minimally invasive method to detect tumor-related materials in blood [109]. A study in 2022 revealed that HOTAIR increases resistance to breast cancer radiation by increasing the expression of heat shock protein family A (Hsp70) member 1A (HSPA1A). MiR-449b-5p inhibits HSPA1A expression by targeting its mRNA, but HOTAIR interferes with this process by sponging miR-449b-5p, leading to the recovery of HSPA1A expression during radiation [110]. Finally, HOTAIR has been identified as a potential target for medication [111], and its expression is related to drug resistance [19]. For example, Tianwen Chen et al. highlighted the upregulation of HOTAIR with resistance against trastuzumab drug in SK-BR-3-TR cell lines and suggested that blocking its expression can restore sensitivity [112]. HOTAIR is known as a promising specific biomarker for breast cancer, and its overexpression has been documented in both primary and metastatic cases. Studies suggest that detecting elevated HOTAIR levels in blood or tissue samples could serve as an indicator of the presence of breast cancer and may offer insights into its aggressiveness and metastatic potential. These findings highlight the potential of HOTAIR as a critical diagnostic and prognostic tool in breast cancer management [105, 113, 114].
MALAT1
Metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) is another long non-coding RNA that plays an instrumental role in cancer progression. MALAT1 is often linked to poor outcomes in many types of cancer, including breast cancer, making it a potential prognostic marker. Studies have shown that MALAT1 is most highly expressed in the later stages of breast cancer (stages III and IV), where it is mostly associated with increased metastasis and worse outcomes. These findings indicate that its levels tend to increase as breast cancer becomes more advanced [115, 116]. A study in 2015 demonstrated that MALAT 1 is downregulated in breast cancer cells and tissue and that its knockdown can increase EMT through the phosphatidylinositide-3 kinase (PI3K)-AKT pathway [117]. Research in 2021 revealed that the expression of MALAT1 is increased in breast cancer cells and that MALAT1 can target miR-570–3p, which is important in the development of resistance in cancer cells to doxorubicin [118]. The results of another study correlate with previous research and indicate that when MALAT1 is knocked down, proliferation is reduced, suggesting that MALAT1 trans-targets the Eukaryotic Translation Elongation Factor 1 Alpha 1 (EEF1A1) promoter, which leads to increased tumor progression in breast cancer cells [119]. MALAT1 can regulate breast cancer progression by downregulating miR-101-3p and stimulating the expression of the mTOR/pyruvate kinase M2 (PKM2) pathway, which is negatively associated with the level of miR-101-3p, suggesting that MALAT1 might control breast cancer progression via upregulation of the mTOR/PKM2 pathway through the sponging of miR-101-3p. This process leads to the proliferation, migration and invasion of breast cancer cells [120]. MALAT1 can act as a competing endogenous RNA (ceRNA) for miR-26a/26b in breast cancer cells and weaken its repressive impact on ST8 alpha-N-acetyl-neuraminide alpha-2,8-sialyltransferase 4 (ST8SIA4) gene expression. The MALAT1/miR-26a/26b/ST8SIA4 axis significantly influences the proliferation, invasion, and migration of breast cancer cells [121]. Furthermore, it has been discovered that MALAT1 activates angiogenesis in breast cancer, possibly as a result of its interaction with miR-145 [122]. Additionally, genetic variations in MALAT1 have been related to an increased risk of breast cancer, whereas other variants have been linked to a reduced risk of the disease [123]. MALAT1 has been shown to play a role in the growth and metastasis of triple-negative and Her-2-positive breast tumors, suggesting that it might be a useful prognostic indicator and therapeutic target [124].
NEAT1
Various studies have highlighted the impact of nuclear paraspeckle assembly transcript 1 (NEAT1) in breast cancer [125,126,127]. NEAT1 targets miR-146b-5p, which stimulates EMT, proliferation and invasion, suggesting that, by silencing the expression of NEAT1, the invasion of cancer cells can be stopped [128]. NEAT1 can also participate in the regulation of the miR-410-3p/cyclin D1 (CCND1) axis and increase breast cancer progression [129]. It plays a role in the inhibition of miR-448 and the increase in ZEB1, leading to breast cancer development [130]. The overexpression of NEAT1 impacts the Wingless-related integration site (Wnt)/beta-catenin (β-catenin) regulatory network by sponging microRNA-148a-3p, resulting in tumor progression in breast cancer [131]. NEAT1 sponges miR-218-5p and stops its repressive function toward TPD52; in other words, NEAT1 upregulates tumor protein D52 (TPD52). This can result in increased migration and proliferation of breast cancer cells [132]. The upregulation of Krüppel-like transcription factor 12 (KLF12), a transcriptional regulator, through the connection of NEAT1 to miR-141-3p is another pathway that promotes metastasis in breast cancer cells and can lead to resistance to chemotherapy [133]. C-X-C motif chemokine ligand 12 (CXCL12) is another factor important for cell growth and metastasis and is activated by NEAT1; it also connects to miR-133b and stimulates drug resistance in cancer cells [134]. NEAT1 is linked to Taxol resistance in breast cancer through NEAT1-mediated miRNA-23a-3p downregulation, and NEAT1 targets miR-23a-3p, which affects Forkhead box A1 (FOXA1), which is a participant in tumor growth. Interestingly, restoring FOXA1 reversed miR-23a-3p-induced Taxol sensitization [126]. Moreover, reducing the expression of NEAT1 increases the susceptibility of cells to chemotherapy, which demonstrates its role in chemoresistance [135]. In conclusion, NEAT1 is a promising biomarker for breast cancer due to its noteworthy overexpression in most stages of breast cancer in breast cancer cells compared with normal tissue. Its elevated expression is closely associated with tumor progression, invasion, metastasis, and poor outcomes, which makes it a valuable indicator for monitoring breast cancer. Additionally, its role as a "microRNA sponge," regulating various genes involved in cancer cell behavior, further shows its potential as a biomarker [136,137,138]. LncRNAs and their interactions are shown in Fig. 3.
More information about lncRNAs and their associations with cancer can be found in Table 3.
MiRNAs
MicroRNAs are a class of non-coding RNAs that are significantly dysregulated in cancer. They are typically 22 nucleotides long and transcribed from DNA into primary miRNAs (pri-miRNAs), which are then transformed into precursor miRNAs (pre-miRNAs) and mature miRNAs [139]. They can regulate gene expression via a posttranscriptional mechanism [140, 141] and play a role in DNA methylation via argonaute proteins [142]. Additionally, they are instrumental in processes such as proliferation, metabolism and apoptosis [143] (Fig. 4). Furthermore, miRNAs can act as tumor suppressors (tumor suppressor miRNAs), but they also have the potential to act as oncogenes [144]. Several miRNA subtypes and their pathways are discussed below.
Tumor suppressors
MiR-34a
MicroRNA-34a (miR-34a) is a tumor suppressor that has recently gained much popularity [145, 146]. Research by Shaban et al. revealed that the expression levels of miR-34a, BRCA1, BRCA2, and p53 were significantly lower in stage III patients than in stage I and II patients both before and after chemoradiotherapy [147]. The ability of miR-34a to distinguish between early and advanced stages of cancer shows its potential as a biomarker for disease progression. MiR-34a has also been linked to the regulation of cancer stem cell functions across different cancer types, such as pancreatic cancer, prostate cancer, glioblastoma, and medulloblastoma [148]. In breast cancer stem cells, miR-34a can reduce the regulation of the neurogenic locus Notch homolog protein 1 pathway (NOTCH) pathway, which is essential for cell maintenance, inhibits proliferation, and can augment sensitivity to paclitaxel (PTX) [149]. Rui Han et al. reported that the overexpression of miR-34a leads to a decrease in 3D spheroid cell mass in both the T-47D and MDA-MB-231 cell lines, demonstrating its potential to diminish spheroid formation by inhibiting the expression of the breast cancer stem cell (BCSC)-associated transcription factors E2F1 and E2F3 and that E2F1 and E2F3 regulate CASP3 and control apoptosis [150]. Proteins of the miR-34 family have various functions, such as inhibiting cell migration and proliferation and regulating the p53 pathway. p53 is a transcription factor important in DNA damage responses and oncogene activation [151, 152]. One of the important functions of miR-34 is the regulation of EMT, which occurs via the targeting of key genes and signaling pathways involved in EMT control [153]. Research has shown that miR-34a affects CD4 + T-cell infiltration indirectly by changing variables such as the tumor microenvironment, immunological checkpoint molecules, and macrophage polarization [154, 155]. By synthesizing poly(lactic-co-glycolic acid) (PLGA) nanoparticles carrying miR-34a, scientists were able to effectively suppress triple-negative breast cancer (TNBC) cell proliferation and induce cell cycle arrest [156]. Ectopic expression of miR-34a can downregulate survivin, an inhibitor of apoptosis [157], in TNBC cells, stimulate caspase-3 activation and reduce cell growth and migration, and in the MDA-MB-231 TNBC model, it improves the anti-proliferative function of selinexor. These findings suggest that miR-34a can positively modulate the function of anticancer medications [158]. This highlights its potential in further therapeutic studies and drug delivery systems.
MiR-200
miR-200 is another microRNA that has been investigated in many types of cancer, including breast cancer, and its role as a tumor suppressor has been widely researched [159,160,161,162]. A study by Feng et al. suggested that miR-200 interacts with a specific cell communication system known as the Notch signaling pathway, creating a regulatory loop. The balance of this interaction may play a crucial role in determining tumor stage [163].
MiR-200 targets transcription factors such as ZEB1 and ZEB2, which are involved in cell invasiveness and metastasis and subsequently limit EMT [164, 165]. Additionally, it can reduce the population of cancer stem cells by affecting genes involved in stemness [166, 167]. MiR-200 is a crucial factor in the inhibition of genes vital in cell migration and invasion [168, 169]. Furthermore, it can eliminate cells at risk of becoming malignant or with undesirable alternations by causing them to undergo programmed cell death (apoptosis) [170, 171]. One study showed that interleukin-8 (IL-8) promotes EMT in breast cancer cells and increases migration by downregulating the miR-200 family; controlling this pathway can be an effective way to hinder invasion and metastasis [172]. In ER-positive breast cancer cells, Cellular Myeloblastosis Oncogene (c-MYB), which is a transcription factor, is downregulated by miR-200 (miR-200b and miR-200c), and c-MYB is stabilized by EMT-inducing Transforming Growth Factor-β (TGF-β) [173]. This study revealed that IL-8 promotes EMT in breast cancer cells and enhances cell migration and invasion by downregulating the expression of the miR-200 family. This pathway plays a significant role in regulating EMT in breast cancer cells, and modulating it could be a potential method to inhibit invasion and metastasis. By decreasing B-cell leukemia/lymphoma 2 (Bcl-2) expression, miR-200 inhibits cell proliferation, and through reducing the regulation of the multidrug resistance 1 (MDR1) gene, it can increase the effect of doxorubicin on breast cancer cells [174]. Another study suggested that by targeting fibronectin 1 (FN1), which is an important factor in migration, miR-200 controls EMT and increases vulnerability to doxorubicin in cancer cells [175]. On the basis of these findings, miR-200 shows great potential as a biomarker in breast cancer because of its role in regulating EMT, metastasis, and drug sensitivity. Its involvement in key pathways shows its diagnostic and therapeutic value.
MiR-15a/16-1
The miR-15/16 cluster encodes microRNAs (miRNAs), which are known as tumor suppressors. The expression of these miRNAs can hinder cell growth and induce cancer cell apoptosis [176]. MiR-15a and miR-16-1 are generally downregulated in advanced stages of cancer [176], making them valuable elements for studying and investigating cancer progression. One of the main mechanisms by which miR-15a/16-1 functions in tumor suppression involves targeting the antiapoptotic gene BCL2. BCL2 is downregulated by miR-15a and miR-16-1, resulting in an increase in apoptosis and limiting excessive cell proliferation [177]. MiR-15/16 reduces cell proliferation, augments cancer cell death, and decreases tumorigenicity [178]. It can regulate cell proliferation and hinder EMT in breast cancer cells by influencing the Enhancer of Zeste Homolog 2 (EZH2) and Twist1. By forced expression of miR-15/16, the expression of mTOR and ribosomal protein S6 kinase beta-1 (RPS6KB1) can be reduced, and its overexpression can lead to G1-cell cycle arrest [178]. MiR-15a/16-1 has been shown to target genes involved in angiogenesis, preventing the creation of new blood vessels and reducing the tumor’s nutritional supply [179, 180]. The miR-15 family can increase radiosensitivity by causing unrepaired DNA damage, disrupting radiation-induced G2 arrest, and reducing proliferation. This process involves checkpoint kinase 1 (chk1) and Wee1 [181]. The miR-15a/miR-16 cluster can reduce the expression of the oncogene B-lymphoma Moloney Murine Leukemia Virus Insertion Region 1 (BMI1), impair the repair of damaged DNA through apoptosis and homologous recombination, and increase sensitivity to doxorubicin [182]. A study showed that a decreased level of miR-16 expression in the MCF-7/A breast cancer cell line can lead to adriamycin resistance. Additionally, the chemosensitivity of these cells was increased when miR-16 was overexpressed making Wild-type p53-induced phosphatase 1 (Wip1 or PPM1D) and Bcl-2 expression reduced and the cell death spurred by Adriamycin, augmented. Furthermore, luciferase activity was decreased by the overexpression of miR-16 in cells transfected with BCL-2 3′UTRs and PPM1D [183]. Therefore, miR-15/16 has potential to be a new cancer biomarker and a promising candidate for cancer treatment. Tumor suppressor miRNAs and their interactions are shown in Fig. 5.
Oncogenes
Genetic mutations in DNA can change gene structure and expression, which drives cancer development and tumor progression. While traditionally linked to mutations in coding regions of oncogenes and tumor-suppressor genes, recent research highlights the role of non-coding RNAs in cancer. Notably, microRNAs regulate key processes such as the cell cycle and apoptosis and are often dysregulated in tumors [184]. While many microRNAs have the ability to control and suppress cancer, many of them are considered oncogenes [185,186,187]. Meaning, that they can drive cancer progression by suppressing tumor suppressor genes or those involved in cell differentiation and apoptosis [188]. We conducted our review on a selection of more researched ones, listed below.
MiR-21
One of the most significant miRNAs in breast cancer, which is linked to increased cell invasion, proliferation, and resistance to apoptosis, is miR-21 [189,190,191,192]. MiR-21 is located on chromosome 21 and in a region that overlaps the gene for the human papillomavirus (HPV16) on chromosome 17 in the tenth intron of the coding gene transmembrane protein 49 [193]. Research has revealed a significant link between high levels of miR-21 expression and advanced breast cancer stages [194], which can potentially serve as a biomarker to differentiate advanced stages from early stages of the disease. Additionally, the overexpression of miR-21 has often been observed in various cancers, such as lung, breast, stomach, prostate, colon, and pancreatic cancers [195]. A study showed that removing miR-21 from MDA-MB-231 breast cancer cells results in the downregulation of mesenchymal markers and a decrease in epithelial‒mesenchymal transition. The findings of this study identified Wnt-11 as an important target of miR-12. Furthermore, the number of extracellular vesicles was reduced in cells with deleted miR-12 [196]. MiR-21 can increase metastasis and proliferation in breast cancer by suppressing leucine zipper transcription factor-like 1 (LZTFL1), which is a vital gene in cancer progression, by removing LZTFL1, these effects can be reversed. Disruption of LZTFL1 can neutralize the effect of miR-21 on the expression of EMT markers through the regulation of β-catenin, snail, and slug by miR-21/LZTFL1 [196]. In triple-negative breast cancer, Nestin (NES) is a protein associated with poor prognosis [197], the thyroid hormone receptor interactor 13 (TRIP13) protein is related to DNA repair [198], and reticulum membrane protein complex subunit 1 (EMC1) is another protein that controls the actin cytoskeleton and enhances angiogenesis [199]. The findings of a previous study indicated that miR-21 upregulates these three proteins and promotes metastasis via an immune cell-independent approach [191]. MiR-21 contributes to breast cancer by downregulating the expression of phosphatase and tensin homolog (PTEN), a tumor suppressor that stimulates TBNC cells to proliferate and invade [200]. These findings suggest that miR-21 may prevent the death of breast cancer cells by encouraging their division and survival. Moreover, miR-21 has been linked to poorer prognosis, lymph node involvement, and larger tumors in individuals with breast cancer [201]. Another study indicated that a high level of miR-21 can lead to a reduction in the expression of PTEN (a tumor suppressor) and trastuzumab resistance in HER2-expressing breast cancer cells. This study further showed that by employing miR-21 antisense oligonucleotides (ASOs) or a PTEN rescue vector via the removal of the miR-21 targeting sequence, it is possible to reverse the resistance created against trastuzumab drugs and reverse PTEN expression [202], demonstrating the therapeutic use of miR-21 in breast cancer treatment.
MiR-221/222
The miR-221/222 cluster is known to act as an oncogene in breast cancer, driving tumor development, inhibiting apoptosis, and fostering medication resistance [203,204,205]. It can be involved in epithelial-to-mesenchymal transition in breast cancer through the negative regulation of Notch3, which is an important factor in limiting EMT [206]. Previous studies reported increased expression levels of miR-221 and miR-222 in breast cancer tissues compared with normal breast tissues, and higher levels of these microRNAs were associated with more advanced clinical stages of the tumor [207]. MiR-221 and miR-222 are highly expressed in various tumors, such as those in the colon, pancreas, and stomach [208]. MiR-221/222 play a role in luminal-like breast cancer by regulating invasion, proliferation, and tumor growth by targeting signal transducer and activator of transcription signal transducer and activator of transcription 5A (STAT5A), disintegrin and metalloproteinase domain-containing protein 17 (ADAM17), and β4-integrin [209]. These microRNAs are linked to S-phase entry, cellular migration, and the G1/S transition of the cell cycle, and they are overexpressed in highly invasive basal-like breast cancer cells [205]. Additionally, they target the growth arrest-specific 5 (GAS5) tumor suppressor gene, which promotes the growth of tumors [204]. Researchers have shown that, through the inhibition of miR-221/222, sensitivity to tamoxifen in tamoxifen-resistant breast cancer cell lines can be restored. It was confirmed that miR-221/222 inhibitors derepress estrogen receptor alpha (Erα) and PTEN expression, which can result in a decrease in cell growth and cause cell cycle arrest [210]. Another study on resistance to tamoxifen suggested that the repression of miRNA-221/222 enhances the sensitivity of estrogen receptor (ER)-positive MCF-7 breast cancer cells to tamoxifen by increasing the expression of tissue inhibitor of metalloproteinases 3 (TIMP3), which is involved in limiting migration and tumor growth [211]. Another study highlighted the importance of miR-221/222 in breast cancer through the activation of the Akt/nuclear factor kappa B (NF-κB)/cyclooxygenase-2 (COX-2) pathway through the downregulation of PTEN [212]. The effect of miR-221/222 on cisplatin drug sensitivity in TNBC was investigated, and it was found that the removal of miR-221/222 can lead to the upregulation of suppressor of cytokine signaling 1 (SOCS1) and cyclin-dependent kinase inhibitor 1B (p27), which in turn inhibits signal transducer and activator of transcription 3 (STAT3) expression. This cascade can reduce the expression of c-Myc and Bcl2 and upregulate p53 and Pten, which participate in increased cell apoptosis in response to cisplatin [213]. Given the role of miR-221 and miR-222 in regulating tumor progression, invasion, proliferation, and drug resistance in breast cancer, these non-coding RNAs have potential as biomarkers for this disease. Many studies have highlighted their diagnostic and therapeutic relevance, supporting their use as indicators for breast cancer prognosis and treatment response [214,215,216].
MiR-155
MiR-155 is found inside the B-cell integration cluster (BIC) on chromosome 21 [217]. miR-155 has been linked to the development of several tumors, their growth, and resistance to treatment [218,219,220]. It was previously shown that increased miR-155 expression in breast cancer is linked to increased tumor grade and advanced disease stage. Additionally, higher miR-155 levels are associated with poorer overall survival, indicating its potential as a clinically significant miRNA [221, 222]. MicroRNA-155 is elevated across various cancer types, including breast, prostate, liver, lung, and pancreatic [223]. In breast cancer patients, the level of circulating miR-155 (in the blood) is highly increased, whereas after treatment, this level is decreased [224]. These findings highlight it as an important biomarker. miR-155 can upregulate MMP16, a cytokine that regulates tumor migration and invasion; additionally, miR155 enhances proliferation through negatively regulating SOCS1 (tumor suppressor gene) and turning on the critical inflammatory Janus kinase/signal transducers and activators of transcription (JAK-STAT) signaling pathway [225]. The nucleotide-binding domain, leucine-rich–containing family, pyrin domain–containing-3 (NLRP3) inflammasome regulates the immune system; once activated, it triggers the release of proinflammatory cytokines [226]. Research has shown that disruption of miRNA-155 can suppress the NLRP3 pathway in MDA-MB-231 cells, increase apoptosis and prevent the proliferation, migration and secretion of inflammatory factors [227]. It was discovered that miR-155 enhances the development of stem cell growth by targeting tetraspanin 5 (TSPAN5), which is important for signaling and cell differentiation, affecting resistance to decitabine drugs in TNBC cells [228]. Another study demonstrated that by inhibiting the expression of miR-155 in ATP-binding cassette subfamily G member 2 (ABCG2), the expression of CD44 and CD90 decreases while that of CD24 increases, and the inhibition of miR-155 enhances the sensitivity of MDA-MB-231 cells to the drug doxorubicin [229]. According to one study, Schizandrin A (SchA) can reduce the migration and proliferation of breast cancer cells by decreasing PI3K/AKT and Wnt/β-catenin through the negative regulation of miR-155 expression [230]. MiR-155 plays an important role in breast cancer progression, regulating processes such as tumor migration, invasion, proliferation, and treatment resistance. Its increased expression in tumor tissues and the circulation, along with its reduction following treatment, shows its potential as a biomarker. Oncogenic miRNAs and their interactions can be observed in Fig. 6.
More information on miRNAs and their associations with cancer can be found in Table 4.
Experimental approaches
To understand the functions, expression patterns and regulatory mechanisms of non-coding RNAs (ncRNAs) in cancer, a variety of experimental methods and approaches have been employed. In this review, we cover some of the most known of them below.
Microarray analysis
Microarray analysis plays a significant role in understanding the function of non-coding RNAs in breast cancer [231]. It can be used to compare the expression of miRNAs and lncRNAs in breast cancer tissues to that in normal tissues. This approach has been utilized in many investigations [231,232,233]. Microarrays consist of glass or silicon slides with DNA probes arranged in a grid-like pattern [234]. Samples of mature miRNAs (reference and test) should be purified and then converted to cDNA via reverse transcription. The samples are labeled with fluorescent dyes, and fluorescent green and red labels are provided for the reference and test samples, respectively. Next, two samples are mixed together and applied to high-affinity probes on the array. By measuring the amount of red and green fluorescence at each location, the relative number of red- and green-labeled fragments can be determined, which allows the comparison of numbers between reference and test samples [235,236,237]. Figure 7. A study focused on epigenetic and genetic factors related to the expression of miRNAs in two populations of Lebanese women and American women. By employing microarray analysis, they identified miRNAs that regulate tumor-related mRNAs, and their results demonstrated ethnic and age-associated differences in miRNA expression [238]. Researchers have used microarrays and unsupervised cluster analysis to measure the expression of miRNAs in breast cancer cell lines. These findings indicate that specific miRNAs are associated with genetic and epigenetic alterations in various subtypes of breast cancer cell lines, such as Erb-B2 receptor tyrosine kinase 2 (ERBB2) overexpression, E-cadherin mutation, and promoter hypermethylation. LessFewer miRNAs are associated with phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha (PIK3CA)/PTEN, breast cancer type 1 susceptibility protein (BRCA1), and tumor protein p53 (TP53) mutation status [239].
Another study focused on non-small cell lung cancer (NSCLC) patients who underwent surgery at a Chinese hospital. Tumors and adjacent tissues were collected, preserved, and analyzed for circular RNA (circRNA) profiles. By employing microarray analysis, total RNA was treated with RNase R to remove linear RNAs and to enrich circRNAs. Next, the circRNAs were labeled and hybridized onto Arraystar Human circRNA Arrays. Scanned images were processed, and circRNA profiles were analyzed via quantile normalization, cluster analysis, fold-change filtering, and student’s t-testing. Two differentially expressed circRNAs (hsa_circ_0014130 and hsa_circ_0016760) were validated via quantitative reverse transcription‒polymerase chain reaction (qRT‒PCR). Network analyses revealed possible interactions between circRNAs, microRNAs, and mRNAs, providing insights into the regulatory mechanisms involved. Functional analyses explored the roles of circRNAs and their associated genes. These findings indicate that hsa_circ_0014130 can play a significant role in the development of non-small cell lung cancer (NSCLC) and can be employed as a marker [240].
Microarray analysis is a promising tool in ncRNA discovery and is praised for its ability to monitor transcript levels fully; however, it has some limitations that are worthy of note. Issues include the high cost of the experiment, dependence on low-specificity probes, and low control of the studied transcript pool since most platforms use manufacturer-designed probes; challenges in accuracy, precision, and specificity; and susceptibility to changes in hybridization settings, genetic material quality, and amplification procedures [241,242,243,244].
Next-generation sequencing (NGS)
Sequencing began a lengthy journey with first-generation sequencing and Sangers' work. Next-generation sequencing refers to second- and third-generation sequencing technologies, which include the Roche 454, Illumina Solexa, ABI-SOLiD, Helicos, and PacBio, and the fourth generation consists of Nanopore sequencers offered by Oxford Nanopore Technologies (ONT) [245]. Transcriptomics or expression profiling is the study of the transcriptome, which is the entire collection of RNA [246]. Next-generation sequencing significantly affects transcriptomics and makes it possible to study the complete collection of RNA molecules, including non-coding RNAs. It can be employed in several approaches, such as sequencing mRNAs, analyzing alternative splicing (a process with transcriptome complexity), studying lncRNAs, and assembling the transcriptome [247]. Next-generation sequencing (NGS) technology is a critical step in identifying ncRNAs and their link to breast cancer [248] and plays a significant role in tumor investigation, enabling breakthroughs in tumor genetic profiling [249]. Any NGS technology results in a large amount of output data. The fundamentals of sequence analysis follow a centralized approach that comprises raw read quality control (QC), preprocessing and mapping, postalignment processing, variant annotation, variant calling, and visualization [250]. For example, one of the well-known NGS approaches is Illumina, and its workflow consists of RNA isolation, preparation of the desired non-coding RNA sequencing library, mapping and identification, functional enrichment analysis, qRT‒PCR and statistical analysis [251]. A study created a complementary DNA (cDNA) library of HCT-8 VCR-resistant colon cancer cells and analyzed it via HiSeq 2500 sequencing and bioinformatics methods to identify differentially expressed lncRNAs in drug-resistant cells and nonresistant cells. The results of this research revealed a greater total transcript number in resistant cells, with the majority being high-quality transcripts, and revealed 121 transcripts whose expression significantly differed between the two cell types [252]. In another study, researchers studied microRNAs in 11 normal samples and 104 cancerous breast tissues via next-generation sequencing (NGS), and the results revealed two novel miRNAs, namely, miR-574-3p and miR-660-5p, related to breast cancer and 10 other miRNAs linked to overall survival (OS) and/or recurrence-free survival (RFS) [253]. Krishnan et al. used the Illumina genome analyzer IIx to study snoRNAs in 104 breast cancer and 11 normal breast tissues. Their findings revealed that 13 snoRNAs are associated with patient outcomes. These results indicate that snoRNAs interact with other RNAs involved in cancer development and can indirectly regulate gene expression associated with tumorigenesis [254].
However, this approach is not without limitations. In technologies such as Roche 454 (GS FLX plus), sequencing the homopolymer regions is difficult [255]. In Illumina, uneven representation of DNA fragments with different guanine‒cytosine (GC) contents in the amplified clusters is the main limitation [256], with a higher error rate in placebo, particularly in terms of insertion and deletion [257]. Despite these challenges, NGS stands as a powerful tool for high-throughput sequencing applications.
Quantitative real-time PCR
Polymerase chain reaction (PCR) is an approach employed to amplify DNA and create many copies from it [258]. It has three main repeated steps consisting of melting through high temperature, annealing to designed primers and replication via enzymes such as Taq DNA polymerase [259]. Real-time PCR or quantitative PCR (qPCR) is a PCR-based technique that detects products in real time and does not require gel electrophoresis for analysis [260, 261]. The fluorescent reporter/dye used in qPCR produces fluorescence related to the amount of amplified DNA product (amplicon). The fluorescent molecules utilized in qPCR can be fluorescent tagged probes such as TaqMan or DNA binding dyes such as SYBR Green. The fluorescence emitted in each qPCR cycle will be detected and quantified [260]. Reverse transcription‒quantitative polymerase chain reaction (RT‒qPCR) is used to determine the amount of RNA in a sample, where RNA is converted to cDNA through reverse transcription and then amplified via qPCR [262]. RT‒qPCR can be a useful method for examining non-coding RNAs (ncRNAs) in the context of breast cancer and has been used in several studies [263,264,265]. Breast cancer tumors and adjacent nontumor tissues from female patients were examined, RNA extraction and cDNA synthesis were performed, and a quantitative real-time PCR technique was used to evaluate differentially expressed genes. The results of this study demonstrated that the lncRNA MCM3AP-AS1 was significantly overexpressed in breast tumor tissues compared with adjacent nontumor tissues. This overexpression is related to estrogen and progesterone receptor expression in tumors [266]. Another study employed a mixture of microarray and PCR to analyze the expression of MALAT1 in 195 cases of benign and malignant thyroid neoplasms; with these two approaches, they discovered a role for MALAT1 in EMT in thyroid tumors and reported that the induction of epithelial-to-mesenchymal transition (EMT) in a papillary thyroid carcinoma (PTC) cell line led to augmented MALAT1 expression [267]. Researchers have used RT‒qPCR to evaluate the expression levels of prostate cancer-associated non-coding RNA 1 (PRNCR1) in breast cancer, and their results demonstrated that this lncRNA is expressed at higher levels in breast cancer tissue than in normal tissue. Furthermore, high expression of PRNCR1 was linked to poor patient prognosis and metastasis, and by reducing PRNCR1 gene expression, the aggressive characteristics of breast cancer cells were decreased [268]. There are several drawbacks to the use of RT‒qPCR in the study of non-coding RNA; for example, miRNA is a short RNA molecule and is difficult to quantify by RT‒qPCR, necessitating the use of special kits for cDNA synthesis. In longer RNA molecules, hairpin loops are a challenge in cDNA synthesis, which demands more temperature to make them available for reverse transcriptase [269, 270]. Moreover, ncRNAs such as lncRNAs are less stable and prone to degradation because of their length [271], which can impact their quantification. Additionally, the localization of lncRNAs in the genome influences their transcript stability [272, 273].
Northern blotting
The northern blot technique is a widely used laboratory method for the analysis of RNA and provides essential information about gene expression. This method involves the separation of purified RNA fragments from a biological sample, such as blood or tissue, by passing them through a sieve-like matrix or gel. The separation process is based on the size of the RNA fragments, with smaller fragments moving more quickly through the matrix or gel than larger fragments. The RNA fragments are transported from the gel or matrix to a solid membrane, which is then exposed to a radioactive, fluorescent, or chemically tagged DNA probe [274]. Figure 8. The tag enables the visualization of any RNA fragments bearing complementary sequences to the DNA probe sequence within the Northern blot. Northern blot analysis can be performed to determine whether a gene is overexpressed or underexpressed in cancer cells compared with normal cells [275]. The size and expression levels of non-coding RNAs in breast cancer tissues have been investigated, and these RNAs have been used in several studies [276,277,278]. One study on the lncRNA DIO3 Opposite Strand Upstream RNA (DIO3OS) revealed that this RNA causes glycolytic-dominant metabolic reprogramming to increase aromatase inhibitor resistance in breast cancer. Northern blotting was used as part of this method in the following way: the RNA was loaded onto a gel, divided by electrophoresis, and transferred to a nylon membrane. The membrane was then treated with buffers and incubated with labeled oligonucleotides overnight, and the resulting hybrids were detected via a luminescent detection kit [267]. Another study employed northern blotting on U2OS osteosarcoma cells. Total RNA was separated and transferred to membranes before UV cross-linking. Digoxin-labeled oligonucleotide probes were made and utilized for hybridization, followed by visualization via an imaging analysis system. The results revealed a main long non-coding RNA (Lnc8) transcript at approximately 1.25 kilobases. This transcript was found to decrease under conditions of glucose deprivation [279]. Researchers have shown the differential expression of the lncRNA LINC00908 in triple-negative breast cancer (TNBC) via northern blot analysis. These findings revealed that the protein ASRPS produced by LINC00908 can reduce the phosphorylation of STAT3 and decrease the level of vascular endothelial growth factor (VEGF), which is involved in the growth of blood vessels. By increasing ASRPS levels in mouse models, angiogenesis was reduced, and the expression of ASRPS was decreased in TNBCs [280].
It is undeniable that northern blotting is an extensively used method for the analysis of genes at the RNA level and guides the detection of transcript size; however, similar to any other approach, it has limitations, such as the need for high-quality RNA and the lengthy time it requires [281, 282].
Fluorescence in situ hybridization (FISH)
Joseph Gall and Mary Lou Pardue pioneered fluorescence in situ hybridization (FISH) in the 1960s [283], which was followed by John et al. in [284]. This method has various applications, such as utilization in gene mapping, locating DNA sequences and recognition of new oncogenes in cancer [285]. RNA-FISH is a powerful method for evaluating RNA transcription and can provide information about degradation rates [286]. RNA fluorescence in situ hybridization (FISH) can be used to determine the cellular localization of lncRNAs by creating a lncRNA-specific antisense chain sequence and tagging it with fluorescent dyes [287]. Therefore, it has been employed in several studies [288,289,290]. The steps involved designing a probe, fixation, permeabilization, hybridization, posthybridization washing, sequence amplification and detection [291]. A study in 2020 revealed that high expression of the lncRNA NONHSAT028712 (Lnc712) is associated with breast cancer. They used RNA-FISH as a method to localize Lnc712 within the cell, and the results demonstrated its existence in the cytoplasm, which might be due to its role in the regulation of cytosolic processes [292]. Hongnan Jiang et al. suggested a correlation between lncRNAs, small nuclear protein RNA host gene 3 (SNHG3), and breast cancer via microRNA-154-3p; they used probes targeting SNHG3 via FISH to identify the location of a given lncRNA, and the results of their experiment confirmed its presence in the cytoplasm within HCC1937 and MCF-7 cells [293]. In another study on the role of small nucleolar RNA host gene 7 (SNHG7) in liver cancer, RNA FISH was performed on hepatocellular carcinoma (HCC) tissue cells and adjacent normal tissues via a fluorescent in situ hybridization kit. The Cy3-conjugated SNHG7 probe was used for detection, and fluorescence images were acquired via a confocal microscope. These results demonstrated that the expression of SNHG7 was increased in hepatocellular carcinoma tissues compared with adjacent normal tissues [294]. The limitations of FISH methods can be considered as follows: long hybridization times are needed for saturation of the signal, making this method time-consuming and demanding experienced personnel. Another challenge is that few commercially available probes and laboratories produce their own probes, and the process of creating, preparing, and labeling these probes still requires much effort. Additionally, adjusting cutoffs for FISH probes can be a technical challenge in laboratories. The cost of probes, especially at a relatively high scale, is relatively high [295, 296]. Enhancing automation and hindering hybridization time are platforms for improving the efficiency of the FISH method.
CRISPR-Cas9
Initially, found in bacteria, clustered regularly interspaced palindromic repeats (CRISPR) is an effective system in bacteria for defending against phages and plasmid DNAs as part of adaptive immunity [297]. Cas9 is a nuclease enzyme in the type II CRISPR system from S. pyogenes that is guided by single guide RNA (sgRNA) to target the DNA site [298]. The sgRNA is composed of two parts, CRISPR RNA (crRNA) and transactivating CRISPR RNA (tracrRNA), and it interacts with the enzyme Cas9 [299]. They work together to find the target DNA sequence, break it, and enable gene editing. CRISPR/Cas gene editing can provide knockdown of certain non-coding RNAs by deleting the transcriptional termination or start site and removing the promoter and exon‒exon junctions [300], making it a valuable approach for harnessing the effects of non-coding RNA in cancer. The CRISPR/Cas9 system is known as a molecular scissor and is widely used in a variety of studies, including cancer research, drug development, mental disease therapy, and plant applications [301]. Increasing evidence suggests that CRISPR/Cas9 can target human ncRNAs as well as the protein-coding genome [302, 303]. This system has been used in a variety of studies [304,305,306]. A study by Hebatalla Said Ali et al. revealed that the levels of the genetic network were reversed when CRISPR Cas9 was used to knock out the lncRNA RP11-156p1.3 in a human liver cancer cell line (HepG2), and the proportions of tumor necrosis factor alpha (TNF-α) and NFκβ were reduced as a result of this experiment, highlighting the use of CRISPR-Cas9 in hepatocellular carcinoma [307]. Another study used CRISPR-Cas9 to knock out mature miR-3662 in TNBC cells, resulting in a reduction in the proliferation and migration of tumors in vivo [308]. Researchers have used the CRISPR/Cas9 system to deactivate maternally expressed gene 3 (MEG3) in gene-expressing breast cancer cell lines. They verified the knockout via PCR and RT‒qPCR. These results showed that deleting MEG3 makes cells resistant to doxorubicin drugs. Additionally, it leads to increased levels of certain proteins, such as N‑cadherin and transforming growth factor β, while reducing the levels of factors such as the collagen type III α1 chain, zinc‑finger E‑box binding homeobox 1, and matrix metallopeptidase 2 [309]. Another in vitro and in vivo study revealed that knocking out miR-23b and miR-27b via CRISPR/Cas9 decreased tumorigenesis [185]. It is undeniable that CRISPR/Cas9 is a revolutionary and valuable tool, but its limitations and challenges should be considered to achieve better results. One notable complexity is the gene locus that must be estimated for sgRNA design and data analysis [310], which can cause off-target results and changes in sequences that are not aimed at [311]. This can be due to the untargeted function of Cas9, which cleaves DNA sites other than the intended regions [312].
CLIP-Seq (crosslinking and immunoprecipitation followed by sequencing)
With cross-linking immunoprecipitation and high-throughput sequencing (CLIP-seq), we can determine RNA targets connected to a certain RNA-binding protein with significant precision [313]. The method isolates particular RNA‒protein complexes by crosslinking them with UV light and then purifying the resulting mixture under strict conditions. After the RNA fragments have been extracted, techniques such as RNA linker ligation, RT‒PCR amplification, and sequencing can be used to search for native binding sites of RNA-binding proteins [314]. This approach has been applied across several investigations [315,316,317]. For example, Junhao Li et al. used CLIP-seq to construct miRNA–mRNA and RNA–lncRNA interaction networks and identified approximately 10,000 ceRNA pairs from miRNA target sites. Additionally, miRNA-mediated regulatory networks can be used to predict the roles of miRNAs and other non-coding RNAs [318]. Michael Lidschreiber et al. identified protein systems for the creation of mature eukaryotic protein-coding and non-coding RNAs via CLIP-seq and reported that sync1/APT (a specific protein involved in the regulatory mechanisms of the cell cycle and transcription in yeast) and cleavage and polyadenylation factor (CPF) have different functions, which might overlap. Syc1/APT is more involved in sn/snoRNA construction, whereas CPF is responsible for handing out the 3'-ends of protein-coding pre-mRNAs [319]. The results of another study that employed CLIP-Seq data demonstrated that the RNA-binding protein FUS/TLS interacts with NEAT1 and that, by hindering fused in sarcoma/translocated in liposarcoma (FUS/TLS) expression, cell apoptosis can be triggered. Additionally, they suggested that MiR-548ar-3p interacts with NEAT1 and downregulates it [320]. One of the limitations of CLIP-seq is its ability to measure in vivo protein binding occupancy on transcripts because of the challenges in quantifying binding occupancy, which is important in understanding the relevance of binding sites. For example, the regulatory impact of a protein binding to all 20 copies of one transcript in a cell varies from protein binding to 20 out of 200 copies of another transcript. Future improvements are needed to improve protein‒RNA capture efficiency and increase protein capture strength [321].
Bioinformatics tools
The study of non-coding RNAs (ncRNAs) in breast cancer is a complex and evolving field, with a variety of bioinformatic tools being used to understand their role. MiRBase, miRWalk, LNCipedia, and NONCODE are four popular bioinformatic tools for non-coding RNAs related to breast cancer. miRBase (https://mirbase.org/) [322] is a comprehensive database for microRNA sequences and annotations that provides valuable information on the expression patterns and potential functions of non-coding RNAs in breast cancer. This approach can aid in the discovery of biomarkers and medicinal targets. The database also aids in the prediction of novel microRNAs and their targets in cancer [323]. MiRWalk (http://mirwalk.umm.uni-heidelberg.de/) [324] is another comprehensive online resource that is employed for predicting and validating microRNA (miRNA) binding sites in genes from humans, mice, rats, dogs, and cows [325]. It includes a comparison platform of miRNA-binding sites, genetic networks of miRNA-gene pathways, and analytically validated information [326]. Since the database is constantly updated and openly accessible, it is a great resource for researchers studying miRNA‒target interactions [327]. LNCipedia (https://lncipedia.org/) [328] is an annotated human long non-coding RNA (lncRNA) sequencing database that provides a high-confidence list of lncRNA transcripts with low coding potential [329]. It has been used to create a custom microarray for profiling lncRNA expression [330]. NONCODE (http://www.noncode.org/) [331] is a database of non-coding RNAs (ncRNAs), in addition to transfer RNAs and ribosomal RNAs, that covers a wide range of these molecules. It contains detailed information on the sequencing, functions, and cellular roles of ncRNAs and offers an easy-to-use interface [332].
Non-coding RNAs as biomarkers for diagnosis and prognosis
In recent years, significant advancements have been made in the diagnosis, subtyping, and complex treatment of breast cancer. Traditional methods for diagnosing breast cancer have been supplemented by modern molecular methodologies. NcRNAs, including miRNAs, lncRNAs, and circRNAs, are recognized as pivotal in tumorigenesis and have shown promise in the early detection, diagnosis, prognosis and treatment response prediction of multiple cancers [333, 334].
MiRNAs regulate various biological processes, such as proliferation, cell adhesion, motility, and apoptosis, all of which are fundamental to tumorigenesis. Studies suggest that the abnormal expression of miRNAs might be clinically useful as predictive markers, potential therapeutic targets, and suppressors of resistance to certain anticancer chemotherapies. Tumor-derived miRNAs exist in the circulating nucleic acids of cancer patients, making their detection a favorable potential diagnostic method, as it is noninvasive [335, 336]. More than 3000 miRNAs associated with the occurrence and progression of tumors have been identified [334]. Some of these findings are summarized in Table 5.
Another group of ncRNAs, lncRNAs, are key regulators of tumor development and progression, and their abnormal expression is closely linked to breast cancer initiation and metastasis [337]. Notable lncRNAs, such as H19, MALAT1, and DANCR, influence cell survival, proliferation, metastasis, and resistance therapy, making them promising biomarkers for breast cancer diagnosis, prognosis, and risk management [338, 339]. For example, MALAT1 overexpression is correlated with poor patient survival, and its expression level can be a potential prognostic factor [340]. Similarly, LINC01787 is upregulated in breast cancer tissues and is associated with advanced stages and poor survival [341].
While ncRNAs show significant potential as biomarkers, their clinical utility faces several challenges. Finding a suitable molecular biomarker remains difficult, as it must be stable, with sufficient specificity, easily detectable, and ideally applicable to all patients. Since the content of ncRNAs in body fluids is not high, they are difficult to detect. Additionally, variability between preclinical models and humans, as well as between individuals, complicates their reliability [333]. Addressing these limitations requires further research to validate the potential of ncRNAs as biomarkers.
Clinical trials
Various endeavors have been made to demonstrate the usage of ncRNAs in clinical settings, including cancer therapy. For example, miRNAs can act as either conventional tumor suppressors or oncogenes and target multiple genes simultaneously; therefore, they are promising cancer therapy agents. Before any testing in humans, preclinical studies are conducted in vitro, in vivo, ex vivo, or in silico to determine a drug, procedure, or treatment’s efficacy and safety. In this section, we outline some of the preclinical studies that have investigated the therapeutic potential of non-coding RNAs in breast cancer, followed by clinical trials that were carried out in humans. One study in 2021 was conducted on multiple human breast cancer cell lines using a miR-100 mimic as well as miR-100 inhibitors. Oligonucleotides were transfected into breast cancer cells. Taken together, the results of this study revealed that, by targeting FOXA1 expression, miR-100 inhibits the proliferation, invasion, and migration of breast cancer cells [342]. SUM159PT (a triple-negative breast cancer cell line) was transfected with plasmid expression vector (pEP)-miR-205-miR205 and pEP-empty vectors, followed by a trypan blue exclusion assay, cell lysis and immunoblotting, quantification of cell migration, determination of invasiveness, and calculation of sphere-forming efficiency. MiR-205 drastically reduced the proliferation, migration and invasion properties of the SUM159 cell line. An analysis of the expression of different markers involved in the epithelial–mesenchymal process and the ability of cells to form mammospheres revealed that miR-205 inhibits cancer stem cell renewal and affects several parameters associated with the initial steps of tumorigenesis, suggesting that miR-205 is a potential therapeutic agent for triple-negative breast cancer. [343]. Kim et al. demonstrated that miR-145 can inhibit the growth and motility of breast cancer cells both in vitro and in vivo [344]. Multiple studies have been conducted using miR-21 antagonists in various cell lines and mouse models via different delivery systems, such as transfection and different types of nanoparticles. The results demonstrated that the combination of these drugs suppressed tumor growth and angiogenesis and reduced cancer cell migration while also improving the efficacy of anticancer drugs [345,346,347,348]. In another study, miR34-a-treated tumors from nude mice were significantly smaller than those from control mice [344, 349]. More studies on ncRNAs as treatment agents as well as human clinical trials will aid in the development of effective ncRNA-based therapeutic agents for treating malignant diseases [350].
MiR-34a is a naturally occurring tumor suppressor that is mutated, inactivated, or expressed at low levels in a wide range of different tumors. In the case of breast cancer, miR-34a is downregulated compared with that in healthy tissue. Studies on the exogenous introduction of miR-34a mimics in vitro have shown decreased cell proliferation, migration, and invasion [351, 352]. When combined with other anticancer therapies, miR-34a mimics have shown synergistic effects. Compared with miRNA-34a or docetaxel alone, codelivery of miRNA-34a with nanocarriers prolonged the blood circulation of docetaxel, improved tumor accumulation, and significantly inhibited tumor growth and metastasis in vivo in mouse models [353].
MiR-34a has been administered via different routes in preclinical animal models, where it inhibits primary tumor growth, blocks metastasis, and improves survival.
MRX34 is a liposomal formulation of miR-34a. A clinical trial was conducted on adult patients with solid tumors refractory to standard treatment in 2017. Among the participants, 3 patients (6 percent) suffered from breast cancer. Patients were given MRX34 intravenously twice weekly for three weeks in 4-week cycles. Dexamethasone premedication was administered to manage infusion-related adverse effects. This treatment showed acceptable safety and antitumor activity in a subset of patients with refractory advanced solid tumors. However, the trial faced significant challenges, including serious immune-mediated adverse events, which led to it ending early after four patients died. The maximal tolerated dose for the twice-weekly schedule was 110 mg/m2 for nonhepatocellular carcinoma patients and 93 mg/m2 for HCC patients. The final report included pharmacodynamics, determination, and evaluation of the recommended phase 2 dose of a different daily treatment schedule (MRX34 daily for 5 days along with dexamethasone premedication twice daily for seven days in week 1 followed by two weeks of rest in 3-week cycles), in which some antitumor activity was observed.
While the trial showed promise, it also highlighted several limitations and challenges that must be addressed in future studies. The unexpected immune-mediated adverse events showed difficulty in translating findings from preclinical models to humans as these events did not occur in preclinical studies. It is important to anticipate toxic effects that may not be evident in preclinical toxicology studies, as such models often fail to capture complex human immune responses. In addition, while MRX34 treatment with dexamethasone premedication showed manageable toxicity in most patients, its effect did not prevent severe reactions in all patients. The lack of adequate preclinical and animal studies on dosing regimens and toxicity profiles made it more difficult to predict MRX34’s safety profile in a clinical setting. While the trial demonstrated that miRNA-based cancer therapy is potentially useful in terms of clinical activity, such as dose-dependent modulation of miR-34a target genes (in patients’ white blood cells) and miR-34a localization in tumors, effectively delivering RNA constructs and further optimizing delivery systems remain critical areas for improvement so that systemic exposure is reduced and side effects become more manageable. Although adverse immune-mediated effects have occurred, the MRX34 trial was the first to study the effects of the administration of a synthetic miRNA to human patients and therefore provides a direction for the application of MRX34 in cancer therapy. Further studies on alternative dosing schedules, combination therapies, and premedication regimens are needed. Overall, the use of RNA constructs requires further development to prevent immune-related toxicity and be clinically useful in humans [352, 354].
Recent clinical trials have been investigating the role of ncRNAs in the diagnosis, prognosis, and treatment of breast cancer. Although results are pending or have yet to be published, these studies significantly contribute to the understanding of ncRNAs as biomarkers for breast cancer Table 6.
Discussion
The evolution of ncRNA research
Research on ncRNAs began a long journey, with the discovery of rRNAs in 1955, which led to their employment in clinical platforms in 2013 [46, 66]. During this process, various endeavors have been made to understand their multiple functions in the body. Projects such as ENCODE have advanced the understanding of their role, shedding light on the regulatory mechanism of ncRNAs [18]. With the increasing number of cancers, efforts to find effective treatments are increasing. Most anticancer treatments eventually lead to some level of resistance among patients and involve serious side effects. Advances in technologies such as microarray analysis, next-generation sequencing (NGS), and real-time quantitative reverse transcription (qPCR) have helped researchers determine the delicate functions of ncRNAs in cancer, and numerous studies have investigated their roles.
NcRNAs: a double-edged sword in breast cancer
MiRNAs and lncRNAs are two of the most studied types of ncRNAs. Famous lncRNAs, including MALAT1, NEAT1 and HOTAIR, have been fully studied. They can act as oncogenes and promote cancer via multiple pathways, including EMT, angiogenesis, autophagy, proliferation and metastasis. Moreover, they play an instrumental role in the development of resistance in cancer cells against radiation and anticancer medications [106, 112, 117, 118, 128, 134]. MicroRNAs are other popular ncRNAs in the literature, and their role has been discussed as both oncogenes and tumor suppressors [144]. For example, miR-21 has been considered an oncogene in various studies [189,190,191,192], whereas ncRNAs such as miR-34a are known as tumor suppressors [145, 146].
The study of ncRNAs is a dynamic field; even among the most well-known types of ncRNAs, notable differences in the results of researchers can be observed. For example, MALAT1 has been identified as an oncogene in various cancers, whereas in a study by Jongchan Kim et al., it was demonstrated to be a tumor suppressor. In this study, MALAT1 was reported to bind to the transcriptional enhanced associate domain (TEAD), which is a prometastatic transcription factor, inactivating it and preventing it from interacting with the target gene promoter and coactivator Yes-Associated Protein (YAP), and a negative correlation between MALAT1 and the spread of breast cancer was shown [355]. The results of a study performed by Zhang Yang et al. were consistent and reported the downregulation of MALAT1 in breast cancer cell lines and tissues. Cell proliferation was augmented when its expression was inhibited. Additionally, they demonstrated that in MALAT-depleted cells, a large increase in cyclin D1 (CCND1) mRNA and protein levels can be observed [356]. In a review paper published by Nadine Smith et al., it was mentioned that NEAT1 can play a role in both parts depending on its specific form. They highlighted the importance of studying different isoforms of NEAT1 with respect to cancer progression and usage in therapy [357]. In Chunfu Zhang et al.’s study, miR-21 is considered to play a dual role, as it can affect both breast cancer genes and tumor suppressor genes. Interestingly, when cancer gene suppression is dominant in tumor cells, it prevents tumor growth, and when suppression of tumor suppressor genes is abundant, these genes act as oncogenes and augment cancer progression [190]. A study by Junfeng Wang focused on the levels of miR-155 in breast cancer and reported a positive relationship between higher levels of miR-155 and the antitumor immune profile. Experiments in mice demonstrated that miR-155 overexpression upregulated CXCL9/10/11 production [358].
Future direction
The inconsistency observed in the literature illustrates the importance of studying ncRNA at a more in-depth level. Notably, despite a vast amount of research examining and validating the use of ncRNAs as potential therapeutic agents in cancer treatment, the results of the most recent clinical trial revealed significant adverse effects, which is a crucial aspect to look at. The large number of studies on cancer cell lines and the small number of studies on mouse models highlight the lack of sufficient information on the side effects of ncRNA drug models in entire body systems.
Conclusion
On the basis of numerous studies, the role of ncRNAs in cancer is undeniable. Their function is similar to that of a double-edged sword since many of them are considered tumor suppressors and many are considered oncogenes. LncRNAs can participate in various pathways leading to cancer promotion, such as angiogenesis, EMT, metastasis and unregulated proliferation, through mediators such as miRNAs, which themselves can act as inducers or suppressors of cancer [104, 359]. With recent advances in technology, studying the role of ncRNAs is more feasible than ever, although challenges remain with each approach, highlighting a place for future research and investigation. Furthermore, bridging the gap between preclinical studies and clinical trials is essential to validate the therapeutic potential of ncRNAs. The development of robust animal models and the execution of large-scale clinical trials will play pivotal roles in translating ncRNA-based therapies into clinical practice. By gaining a deeper understanding of the functions and mechanisms of ncRNAs, we may be able to effectively control cancer progression, reverse adverse outcomes, and design more targeted and effective therapeutic strategies. However, significant gaps and unanswered questions remain, necessitating further comprehensive research to fully unlock their potential.
Availability of data and materials
No datasets were generated or analysed during the current study.
Abbreviations
- BC:
-
Breast cancer
- ER:
-
Estrogen receptor
- PR:
-
Progesterone receptor
- HER2:
-
Human epithelial growth factor receptor 2
- NcRNA:
-
Non-coding RNA
- lncRNA:
-
Long non-coding RNA
References
Breast Cancer 2023. https://www.who.int/news-room/fact-sheets/detail/breast-cancer. Accessed 13 Dec 2023.
Arnold M, et al. Current and future burden of breast cancer: Global statistics for 2020 and 2040. Breast. 2022;66:15–23.
Watkins EJ. Overview of breast cancer. JAAPA. 2019;32(10):13–7.
Yersal O, Barutca S. Biological subtypes of breast cancer: Prognostic and therapeutic implications. World J Clin Oncol. 2014;5(3):412–24.
Carey LA, et al. Race, breast cancer subtypes, and survival in the Carolina breast cancer study. JAMA. 2006;295(21):2492–502.
Orrantia-Borunda E, et al. Subtypes of breast cancer. Breast Cancer. Brisbane: Exon Publications; 2022.
Tran B, Bedard PL. Luminal-B breast cancer and novel therapeutic targets. Breast Cancer Res. 2011;13(6):221.
Ellis MJ, et al. Outcome prediction for estrogen receptor–positive breast cancer based on postneoadjuvant endocrine therapy tumor characteristics. J Natl Cancer Inst. 2008;100(19):1380–8.
Marei HE, et al. p53 signaling in cancer progression and therapy. Cancer Cell Int. 2021;21(1):703.
Bergin ART, Loi S. Triple-negative breast cancer: recent treatment advances. F1000Res. 2019. https://doiorg.publicaciones.saludcastillayleon.es/10.12688/f1000research.18888.1.
Foulkes WD, Smith IE, Reis-Filho JS. Triple-negative breast cancer. N Engl J Med. 2010;363(20):1938–48.
Jitariu AA, et al. Triple negative breast cancer: the kiss of death. Oncotarget. 2017;8(28):46652–62.
Iacopetta D, et al. Targeting breast cancer: an overlook on current strategies. Int J Mol Sci. 2023;24(4):3643.
Zhang Y, et al. Albumin stabilized Pt nanoparticles as radiosensitizer for sensitization of breast cancer cells under X-ray radiation therapy. Inorg Chem Commun. 2022;140: 109423.
Costa B, et al. Understanding Breast cancer: from conventional therapies to repurposed drugs. Eur J Pharm Sci Off J Eur Fed Pharm Sci. 2020;151:105401.
Salata C, et al. Preliminary pre-clinical studies on the side effects of breast cancer treatment. Int J Radiat Biol. 2021;97:877–87.
Palesh OG, et al. Management of side effects during and post-treatment in breast cancer survivors. Breast J. 2018;24:167–75.
Li J, Liu C. Coding or non-coding, the converging concepts of RNAs. Front Genet. 2019;10:496.
Cantile M, et al. Long Non-Coding RNA HOTAIR in Breast Cancer therapy. Cancers. 2020;12:1197.
Ling H, et al. Non-coding RNAs: the cancer genome dark matter that matters! Clin Chem Lab Med (CCLM). 2017;55(5):705–14.
Zhang P, et al. Non-coding RNAs and their integrated networks. J Integr Bioinform. 2019;16(3):20190027.
Salmena L, et al. A ceRNA hypothesis: the Rosetta stone of a hidden RNA language? Cell. 2011;146(3):353–8.
Calin GA, Croce CM. MicroRNA signatures in human cancers. Nat Rev Cancer. 2006;6(11):857–66.
Peng Y, Croce CM. The role of MicroRNAs in human cancer. Signal Transduct Target Ther. 2016;1:15004.
Abasi M, et al. Differential maturation of miR-17–92 cluster members in human cancer cell lines. Appl Biochem Biotechnol. 2017;182:1540–7.
Abasi M, Ranjbari J, Ghanbarian H. 7SK small nuclear RNA (Rn7SK) induces apoptosis through intrinsic and extrinsic pathways in human embryonic kidney cell line. Mol Biol Rep. 2024;51(1):96.
Bazi Z, et al. Rn7SK small nuclear RNA is involved in neuronal differentiation. J Cell Biochem. 2018;119(4):3174–82.
Musavi M, et al. Rn7SK small nuclear RNA is involved in cellular senescence. J Cell Physiol. 2019;234(8):14234–45.
Fabbri M, et al. A new role for microRNAs, as ligands of Toll-like receptors. RNA Biol. 2013;10(2):169–74.
Klinge CM. Non-coding RNAs: long non-coding RNAs and microRNAs in endocrine-related cancers. Endocr Relat Cancer. 2018;25(4):R259–82.
Baradaran B, Shahbazi R, Khordadmehr M. Dysregulation of key microRNAs in pancreatic cancer development. Biomed Pharmacother. 2019;109:1008–15.
Bautista-Sánchez D, et al. The promising role of miR-21 as a cancer biomarker and its importance in RNA-based therapeutics. Mol Ther Nucleic Acids. 2020;20:409–20.
Liu J, Zhang S, Cheng B. Epigenetic roles of PIWI-interacting RNAs (piRNAs) in cancer metastasis. Oncol Rep. 2018;40(5):2423–34.
Shen S, et al. PIWIL1/piRNA-DQ593109 regulates the permeability of the blood-tumor barrier via the MEG3/miR-330-5p/RUNX3 axis. Mol Ther Nucleic Acids. 2018;10:412–25.
Huang G-L, et al. Altered expression of piRNAs and their relation with clinicopathologic features of breast cancer. Clin Transl Oncol. 2013;15:563–8.
Cheng J, et al. piR-823, a novel non-coding small RNA, demonstrates in vitro and in vivo tumor suppressive activity in human gastric cancer cells. Cancer Lett. 2012;315(1):12–7.
Su J-F, et al. piR-823 demonstrates tumor oncogenic activity in esophageal squamous cell carcinoma through DNA methylation induction via DNA methyltransferase 3B. Pathol Res Pract. 2020;216(4): 152848.
Yin J, et al. piR-823 contributes to colorectal tumorigenesis by enhancing the transcriptional activity of HSF 1. Cancer Sci. 2017;108(9):1746–56.
Abasi M, et al. 7SK small nuclear RNA transcription level down-regulates in human tumors and stem cells. Med Oncol. 2016;33:1–5.
Hajjari M, Salavaty A. HOTAIR: an oncogenic long non-coding RNA in different cancers. Cancer Biol Med. 2015;12(1):1.
Rinn JL, et al. Functional demarcation of active and silent chromatin domains in human HOX loci by non-coding RNAs. Cell. 2007;129(7):1311–23.
Raveh E, et al. The H19 Long non-coding RNA in cancer initiation, progression and metastasis–a proposed unifying theory. Mol Cancer. 2015;14:1–14.
Madhi H, Kim MH. Beyond X-chromosome inactivation: the oncogenic facet of XIST in human cancers. Biomed Sci Lett. 2019;25(2):113–22.
Liu B, et al. Long non-coding RNA XIST acts as a ceRNA of miR-362-5p to suppress breast cancer progression. Cancer Biother Radiopharm. 2021;36(6):456–66.
Li X, et al. LncRNA XIST interacts with miR-454 to inhibit cells proliferation, epithelial mesenchymal transition and induces apoptosis in triple-negative breast cancer. J Biosci. 2020;45(1):45.
Palade GE. A small particulate component of the cytoplasm. J Biophys Biochem Cytol. 1955;1(1):59.
Crick FH. On protein synthesis. Symp Soc Exp Biol. 1958;12(138–63):8.
Cobb M. 60 years ago, Francis Crick changed the logic of biology. PLoS Biol. 2017;15(9): e2003243.
Hoagland MB, et al. A soluble ribonucleic acid intermediate in protein synthesis. J biol Chem. 1958;231(1):241–57.
Hodnett J, Busch H. Isolation and characterization of uridylic acid-rich 7 S ribonucleic acid of rat liver nuclei. J Biol Chem. 1968;243(24):6334–42.
Weinberg RA, Penman S. Small molecular weight monodisperse nuclear RNA. J Mol Biol. 1968;38(3):289–304.
Patop IL, Wüst S, Kadener S. Past, present, and future of circRNAs. Embo j. 2019;38(16): e100836.
Brannan CI, et al. The product of the H19 gene may function as an RNA. Mol Cell Biol. 1990. https://doiorg.publicaciones.saludcastillayleon.es/10.1128/mcb.10.1.28-36.1990.
Lee RC, Feinbaum RL, Ambros V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75(5):843–54.
Wightman B, Ha I, Ruvkun G. Posttranscriptional regulation of the heterochronic gene lin-14 by lin-4 mediates temporal pattern formation in C. elegans. Cell. 1993;75(5):855–62.
Lagos-Quintana M, et al. Identification of novel genes coding for small expressed RNAs. Science. 2001;294(5543):853–8.
Fire A, et al. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391(6669):806–11.
Hamilton AJ, Baulcombe DC. A species of small antisense RNA in posttranscriptional gene silencing in plants. Science. 1999;286(5441):950–2.
Calin GA, et al. Frequent deletions and down-regulation of micro-RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia. Proc Natl Acad Sci. 2002;99(24):15524–9.
Lee Y, et al. The nuclear RNase III Drosha initiates microRNA processing. Nature. 2003;425(6956):415–9.
Margulies M, et al. Genome sequencing in microfabricated high-density picolitre reactors. Nature. 2005;437(7057):376–80.
Cheung F, et al. Sequencing Medicago truncatula expressed sequenced tags using 454 Life Sciences technology. BMC Genomics. 2006;7:1–10.
Zhang X, et al. Genome-wide high-resolution mapping and functional analysis of DNA methylation in Arabidopsis. Cell. 2006;126(6):1189–201.
Skipper M. Users’ guide to the human genome. Nat Rev Genet. 2012;13(10):678–678.
Djebali S, et al. Landscape of transcription in human cells. Nature. 2012;489(7414):101–8.
Austin TBM. First microRNA mimic enters clinic. Nat Biotechnol. 2013;31(7):577.
Hongkuan Z, et al. The functional roles of the non-coding RNAs in molluscs. Gene. 2021;768: 145300.
Szymański M, Erdmann VA, Barciszewski J. Non-coding regulatory RNAs database. Nucleic Acids Res. 2003;31(1):429–31.
Bhatti GK, et al. Emerging role of non-coding RNA in health and disease. Metab Brain Dis. 2021;36(6):1119–34.
Palazzo AF, Lee ES. Non-coding RNA: what is functional and what is junk? Front Genet. 2015;6:2.
Morris KV, Mattick JS. The rise of regulatory RNA. Nat Rev Genet. 2014;15(6):423–37.
Kung JT, Colognori D, Lee JT. Long non-coding RNAs: past, present, and future. Genetics. 2013;193(3):651–69.
Frankish A, et al. GENCODE reference annotation for the human and mouse genomes. Nucleic Acids Res. 2019;47(D1):D766–73.
Bester AC. Deciphering the Hidden Language of Long Non-Coding RNAs: Recent Findings and Challenges. 2023.
Gil N, Ulitsky I. Production of spliced long non-coding RNAs specifies regions with increased enhancer activity. Cell Syst. 2018;7(5):537-547. e3.
Ali T, Grote P. Beyond the RNA-dependent function of LncRNA genes. Elife. 2020;9: e60583.
Ghafouri-Fard S, et al. HOX transcript antisense RNA: an oncogenic lncRNA in diverse malignancies. Exp Mol Pathol. 2021;118: 104578.
Luo Y, et al. MicroRNA-592 suppresses the malignant phenotypes of thyroid cancer by regulating lncRNA NEAT1 and downregulating NOVA1. Int J Mol Med. 2019;44(3):1172–82.
Chen M-Y, et al. Long non-coding RNA nuclear enriched abundant transcript 1 (NEAT1) sponges microRNA-124-3p to up-regulate phosphodiesterase 4B (PDE4B) to accelerate the progression of Parkinson’s disease. Bioengineered. 2021;12(1):708–19.
An Q, et al. Knockdown of long non-coding RNA NEAT1 relieves the inflammatory response of spinal cord injury through targeting miR-211-5p/MAPK1 axis. Bioengineered. 2021;12(1):2702–12.
Chen R, et al. Quantitative proteomics reveals that long non-coding RNA MALAT1 interacts with DBC1 to regulate p53 acetylation. Nucleic Acids Res. 2017;45(17):9947–59.
Vasudevan S. Posttranscriptional upregulation by microRNAs. Wiley Interdiscip Rev RNA. 2012;3(3):311–30.
Makarova JA, et al. Intracellular and extracellular microRNA: an update on localization and biological role. Prog Histochem Cytochem. 2016;51(3–4):33–49.
Peng Y, Croce CM. The role of MicroRNAs in human cancer. Signal Transduct Target Ther. 2016;1(1):1–9.
Paul P, et al. Interplay between miRNAs and human diseases. J Cell Physiol. 2018;233(3):2007–18.
Hayes J, Peruzzi PP, Lawler S. MicroRNAs in cancer: biomarkers, functions and therapy. Trends Mol Med. 2014;20(8):460–9.
Wang J, Chen J, Sen S. MicroRNA as biomarkers and diagnostics. J Cell Physiol. 2016;231(1):25–30.
Huang W. MicroRNAs biomarkers, diagnostics, and therapeutics. Bioinformatics in MicroRNA research. New York: Springer New York; 2017. p. 57–67.
Sciences N.I.o.E.H. Biomarkers. 2025. https://www.niehs.nih.gov/health/topics/science/biomarkers. Accessed 22 Jan 2025.
Valadkhan S, Gunawardane LS. Role of small nuclear RNAs in eukaryotic gene expression. Essays Biochem. 2013;54:79–90.
Hari R, Parthasarathy S, et al. Prediction of coding and Non-Coding RNA. In: Ranganathan S, et al., editors. Encyclopedia of bioinformatics and computational biology. Oxford: Academic Press; 2019. p. 230–40.
Kanakis I, et al. Small-RNA sequencing reveals altered skeletal muscle microRNAs and snoRNAs signatures in weanling male offspring from mouse dams fed a low protein diet during lactation. Cells. 2021;10(5):1166.
Huang Z-H, et al. snoRNAs: functions and mechanisms in biological processes, and roles in tumor pathophysiology. Cell Death Discov. 2022;8(1):259.
Yamashiro H, Siomi MC. PIWI-interacting RNA in drosophila: biogenesis, transposon regulation, and beyond. Chem Rev. 2017;118(8):4404–21.
Yao J, et al. PIWI-interacting RNAs in cancer: biogenesis, function, and clinical significance. Front Oncol. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fonc.2022.965684.
Moin M, et al. Small interfering RNA-mediated regulation of gene expression and its role as a plant reverse genetic tool. Indian J Plant Physiol. 2017;22:549–57.
Piatek MJ, Werner A. Endogenous siRNAs: regulators of internal affairs. Biochem Soc Trans. 2014;42(4):1174–9.
Wang X, et al. Identification of competitive endogenous RNAs network in breast cancer. Cancer Med. 2019;8(5):2392–403.
Jin X, et al. LncRNA TROJAN promotes proliferation and resistance to CDK4/6 inhibitor via CDK2 transcriptional activation in ER+ breast cancer. Mol Cancer. 2020;19:1–18.
Si H, et al. Long non-coding RNA H19 regulates cell growth and metastasis via miR-138 in breast cancer. Am J Trans Res. 2019;11(5):3213.
Zhao J, et al. Long non-coding RNA HOTAIR promotes breast cancer development through the lncRNA HOTAIR/miR-1/GOLPH3 axis. Clin Trans Oncol. 2023;25(12):3420–30.
Manfioletti G, Fedele M. Epithelial-mesenchymal transition (EMT). Int J Mol Sci. 2023;24(14):11386.
Akhmetkaliyev A, et al. EMT/MET plasticity in cancer and Go-or-Grow decisions in quiescence: the two sides of the same coin? Mol Cancer. 2023;22(1):90.
He W, Li D, Zhang X. LncRNA HOTAIR promotes the proliferation and invasion/metastasis of breast cancer cells by targeting the miR-130a-3p/Suv39H1 axis. Biochem Biophys Reports. 2022;30: 101279.
Raju GSR, et al. HOTAIR: a potential metastatic, drug-resistant and prognostic regulator of breast cancer. Mol Cancer. 2023;22(1):65.
Tang Y, et al. Silencing of long non-coding RNA HOTAIR alleviates epithelial–mesenchymal transition in pancreatic cancer via the Wnt/β-catenin signaling pathway. Cancer Manag Res. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.2147/CMAR.S265578.
Wang Y, et al. Long non-coding RNA HOTAIR promotes breast cancer development by targeting ZEB1 via sponging miR-601. Cancer Cell Int. 2020;20:320.
Pawłowska E, Szczepanska J, Błasiak J. The long non-coding RNA HOTAIR in breast cancer: does autophagy play a role? Int J Mol Sci. 2017. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/ijms18112317.
Wang Y-L, et al. Long non-coding RNA HOTAIR in circulatory exosomes is correlated with ErbB2/HER2 positivity in breast cancer. Breast. 2019;46:64–9.
Zhang S, et al. LncRNA HOTAIR enhances breast cancer radioresistance through facilitating HSPA1A expression via sequestering miR-449b-5p. Thoracic Cancer. 2020;11(7):1801–16.
Shi Y, et al. Current knowledge of long non-coding RNA HOTAIR in breast cancer progression and its application. Life. 2021;11:483.
Chen T, et al. Down-regulation of long non-coding RNA HOTAIR sensitizes breast cancer to trastuzumab. Sci Rep. 2019;9(1):19881.
Abdel-hamid NR, et al. Circulating ESR1, long non-coding RNA HOTAIR and microRNA-130a gene expression as biomarkers for breast cancer stage and metastasis. Sci Rep. 2023;13(1):22654.
Abba MC, et al. HOTAIR modulated pathways in early-stage breast cancer progression. Front Oncol. 2021;11: 783211.
Yu J, Jin T, Zhang T. Suppression of long non-coding RNA metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) potentiates cell apoptosis and drug sensitivity to Taxanes and Adriamycin in breast cancer. Med Sci Monit. 2020;26: e922672.
Tufail M. The MALAT1-breast cancer interplay: insights and implications. Expert Rev Mol Diagn. 2023;23(8):665–78.
Xu S, et al. Downregulation of long non-coding RNA MALAT1 induces epithelial-to-mesenchymal transition via the PI3K-AKT pathway in breast cancer. Int J Clin Exp Pathol. 2015;8(5):4881–91.
Yue X, et al. LncRNA MALAT1 promotes breast cancer progression and doxorubicin resistance via regulating miR-570-3p. Biomed J. 2021;44(6 Suppl 2):S296-s304.
Li X, et al. Genome-wide target interactome profiling reveals a novel EEF1A1 epigenetic pathway for oncogenic lncRNA MALAT1 in breast cancer. Am J Cancer Res. 2019;9(4):714–29.
Shao J, et al. LncRNA MALAT1 promotes breast cancer progression by sponging miR101-3p to mediate mTOR/PKM2 signal transmission. Am J Transl Res. 2021;13(9):10262–75.
Wang N, et al. lncRNA MALAT1/miR-26a/26b/ST8SIA4 axis mediates cell invasion and migration in breast cancer cell lines. Oncol Rep. 2021;46(2):1–12.
Huang X, et al. MALAT1 promotes angiogenesis of breast cancer. Oncol Rep. 2018;40(5):2683–9.
Peng R, et al. Association analyses of genetic variants in long non-coding RNA MALAT1 with breast cancer susceptibility and mRNA expression of MALAT1 in Chinese Han population. Gene. 2018;642:241–8.
Zhang X, et al. Roles of MALAT1 in development and migration of triple negative and Her-2 positive breast cancer. Oncotarget. 2017;9:2255–67.
Park MK, et al. NEAT1 is essential for metabolic changes that promote breast cancer growth and metastasis. Cell Metab. 2021;33(12):2380–23979.
Zhu L, et al. lncRNA NEAT1 promotes the Taxol resistance of breast cancer via sponging the miR-23a-3p-FOXA1 axis. Acta Biochim Biophys Sin. 2021;53(9):1198–206.
Yan L, et al. lncRNA NEAT1 facilitates cell proliferation, invasion and migration by regulating CBX7 and RTCB in breast cancer. Onco Targets Ther. 2020;13:2449.
Li S, et al. Long non-coding RNA NEAT1 promotes the proliferation, migration, and metastasis of human breast-cancer cells by inhibiting miR-146b-5p expression. Cancer Manag Res. 2020;12:6091–101.
Liu X, et al. LncRNA NEAT1 accelerates breast cancer progression through regulating miR-410-3p/CCND1 axis. Cancer Biomark. 2020;29(2):277–90.
Jiang X, et al. NEAT1 contributes to breast cancer progression through modulating miR-448 and ZEB1. J Cell Physiol. 2018;233(11):8558–66.
An S, et al. LncRNA-NEAT1/miR-148a-3p axis regulates cell viability, apoptosis and autophagy through wnt/β-catenin signaling pathway in Breast Cancer. Trop J Pharm Res. 2021;20(5):899–910.
Ren J, et al. Tumor protein D52 promotes breast cancer proliferation and migration via the long non-coding RNA NEAT1/microRNA-218-5p axis. Ann Transl Med. 2021;9(12):1008.
Zhou D, et al. Long non-coding RNA NEAT1 transported by extracellular vesicles contributes to breast cancer development by sponging microRNA-141-3p and regulating KLF12. Cell Biosci. 2021;11:1–16.
Wei X, et al. Exosomal lncRNA NEAT1 induces paclitaxel resistance in breast cancer cells and promotes cell migration by targeting miR-133b. Gene. 2023;860: 147230.
Shin VY, et al. Long non-coding RNA NEAT1 confers oncogenic role in triple-negative breast cancer through modulating chemoresistance and cancer stemness. Cell Death Dis. 2019;10(4):270.
Yang C, et al. Long non-coding RNA NEAT1 overexpression is associated with poor prognosis in cancer patients: a systematic review and meta-analysis. Oncotarget. 2017;8(2):2672–80.
Azadeh M, et al. NEAT1 can be a diagnostic biomarker in the breast cancer and gastric cancer patients by targeting XIST, hsa-miR-612, and MTRNR2L8: integrated RNA targetome interaction and experimental expression analysis. Genes Environ. 2022;44(1):16.
Knutsen E, Harris AL, Perander M. Expression and functions of long non-coding RNA NEAT1 and isoforms in breast cancer. Br J Cancer. 2022;126(4):551–61.
O’Brien J, et al. Overview of MicroRNA biogenesis, mechanisms of actions, and circulation. Front Endocrinol. 2018. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fendo.2018.00402.
Lau NC, et al. An abundant class of tiny RNAs with probable regulatory roles in Caenorhabditis elegans. Science. 2001;294(5543):858–62.
Peng Y, Croce CM. The role of MicroRNAs in human cancer. Signal Transduct Target Ther. 2016;1(1):15004.
Macfarlane LA, Murphy PR. MicroRNA: biogenesis, function and role in cancer. Curr Genomics. 2010;11(7):537–61.
Ardekani AM, Naeini MM. The role of MicroRNAs in human diseases. Avicenna J Med Biotechnol. 2010;2(4):161–79.
Otmani K, Lewalle P. Tumor suppressor miRNA in cancer cells and the tumor microenvironment: mechanism of deregulation and clinical implications. Front Oncol. 2021;11: 708765.
Ito Y, et al. Identification of targets of tumor suppressor microRNA-34a using a reporter library system. Proc Natl Acad Sci. 2017;114(15):3927–32.
Wen D, et al. Micellar delivery of miR-34a modulator rubone and paclitaxel in resistant prostate cancer. Can Res. 2017;77(12):3244–54.
Shaban NZ, et al. miR-34a and miR-21 as biomarkers in evaluating the response of chemo-radiotherapy in Egyptian breast cancer patients. J Radiat Res Appl Sci. 2022;15(3):285–92.
Misso G, et al. Mir-34: a new weapon against cancer? Mol Ther Nucleic Acids. 2014;3: e195.
Kang L, et al. Retracted: Micro RNA-34a suppresses the breast cancer stem cell-like characteristics by downregulating Notch1 pathway. Cancer Sci. 2015;106(6):700–8.
Han R, Zhao J, Lu L. MicroRNA-34a expression affects breast cancer invasion in vitro and patient survival via downregulation of E2F1 and E2F3 expression. Oncol Rep. 2020;43(6):2062–72.
Engkvist M, et al. Analysis of the miR-34 family functions in breast cancer reveals annotation error of miR-34b. Sci Rep. 2017;7(1):9655.
Chang T-C, et al. Transactivation of miR-34a by p53 broadly influences gene expression and promotes apoptosis. Mol Cell. 2007;26(5):745–52.
Weng Y-S, et al. MCT-1/miR-34a/IL-6/IL-6R signaling axis promotes EMT progression, cancer stemness and M2 macrophage polarization in triple-negative breast cancer. Mol Cancer. 2019;18(1):1–15.
Yin M, Zhang Z, Wang Y. Anti-tumor effects of miR-34a by regulating immune cells in the tumor microenvironment. Cancer Med. 2023;12(10):11602–10.
Gu Y, Zhang Z, Ten Dijke P. Harnessing epithelial-mesenchymal plasticity to boost cancer immunotherapy. Cell Mol Immunol. 2023;20(4):318–40.
Kapadia CH, Ioele SA, Day ES. Layer-by-layer assembled PLGA nanoparticles carrying miR-34a cargo inhibit the proliferation and cell cycle progression of triple-negative breast cancer cells. J Biomed Mater Res Part A. 2020;108(3):601–13.
Jaiswal PK, Goel A, Mittal RD. Survivin: a molecular biomarker in cancer. Indian J Med Res. 2015;141(4):389–97.
Martini S, et al. miR-34a-mediated survivin inhibition improves the antitumor activity of Selinexor in triple-negative breast cancer. Pharmaceuticals. 2021;14(6):523.
Song C, et al. miR-200c inhibits breast cancer proliferation by targeting KRAS. Oncotarget. 2015;6:34968–78.
Mansoori B, et al. miR-34a and miR-200c have an additive tumor-suppressive effect on breast cancer cells and patient prognosis. Genes. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/genes12020267.
Zhao Y-X, et al. microRNA-372 inhibits proliferation and induces apoptosis in human breast cancer cells by directly targeting E2F1. Mol Med Rep. 2017;16:8069–75.
Klicka K, et al. The role of miR-200 family in the regulation of hallmarks of cancer. Front Oncol. 2022;12: 965231.
Feng X, et al. MiR-200, a new star miRNA in human cancer. Cancer Lett. 2014;344(2):166–73.
Piperigkou Z, et al. miR-200b restrains EMT and aggressiveness and regulates matrix composition depending on ER status and signaling in mammary cancer. Matrix Biol Plus. 2020. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.mbplus.2020.100024.
Park S-M, et al. The miR-200 family determines the epithelial phenotype of cancer cells by targeting the E-cadherin repressors ZEB1 and ZEB2. Genes Dev. 2008;22(7):894–907.
Song C, et al. miR-200c suppresses stemness and increases cellular sensitivity to trastuzumab in HER2+ breast cancer. bioRxiv. 2017;33:2040.
Samaeekia R, et al. miR-206 inhibits stemness and metastasis of breast cancer by targeting MKL1/IL11 pathway. Clin Cancer Res. 2016;23:1091–103.
Hong H, et al. MicroRNA-200b impacts breast cancer cell migration and invasion by regulating Ezrin-Radixin-Moesin. Med Sci Monitor Int Med J Exp Clin Res. 2016;22:1946–52.
Yuan J, et al. miR-200b regulates breast cancer cell proliferation and invasion by targeting radixin. Exp Ther Med. 2020;19(4):2741–50.
Kozak J, Jonak K, Maciejewski R. The function of miR-200 family in oxidative stress response evoked in cancer chemotherapy and radiotherapy. Biomed Pharmacother Biomed Pharmacother. 2020;125:110037.
Klicka K, et al. The role of miR-200 family in the regulation of hallmarks of cancer. Front Oncol. 2022;12:965231.
Zhang J, et al. Interleukin-8 promotes epithelial-to-mesenchymal transition via downregulation of Mir-200 family in breast cancer cells. Technol Cancer Res Treat. 2020;19:1533033820979672.
Cesi V, et al. TGFβ-induced c-Myb affects the expression of EMT-associated genes and promotes invasion of ER+ breast cancer cells. Cell Cycle. 2011;10(23):4149–61.
Safaei S, et al. miR-200c increases the sensitivity of breast cancer cells to doxorubicin through downregulating MDR1 gene. Exp Mol Pathol. 2022;125: 104753.
Yang X, et al. miR-200b regulates epithelial-mesenchymal transition of chemo-resistant breast cancer cells by targeting FN1. Discov Med. 2017;24(131):75–85.
Aqeilan RI, Calin GA, Croce CM. miR-15a and miR-16-1 in cancer: discovery, function and future perspectives. Cell Death Differ. 2010;17(2):215–20.
Cimmino A, et al. miR-15 and miR-16 induce apoptosis by targeting BCL2. Proc Natl Acad Sci U S A. 2005;102(39):13944–9.
Ramaiah MJ. Functions and epigenetic aspects of miR-15/16: possible future cancer therapeutics. Gene Reports. 2018;12:149–64.
Sun C, et al. miR-15a and miR-16 affect the angiogenesis of multiple myeloma by targeting VEGF. Carcinogenesis. 2013;34(2):426–35.
Yin K-J, et al. Abstract Tp117: Mir-15a/16–1 cluster inhibits angiogenesis in mouse after ischemic stroke. Stroke. 2013. https://doiorg.publicaciones.saludcastillayleon.es/10.1161/str.44.suppl_1.ATP117.
Mei Z, et al. The miR-15 family enhances the radiosensitivity of breast cancer cells by targeting G2 checkpoints. Radiat Res. 2015;183(2):196–207.
Patel N, et al. miR-15a/miR-16 down-regulates BMI1, impacting Ub-H2A mediated DNA repair and breast cancer cell sensitivity to doxorubicin. Sci Rep. 2017;7(1):4263.
Gao X, et al. MicroRNA-16 sensitizes drug-resistant breast cancer cells to Adriamycin by targeting Wip1 and Bcl-2. Oncol Lett. 2019;18(3):2897–906.
Frixa T, Donzelli S, Blandino G. Oncogenic MicroRNAs: key players in malignant transformation. Cancers (Basel). 2015;7(4):2466–85.
Hannafon BN, et al. miR-23b and miR-27b are oncogenic microRNAs in breast cancer: evidence from a CRISPR/Cas9 deletion study. BMC Cancer. 2019;19(1):642.
M Maryam, M Naemi, SS Hasani. A comprehensive review on oncogenic miRNAs in breast cancer. J Genet. 2021;100.
Wang W, Luo YP. MicroRNAs in breast cancer: oncogene and tumor suppressors with clinical potential. J Zhejiang Univ Sci B. 2015;16(1):18–31.
Zhang B, et al. microRNAs as oncogenes and tumor suppressors. Dev Biol. 2007;302(1):1–12.
Wang H, et al. microRNA-21 promotes breast cancer proliferation and metastasis by targeting LZTFL1. BMC Cancer. 2019;19(1):738.
Zhang C, et al. miR-21: a gene of dual regulation in breast cancer. Int J Oncol. 2016;48(1):161–72.
Chi LH, et al. MicroRNA-21 is immunosuppressive and pro-metastatic via separate mechanisms. Oncogenesis. 2022;11(1):38.
Han M, et al. Re-expression of miR-21 contributes to migration and invasion by inducing epithelial-mesenchymal transition consistent with cancer stem cell characteristics in MCF-7 cells. Mol Cell Biochem. 2012;363:427–36.
Fujita S, et al. miR-21 Gene expression triggered by AP-1 is sustained through a double-negative feedback mechanism. J Mol Biol. 2008;378(3):492–504.
Yan LX, et al. MicroRNA miR-21 overexpression in human breast cancer is associated with advanced clinical stage, lymph node metastasis and patient poor prognosis. RNA. 2008;14(11):2348–60.
Rhim J, et al. From molecular mechanisms to therapeutics: understanding MicroRNA-21 in cancer. Cells. 2022;11(18):2791.
Arisan ED, et al. MiR-21 is required for the epithelial-mesenchymal transition in MDA-MB-231 breast cancer cells. Int J Mol Sci. 2021;22(4):1557.
Piras F, et al. Nestin expression associates with poor prognosis and triple negative phenotype in locally advanced (T4) breast cancer. Eur J Histochem EJH. 2011. https://doiorg.publicaciones.saludcastillayleon.es/10.4081/ejh.2011.e39.
Clairmont CS, et al. TRIP13 regulates DNA repair pathway choice through REV7 conformational change. Nat Cell Biol. 2020;22(1):87–96.
Gomez-Contreras P, et al. Extracellular matrix 1 (ECM1) regulates the actin cytoskeletal architecture of aggressive breast cancer cells in part via S100A4 and Rho-family GTPases. Clin Exp Metas. 2017;34:37–49.
Fang H, et al. miRNA-21 promotes proliferation and invasion of triple-negative breast cancer cells through targeting PTEN. Am J Trans Res. 2017;9(3):953–61.
Badr FM. Potential role of miR-21 in breast cancer diagnosis and therapy. SciMed Central. 2016;3(5):1068–75.
Gong C, et al. Up-regulation of miR-21 mediates resistance to trastuzumab therapy for breast cancer. J Biol Chem. 2011;286(21):19127–37.
Najjary S, et al. Role of miR-21 as an authentic oncogene in mediating drug resistance in breast cancer. Gene. 2020. https://doiorg.publicaciones.saludcastillayleon.es/10.1016/j.gene.2020.144453.
Zong Y, et al. miR-221/222 promote tumor growth and suppress apoptosis by targeting lncRNA GAS5 in breast cancer. Biosci Reports. 2018;39:BSR20181859.
Li Y, et al. miR-221/222 promotes S-phase entry and cellular migration in control of basal-like breast cancer. Molecules. 2014;19(6):7122–37.
Liang Y-K, et al. MiR-221/222 promote epithelial-mesenchymal transition by targeting Notch3 in breast cancer cell lines. npj Breast Cancer. 2018;4(1):20.
Kim JY, et al. MiR-221 and miR-222 regulate cell cycle progression and affect chemosensitivity in breast cancer by targeting ANXA3. Exp Ther Med. 2023;25(3):127.
Galardi S, et al. miR-221 and miR-222 expression affects the proliferation potential of human prostate carcinoma cell lines by targeting p27Kip1*. J Biol Chem. 2007;282(32):23716–24.
Dentelli P, et al. miR-221/222 control luminal breast cancer tumor progression by regulating different targets. Cell Cycle. 2014;13(11):1811.
Ouyang YX, et al. miR-221/222 sponge abrogates tamoxifen resistance in ER-positive breast cancer cells through restoring the expression of ERα. Mol Biomed. 2021;2:1–11.
Gan R, et al. Downregulation of miR-221/222 enhances sensitivity of breast cancer cells to tamoxifen through upregulation of TIMP3. Cancer Gene Ther. 2014;21(7):290–6.
Li B, et al. miR-221/222 promote cancer stem-like cell properties and tumor growth of breast cancer via targeting PTEN and sustained Akt/NF-κB/COX-2 activation. Chem Biol Interact. 2017;277:33–42.
Li S, et al. Targeted inhibition of miR-221/222 promotes cell sensitivity to cisplatin in triple-negative breast cancer MDA-MB-231 cells. Front Genet. 2020;10:1278.
Chen W-X, et al. miR-221/222: promising biomarkers for breast cancer. Tumor Biol. 2013;34(3):1361–70.
Di Martino MT, et al. miR-221/222 as biomarkers and targets for therapeutic intervention on cancer and other diseases: a systematic review. Mol Ther Nucleic Acids. 2022;27:1191–224.
Song J, et al. Potential value of miR-221/222 as diagnostic, prognostic, and therapeutic biomarkers for diseases. Front Immunol. 2017;8:56.
Das A, Roy S. Chapter 21—micromanaging inflammation and tissue repair. In: Sen CK, editor. MicroRNA in regenerative medicine (Second Edition). Cambridge: Academic Press; 2023. p. 573–91.
Van Roosbroeck K, et al. Combining anti-miR-155 with chemotherapy for the treatment of lung cancers. Clin Cancer Res. 2017;23(11):2891–904.
Park S, et al. MiR-9, miR-21, and miR-155 as potential biomarkers for HPV positive and negative cervical cancer. BMC Cancer. 2017;17:1–8.
Li N, et al. MiR-155-5p accelerates the metastasis of cervical cancer cell via targeting TP53INP1. Onco Targets Ther. 2019;12:3181.
Chen J, Wang B-C, Tang J-H. Clinical significance of MicoRNA-155 expression in human breast cancer. J Surg Oncol. 2012;106(3):260–6.
Mattiske S, et al. The oncogenic role of miR-155 in breast cancer. Cancer Epidemiol Biomark Prev. 2012;21(8):1236–43.
Moutabian H, et al. MicroRNA-155 and cancer metastasis: regulation of invasion, migration, and epithelial-to-mesenchymal transition. Pathol Res Pract. 2023;250: 154789.
Anwar SL, et al. Dynamic changes of circulating Mir-155 expression and the potential application as a non-invasive biomarker in breast cancer. Asian Pac J Cancer Prev. 2020;21(2):491–7.
Zhang W, Chen C-J, Guo G-L. MiR-155 promotes the proliferation and migration of breast cancer cells via targeting SOCS1 and MMP16. Eur Rev Med Pharmacol Sci. 2018;22(21):7323–32.
Blevins HM, et al. The NLRP3 inflammasome pathway: a review of mechanisms and inhibitors for the treatment of inflammatory diseases. Front Aging Neurosci. 2022;14: 879021.
Hu Y, et al. Research on the effect of interfering with miRNA-155 on triple-negative breast cancer cells. Genes Genomics. 2022;44(9):1117–24.
Yang L-W, et al. miR-155 increases stemness and decitabine resistance in triple-negative breast cancer cells by inhibiting TSPAN5. Mol Carcinog. 2020;59(4):447–61.
Zuo J, et al. Inhibition of miR-155, a therapeutic target for breast cancer, prevented in cancer stem cell formation. Cancer Biomark. 2018;21:383–92.
Yan H, Guo M. Schizandrin A inhibits cellular phenotypes of breast cancer cells by repressing miR-155. IUBMB Life. 2020;72(8):1640–8.
Zhang F-C, et al. Long non-coding RNA profile revealed by microarray indicates that lncCUEDC1 serves a negative regulatory role in breast cancer stem cells. Int J Oncol. 2020;56(3):807–20.
Xu N, et al. Microarray expression profile analysis of long non-coding RNAs in human breast cancer: a study of Chinese women. Biomed Pharmacother Biomed Pharmacother. 2015;69:221–7.
Chen C, et al. Microarray expression profiling of dysregulated long non-coding RNAs in triple-negative breast cancer. Cancer Biol Ther. 2015;16:856–65.
Hüttenhofer A, Vogel J. Experimental approaches to identify non-coding RNAs. Nucleic Acids Res. 2006;34(2):635–46.
Brzezinski M, Michelotti GA, Schwinn DA. Chapter 5—Genomics and proteomics. In: Hemmings HC, Hopkins PM, editors. Foundations of anesthesia (Second Edition). Edinburgh: Mosby; 2006. p. 71–8.
Siddika T, Heinemann IU. Bringing MicroRNAs to light: methods for MicroRNA quantification and visualization in live cells. Front Bioeng Biotechnol. 2021;8:619583.
Stoughton RB. Applications of DNA microarrays in biology. Annu Rev Biochem. 2005;74:53–82.
Nassar FJ, et al. microRNA expression in ethnic specific early stage breast cancer: an integration and comparative analysis. Sci Rep. 2017;7(1):16829.
Riaz M, et al. miRNA expression profiling of 51 human breast cancer cell lines reveals subtype and driver mutation-specific miRNAs. Breast Cancer Res. 2013;15(2):R33.
Zhang S, et al. Microarray profile of circular RNAs identifies hsa_circ_0014130 as a new circular RNA biomarker in non-small cell lung cancer. Sci Rep. 2018;8(1):2878.
Jaksik R, et al. Microarray experiments and factors which affect their reliability. Biol Direct. 2015;10:46.
Fathallah-Shaykh HM. Microarrays: applications and pitfalls. Arch Neurol. 2005;62(11):1669–72.
Ewis AA, et al. A history of microarrays in biomedicine. Expert Rev Mol Diagn. 2005;5(3):315–28.
Aparna GM, Tetala KKR. Recent progress in development and application of DNA, protein, peptide, glycan, antibody, and aptamer microarrays. Biomolecules. 2023;13(4):602.
N Gupta, VK Verma. Next-generation sequencing and its application: empowering in public health beyond reality. Microbial Technol Welfare Soc. 2019; 313–341.
Munshi A, Sharma V. Chapter 8—omics and edible vaccines. In: Barh D, Azevedo V, editors. Omics technologies and bio-engineering. Cambridge: Academic Press; 2018. p. 129–41.
Satam H, et al. Next-generation sequencing technology: current trends and advancements. Biology. 2023;12(7):997.
Shin T-J, et al. Concise approach for screening long non-coding RNAs functionally linked to human breast cancer associated genes. Exp Mol Pathol. 2019;108:89–96.
Evans T, Matulonis UA. Next-generation sequencing: role in gynecologic cancers. J Natl Compr Cancer Network JNCCN. 2016;14(9):1165–73.
A Gkazi A. An overview of next-generation sequencing. 2021. https://www.technologynetworks.com/genomics/articles/an-overview-of-next-generation-sequencing-346532.
Zhong W, et al. Long non-coding RNA expression profiles in peripheral blood mononuclear cells of patients with coronary artery disease. J Thorac Dis. 2020;12(11):6813–25.
Sun Q-L, et al. Expression profile analysis of long non-coding RNA associated with vincristine resistance in colon cancer cells by next-generation sequencing. Gene. 2015;572(1):79–86.
Krishnan P, et al. Next generation sequencing profiling identifies miR-574-3p and miR-660-5p as potential novel prognostic markers for breast cancer. BMC Genomics. 2015;16(1):735.
Krishnan P, et al. Profiling of small nucleolar RNAs by next generation sequencing: potential new players for breast cancer prognosis. PLoS ONE. 2016;11(9): e0162622.
Mardis ER. Next-generation DNA sequencing methods. Annu Rev Genomics Hum Genet. 2008;9:387–402.
Mardis ER. Next-generation sequencing platforms. Annu Rev Anal Chem. 2013;6:287–303.
Mardis ER. A decade’s perspective on DNA sequencing technology. Nature. 2011;470(7333):198–203.
Dixon P. Chapter One—Structure and function of the genome. In: Bennett P, Williamson C, editors. Basic science in obstetrics and gynaecology (Fourth Edition). London: Churchill Livingstone; 2010. p. 1–11.
Caetano-Anollés D. Polymerase chain reaction. In: Maloy S, Hughes K, editors. Brenner’s Encyclopedia of Genetics (Second Edition). San Diego: Academic Press; 2013. p. 392–5.
Tang JYH. Chapter 38—detection of microbiological hazards. In: Andersen V, Lelieveld H, Motarjemi Y, editors. Food safety management (Second Edition). San Diego: Academic Press; 2023. p. 835–50.
Allard MW, et al. Molecular techniques in foodborne disease surveillance. In: Smithers GW, editor., et al., Encyclopedia of food safety (Second Edition). Oxford: Academic Press; 2024. p. 61–85.
Ho-Pun-Cheung A, et al. Reverse transcription-quantitative polymerase chain reaction: description of a RIN-based algorithm for accurate data normalization. BMC Mol Biol. 2009;10:31.
Shen Y, et al. Prognostic and predictive values of long non-coding RNA LINC00472 in breast cancer. Oncotarget. 2015;6(11):8579–92.
Redis RS, et al. CCAT2, a novel long non-coding RNA in breast cancer: expression study and clinical correlations. Oncotarget. 2013;4(10):1748–62.
Fan S, et al. Downregulation of the long non-coding RNA ZFAS1 is associated with cell proliferation, migration and invasion in breast cancer. Mol Med Rep. 2018;17(5):6405–12.
Riahi A, et al. Overexpression of long non-coding RNA MCM3AP-AS1 in breast cancer tissues compared to adjacent non-tumour tissues. Br J Biomed Sci. 2021;78(2):53–7.
Zhang R, et al. MALAT1 Long Non-coding RNA expression in thyroid tissues: analysis by in situ hybridization and real-time PCR. Endocr Pathol. 2017;28(1):7–12.
Guo Q, et al. Long non-coding RNA PRNCR1 has an oncogenic role in breast cancer. Exp Ther Med. 2019;18(6):4547–54.
Harrison GP, et al. Pausing of reverse transcriptase on retroviral RNA templates is influenced by secondary structures both 5′ and 3′ of the catalytic site. Nucleic Acids Res. 1998;26(14):3433–42.
Brooks EM, Sheflin LG, Spaulding SW. Secondary structure in the 3’UTR of EGF and the choice of reverse transcriptases affect the detection of message diversity by RT-PCR. Biotechniques. 1995;19(5):806–12.
Gareev I, et al. Long non-coding RNAs in oncourology. Non-coding RNA Res. 2021;6(3):139–45.
Tang H, et al. Salivary lncRNA as a potential marker for oral squamous cell carcinoma diagnosis. Mol Med Rep. 2013;7(3):761–6.
Clark MB, et al. Genome-wide analysis of long non-coding RNA stability. Genome Res. 2012;22(5):885–98.
He SL, Green R. Northern blotting. Methods Enzymol. 2013;530:75–87.
WD Northern Blot. 2023. https://www.genome.gov/genetics-glossary/Northern-Blot.
Zambalde ÉP et al. TLNC-UC.147, a novel long RNA (lncRNA) from an ultraconserved region as potential biomarker in luminal a breast cancer. Mastology 2021.
Goyal B, et al. Diagnostic, prognostic, and therapeutic significance of long non-coding RNA MALAT1 in cancer. Biochim Biophys Acta Rev Cancer. 2021;1875(2): 188502.
Pereira Zambalde E, et al. A novel lncRNA derived from an ultraconserved region: lnc-uc147, a potential biomarker in luminal A breast cancer. RNA Biol. 2021;18(sup1):416–29.
Zhang B, et al. Vanguard is a glucose deprivation-responsive long non-coding RNA essential for chromatin remodeling-reliant DNA repair. Adv Sci. 2022;9(30):2201210.
Wang Y, et al. LncRNA-encoded polypeptide ASRPS inhibits triple-negative breast cancer angiogenesis. J Exp Med. 2019. https://doiorg.publicaciones.saludcastillayleon.es/10.1084/jem.20190950.
Yang T, Zhang M, Zhang N. Modified Northern blot protocol for easy detection of mRNAs in total RNA using radiolabeled probes. BMC Genomics. 2022;23(1):66.
Streit S, et al. Northern blot analysis for detection and quantification of RNA in pancreatic cancer cells and tissues. Nat Protoc. 2009;4(1):37–43.
Pardue ML, Gall JG. Molecular hybridization of radioactive DNA to the DNA of cytological preparations. Proc Natl Acad Sci. 1969;64(2):600–4.
John H, Birnstiel M, Jones K. RNA-DNA hybrids at the cytological level. Nature. 1969;223(5206):582.
Shakoori AR. Fluorescence in situ hybridization (FISH) and its applications. Chromosome structure and aberrations. Berlin: Springer; 2017. p. 343–67.
Vautrot V, et al. Fluorescence in situ hybridization of small non-coding RNAs. Methods Mol Biol. 2021;2300:73–85.
Chang J, et al. RNA fluorescence in situ hybridization for long non-coding RNA localization in human osteosarcoma cells. JOVE. 2023. https://doiorg.publicaciones.saludcastillayleon.es/10.3791/65545.
Guo S, et al. Novel breast-specific long non-coding RNA LINC00993 Acts as a tumor suppressor in triple-negative breast cancer. Front Oncol. 2019;9:1325.
Xia W, et al. Extracellular vesicles carry lncRNA SNHG16 to promote metastasis of breast cancer cells via the miR-892b/PPAPDC1A Axis. Front Cell Dev Biol. 2021;9:628573.
Chang J, et al. RNA fluorescence in situ hybridization for long non-coding RNA localization in human osteosarcoma cells. J Vis Exp. 2023. https://doiorg.publicaciones.saludcastillayleon.es/10.3791/65545.
Urbanek MO, Nawrocka AU, Krzyzosiak WJ. Small RNA detection by in situ hybridization methods. Int J Mol Sci. 2015;16(6):13259–86.
Cui Y, et al. A long non-coding RNA Lnc712 regulates breast cancer cell proliferation. Int J Biol Sci. 2020;16(1):162–71.
Jiang H, et al. Long non-coding RNA SNHG3 promotes breast cancer cell proliferation and metastasis by binding to microRNA-154-3p and activating the notch signaling pathway. BMC Cancer. 2020;20(1):1–13.
Chen Z, et al. Long non-coding RNA SNHG7 inhibits NLRP3-dependent pyroptosis by targeting the miR-34a/SIRT1 axis in liver cancer. Oncol Lett. 2020;20(1):893–901.
Huber D, Voith von Voithenberg L, Kaigala GV. Fluorescence in situ hybridization (FISH): History, limitations and what to expect from micro-scale FISH? Micro Nano Eng. 2018;1:15–24.
Ikbal Atli E, et al. Pros and cons for fluorescent in situ hybridization, karyotyping and next generation sequencing for diagnosis and follow-up of multiple myeloma. Balkan J Med Genet. 2020;23(2):59–64.
Bhaya D, Davison M, Barrangou R. CRISPR-Cas systems in bacteria and archaea: versatile small RNAs for adaptive defense and regulation. Annu Rev Genet. 2011;45:273–97.
Wang H, La Russa M, Qi LS. CRISPR/Cas9 in genome editing and beyond. Annu Rev Biochem. 2016;85:227–64.
Ede DR, et al. 4.32 Gene editing tools. In: Ducheyne P, editor., et al., Comprehensive biomaterials II. Oxford: Elsevier; 2017. p. 589–99.
Rosenlund IA, et al. CRISPR/Cas9 to silence long non-coding RNAs. In: Long non-coding RNAs in cancer. Springer; 2021. p. 175–87.
Yang J, et al. CRISPR/Cas9-mediated non-coding RNA editing in human cancers. RNA Biol. 2018;15(1):35–43.
Fellmann C, et al. Cornerstones of CRISPR–Cas in drug discovery and therapy. Nat Rev Drug Discov. 2017;16(2):89–100.
Canver MC, Bauer DE, Orkin SH. Functional interrogation of non-coding DNA through CRISPR genome editing. Methods. 2017;121:118–29.
DeOcesano-Pereira C, et al. Functional impact of the long non-coding RNA MEG3 deletion by CRISPR/Cas9 in the human triple negative metastatic Hs578T cancer cell line. Oncol Lett. 2019;18:5941–51.
Diermeier SD, et al. The long non-coding RNA MaTAR20 promotes mammary tumor growth by regulating angiogenesis pathways. biorxiv. 2021;71:7.
Lin L-C, et al. NAD(P)H:quinone oxidoreductase 1 determines radiosensitivity of triple negative breast cancer cells and is controlled by long non-coding RNA NEAT1. Int J Med Sci. 2020;17:2214–24.
Ali HS, et al. lncRNA- RP11–156p1.3, novel diagnostic and therapeutic targeting via CRISPR/Cas9 editing in hepatocellular carcinoma. Genomics. 2020;112(5):3306–14.
Yi B, et al. CRISPR interference and activation of the microRNA-3662-HBP1 axis control progression of triple-negative breast cancer. Oncogene. 2022;41(2):268–79.
Deocesano-Pereira C, et al. Functional impact of the long non-coding RNA MEG3 deletion by CRISPR/Cas9 in the human triple negative metastatic Hs578T cancer cell line. Oncol Lett. 2019;18(6):5941–51.
Goyal A, et al. Challenges of CRISPR/Cas9 applications for long non-coding RNA genes. Nucleic Acids Res. 2017;45(3): e12.
Sato G, Kuroda K. Overcoming the limitations of CRISPR-Cas9 systems in saccharomyces cerevisiae: off-target effects, epigenome, and mitochondrial editing. Microorganisms. 2023;11(4):1040.
Guo C, et al. Off-target effects in CRISPR/Cas9 gene editing. Front Bioeng Biotechnol. 2023;11:1143157.
Sahadevan S, Pérez-Berlanga M, Polymenidou M. Identification of RNA–RBP interactions in subcellular compartments by CLIP-Seq. In: The integrated stress response: methods and protocols. Springer; 2022. p. 305–23.
Jensen KB, Darnell RB. CLIP: crosslinking and immunoprecipitation of in vivo RNA targets of RNA-binding proteins. Methods Mol Biol. 2008;488:85–98.
Fallatah A et al. HNRNPK is retained in the cytoplasm by Keratin 19 to stabilize target mRNAs. Biorxiv. 2022.
Mato Prado M, et al. Investigating miRNA-mRNA regulatory networks using crosslinking immunoprecipitation methods for biomarker and target discovery in cancer. Expert Rev Mol Diagn. 2016;16:1155–62.
Costales MG, et al. A designed small molecule inhibitor of a non-coding RNA sensitizes HER2 negative cancers to herceptin. J Am Chem Soc. 2019;141(7):2960–74.
Li J, et al. starBase v2.0: decoding miRNA-ceRNA, miRNA-ncRNA and protein–RNA interaction networks from large-scale CLIP-Seq data. Nucleic Acids Res. 2013;42:D92–7.
Lidschreiber M, et al. The APT complex is involved in non-coding RNA transcription and is distinct from CPF. Nucleic Acids Res. 2018;46:11528–38.
Ke H, et al. NEAT1 is required for survival of breast cancer cells through FUS and miR-548. Gene Regul Syst Biol. 2016;10s1:GRSB.S29414.
Wheeler EC, Van Nostrand EL, Yeo GW. Advances and challenges in the detection of transcriptome-wide protein-RNA interactions. Wiley Interdiscip Rev RNA. 2018. https://doiorg.publicaciones.saludcastillayleon.es/10.1002/wrna.1436.
mirbase. https://mirbase.org/. Accessed 01 Feb 2024.
Kozomara A, Birgaoanu M, Griffiths-Jones S. miRBase: from microRNA sequences to function. Nucleic Acids Res. 2018;47:D155–62.
mirwalk. http://mirwalk.umm.uni-heidelberg.de/. Accessed 01 Feb 2024.
Sticht C, et al. miRWalk: an online resource for prediction of microRNA binding sites. PLoS ONE. 2018;13(10):e0206239.
Dweep H, Gretz N, Sticht C. miRWalk database for miRNA-target interactions. Methods Mol Biol. 2014;1182:289–305.
Parveen A, Gretz N, Dweep H. Obtaining miRNA-target interaction information from miRWalk20. Curr Protocols Bioinform. 2016;55:12.15.1-12.15.27.
lncipedia. https://lncipedia.org/. Accessed 01 Feb 2024.
Volders P-J, et al. An update on LNCipedia: a database for annotated human lncRNA sequences. Nucleic Acids Res. 2015;43:4363–4.
Volders P-J, et al. LNCipedia: a database for annotated human lncRNA transcript sequences and structures. Nucleic Acids Res. 2012;41:D246–51.
noncode. http://www.noncode.org/. Accessed 01 Feb 2024.
Liu C, et al. NONCODE: an integrated knowledge database of non-coding RNAs. Nucleic Acids Res. 2004;33:D112–5.
Beňačka R, et al. Classic and new markers in diagnostics and classification of breast cancer. Cancers (Basel). 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.3390/cancers14215444.
Xu J, et al. Roles of miRNA and lncRNA in triple-negative breast cancer. J Zhejiang Univ Sci B. 2020;21(9):673–89.
Tie Y, et al. Circulating miRNA and cancer diagnosis. Sci China Ser C Life Sci. 2009;52(12):1117–22.
Iorio MV, Croce CM. MicroRNA dysregulation in cancer: diagnostics, monitoring and therapeutics. A comprehensive review. EMBO Mol Med. 2012;4(3):143–59.
Zhang T, et al. Long non-coding RNA and breast cancer. Technol Cancer Res Treat. 2019;18:1533033819843889.
Elias-Rizk T, et al. The long non coding RNA H19 as a biomarker for breast cancer diagnosis in Lebanese women. Sci Rep. 2020;10(1):22228.
Tao W, et al. LncRNA DANCR contributes to tumor progression via targetting miR-216a-5p in breast cancer: lncRNA DANCR contributes to tumor progression. Biosci Rep. 2019;39(4):BSR20181618.
Tsyganov MM, Ibragimova MK. MALAT1 long non-coding RNA and its role in breast carcinogenesis. Acta Nat. 2023;15(2):32–41.
Li Y, et al. Long non-coding RNA LINC01787 drives breast cancer progression via disrupting miR-125b generation. Front Oncol. 2019;9:1140.
Xie H, et al. MicroRNA-100 inhibits breast cancer cell proliferation, invasion and migration by targeting FOXA1. Oncol Lett. 2021;22(6):1.
Mayoral-Varo V, et al. miR205 inhibits stem cell renewal in SUM159PT breast cancer cells. PLoS ONE. 2017;12(11): e0188637.
Kim S-J, et al. Development of microRNA-145 for therapeutic application in breast cancer. J Control Release. 2011;155(3):427–34.
Zhao D, et al. In vivo monitoring of angiogenesis inhibition via down-regulation of mir-21 in a VEGFR2-luc murine breast cancer model using bioluminescent imaging. PLoS ONE. 2013;8(8): e71472.
Devulapally R, et al. Polymer nanoparticles mediated codelivery of antimiR-10b and antimiR-21 for achieving triple negative breast cancer therapy. ACS Nano. 2015;9(3):2290–302.
Yin H, et al. Delivery of Anti-miRNA for triple-negative breast cancer therapy using RNA nanoparticles targeting stem cell marker CD133. Mol Ther. 2019;27(7):1252–61.
Ren Y, et al. Sequential co-delivery of miR-21 inhibitor followed by burst release doxorubicin using NIR-responsive hollow gold nanoparticle to enhance anticancer efficacy. J Control Release. 2016;228:74–86.
Park EY, et al. Targeting of miR34a–NOTCH1 axis reduced breast cancer stemness and chemoresistance. Can Res. 2014;74(24):7573–82.
Ito M, Miyata Y, Okada M. Current clinical trials with non-coding RNA-based therapeutics in malignant diseases: a systematic review. Trans Oncol. 2023;31: 101634.
Zhang L, Liao Y, Tang L. MicroRNA-34 family: a potential tumor suppressor and therapeutic candidate in cancer. J Exp Clin Cancer Res. 2019;38(1):1–13.
Hong DS, et al. Phase 1 study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br J Cancer. 2020;122(11):1630–7.
Zhang L, et al. Cytosolic co-delivery of miRNA-34a and docetaxel with core-shell nanocarriers via caveolae-mediated pathway for the treatment of metastatic breast cancer. Sci Rep. 2017;7(1):46186.
Beg MS, et al. Phase I study of MRX34, a liposomal miR-34a mimic, administered twice weekly in patients with advanced solid tumors. Invest New Drugs. 2017;35:180–8.
Kim J, et al. Long non-coding RNA MALAT1 suppresses breast cancer metastasis. Nat Genet. 2018;50(12):1705–15.
Yang Z, et al. Downregulation of long non-coding RNA MALAT1 induces tumor progression of human breast cancer through regulating CCND1 expression. Open Life Sci. 2016;11(1):232–6.
Smith NE, et al. The long and the short of it: NEAT1 and cancer cell metabolism. Cancers. 2022;14(18):4388.
Wang J, et al. Breast cancer cell–derived microRNA-155 suppresses tumor progression via enhancing immune cell recruitment and antitumor function. J Clin Investig. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.1172/JCI157248.
İlhan A, et al. The dual role of microRNA (miR)-20b in cancers: friend or foe? Cell Commun Signal. 2023;21(1):26.
Wu T, et al. RETRACTED ARTICLE: LncRNA BCAR4 promotes migration, invasion, and chemo-resistance by inhibiting miR-644a in breast cancer. J Exp Clin Cancer Res. 2023;42(1):14.
Shui X, et al. Long non-coding RNA BCAR4 promotes chondrosarcoma cell proliferation and migration through activation of mTOR signaling pathway. Exp Biol Med. 2017;242(10):1044–50.
Liang W-H, et al. DSCAM-AS1 promotes tumor growth of breast cancer by reducing miR-204-5p and up-regulating RRM2. Mol Carcinog. 2019;58(4):461–73.
Niknafs YS, et al. The lncRNA landscape of breast cancer reveals a role for DSCAM-AS1 in breast cancer progression. Nat Commun. 2016;7(1):12791.
Yadav N, et al. Progesterone modulates the DSCAM-AS1/miR-130a/ESR1 axis to suppress cell invasion and migration in breast cancer. Breast Cancer Res. 2022;24(1):97.
Li L, et al. lncRNA DSCAM-AS1 facilitates the progression of endometrial cancer via miR-136-5p. Oncol Lett. 2021;22(6):825.
Feng Y, et al. A novel lncRNA SOX2OT promotes the malignancy of human colorectal cancer by interacting with miR-194-5p/SOX5 axis. Cell Death Dis. 2021;12(5):499.
Zhang W, et al. SOX2-OT induced by PAI-1 promotes triple-negative breast cancer cells metastasis by sponging miR-942-5p and activating PI3K/Akt signaling. Cell Mol Life Sci. 2022;79(1):59.
Wang Y, et al. SOX2OT, a novel tumor-related long non-coding RNA. Biomed Pharmacother. 2020;123: 109725.
Li Y, et al. Long non-coding RNA UCA1 promotes breast cancer by upregulating PTP1B expression via inhibiting miR-206. Cancer Cell Int. 2019;19(1):275.
Choudhry H. UCA1 overexpression promotes hypoxic breast cancer cell proliferation and inhibits apoptosis via HIF-1α activation. J Oncol. 2021;2021:5512156.
Wo L, et al. Up-regulation of LncRNA UCA1 by TGF-β promotes doxorubicin resistance in breast cancer cells. Immunopharmacol Immunotoxicol. 2022;44(4):492–9.
Sun Y, et al. ETS-1-activated LINC01016 over-expression promotes tumor progression via suppression of RFFL-mediated DHX9 ubiquitination degradation in breast cancers. Cell Death Dis. 2023;14(8):507.
Li Y, et al. LncRNA H19 promotes triple-negative breast cancer cells invasion and metastasis through the p53/TNFAIP8 pathway. Cancer Cell Int. 2020;20(1):200.
Si X, et al. LncRNA H19 confers chemoresistance in ERα-positive breast cancer through epigenetic silencing of the pro-apoptotic gene BIK. Oncotarget. 2016;7(49):81452–62.
Pickard MR, Williams GT. The hormone response element mimic sequence of GAS5 lncRNA is sufficient to induce apoptosis in breast cancer cells. Oncotarget. 2016;7(9):10104–16.
Li W, et al. Downregulation of LncRNA GAS5 causes trastuzumab resistance in breast cancer. Oncotarget. 2016;7(19):27778–86.
Ma C, et al. The growth arrest-specific transcript 5 (GAS5): a pivotal tumor suppressor long non-coding RNA in human cancers. Tumour Biol. 2016;37(2):1437–44.
Cai Y, He J, Zhang D. Long non-coding RNA CCAT2 promotes breast tumor growth by regulating the Wnt signaling pathway. Onco Targets Ther. 2015;8:2657–64.
Liu B, et al. A cytoplasmic NF-κB interacting long non-coding RNA blocks IκB phosphorylation and suppresses breast cancer metastasis. Cancer Cell. 2015;27(3):370–81.
Hu P, et al. NBAT1 suppresses breast cancer metastasis by regulating DKK1 via PRC2. Oncotarget. 2015;6(32):32410–25.
Sun L, Li Y, Yang B. Downregulated long non-coding RNA MEG3 in breast cancer regulates proliferation, migration and invasion by depending on p53’s transcriptional activity. Biochem Biophys Res Commun. 2016;478(1):323–9.
Sharma U, et al. LncRNA ZFAS1 inhibits triple-negative breast cancer by targeting STAT3. Biochimie. 2021;182:99–107.
Li H, et al. miR-17-5p promotes human breast cancer cell migration and invasion through suppression of HBP1. Breast Cancer Res Treat. 2011;126(3):565–75.
Kong W, et al. Prognostic value of miR-17–5p in cancers: a meta-analysis. OncoTargets Ther. 2018;11(null):3541–9.
An G, et al. Effects of miR-93 on epithelial-to-mesenchymal transition and vasculogenic mimicry in triple-negative breast cancer cells. Mol Med Rep. 2021;23(1):30.
Ren L, et al. EYA2 upregulates miR-93 to promote tumorigenesis of breast cancer by targeting and inhibiting the STING signaling pathway. Carcinogenesis. 2021;43(12):1121–30.
Liu Y, et al. Association of high miR-27a, miR-206, and miR-214 expression with poor patient prognosis and increased chemoresistance in triple-negative breast cancer. Am J Cancer Res. 2023;13(6):2471–87.
Ouyang Q, et al. lncRNA MT1JP suppresses biological activities of breast cancer cells in vitro and in vivo by regulating the miRNA-214/RUNX3 axis. OncoTargets Ther. 2020;13(null):5033–46.
Jia Y, et al. miR-301 regulates the SIRT1/SOX2 pathway via CPEB1 in the breast cancer progression. Mol Ther Oncolyt. 2021;22:13–26.
Rahmani F, et al. Role of regulatory miRNAs of the PI3K/AKT signaling pathway in the pathogenesis of breast cancer. Gene. 2020;737: 144459.
Gwak JM, et al. MicroRNA-9 is associated with epithelial-mesenchymal transition, breast cancer stem cell phenotype, and tumor progression in breast cancer. Breast Cancer Res Treat. 2014;147(1):39–49.
Yao X, Xie L, Zeng Y. MiR-9 Promotes angiogenesis via targeting on sphingosine-1- phosphate receptor 1. Front Cell Dev Biol. 2020;8:755.
Zhuang G, et al. Tumour-secreted miR-9 promotes endothelial cell migration and angiogenesis by activating the JAK-STAT pathway. EMBO J. 2012;31(17):3513–23.
Chen D, et al. MiR-373 drives the epithelial-to-mesenchymal transition and metastasis via the miR-373-TXNIP-HIF1α-TWIST signaling axis in breast cancer. Oncotarget. 2015;6(32):32701–12.
Bakr NM, et al. Impact of circulating miRNA-373 on breast cancer diagnosis through targeting VEGF and cyclin D1 genes. J Genet Eng Biotechnol. 2021;19(1):84.
Liu K, et al. Let-7a inhibits growth and migration of breast cancer cells by targeting HMGA1. Int J Oncol. 2015;46(6):2526–34.
Sun X, et al. Let-7c blocks estrogen-activated Wnt signaling in induction of self-renewal of breast cancer stem cells. Cancer Gene Ther. 2016;23(4):83–9.
He T, et al. MicroRNA-542-3p inhibits tumour angiogenesis by targeting angiopoietin-2. J Pathol. 2014;232(5):499–508.
Chao C-H, et al. MicroRNA-205 signaling regulates mammary stem cell fate and tumorigenesis. J Clin Investig. 2014;124(7):3093–106.
Xu Q, et al. A regulatory circuit of miR-148a/152 and DNMT1 in modulating cell transformation and tumor angiogenesis through IGF-IR and IRS1. J Mol Cell Biol. 2013;5(1):3–13.
Sossey-Alaoui K, et al. WAVE3, an actin remodeling protein, is regulated by the metastasis suppressor microRNA, miR-31, during the invasion-metastasis cascade. Int J Cancer. 2011;129(6):1331–43.
Goldberger N, et al. Inherited variation in miR-290 expression suppresses breast cancer progression by targeting the metastasis susceptibility gene Arid4b. Can Res. 2013;73(8):2671–81.
Li W, et al. MicroRNA-34a: potent tumor suppressor, cancer stem cell inhibitor, and potential anticancer therapeutic. Front Cell Dev Biol. 2021. https://doiorg.publicaciones.saludcastillayleon.es/10.3389/fcell.2021.640587.
Chen H, et al. MiR-25-3p promotes the proliferation of triple negative breast cancer by targeting BTG2. Mol Cancer. 2018;17(1):4.
Abdallah M, Aziz IH, Alsammarraie AZ. Assessment of miRNA-10b expression levels as a potential precursor to metastasis in localized and locally advanced/metastatic breast cancer among Iraqi patients. Int J Breast Cancer. 2024;2024:2408355.
Nie J, et al. MiR-125b regulates the proliferation and metastasis of triple negative breast cancer cells via the Wnt/β-catenin pathway and EMT. Biosci Biotechnol Biochem. 2019;83(6):1062–71.
Rajarajan D, et al. miR-145-5p as a predictive biomarker for breast cancer stemness by computational clinical investigation. Comput Biol Med. 2021;135: 104601.
Wang J, et al. Breast cancer cell-derived microRNA-155 suppresses tumor progression via enhancing immune cell recruitment and antitumor function. J Clin Invest. 2022. https://doiorg.publicaciones.saludcastillayleon.es/10.1172/JCI157248.
MicroRNAs methylation and expression profiling for identification of breast cancer patients at high risk to develop distant metastases. 2024.
Studying the association of genetic variations in long intergenic non-coding RNA 00511 (LINC00511) with breast cancer among the Egyptian population, E. British University In, Editor. 2024.
"Investigating the role of key non-coding RNA(s) in breast cancer patients and their possible crosstalk with chromatin remodeling machinery". 2024.
Efficacy and Safety Study of TA(E)C-GP Versus A(E)C-T for the high risk triple-negative breast cancer patients predicted by the messenger RNA (mRNA)-Long Non-coding RNA (lncRNA) signature and validation of the signature's efficacy. 2015.
Impact of somatic PIK3CA mutations on pathological complete response (pCR) in HER2-positive early breast cancer. 2023.
Acknowledgements
We would like to thank Mazandaran University of Medical Sciences and Amol University of Special Modern Technologies.
Funding
There was no external funding obtained for this study.
Author information
Authors and Affiliations
Contributions
M.S. and S.H. Wrote the manuscript and designed figures. M.A. Assembly of the paper and guided the writing process with editing. The authors read and approved the final manuscript for publication.
Corresponding author
Ethics declarations
Ethics approval and consent to participate
Not applicable.
Competing interests
The authors declare no competing interests.
Additional information
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Solaimani, M., Hosseinzadeh, S. & Abasi, M. Non-coding RNAs, a double-edged sword in breast cancer prognosis. Cancer Cell Int 25, 123 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12935-025-03679-0
Received:
Accepted:
Published:
DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s12935-025-03679-0