Next Article in Journal
Bridging Metabolic-Associated Steatotic Liver Disease and Cardiovascular Risk: A Potential Role for Ketogenesis
Previous Article in Journal
Short-Term Periodic Fasting Reduces Ischemia-Induced Necrosis in Musculocutaneous Flap Tissue
Previous Article in Special Issue
Predictive Value and Therapeutic Significance of Somatic BRCA Mutation in Solid Tumors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functions of Differentially Regulated miRNAs in Breast Cancer Progression: Potential Markers for Early Detection and Candidates for Therapy

by
Kumar Subramanian
and
Raghu Sinha
*
Department of Biochemistry and Molecular Biology, Penn State College of Medicine, Hershey, PA 17033, USA
*
Author to whom correspondence should be addressed.
Current address: Department of Surgery, Georgetown University School of Medicine, Washington, DC 20007, USA
Biomedicines 2024, 12(3), 691; https://doi.org/10.3390/biomedicines12030691
Submission received: 13 December 2023 / Revised: 15 February 2024 / Accepted: 21 February 2024 / Published: 20 March 2024
(This article belongs to the Special Issue Feature Reviews in Cancer Biomarkers)

Abstract

:
Breast cancer remains a major global health concern, emphasizing the need for reliable biomarkers to enhance early detection and therapeutic interventions. MicroRNAs (miRNAs) are evolutionarily conserved small non-coding RNA (~22 nt in length) molecules, which are aberrantly expressed in cancer and seem to influence tumor behavior and progression. Specific miRNA dysregulation has been associated with breast cancer initiation, proliferation, invasion, and metastasis. Understanding the functional roles of these miRNAs provides valuable insights into the intricate molecular mechanisms underlying breast cancer progression. The diagnostic potential of miRNAs as non-invasive biomarkers for early breast cancer detection is a burgeoning area of research. This review aims to elucidate the functions of differentially regulated miRNAs in breast cancer progression and assess their potential as markers for early detection, stage-specific biomarkers, and therapeutic targets. Furthermore, the ability of specific miRNAs to serve as prognostic indicators and predictors of treatment response highlights their potential clinical utility in guiding personalized therapeutic interventions.

Graphical Abstract

1. Introduction

Breast cancer is a disease caused by genetic aberrations and deleterious environmental exposure [1]. Breast cancer is a complex and enigmatic disease caused by a series of alterations in genes that control cell growth and proliferation. Breast cancer is the most common cancer in women and is a heterogeneous disease with diverse molecular subtypes and clinical presentations. Despite advances in treatment, it is the leading cause of cancer mortality in women worldwide. Globally, it accounts for 16 percent of cancer deaths, and has over 20 distinct subtypes that differ genetically, morphologically, and clinically [2,3]. About 90% of these deaths are due to metastases. Currently, breast cancer is the fourth leading cause of cancer-related death, with 2.3 million new cases in 2022 worldwide [4]. The metastatic spread of breast cancer, typically to the bone, lung, liver, and brain, accounts for most cancer-related deaths [5]. There are few treatment strategies for metastatic breast cancer, which is incurable with a median survival rate of 2–3 years [6]. Breast cancer cases are increasing alarmingly, underscoring the importance of treating the disease in many ways. Several trials have been conducted for the successful development of drugs for breast cancers over the past few decades. As a result of inadequate early detection, breast cancer patients have a high mortality rate and recurrence rate. Thus, to improve disease outcomes and prolong patient survival, it is vital to find novel early prognostic and diagnostic biomarkers and effective therapeutic methods [7].
Recent research studies have shed light on the progression of breast cancer, its pathogenesis, and the molecular pathways involved in proliferation. As evidenced by recent research, molecular marker-based targeted therapies using miRNAs may improve the prognosis and diagnosis of a wide range of diseases, including breast cancer [8]. A growing body of evidence suggests that miRNAs play a critical role in tumorigenesis and breast cancer development. These molecules are altered in different tumorigenic processes of breast cancer. miRNAs are small, noncoding RNA molecules that regulate gene expression through interfering with transcription. miRNAs play a significant role in regulating a variety of cellular processes, such as angiogenesis, apoptosis, and the cell cycle. Due to their ability to modulate multiple targets within these pathways, they play a significant role in maintaining cellular homeostasis. It is often observed that dysregulation of miRNAs in these processes contributes to various diseases, including breast cancer, which makes them potential therapeutic targets or diagnostic markers for cancers [9]. It is important to understand the complex network of miRNA-mediated regulation to gain insight into the molecular mechanisms that govern these vital processes within cells.
miRNA was discovered by Ambros and co-workers in Caenorhabditis elegans (Nematode), during their genetic study to investigate defects in the temporal control of C. elegans development [10]. Recent studies have demonstrated that miRNAs play critical roles in the development of breast cancer, differentiation, proliferation, and other physiological processes [11]. A growing body of evidence suggests that miRNAs might have very important clinical implications.
It is well established that miRNAs are critical regulators of mRNA expression and cell activity, both in normal and abnormal biological processes, including breast cancer. As miRNA dysregulation occurs in various types of cancers including breast cancer and leads to tumor initiation, drug resistance, and metastasis, the therapeutic strategies aimed at modulating miRNA expression levels and identifying their targets are promising strategies [9,11]. miRNAs are secreted by a variety of cells and transported to a variety of bodily fluids in remarkably stable forms (i.e., peripheral blood, saliva, cerebrospinal fluid, ascites, urine, and breast milk) through extracellular vesicles, and interestingly, the levels of miRNAs circulating in cancer patients differ from those of healthy donors [12,13,14,15,16]. Therefore, a quantitative analysis of such potential circulatory miRNAs would be an ideal approach for detecting breast cancer via liquid biopsy in the early stages. This review summarizes the major functions of miRNAs in breast cancer progression and discusses the clinical applications of differentially regulated miRNAs, especially circulating miRNAs in early diagnosis and as targets for therapy.

2. miRNA Biogenesis and Maturation

Pri-miRNAs are processed step-by-step in the cytoplasm and nucleus during the biogenesis of miRNAs. In humans, miRNA gene transcription takes place within the nucleus following the cleavage of the ~80 nucleotide stem-loop pre-microRNA precursor performed by the microprocessor complex consisting of Drosha, an RNase III-type nuclease, a double strand RNA-binding protein co-factor, and the DiGeorge syndrome critical region 8 gene (DGCR8) (Figure 1). The complex cleaves the pri-miRNA to generate a shorter hairpin-shaped precursor called precursor miRNA (pre-miRNA). The pri-miRNAs are processed into 60–70 nucleotide hairpin structures (pre-miRNAs) and are exported from the nucleus to the cytoplasm, supported by the nucleocytoplasmic shuttle protein Exportin-5 in a Ran-GTP-dependent manner. The enzyme Dicer further breaks down the pre-miRNA into a double-stranded RNA duplex in the cytoplasm. The mature miRNA is one of the two strands of the duplex that is chosen and loaded into the RNA-induced silencing complex (RISC) [17,18].
By base-pairing with complementary sequences in the 3’ untranslated region (UTR) of the target mRNA, the mature miRNA directs the RISC complex to its target mRNAs. As a result of this interaction, the target gene may be downregulated due to translational repression or mRNA degradation. Multiple steps and molecular players are involved in miRNA biogenesis [17]. Breast cancer has been associated with dysregulation of miRNA biogenesis or its function.

3. miRNA Targets and Their Role in Breast Cancer

Discovering the regulatory network regulated by miRNAs requires identification of miRNA-mRNA target interactions. The development of miRNA target prediction algorithms is based on several approaches. Statistical inference based on machine learning and algorithms derived from characteristics of the mRNA sequence and/or based on the miRNA-mRNA interaction are the two major categories. Pairings of miRNA and mRNA seed sequences can be analyzed and evaluated. Instead of making “de novo” predictions based on sequence features, the aim of machine learning is to classify miRNA targets that reference miRNA-mRNA duplexes with known biological significance [19]. miRNA sequences are complementary to 3′-UTR sequences of mRNA targets. For miRNA to bind to target mRNAs, the seed sequence of the miRNA 5′ region is essential. Specific seed region characteristics, as well as those in proximity, have been linked to specific effects on miRNA-induced gene repression [19].
Watson–Crick Pairing between miRNA and mRNA is needed for most target prediction algorithms. miRNA prediction tools include miRNAFinder, miRscan, miRbase, miRTarBase, and SSC profiler [20,21]. Most of them are focused on miRNA conservation characteristics across ecosystems. The prediction of miRNA in a wide range of species from the animal and plant kingdoms have proven successful with this technique.
The miRNAs play an important role in the regulation of gene expression, and the dysregulation of these molecules has been implicated in various stages of breast cancer development. Breast cancer miRNAs are classified based on their expression patterns, functional roles, and clinical implications [22,23]. miRNAs can be classified into two main categories established on their effects on tumorigenesis:
  • Oncogenic miRNAs (OncomiRs): these miRNAs are often upregulated in breast cancer and promote tumorigenesis by inhibiting tumor-suppressor genes or regulatory pathways.
  • Tumor-Suppressive miRNAs: conversely, these miRNAs are downregulated in breast cancer and typically act to inhibit oncogenes or other pro-tumorigenic processes.
To maintain normal cellular function, it is essential to maintain a balance between oncogenic and tumor-suppressive miRNAs as reviewed earlier [24]. In breast cancer, dysregulation of miRNA expression contributes to tumorigenesis (Table 1). Therapeutic potential exists in manipulating miRNA expression. OncomiRs can be inhibited or replaced with miRNAs (for tumor-suppressive miRNAs).
Studying OncomiRs and tumor-suppressive miRNAs in breast cancer is of paramount importance due to several reasons. Interactions between OncomiRs and tumor-suppressive miRNAs provide insight into the molecular mechanisms underlying breast cancer progression. The level of dysregulation of these miRNAs in tumor tissue or bodily fluids correlates with the cancer stage, aggressiveness, and clinical outcome. It is possible to tailor personalized treatment strategies for patients by identifying specific OncomiRs and tumor-suppressive miRNAs associated with breast cancer subtypes or treatment responses.

4. Significance of miRNAs in Breast Cancer Development

miRNAs function by binding to the 3′ UTR of target mRNAs, leading to translational repression or mRNA degradation. Breast cancer is characterized by dysregulation of miRNAs and their targets, which contribute to various aspects of cancer initiation, progression, and metastasis. Below is an overview of some key miRNAs and their known roles in the regulation of their targets in breast cancer.

4.1. Breast Cancer Initiation and Progression

As a multistep process, cancer initiates and progresses with the gradual transformation of human cells into highly malignant forms through progressive genetic changes [70]. It has been recognized that malignant transformation occurs through successive mutations in specific genes, leading to the activation of oncogenes and the inactivation of tumor-suppressor genes. The activities of these genes may represent the final common pathway using which many carcinogens act [71]. There are three main types of genes that play a role in tumor initiation: proto-oncogenes, tumor-suppressor genes, and genes involved in DNA repair. Changes in the genes, like mutations, amplifications, or deletions, may lead to decoupling of biological mechanism involved in the regulation of normal cell growth and differentiation [72].
Accumulating evidence suggests that cancer stem cells (CSCs) or tumor-initiating cells (TICs) with stem cell-like properties can propagate human tumors with heterogeneous tumor populations in immunodeficient mice, such as human-breast-tumor-initiating cells (BTICs) or breast cancer stem cells (BCSCs) that can regenerate breast tumors in an in vivo model. Additionally, BTICs self-renew and asymmetrically divide into differentiated cancer cells, and these are trusted to be accountable for cancer stem-like cells that drive breast tumor formation, recurrence, metastasis, and drug resistance [66,73]. Liu et al. proved that BCSCs are involved in the spontaneous metastases of human breast cancer in mouse breast cancer orthotopic models [74]. Several miRNAs have been implicated in the regulation of CSC properties. The dysregulation of miRNA might contribute to the self-renewal of BCSCs and cancer progression.
The pivotal roles of miRNA-200 are well characterized in breast cancer initiation. For example, miRNA-200 family members are significantly downregulated in CD44+, CD24−low-lineage human primary BCSCs when compared to non-tumorigenic cancer cells. Moreover, miRNA-200b regulates BCSC growth by directly targeting Suz12, a subunit of a polycomb repressor complex (PRC2) and regulates EMT by repressing the E-cadherin gene. Higher expression of the polycomb protein EZH2, which is important in the stem cell self-renewal capability of embryonic and adult stem cells, has been associated with breast cancer progression [75]. Shimono et al. reported that the expression of the tumor-suppressor miRNA-200c decreased the self-renewal ability of BCSCs in vitro and tumor formation ability in vivo [76].
Yu et al. showed that let-7 and miRNA-30e were downregulated in BCSCs and SK-3rd mammospheres compared to non-tumorigenic cells or more differentiated cells, respectively. Enforced expression of miRNA-30e inhibited mammosphere formation and tumorigenesis of SK-3rd cells in vitro and in vivo, respectively, by targeting ITGB3 and UBC9 [51]. The miRNA-128 was found to be downregulated in BCSCs (in CD44+, CD24−/low, and mammospheres) compared to non-cancerous cells [77]. Furthermore, miRNA-128 is significantly low in chemoresistant BTICs enriched from breast cancer cells (SK-3rd and MCF-7) and primary breast tumors, and that its target Bmi-1 and ABCC5 are overexpressed in these cancers. The ectopic expression of miRNA-128 sensitizes BTICs to the proapoptotic and DNA-damaging effects of doxorubicin, showing its therapeutic potential [77]. In another finding, miRNA-30c regulates invasion of breast cancer by targeting the cytoskeleton network genes encoding Twinfilin 1 (TWF1) and Vimentin (VIM), both of which regulate EMT [78].
The miRNA-34c expression level and function were reduced in the BTICs of MCF-7 and SK-3rd cells, and these cell lines were enriched for BTICs [79]. When ectopic expression of miRNA-34c decreased the self-renewal of BTICs, inhibited EMT, and suppressed migration of the tumor cells through inhibition of the target gene Notch4, miRNA-495 was significantly upregulated in CD44+/CD24−/low BCSCs, reflecting its potential importance in maintaining common BCSC properties, and its ectopic expression promoted colony formation in vitro and tumorigenesis in mice. In addition, the overexpression of miRNA-495 is associated with the lowered expression of E-cadherin, which enhances the stem-like phenotype in BCSCs [80]. miRNA-181 family members and miRNA-155 oncogenic miRNAs promote self-renewal, sphere formation, colony formation, or tumor development in breast cancer cells [81].

4.2. miRNA in Metastatic Breast Cancer

Metastatic breast cancer is a complex, multistep malignant process through which tumor cells migrate from their primary tumor (site of origin) to colonize distant tissues (e.g., liver, brain, bones, or lungs) and is often responsible for 90% of cancer-related mortality [82,83]. EMT is one of the important processes by which cancer metastasis starts, and EMT induces morphological changes in epithelial cells by which epithelial cells transform into mesenchymal cells. During this process, cancer cells lose their cell–cell communication and become mobile and invasive, spreading into distant organs and tissues [84]. Suppression of E-cadherin expression in epithelial cancer cells is a hallmark of EMT. Various miRNAs have been linked to the control of EMT in cancer, and several genes, including ZEB [85], Twist [86], Snail [87], and Slug [88] are known to restrict the expression of E-cadherin.
One of the most important miRNAs that control metastasis is the miRNA-200 family (miRNA-200a/200b/200c/141/429), and this prevents cell migration and invasion by targeting ZEB in several cancer types, including breast cancer [89], and miRNA-200/ZEB plays a central role in the EMT/MET processes. In the meantime, inhibition of miRNA-200 reduces the E-cadherin level while supporting VIM expression, thereby increasing cell motility [89]. Many studies also demonstrated that TGF-β-induced EMT might be inhibited only with the ectopic expression of the miRNA-200 family [90]. Xie et al. reported that miRNA-193a-WT1 interaction plays an important role in breast cancer metastasis and indicates that restoring miRNA-193a expression is a therapeutic strategy in breast cancer [91].
Ma and co-workers discovered that upregulation of miRNA-10b promotes invasion and metastasis by indirectly activating the pro-metastatic gene RHOC through repressing HOXD10 [25]. In addition, miRNA-373 regulates the CD44 gene, which promotes tumor invasion and metastasis [44]. miRNA-9, a MYC/MYCN-induced miRNA, has been demonstrated to directly target E-cadherin to promote breast cancer metastasis via activation of β-catenin signaling, through increasing tumor invasiveness and angiogenesis [92].
Recently, several studies have reported that Dicer, an endonuclease that processes miRNAs, is also associated with EMT and metastatic progression. Dicer inhibition promotes metastasis, and restoring its expression suppresses metastasis. Certain miRNAs also target Dicer for controlling metastasis. For example, miRNA-103/107 induced EMT by targeting Dicer expression [93]. In addition, the miRNA-221/222 cluster has been shown to induce EMT in breast cancer cells by targeting Dicer, and estrogen receptor 1 (ESR1) [94]. In breast cancer, miRNA-107 acts as an endogenous suppressor of let-7, and it interacts with mature let-7 to inhibit its function, which plays a major role in determining metastatic progression [95]. According to Huang et al., both in vitro and in vivo cancer cell migration and invasion were promoted by miRNA-373 and miRNA-520c [44]. Wu et al. reported that in breast cancer cells, miRNA-29a regulates the critical roles of EMT and metastasis by targeting SUV420H2 [96].
The overexpression of miRNA-17/92 is also involved in metastatic breast cancer [27]. miRNA-454-3p plays an important role in breast cancer’s early metastatic events, promotes the stemness of breast cancer cells, and promotes early distant relapse in both in vitro and in vivo conditions. The higher expression of miRNA-454-3p was found to be significantly associated with both a poor prognosis and early recurrence in breast cancer through Wnt/β-catenin signaling activation [97]. Higher expression of miRNA-373 was found in breast cancer samples from tumors exhibiting lymph node metastasis [44]. Dobson et al. identified a novel target of miRNA-30c, the nephroblastoma overexpressed gene (NOV), which is an inhibitor of the invasiveness of metastatic TNBC (MDA-MB-231) cells [98]. The miRNA-34a plays a key role in the proliferation, invasion, and metastasis of breast cancer cells [53]. Moreover, miRNA-373, miRNA-520c, miRNA-210, and miRNA-29b were also shown to influence the invasion and migration of breast cancer [44]. In addition, oncogenic miRNA-224 expression is significantly upregulated in highly invasive MDA-MB-231 cells and correlated with increased metastasis [99]. Another important oncogenic miRNA, miRNA-155, is frequently overexpressed in invasive breast cancer tissues and is directly targeted to RhoA and contributes to breast cancer metastasis. Inhibition of miRNA-155 suppressed TGF-β-induced EMT and tight junction dissolution, along with cell migration and invasion. Further, the ectopic expression of miRNA-155 reduced RhoA protein and disrupted tight junction formation [100]. miRNA-21 is an important oncogene and has a role in breast cancer tumorigenesis. It regulates invasion and tumor metastasis by targeting multiple tumor and metastasis suppressor genes. Also, inhibiting miRNA-21 reduced the invasion and lung metastasis of MDA-MB-231 cells [30]. All these studies demonstrated that some of the miRNAs have a metastatic role in breast cancer cells.

5. miRNA Expression in Breast Cancer Subtypes

Breast cancer comprises a diverse pathophysiology, with several molecular subtypes exhibiting varying clinical features, therapeutic responses, and prognoses. The immunohistochemical expression of hormone receptors such as the estrogen receptor (ER), progesterone receptor (PR), and human epidermal growth factor receptor 2 (HER2), is a widely accepted and commonly used classification system for breast cancer [101,102,103]. There are ER+/PR+, HER+, or triple negative (ER, PR, HER2, i.e., does not express these three receptors) clinical definitions for breast tumors. Further, breast cancer can be subtyped including Luminal A (ER+ PRhigh HER2 Ki-67low), Luminal B (ER+ PR+ HER2+ or HER2 Ki-67high), HER2-enriched (ER PR HER2+), and triple negative or basal-like (ER PR HER2) [104]. Dysregulation of miRNAs has been associated with breast cancer development at various stages. In addition to these markers, miRNAs have emerged as additional biomarkers that contribute to the molecular and functional diversity of breast cancers [22,23]. miRNAs in breast cancer are classified based on their expression patterns, functional roles, and clinical implications. Table 2 lists some miRNA biomarkers associated with breast cancer molecular subtypes. This classification allows us to stratify breast cancers according to their molecular subtype, which guides the diagnosis and treatment decisions and predicts prognosis. As far as the treatment of breast cancer is concerned, two main targeted therapies are commonly used: hormone therapy for hormone receptor-positive tumors [105] and anti-HER2 therapy for HER2+ tumors [106]. It is anticipated that future studies will identify miRNA biomarkers and their associations with specific subtypes of cancer. In breast cancer, miRNA profiling can improve molecular subtyping accuracy and guide better personalized treatment approaches.

6. miRNAs as Potential Diagnostic Biomarkers for Breast Cancer

In recent years, investigating biomarkers for early detection has rapidly expanded for better diagnoses and prognoses and the prediction of treatment responses in breast cancer. Using significant features, including stability, tissue specificity, ease of detection, and manipulation, will attract the clinician’s attention to achieve personalized breast cancer treatment [127]. miRNA are emerging, attractive, and promising biomarkers owing to their detectability and stability in the bloodstream. Usually, a significant increase in miRNA level is observed in cancer compared to controls; therefore, diagnostic and monitoring applications for miRNAs must be examined. Determining the expression ratios of genes and miRNAs has been a useful technique to improve diagnostic potential. The circulating miRNAs are widely known to be relatively stable, accessible, less invasive, easy to test, promising, and stage-specific biomarkers for the non-invasive diagnosis of breast cancer [128]. miRNAs are exceptionally stable in the bloodstream against the action of endogenous RNases, suggesting that miRNAs may serve as an effective blood-based biomarker for the detection and diagnosis of cancer [129]. miRNA expression has been observed to be higher in tumors than in normal tissues, leading us to hypothesize that global miRNA expression reflects the level of differentiation of cells. Tumor tissues release miRNAs into the bloodstream and other body fluids. miRNA levels may reflect abnormal conditions in breast cancer [130]. Therefore, miRNA status in the body fluids could be a novel, specific, and very sensitive blood-based diagnostic tool for the early detection of breast cancer. Cancer-specific miRNAs are shedding light on the molecular basis of cancer by being able to identify cancer-specific expression of miRNAs in the blood, which plays a vital role in the early-stage diagnosis of breast cancer. Currently, blood-based biomarkers are widely used for the early diagnosis of breast cancer and in response to treatment [130,131].
Using microarrays and other conventional methods, it is possible to detect differences in miRNA expression patterns between normal and cancerous tissue. Studies have shown that the profiles of miRNAs can be correlated with different types and grades of tumors, and that these profiles may be different from those of the normal breast tissue and cells [132]. Furthermore, exosome-derived miRNAs may influence the evolution of tumors by serving as messengers between tumor cells and non-tumor cells like immune and stromal cells. This phenomenon was extensively covered in a recent review on the emerging role of exosomes in breast cancer progression [133].
The detection of circulating miRNA can be a useful method for the non-invasive detection of breast cancer biomarkers. Because circulating miRNAs are very stable in clinical sources such as sputum, plasma, serum, urine, and saliva, they have been extensively used to classify breast cancer subtypes more accurately than circulatory cell-free DNA or RNA [134].
The miRNA-21 gene is an important regulator of breast carcinogenesis and is the most sensitive (87.6%) and specific (87.3%) biomarker for breast cancer diagnoses at early stages compared to other biomarkers like CEA and CA153 [135]. Canatan et al. reported that miRNA-21, miRNA-125, and are reliable candidates for circulating miRNA biomarkers for the detection of breast cancer [136]. On the other hand, the serum miRNA expression profile analysis using highly sensitive microarrays revealed five miRNA signatures (miRNA-1246, miRNA-1307-3p, miRNA-4634, miRNA-6861-5p, and miRNA-6876-5p) that can be used to detect early-stage breast cancer [137]. miRNA-195 and let-7a had significantly higher levels in the circulating blood of the breast cancer cohort than in the healthy controls. However, circulating levels of miRNA-195 and let-7a decreased in cancer patients following curative tumor resection [138]. Similarly, preoperative serum miRNA-20a and miRNA-21 expression levels were significantly higher in patients with breast cancer and benign disease than in healthy women. Serum miRNA-214 levels, on the other hand, could distinguish between benign and malignant tumors and healthy controls. In addition, in postoperative serum samples, miRNA-214 levels significantly decreased as compared to the preoperative sample [139].
Breast cancer patients with lymph node metastasis have high levels of miRNA-10b and miRNA-373 circulating in their blood, and their expression is associated with promoting the migration and invasion of breast cancer cells. Furthermore, such miRNAs may serve as viable biomarkers for detecting lymph node metastases in individuals with breast cancer [140]. Some studies revealed that groups of circulating miRNAs such as miRNA-299-5p, miRNA-411, miRNA-215, and miRNA-452 were differentially expressed in metastatic patients and increased the expression of miRNA-20a, miRNA-214, and miRNA-210 in lymph node-positive patient subgroups [139,141,142].
A few studies revealed significant high expression levels of circulatory miRNA-10b, miRNA-34a, miRNA-155, and miRNA-122 in breast cancer patients, which are associated with primary metastatic breast cancer [143]. Similarly, there were significantly higher expression levels of serum miRNA-21, miRNA-29a, miRNA-130b-5p, miRNA-145, miRNA-151a-5p, miRNA-206, miRNA-222-3p, and miRNA-451 in the breast cancer group than in the control group [144,145]. Cuk et al. also identified dysregulated miRNAs (miRNA-148b, miRNA-376c, miRNA-409-3p, and miRNA-801) in the plasma of 127 sporadic breast cancer patients and 80 healthy controls using RT-qPCR [146]. Other studies revealed some of the candidate biomarkers in plasma, such as miRNA-16, miRNA-21, miRNA-451, miRNA-409-3p, and miRNA-652, in the plasma of breast cancer patients [147]. But interestingly, these miRNAs were downregulated after surgery.

7. miRNA-Based Therapeutic Approaches to Breast Cancer Treatment

In the field of medicine, miRNAs have gained increasing interest in recent years due to their potential as therapeutic targets. Despite their promise, miRNA-based therapeutics face significant challenges, including issues related to delivery, off-target effects, and the need for rigorous clinical trials. miRNAs are currently being researched to unlock their full therapeutic potential.
For integrated cancer care, miRNA-based gene therapy provides an appealing anti-tumor approach. miRNA expression imbalance in cancer is related to tumorigenesis, so miRNA-based therapies are based on restoring miRNA function or inhibiting overexpressed miRNAs as reviewed earlier [148]. Recent experiments have shown that the phenotype in cancer cells can be reversed and the gene regulation network and signaling pathways can be restored by correcting miRNA changes with miRNA mimics or antagomiRs [149].
There are two different strategies and challenges to the clinical application of miRNAs. One strategy is designed to inhibit oncogenic miRNAs using miRNA antagonists, such as locked-nucleic acids (LNA), or antagomiRs, and is aimed at gaining function. To inhibit miRNA activity, small-molecule inhibitors specific to certain miRNAs are also being developed. The second strategy, miRNA replacement, includes the reintroduction of tumor-suppressor miRNA mimics to restore the loss of function [149,150,151,152].

7.1. miRNA Inhibition Therapy

By repressing oncogenic miRNAs, we can inhibit tumorigenesis, which is a promising cancer treatment strategy. Oncogenic miRNAs are often overexpressed in breast cancer and need to be suppressed to restore the normal expression of their target tumor-suppressor genes. In the recent past, multiple strategies have been developed to inhibit oncogenic miRNAs [153,154].
The decision to use miRNA-based therapies is because miRNA expression in tumor cells is altered and the tumor phenotype can be altered with miRNA expression modulation. miRNA inhibition therapy consists of the following agents: antisense anti-miRNA oligonucleotides (AMOs), LNA, antagomiRs, miRNA sponges, and miRNA small-molecule inhibitors (SMIRs) [155]. The fundamentals of these approaches are that endogenous miRNAs are isolated in an unrecognizable configuration, resulting in the inactivation and removal of mature miRNAs from the RISC [151,152]. Figure 2 shows the various miRNA inhibition therapies discussed here.

7.1.1. Anti-miRNA Oligonucleotides (AMOs)

AMOs are single stranded chemically modified anti-sense oligonucleotides with a length of 17 to 22 nt and complement the miRNA target of interest. The AMOs target and bind specifically to miRNAs, preventing them from interacting with the mRNAs they are targeting. Inhibiting the function of miRNAs can modulate gene expression and alleviate dysregulated pathways associated with diseases. A significant method for uncovering miRNA biology has been the antisense inhibition of miRNA action. This antisense oligonucleotide was designed to bind to and inhibit the interaction of that miRNA with its unique mRNA targets and the complementary mature miRNA, thus enabling normal translation. Unmodified AMOs, by comparison, are unable to inhibit in vitro miRNA function. Okumura et al. stated that modified AMOs targeting miRNA-21 (CL-miR21) interstrand cross-linked duplexes (CLDs) suppressed breast cancer cell proliferation with greater efficiency compared to other types of AMOs. Furthermore, the miRNA-21-controlled expression of tumor-suppressor genes was probably upregulated [156]. In the regulation of miRNAs, CLD-modified AMOs are substantially successful and could be a promising strategy for treating breast cancer. Clinical trials showed some progress for experimental AMO-based drugs directly targeting miRNA-122 in chronic hepatitis C patients (Santaris Pharma-developed Miravirsen and Regulus®-developed RG-101) and might be worth considering for therapy of breast cancer [157].

7.1.2. Locked Nucleic Acid (LNA)

LNA is a modified nucleotide with a methylene bridge connecting the 2′-oxygen and 4′-carbon of the ribose sugar ring. As a result, the oligonucleotide becomes more stable and binds better to complementary RNA sequences. LNA-modified AMOs specifically target and inhibit miRNAs. LNA is an example of a modified AMO that binds to miRNA, can modulate miRNA as antisense-based gene silencing, and has recently emerged as a possible miRNA-targeting therapeutic alternative [158]. These are very similar to RNAs in that the methylene bridge in the ribose ring binds the 2′-oxygen atom and the 4′-carbon atom (2′-O 4′-C methylene). This bridge forms a bi-cyclic structure that locks the conformation of the ribose and is integral to its complementary RNA sequences due to the high stability and low toxicity of the biological systems and the affinity of the LNA. In the regulation of RNAs, LNA oligonucleotides are stable, have greater aqueous solubility, and may not induce an immune response, indicating their utility in gene therapy for antisense-based gene silencing [28,158,159]. An in vivo study demonstrated that LNA-modified antisense oligonucleotides were capable of silencing upregulated miRNA-205-5p in breast cancer, leading to a significant reduction in tumor growth and metastatic spread in mouse models, thus suggesting the possible future use of this approach in therapy [160].

7.1.3. AntagomiRs

AntagomiRs, also known as anti-miRs, are another class of synthetic oligonucleotides that inhibit the activity of miRNAs, like AMOs. Chemically modified antagomiRs complement the targeted miRNAs with cholesterol-conjugated single-stranded 23-nt RNA molecules. Their backbone consists of single-stranded oligoribonucleotides with 2′-O-methyl (2′-O-Me) and partially modified phosphorothioate linkers [161]. The modifications were made to improve the RNA’s stability and protect it from degradation. The strategy of antagomiRs seems promising for suppressing miRNAs in therapeutic strategies.
The use of antisense miRNAs (antagomiRs) to knockdown miRNAs is one of the most common approaches. Ma et al. demonstrated that systemic treatment of tumor-bearing mice with miRNA-10b antagomiRs suppressed the metastasis of breast cancer, both in vitro and in vivo. This achieved significant reduction in the levels of miRNA-10b and increased levels of a functionally important miRNA-10b target, Hoxd10 [25]. In another study, the effect of miRNA-10b antagomiR in a 4T1 mouse model of mammary tumor metastasis was analyzed. AntagomiR-10b systemic delivery had a potent and highly specific metastasis-suppressing effect on these malignant breast cancer cells without affecting their ability to develop as primary tumors; in fact, antagomiR-10b prevented the spread of cancer cells from the primary tumor but did not influence the late stages of the metastatic phase after tumor cells had already disseminated [92]. AntagomiR-21 can affect breast cancer cells by inducing apoptosis and reducing cell proliferation [29].

7.1.4. miRNA Sponges

A miRNA sponge is an artificial RNA molecule which binds to miRNAs and prevents them from interacting with their target mRNAs. “miRNA sponges” or “miRNA decoys” offer many synthetic miRNA binding sites which are compete with natural miRNA targets for miRNA binding [162]. The vectors of expression constitute the source of the miRNA sponge transcription, thus reducing the effects of miRNA and increasing the expression of the native targets of miRNA [162]. Because these compounds strongly inhibit target miRNAs and have a high specificity for miRNA sponges, they have been used more frequently in miRNA loss-of-function studies [163]. In one study, a sponge plasmid against miRNA-10b was transiently transfected into MDA-MB-231 and MCF-7, high and low metastatic cell lines. This resulted in the miRNA-10b sponge effectively inhibiting the growth and proliferation of these cell lines [164]. Furthermore, miRNA-21 sponges have been effectively applied to the MDA-MB-231 and MCF-7 breast cancer cell lines [165], while miRNA-9 sponges have been applied to the 4T1 metastatic breast cancer cell line, resulting in a nearly 50% reduction in metastatic activity [92]. Clinically, miRNA sponges and decoys have been developed for more stable suppression and targeted delivery of miRNAs.

7.1.5. miRNA Small-Molecule Inhibitors (SMIRs)

A small-molecule inhibitor of miRNA is a compound designed to modulate the activity of miRNAs. This differs from traditional oligonucleotide-based approaches, which include AMOS and miRNA sponges [166,167]. Due to the decreased time of development, acceptance, and cost, inhibition therapy using SMIRs is an encouraging one. Instead of inhibiting target recognition from miRNA-21, the small molecule’s mode of action is primarily through the transcriptional regulation of miRNA-21. The small molecule enoxacin, an antibacterial fluoroquinolone, has been documented to bind to the miRNA biosynthesis protein TRBP and to increase tumor-suppressor miRNA output [168]. Enoxacin was also shown to inhibit MDA-MB-231 and MCF-7 cell growth [168]. Linifanib, a multi-tyrosine kinase inhibitor, could substantially inhibit miRNA-10b and reverse its oncogenic function in both in vitro and in vivo breast and liver cancers [169]. A novel strategy for restoring dysregulated miRNAs in cancer is thus defined using small-molecule modulators of miRNAs [153]. Overall, targeting miRNAs for cancer care with SMIRs is an evidence-based approach with high potential and ability for success.

7.2. miRNA Replacement Therapy

Synthetic miRNA mimics are introduced into cells to restore or enhance the function of a specific miRNA that is deficient or downregulated in breast cancer. By mimicking the function of endogenous miRNAs, this therapeutic approach regulates gene expression and cellular pathways. Synthetic miRNA mimics must be delivered efficiently for target cells for miRNA replacement therapy to be effective. There are a variety of delivery systems that can be used to ensure effective uptake by target tissues, including viral vectors, lipid nanoparticles, and other nanocarriers [150,151,152,170]. A synthetic miRNA mimic is incorporated into the RISC and acts similarly to endogenous miRNAs once inside a cell. Once they bind to target mRNAs, they inhibit the translation or induce the degradation of mRNA, thus regulating gene expression (Figure 3). As a potential treatment strategy for breast cancer, miRNA replacement therapy is being actively explored.
The effects of these therapies may include the suppression of cancer cell growth, modulation of immune responses, and restoration of tissue homeostasis. Tumor cell proliferation can be inhibited, or apoptosis can be induced using miRNA restoration therapy by restoring exogenous tumor-suppressor miRNAs that are downregulated in tumor cells [153]. Different studies have shown the efficacy of in vitro and in vivo miRNA restoration therapies. For example, Liang et al. introduced miRNA replacement in radioresistant breast cancer cells. For the first time, they showed that miRNA-302a’s enforced expression effectively sensitizes radioresistant tumor cells to irradiation by directly downregulating both AKT1 and RAD52 expression [171]. These results suggested that decreased miRNA-302 expression confers radioresistance, and the miRNA-302 baseline expression restoration sensitizes breast cancer cells to radiotherapy.
When miRNA mimics were used to force the expression of miRNA-365 and miRNA-22, breast cancer cell proliferation was inhibited and sensitivity to paclitaxel and fluorouracil was boosted, respectively. By targeting GALNT4, miRNA-365 works on overcoming chemoresistance [172]. In comparison, Jiang et al. found that in breast cancer cells and tissues, miRNA-148a was downregulated, and its overexpression mimics the reduced migration and invasion of breast cancer cells [58]. Also, in breast cancer cells, miRNA-33b expression was downregulated, and it had a negative correlation with the lymph node metastatic status of breast cancer patients. Ectopic overexpression of miRNA-33b inhibited lung metastasis in vivo and compromised stem cell characteristics, migration, and invasion in vitro in highly metastatic breast cancer cells [52]. MRX34 was involved in the first phase I clinical trial for miRNA replacement therapy and was intended to restore miRNA-34 expression in patients with different solid tumors including TNBC [173]. The use of miRNA-mimetic agents in patients is a promising new way of treating clinical breast cancer.

8. Conclusions and Future Directions

To conclude, being able to understand how differentially regulated miRNAs influence breast cancer progression provides valuable insight into their potential diagnostic and therapeutic applications. Unraveling the intricate web of miRNA functions offers valuable insights into their utility as both diagnostic tools and therapeutic targets for breast cancer. By identifying stage- and subtype-specific miRNA signatures, early-detection strategies can be improved and personalized therapeutic interventions can be implemented. Moreover, it is important to determine which miRNA or group of miRNAs is most dysregulated in breast cancer, specifically, at different stages of the disease. Such signatures could serve as robust biomarkers, facilitating more accurate and timely diagnoses, which is crucial for improving patient outcomes. This will assist in identifying and prioritizing the most promising treatment targets, with an emphasis on developing early-detection and -treatment strategies for breast cancer. Better understanding of miRNA-guided networks is essential for improving breast cancer diagnoses and treatment. Ongoing research in this field holds the promise of translating miRNA-based discoveries into clinically relevant applications, fostering advancements in breast cancer management.

Author Contributions

Conceptualization, K.S. and R.S.; investigation, K.S. and R.S.; resources, K.S.; writing—original draft preparation, K.S.; writing—review and editing, K.S. and R.S.; visualization, K.S.; supervision, R.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lewandowska, A.; Rudzki, M.; Rudzki, S.; Lewandowski, T.; Laskowska, B. Environmental Risk Factors for Cancer—Review Paper. Ann. Agric. Environ. Med. 2019, 26, 1–7. [Google Scholar] [CrossRef] [PubMed]
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  3. Lakhani, S.R.; Ellis, I.O.; Schnitt, S.; Tan, P.H.; van de Vijver, M. (Eds.) WHO Classification of Tumours of the Breast; IARC: Lyon, France, 2012. [Google Scholar]
  4. Ferlay, J.; Ervik, M.; Lam, F.; Laversanne, M.; Colombet, M.; Mery, L.; Piñeros, M.; Znaor, A.; Soerjomataram, I.; Bray, F. Global Cancer Observatory: Cancer Today; International Agency for Research on Cancer: Lyon, France, 2024; Available online: https://gco.iarc.who.int/today (accessed on 12 December 2023).
  5. Nguyen, D.X.; Bos, P.D.; Massagué, J. Metastasis: From Dissemination to Organ-Specific Colonization. Nat. Rev. Cancer 2009, 9, 274–284. [Google Scholar] [CrossRef] [PubMed]
  6. Cardoso, F.; Costa, A.; Norton, L.; Cameron, D.; Cufer, T.; Fallowfield, L.; Francis, P.; Gligorov, J.; Kyriakides, S.; Lin, N.; et al. 1st International Consensus Guidelines for Advanced Breast Cancer (ABC 1). Breast 2012, 21, 242–252. [Google Scholar] [CrossRef] [PubMed]
  7. Yedjou, C.G.; Sims, J.N.; Miele, L.; Noubissi, F.; Lowe, L.; Fonseca, D.D.; Alo, R.A.; Payton, M.; Tchounwou, P.B. Health and Racial Disparity in Breast Cancer. Adv. Exp. Med. Biol. 2019, 1152, 31–49. [Google Scholar] [CrossRef] [PubMed]
  8. Bertoli, G.; Cava, C.; Castiglioni, I. MicroRNAs: New Biomarkers for Diagnosis, Prognosis, Therapy Prediction and Therapeutic Tools for Breast Cancer. Theranostics 2015, 5, 1122–1143. [Google Scholar] [CrossRef]
  9. Rohan, T.; Ye, K.; Wang, Y.; Glass, A.G.; Ginsberg, M.; Loudig, O. MicroRNA expression in benign breast tissue and risk of subsequent invasive breast cancer. PLoS ONE 2018, 13, e0191814. [Google Scholar] [CrossRef]
  10. Lee, R.C.; Feinbaum, R.L.; Ambros, V. The C. elegans Heterochronic Gene Lin-4 Encodes Small RNAs with Antisense Complementarity to Lin-14. Cell 1993, 75, 843–854. [Google Scholar] [CrossRef]
  11. Khordadmehr, M.; Shahbazi, R.; Ezzati, H.; Jigari-Asl, F.; Sadreddini, S.; Baradaran, B. Key MicroRNAs in the Biology of Breast Cancer; Emerging Evidence in the Last Decade. J. Cell Physiol. 2019, 234, 8316–8326. [Google Scholar] [CrossRef]
  12. Weber, J.A.; Baxter, D.H.; Zhang, S.; Huang, D.Y.; Huang, K.H.; Lee, M.J.; Galas, D.J.; Wang, K. The microRNA spectrum in 12 body fluids. Clin. Chem. 2010, 56, 1733–1741. [Google Scholar] [CrossRef]
  13. Kumar, S.; Keerthana, R.; Pazhanimuthu, A.; Perumal, P. Overexpression of Circulating MiRNA-21 and MiRNA-146a in Plasma Samples of Breast Cancer Patients. Indian J. Biochem. Biophys. 2013, 50, 210–214. [Google Scholar]
  14. Swellam, M.; El Magdoub, H.M.; Hassan, N.M.; Hefny, M.M.; Sobeih, M.E. Potential diagnostic role of circulating MiRNAs in breast cancer: Implications on clinicopathological characters. Clin. Biochem. 2018, 56, 47–54. [Google Scholar] [CrossRef]
  15. Savari, B.; Boozarpour, S.; Tahmasebi-Birgani, M.; Sabouri, H.; Hosseini, S.M. Overex-pression of MicroRNA-21 in the Serum of Breast Cancer Patients. MicorRNA 2020, 9, 58–63. [Google Scholar] [CrossRef]
  16. Clancy, J.W.; D’Souza-Schorey, C. Tumor-Derived Extracellular Vesicles: Multifunctional Entities in the Tumor Microenvironment. Annu. Rev. Pathol. Mech. Dis. 2023, 18, 205–229. [Google Scholar] [CrossRef]
  17. O’Brien, J.; Hayder, H.; Zayed, Y.; Peng, C. Overview of MicroRNA Biogenesis, Mechanisms of Actions, and Circulation. Front. Endocrinol. 2018, 9, 402. [Google Scholar] [CrossRef]
  18. Ghamlouche, F.; Yehya, A.; Zeid, Y.; Fakhereddine, H.; Fawaz, J.; Liu, Y.-N.; Al-Sayegh, M.; Abou-Kheir, W. MicroRNAs as Clinical Tools for Diagnosis, Prognosis, and Therapy in Prostate Cancer. Transl. Oncol. 2023, 28, 101613. [Google Scholar] [CrossRef] [PubMed]
  19. Riolo, G.; Cantara, S.; Marzocchi, C.; Ricci, C. miRNA Targets: From Prediction Tools to Experimental Validation. Methods Protoc. 2020, 4, 1. [Google Scholar] [CrossRef] [PubMed]
  20. Saçar, M.D.; Hamzeiy, H.; Allmer, J. Can MiRBase provide positive data for machine learning for the detection of MiRNA hairpins? J. Integr. Bioinform. 2013, 10, 215. [Google Scholar] [CrossRef]
  21. Skoufos, G.; Kakoulidis, P.; Tastsoglou, S.; Zacharopoulou, E.; Kotsira, V.; Miliotis, M.; Mavromati, G.; Grigoriadis, D.; Zioga, M.; Velli, A.; et al. TarBase-v9.0 extends experimentally supported miRNA-gene interactions to cell-types and virally encoded miRNAs. Nucleic Acids Res. 2024, 52, D304–D310. [Google Scholar] [CrossRef]
  22. Loh, H.-Y.; Norman, B.P.; Lai, K.-S.; Rahman, N.M.A.N.A.; Alitheen, N.B.M.; Osman, M.A. The Regulatory Role of MicroRNAs in Breast Cancer. Int. J. Mol. Sci. 2019, 20, 4940. [Google Scholar] [CrossRef]
  23. Richard, V.; Davey, M.G.; Annuk, H.; Miller, N.; Dwyer, R.M.; Lowery, A.; Kerin, M.J. MicroRNAs in Molecular Classification and Pathogenesis of Breast Tumors. Cancers 2021, 13, 5332. [Google Scholar] [CrossRef] [PubMed]
  24. Peng, Y.; Croce, C.M. The Role of MicroRNAs in Human Cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef] [PubMed]
  25. Ma, L.; Teruya-Feldstein, J.; Weinberg, R.A. Tumour Invasion and Metastasis Initiated by MicroRNA-10b in Breast Cancer. Nature 2007, 449, 682–688. [Google Scholar] [CrossRef] [PubMed]
  26. Jin, L.; Lim, M.; Zhao, S.; Sano, Y.; Simone, B.A.; Savage, J.E.; Wickstrom, E.; Camphausen, K.; Pestell, R.G.; Simone, N.L. The Metastatic Potential of Triple-Negative Breast Cancer Is Decreased via Caloric Restriction-Mediated Reduction of the MiR-17~92 Cluster. Breast Cancer Res. Treat. 2014, 146, 41–50. [Google Scholar] [CrossRef] [PubMed]
  27. Hossain, M.M.; Sultana, A.; Barua, D.; Islam, M.N.; Gupta, A.; Gupta, S. Differential Expression, Function and Prognostic Value of MiR-17–92 Cluster in ER-Positive and Triple-Negative Breast Cancer. Cancer Treat. Res. Commun. 2020, 25, 100224. [Google Scholar] [CrossRef]
  28. Frankel, L.B.; Christoffersen, N.R.; Jacobsen, A.; Lindow, M.; Krogh, A.; Lund, A.H. Programmed Cell Death 4 (PDCD4) Is an Important Functional Target of the MicroRNA MiR-21 in Breast Cancer Cells. J. Biol. Chem. 2008, 283, 1026–1033. [Google Scholar] [CrossRef]
  29. Zhao, D.; Tu, Y.; Wan, L.; Bu, L.; Huang, T.; Sun, X.; Wang, K.; Shen, B. 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, e71472. [Google Scholar] [CrossRef]
  30. Zhu, S.; Wu, H.; Wu, F.; Nie, D.; Sheng, S.; Mo, Y.-Y. MicroRNA-21 Targets Tumor Suppressor Genes in Invasion and Metastasis. Cell Res. 2008, 18, 350–359. [Google Scholar] [CrossRef]
  31. Song, B.; Wang, C.; Liu, J.; Wang, X.; Lv, L.; Wei, L.; Xie, L.; Zheng, Y.; Song, X. MicroRNA-21 regulates breast cancer invasion partly by targeting tissue inhibitor of metalloproteinase 3 expression. J. Exp. Clin. Cancer Res. 2010, 29, 29. [Google Scholar] [CrossRef]
  32. Roscigno, G.; Puoti, I.; Giordano, I.; Donnarumma, E.; Russo, V.; Affinito, A.; Adamo, A.; Quintavalle, C.; Todaro, M.; Vivanco, M.D.; et al. MiR-24 Induces Chemotherapy Resistance and Hypoxic Advantage in Breast Cancer. Oncotarget 2017, 8, 19507–19521. [Google Scholar] [CrossRef]
  33. Lei, P.; Wang, W.; Sheldon, M.; Sun, Y.; Yao, F.; Ma, L. Role of Glucose Metabolic Reprogramming in Breast Cancer Progression and Drug Resistance. Cancers 2023, 15, 3390. [Google Scholar] [CrossRef]
  34. Hua, K.; Jin, J.; Zhao, J.; Song, J.; Song, H.; Li, D.; Maskey, N.; Zhao, B.; Wu, C.; Xu, H.; et al. MiR-135b, Upregulated in Breast Cancer, Promotes Cell Growth and Disrupts the Cell Cycle by Regulating LATS2. Int. J. Oncol. 2016, 48, 1997–2006. [Google Scholar] [CrossRef]
  35. Jiang, S.; Zhang, H.-W.; Lu, M.-H.; He, X.-H.; Li, Y.; Gu, H.; Liu, M.-F.; Wang, E.-D. MicroRNA-155 Functions as an OncomiR in Breast Cancer by Targeting the Suppressor of Cytokine Signaling 1 Gene. Cancer Res. 2010, 70, 3119–3127. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, C.-M.; Zhao, J.; Deng, H.-Y. MiR-155 Promotes Proliferation of Human Breast Cancer MCF-7 Cells through Targeting Tumor Protein 53-Induced Nuclear Protein 1. J. Biomed. Sci. 2013, 20, 79. [Google Scholar] [CrossRef] [PubMed]
  37. Kong, W.; He, L.; Coppola, M.; Guo, J.; Esposito, N.N.; Coppola, D.; Cheng, J.Q. MicroRNA-155 Regulates Cell Survival, Growth, and Chemosensitivity by Targeting FOXO3a in Breast Cancer. J. Biol. Chem. 2010, 285, 17869–17879. [Google Scholar] [CrossRef] [PubMed]
  38. Taylor, M.A.; Sossey-Alaoui, K.; Thompson, C.L.; Danielpour, D.; Schiemann, W.P. TGF-β Upregulates MiR-181a Expression to Promote Breast Cancer Metastasis. J. Clin. Investig. 2013, 123, 150–163. [Google Scholar] [CrossRef]
  39. Sharma, S.; Nagpal, N.; Ghosh, P.C.; Kulshreshtha, R. P53-MiR-191-SOX4 Regulatory Loop Affects Apoptosis in Breast Cancer. RNA 2017, 23, 1237–1246. [Google Scholar] [CrossRef] [PubMed]
  40. Hong, H.; Yu, H.; Yuan, J.; Guo, C.; Cao, H.; Li, W.; Xiao, C. MicroRNA-200b Impacts Breast Cancer Cell Migration and Invasion by Regulating Ezrin-Radixin-Moesin. Med. Sci. Monit. 2016, 22, 1946–1952. [Google Scholar] [CrossRef]
  41. Zhou, Y.; Wang, M.; Tong, Y.; Liu, X.; Zhang, L.; Dong, D.; Shao, J.; Zhou, Y. miR-206 Promotes Cancer Progression by Targeting Full-Length Neurokinin-1 Receptor in Breast Cancer. Technol. Cancer Res. Treat. 2019, 18, 1533033819875168. [Google Scholar] [CrossRef]
  42. Harquail, J.; LeBlanc, N.; Ouellette, R.J.; Robichaud, G.A. MiRNAs 484 and 210 Regulate Pax-5 Expression and Function in Breast Cancer Cells. Carcinogenesis 2019, 40, 1010–1020. [Google Scholar] [CrossRef]
  43. McAnena, P.; Tanriverdi, K.; Curran, C.; Gilligan, K.; Freedman, J.E.; Brown, J.A.L.; Kerin, M.J. Circulating MicroRNAs MiR-331 and MiR-195 Differentiate Local Luminal a from Metastatic Breast Cancer. BMC Cancer 2019, 19, 436. [Google Scholar] [CrossRef] [PubMed]
  44. Huang, Q.; Gumireddy, K.; Schrier, M.; le Sage, C.; Nagel, R.; Nair, S.; Egan, D.A.; Li, A.; Huang, G.; Klein-Szanto, A.J.; et al. The MicroRNAs MiR-373 and MiR-520c Promote Tumour Invasion and Metastasis. Nat. Cell Biol. 2008, 10, 202–210. [Google Scholar] [CrossRef]
  45. Li, Z.; Meng, Q.; Pan, A.; Wu, X.; Cui, J.; Wang, Y.; Li, L. MicroRNA-455-3p Promotes Invasion and Migration in Triple Negative Breast Cancer by Targeting Tumor Suppressor EI24. Oncotarget 2017, 8, 19455–19466. [Google Scholar] [CrossRef]
  46. Matamala, N.; Vargas, M.T.; González-Cámpora, R.; Arias, J.I.; Menéndez, P.; Andrés-León, E.; Yanowsky, K.; Llaneza-Folgueras, A.; Miñambres, R.; Martínez-Delgado, B.; et al. MicroRNA Deregulation in Triple Negative Breast Cancer Reveals a Role of MiR-498 in Regulating BRCA1 Expression. Oncotarget 2016, 7, 20068–20079. [Google Scholar] [CrossRef]
  47. Jiang, H.; Wang, P.; Li, X.; Wang, Q.; Deng, Z.-B.; Zhuang, X.; Mu, J.; Zhang, L.; Wang, B.; Yan, J.; et al. Restoration of MiR17/20a in Solid Tumor Cells Enhances the Natural Killer Cell Antitumor Activity by Targeting Mekk2. Cancer Immunol. Res. 2014, 2, 789–799. [Google Scholar] [CrossRef]
  48. Yu, Z.; Wang, C.; Wang, M.; Li, Z.; Casimiro, M.C.; Liu, M.; Wu, K.; Whittle, J.; Ju, X.; Hyslop, T.; et al. A Cyclin D1/MicroRNA 17/20 Regulatory Feedback Loop in Control of Breast Cancer Cell Proliferation. J. Cell Biol. 2008, 182, 509–517. [Google Scholar] [CrossRef]
  49. Zhang, H.; Cai, K.; Wang, J.; Wang, X.; Cheng, K.; Shi, F.; Jiang, L.; Zhang, Y.; Dou, J. MiR-7, Inhibited Indirectly by LincRNA HOTAIR, Directly Inhibits SETDB1 and Reverses the EMT of Breast Cancer Stem Cells by Downregulating the STAT3 Pathway. Stem Cells 2014, 32, 2858–2868. [Google Scholar] [CrossRef] [PubMed]
  50. Sun, H.; Ding, C.; Zhang, H.; Gao, J. Let-7 MiRNAs Sensitize Breast Cancer Stem Cells to Radiation-Induced Repression through Inhibition of the Cyclin D1/Akt1/Wnt1 Signaling Pathway. Mol. Med. Rep. 2016, 14, 3285–3292. [Google Scholar] [CrossRef]
  51. Yu, F.; Deng, H.; Yao, H.; Liu, Q.; Su, F.; Song, E. Mir-30 Reduction Maintains Self-Renewal and Inhibits Apoptosis in Breast Tumor-Initiating Cells. Oncogene 2010, 29, 4194–4204. [Google Scholar] [CrossRef] [PubMed]
  52. Lin, Y.; Liu, A.Y.; Fan, C.; Zheng, H.; Li, Y.; Zhang, C.; Wu, S.; Yu, D.; Huang, Z.; Liu, F.; et al. MicroRNA-33b Inhibits Breast Cancer Metastasis by Targeting HMGA2, SALL4 and Twist1. Sci. Rep. 2015, 5, 9995. [Google Scholar] [CrossRef]
  53. Huang, Q.-D.; Zheng, S.-R.; Cai, Y.-J.; Chen, D.-L.; Shen, Y.-Y. IMP3 Promotes TNBC Stem Cell Property through MiRNA-34a Regulation. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 2688–2696. [Google Scholar] [CrossRef]
  54. Ferracin, M.; Bassi, C.; Pedriali, M.; Pagotto, S.; D’Abundo, L.; Zagatti, B.; Corrà, F.; Musa, G.; Callegari, E.; Lupini, L.; et al. MiR-125b Targets Erythropoietin and Its Receptor and Their Expression Correlates with Metastatic Potential and ERBB2/HER2 Expression. Mol. Cancer 2013, 12, 130. [Google Scholar] [CrossRef]
  55. Feliciano, A.; Castellvi, J.; Artero-Castro, A.; Leal, J.A.; Romagosa, C.; Hernández-Losa, J.; Peg, V.; Fabra, A.; Vidal, F.; Kondoh, H.; et al. MiR-125b Acts as a Tumor Suppressor in Breast Tumorigenesis via Its Novel Direct Targets ENPEP, CK2-α, CCNJ, and MEGF9. PLoS ONE 2013, 8, e76247. [Google Scholar] [CrossRef]
  56. Chen, F.; Luo, N.; Hu, Y.; Li, X.; Zhang, K. MiR-137 Suppresses Triple-Negative Breast Cancer Stemness and Tumorigenesis by Perturbing BCL11A-DNMT1 Interaction. Cell. Phys. Biochem. 2018, 47, 2147–2158. [Google Scholar] [CrossRef]
  57. Zhou, L.L.; Dong, J.L.; Huang, G.; Sun, Z.L.; Wu, J. MicroRNA-143 Inhibits Cell Growth by Targeting ERK5 and MAP3K7 in Breast Cancer. Braz. J. Med. Biol. Res. 2017, 50, e5891. [Google Scholar] [CrossRef]
  58. Jiang, Q.; He, M.; Ma, M.-T.; Wu, H.-Z.; Yu, Z.-J.; Guan, S.; Jiang, L.-Y.; Wang, Y.; Zheng, D.-D.; Jin, F.; et al. MicroRNA-148a Inhibits Breast Cancer Migration and Invasion by Directly Targeting WNT-1. Oncol. Rep. 2016, 35, 1425–1432. [Google Scholar] [CrossRef]
  59. Xue, J.; Chen, Z.; Gu, X.; Zhang, Y.; Zhang, W. MicroRNA-148a inhibits migration of breast cancer cells by targeting MMP-13. Tumour Biol. 2016, 37, 1581–1590. [Google Scholar] [CrossRef]
  60. Yao, J.; Zhou, E.; Wang, Y.; Xu, F.; Zhang, D.; Zhong, D. microRNA-200a inhibits cell proliferation by targeting mitochondrial transcription factor A in breast cancer. DNA Cell Biol. 2014, 33, 291–300. [Google Scholar] [CrossRef]
  61. DeCastro, A.J.; Dunphy, K.A.; Hutchinson, J.; Balboni, A.L.; Cherukuri, P.; Jerry, D.J.; DiRenzo, J. MiR203 Mediates Subversion of Stem Cell Properties during Mammary Epithelial Differentiation via Repression of ΔNP63α and Promotes Mesenchymal-to-Epithelial Transition. Cell Death Dis. 2013, 4, e514. [Google Scholar] [CrossRef]
  62. Li, J.; Peng, S.; Zou, X.; Geng, X.; Wang, T.; Zhu, W.; Xia, T. Value of Negatively Correlated MiR-205-5p/HMGB3 and MiR-96-5p/FOXO1 on the Diagnosis of Breast Cancer and Benign Breast Diseases. Cancer Pathog. Ther. 2023, 1, 159–167. [Google Scholar] [CrossRef]
  63. Chao, C.-H.; Chang, C.-C.; Wu, M.-J.; Ko, H.-W.; Wang, D.; Hung, M.-C.; Yang, J.-Y.; Chang, C.-J. MicroRNA-205 Signaling Regulates Mammary Stem Cell Fate and Tumorigenesis. J. Clin. Investig. 2014, 124, 3093–3106. [Google Scholar] [CrossRef]
  64. Zhou, J.; Tian, Y.; Li, J.; Lu, B.; Sun, M.; Zou, Y.; Kong, R.; Luo, Y.; Shi, Y.; Wang, K.; et al. MiR-206 Is down-Regulated in Breast Cancer and Inhibits Cell Proliferation through the up-Regulation of CyclinD2. Biochem. Biophys. Res. Commun. 2013, 433, 207–212. [Google Scholar] [CrossRef]
  65. Lin, Z.-J.; Ming, J.; Yang, L.; Du, J.-Z.; Wang, N.; Luo, H.-J. Mechanism of Regulatory Effect of MicroRNA-206 on Connexin 43 in Distant Metastasis of Breast Cancer. Chin. Med. J. 2016, 129, 424–434. [Google Scholar] [CrossRef]
  66. Sun, X.; Li, Y.; Zheng, M.; Zuo, W.; Zheng, W. MicroRNA-223 Increases the Sensitivity of Triple-Negative Breast Cancer Stem Cells to TRAIL-Induced Apoptosis by Targeting HAX-1. PLoS ONE 2016, 11, e0162754. [Google Scholar] [CrossRef]
  67. Huang, X.; Lyu, J. Tumor Suppressor Function of MiR-483-3p on Breast Cancer via Targeting of the Cyclin E1 Gene. Exp. Ther. Med. 2018, 16, 2615–2620. [Google Scholar] [CrossRef]
  68. Luo, Q.; Li, X.; Gao, Y.; Long, Y.; Chen, L.; Huang, Y.; Fang, L. MiRNA-497 Regulates Cell Growth and Invasion by Targeting Cyclin E1 in Breast Cancer. Cancer Cell Int. 2013, 13, 95. [Google Scholar] [CrossRef]
  69. Li, D.; Song, H.; Wu, T.; Xie, D.; Hu, J.; Zhao, J.; Shen, Q.; Fang, L. MiR-519d-3p Suppresses Breast Cancer Cell Growth and Motility via Targeting LIM Domain Kinase 1. Mol. Cell Biochem. 2018, 444, 169–178. [Google Scholar] [CrossRef]
  70. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  71. Polyak, K. Breast Cancer: Origins and Evolution. J. Clin. Investig. 2007, 117, 3155–3163. [Google Scholar] [CrossRef]
  72. Sever, R.; Brugge, J.S. Signal Transduction in Cancer. Cold Spring Harb. Perspect. Med. 2015, 5, a006098. [Google Scholar] [CrossRef]
  73. Mai, Y.; Su, J.; Yang, C.; Xia, C.; Fu, L. The Strategies to Cure Cancer Patients by Eradicating Cancer Stem-like Cells. Mol. Cancer 2023, 22, 171. [Google Scholar] [CrossRef]
  74. Liu, H.; Patel, M.R.; Prescher, J.A.; Patsialou, A.; Qian, D.; Lin, J.; Wen, S.; Chang, Y.-F.; Bachmann, M.H.; Shimono, Y.; et al. Cancer Stem Cells from Human Breast Tumors Are Involved in Spontaneous Metastases in Orthotopic Mouse Models. Proc. Natl. Acad. Sci. USA 2010, 107, 18115–18120. [Google Scholar] [CrossRef]
  75. Chang, C.-J.; Yang, J.-Y.; Xia, W.; Chen, C.-T.; Xie, X.; Chao, C.-H.; Woodward, W.A.; Hsu, J.-M.; Hortobagyi, G.N.; Hung, M.-C. EZH2 Promotes Expansion of Breast Tumor Initiating Cells through Activation of RAF1-β-Catenin Signaling. Cancer Cell 2011, 19, 86–100. [Google Scholar] [CrossRef]
  76. Shimono, Y.; Zabala, M.; Cho, R.W.; Lobo, N.; Dalerba, P.; Qian, D.; Diehn, M.; Liu, H.; Panula, S.P.; Chiao, E.; et al. Downregulation of MiRNA-200c Links Breast Cancer Stem Cells with Normal Stem Cells. Cell 2009, 138, 592–603. [Google Scholar] [CrossRef]
  77. Zhu, Y.; Yu, F.; Jiao, Y.; Feng, J.; Tang, W.; Yao, H.; Gong, C.; Chen, J.; Su, F.; Zhang, Y.; et al. Reduced MiR-128 in Breast Tumor–Initiating Cells Induces Chemotherapeutic Resistance via Bmi-1 and ABCC5. Clin. Cancer Res. 2011, 17, 7105–7115. [Google Scholar] [CrossRef]
  78. Bockhorn, J.; Yee, K.; Chang, Y.-F.; Prat, A.; Huo, D.; Nwachukwu, C.; Dalton, R.; Huang, S.; Swanson, K.E.; Perou, C.M.; et al. MicroRNA-30c Targets Cytoskeleton Genes Involved in Breast Cancer Cell Invasion. Breast Cancer Res. Treat. 2013, 137, 373–382. [Google Scholar] [CrossRef]
  79. Yu, F.; Jiao, Y.; Zhu, Y.; Wang, Y.; Zhu, J.; Cui, X.; Liu, Y.; He, Y.; Park, E.-Y.; Zhang, H.; et al. MicroRNA 34c Gene Down-Regulation via DNA Methylation Promotes Self-Renewal and Epithelial-Mesenchymal Transition in Breast Tumor-Initiating Cells. J. Biol. Chem. 2012, 287, 465–473. [Google Scholar] [CrossRef] [PubMed]
  80. Hwang-Verslues, W.W.; Chang, P.-H.; Wei, P.-C.; Yang, C.-Y.; Huang, C.-K.; Kuo, W.-H.; Shew, J.-Y.; Chang, K.-J.; Lee, E.Y.-H.P.; Lee, W.-H. MiR-495 Is Upregulated by E12/E47 in Breast Cancer Stem Cells, and Promotes Oncogenesis and Hypoxia Resistance via Downregulation of E-Cadherin and REDD1. Oncogene 2011, 30, 2463–2474. [Google Scholar] [CrossRef] [PubMed]
  81. Wang, Y.; Yu, Y.; Tsuyada, A.; Ren, X.; Wu, X.; Stubblefield, K.; Rankin-Gee, E.K.; Wang, S.E. Transforming Growth Factor-β Regulates the Sphere-Initiating Stem Cell-like Feature in Breast Cancer through MiRNA-181 and ATM. Oncogene 2011, 30, 1470–1480. [Google Scholar] [CrossRef] [PubMed]
  82. Riggio, A.I.; Varley, K.E.; Welm, A.L. The Lingering Mysteries of Metastatic Recurrence in Breast Cancer. Br. J. Cancer 2021, 124, 13–26. [Google Scholar] [CrossRef] [PubMed]
  83. Baumann, Z.; Auf der Maur, P.; Bentires-Alj, M. Feed-forward Loops between Metastatic Cancer Cells and Their Microenvironment—The Stage of Escalation. EMBO Mol. Med. 2022, 14, e14283. [Google Scholar] [CrossRef]
  84. Huang, Y.; Hong, W.; Wei, X. The Molecular Mechanisms and Therapeutic Strategies of EMT in Tumor Progression and Metastasis. J. Hematol. Oncol. 2022, 15, 129. [Google Scholar] [CrossRef]
  85. Chua, H.L.; Bhat-Nakshatri, P.; Clare, S.E.; Morimiya, A.; Badve, S.; Nakshatri, H. NF-kappaB represses E-cadherin expression and enhances epithelial to mesenchymal transition of mammary epithelial cells: Potential involvement of ZEB-1 and ZEB-2. Oncogene 2007, 26, 711–724. [Google Scholar] [CrossRef]
  86. De, S.; Das, S.; Mukherjee, S.; Das, S.; Bandyopadhyay, S.S. Establishment of twist-1 and TGFBR2 as direct targets of microRNA-20a in mesenchymal to epithelial transition of breast cancer cell-line MDA-MB-231. Exp. Cell Res. 2017, 361, 85–92. [Google Scholar] [CrossRef]
  87. Kwak, S.Y.; Yoo, J.O.; An, H.J.; Bae, I.H.; Park, M.J.; Kim, J.; Han, Y.H. miR-5003-3p promotes epithelial-mesenchymal transition in breast cancer cells through Snail stabilization and direct targeting of E-cadherin. J. Mol. Cell Biol. 2016, 8, 372–383. [Google Scholar] [CrossRef] [PubMed]
  88. Pan, Y.; Li, J.; Zhang, Y.; Wang, N.; Liang, H.; Liu, Y.; Zhang, C.Y.; Zen, K.; Gu, H. Slug-upregulated miR-221 promotes breast cancer progression through suppressing E-cadherin expression. Sci. Rep. 2016, 6, 25798. [Google Scholar] [CrossRef] [PubMed]
  89. Park, S.-M.; Gaur, A.B.; Lengyel, E.; Peter, M.E. The MiR-200 Family Determines the Epithelial Phenotype of Cancer Cells by Targeting the E-Cadherin Repressors ZEB1 and ZEB2. Genes Dev. 2008, 22, 894–907. [Google Scholar] [CrossRef] [PubMed]
  90. Gregory, P.A.; Bert, A.G.; Paterson, E.L.; Barry, S.C.; Tsykin, A.; Farshid, G.; Vadas, M.A.; Khew-Goodall, Y.; Goodall, G.J. The miR-200 family and miR-205 regulate epithelial to mesenchymal transition by targeting ZEB1 and SIP1. Nature Cell Biol. 2008, 10, 593–601. [Google Scholar] [CrossRef] [PubMed]
  91. Xie, F.; Hosany, S.; Zhong, S.; Jiang, Y.; Zhang, F.; Lin, L.; Wang, X.; Gao, S.; Hu, X. MicroRNA-193a Inhibits Breast Cancer Proliferation and Metastasis by Downregulating WT1. PLoS ONE 2017, 12, e0185565. [Google Scholar] [CrossRef] [PubMed]
  92. Ma, L.; Young, J.; Prabhala, H.; Pan, E.; Mestdagh, P.; Muth, D.; Teruya-Feldstein, J.; Reinhardt, F.; Onder, T.T.; Valastyan, S.; et al. miR-9, a MYC/MYCN-activated microRNA, regulates E-cadherin and cancer metastasis. Nature Cell Biol. 2010, 12, 247–256. [Google Scholar] [CrossRef] [PubMed]
  93. Martello, G.; Rosato, A.; Ferrari, F.; Manfrin, A.; Cordenonsi, M.; Dupont, S.; Enzo, E.; Guzzardo, V.; Rondina, M.; Spruce, T.; et al. A MicroRNA targeting dicer for metastasis control. Cell 2010, 141, 1195–1207. [Google Scholar] [CrossRef]
  94. Cochrane, D.R.; Cittelly, D.M.; Howe, E.N.; Spoelstra, N.S.; McKinsey, E.L.; LaPara, K.; Elias, A.; Yee, D.; Richer, J.K. MicroRNAs Link Estrogen Receptor Alpha Status and Dicer Levels in Breast Cancer. Horm. Cancer 2010, 1, 306–319. [Google Scholar] [CrossRef]
  95. Chen, P.-S.; Su, J.-L.; Cha, S.-T.; Tarn, W.-Y.; Wang, M.-Y.; Hsu, H.-C.; Lin, M.-T.; Chu, C.-Y.; Hua, K.-T.; Chen, C.-N.; et al. MiR-107 Promotes Tumor Progression by Targeting the Let-7 MicroRNA in Mice and Humans. J. Clin. Investig. 2011, 121, 3442–3455. [Google Scholar] [CrossRef]
  96. Wu, Y.; Shi, W.; Tang, T.; Wang, Y.; Yin, X.; Chen, Y.; Zhang, Y.; Xing, Y.; Shen, Y.; Xia, T.; et al. MiR-29a Contributes to Breast Cancer Cells Epithelial–Mesenchymal Transition, Migration, and Invasion via down-Regulating Histone H4K20 Trimethylation through Directly Targeting SUV420H2. Cell Death Dis. 2019, 10, 176. [Google Scholar] [CrossRef]
  97. Ren, L.; Chen, H.; Song, J.; Chen, X.; Lin, C.; Zhang, X.; Hou, N.; Pan, J.; Zhou, Z.; Wang, L.; et al. MiR-454-3p-Mediated Wnt/β-Catenin Signaling Antagonists Suppression Promotes Breast Cancer Metastasis. Theranostics 2019, 9, 449–465. [Google Scholar] [CrossRef]
  98. Dobson, J.R.; Taipaleenmäki, H.; Hu, Y.-J.; Hong, D.; van Wijnen, A.J.; Stein, J.L.; Stein, G.S.; Lian, J.B.; Pratap, J. Hsa-Mir-30c Promotes the Invasive Phenotype of Metastatic Breast Cancer Cells by Targeting NOV/CCN3. Cancer Cell Int. 2014, 14, 73. [Google Scholar] [CrossRef]
  99. Huang, L.; Dai, T.; Lin, X.; Zhao, X.; Chen, X.; Wang, C.; Li, X.; Shen, H.; Wang, X. MicroRNA-224 Targets RKIP to Control Cell Invasion and Expression of Metastasis Genes in Human Breast Cancer Cells. Biochem. Biophys. Res. Commun. 2012, 425, 127–133. [Google Scholar] [CrossRef]
  100. Kong, W.; Yang, H.; He, L.; Zhao, J.; Coppola, D.; Dalton, W.S.; Cheng, J.Q. MicroRNA-155 Is Regulated by the Transforming Growth Factor β/Smad Pathway and Contributes to Epithelial Cell Plasticity by Targeting RhoA. Mol. Cell. Biol. 2008, 28, 6773–6784. [Google Scholar] [CrossRef]
  101. Sørlie, T.; Perou, C.M.; Tibshirani, R.; Aas, T.; Geisler, S.; Johnsen, H.; Hastie, T.; Eisen, M.B.; van de Rijn, M.; Jeffrey, S.S.; et al. Gene expression patterns of breast carcinomas distinguish tumor subclasses with clinical implications. Proc. Natl. Acad. Sci. USA 2001, 98, 10869–10874. [Google Scholar] [CrossRef]
  102. Tang, P.; Tse, G.M. Immunohistochemical Surrogates for Molecular Classification of Breast Carcinoma: A 2015 Update. Arch. Pathol. Lab. Med. 2016, 140, 806–814. [Google Scholar] [CrossRef]
  103. Russnes, H.G.; Lingjærde, O.C.; Børresen-Dale, A.-L.; Caldas, C. Breast Cancer Molecular Stratification. Am. J. Pathol. 2017, 187, 2152–2162. [Google Scholar] [CrossRef]
  104. Parise, C.A.; Caggiano, V. Breast Cancer Survival Defined by the ER/PR/HER2 Subtypes and a Surrogate Classification according to Tumor Grade and Immunohistochemical Biomarkers. J. Cancer Epidemiol. 2014, 2014, 469251. [Google Scholar] [CrossRef]
  105. Davies, C.; Godwin, J.; Gray, R.; Clarke, M.; Cutter, D.; Darby, S.; McGale, P.; Pan, H.C.; Taylor, C.; Wang, Y.C.; et al. Early Breast Cancer Trialists’ Collaborative Group (EBCTCG). Relevance of breast cancer hormone receptors and other factors to the efficacy of adjuvant tamoxifen: Patient-level meta-analysis of randomised trials. Lancet 2011, 378, 771–784. [Google Scholar] [CrossRef]
  106. Normann, L.S.; Aure, M.R.; Leivonen, S.K.; Haugen, M.H.; Hongisto, V.; Kristensen, V.N.; Mælandsmo, G.M.; Sahlberg, K.K. MicroRNA in combination with HER2-targeting drugs reduces breast cancer cell viability in vitro. Sci. Rep. 2021, 11, 10893. [Google Scholar] [CrossRef]
  107. Søkilde, R.; Persson, H.; Ehinger, A.; Pirona, A.C.; Fernö, M.; Hegardt, C.; Larsson, C.; Loman, N.; Malmberg, M.; Rydén, L.; et al. Refinement of Breast Cancer Molecular Classification by MiRNA Expression Profiles. BMC Genom. 2019, 20, 503. [Google Scholar] [CrossRef]
  108. Amiruddin, A.; Massi, M.N.; Islam, A.A.; Patellongi, I.; Pratama, M.Y.; Sutandyo, N.; Natzir, R.; Hatta, M.; Md Latar, N.H.; Wahid, S. MicroRNA-221 and Tamoxifen Resistance in Luminal-Subtype Breast Cancer Patients: A Case-Control Study. Ann. Med. Surg. 2022, 73, 103092. [Google Scholar] [CrossRef]
  109. Fan, T.; Mao, Y.; Sun, Q.; Liu, F.; Lin, J.; Liu, Y.; Cui, J.; Jiang, Y. Branched Rolling Circle Amplification Method for Measuring Serum Circulating MicroRNA Levels for Early Breast Cancer Detection. Cancer Sci. 2018, 109, 2897–2906. [Google Scholar] [CrossRef]
  110. McDermott, A.M.; Miller, N.; Wall, D.; Martyn, L.M.; Ball, G.; Sweeney, K.J.; Kerin, M.J. Identification and Validation of Oncologic MiRNA Biomarkers for Luminal A-like Breast Cancer. PLoS ONE 2014, 9, e87032. [Google Scholar] [CrossRef]
  111. Chen, Y.; Wu, N.; Liu, L.; Dong, H.; Wu, C. Correlation between MicroRNA-21, MicroRNA-206 and Estrogen Receptor, Progesterone Receptor, Human Epidermal Growth Factor Receptor 2 in Breast Cancer. Clin. Biochem. 2019, 71, 52–57. [Google Scholar] [CrossRef]
  112. Lowery, A.J.; Miller, N.; Devaney, A.; McNeill, R.E.; Davoren, P.A.; Lemetre, C.; Benes, V.; Schmidt, S.; Blake, J.; Ball, G.; et al. MicroRNA Signatures Predict Oestrogen Receptor, Progesterone Receptor and HER2/Neureceptor Status in Breast Cancer. Breast Cancer Res. 2009, 11, R27. [Google Scholar] [CrossRef]
  113. Amorim, M.; Lobo, J.; Fontes-Sousa, M.; Estevão-Pereira, H.; Salta, S.; Lopes, P.; Coimbra, N.; Antunes, L.; Palma de Sousa, S.; Henrique, R.; et al. Predictive and Prognostic Value of Selected MicroRNAs in Luminal Breast Cancer. Front. Genet. 2019, 10, 815. [Google Scholar] [CrossRef]
  114. Ulianova, E.P.; Tokmakov, V.V.; Shatova, I.S.; Sagakyants, A.B.; Goncharova, A.S.; Zaikina, E.V.; Chernikova, E.N.; Bakulina, S.M.; Pushkareva, T.F.; Kit, O.I.; et al. Evaluation of Prognostic Significance of MicroRNA in Tumors of Luminal, Primary Operable Breast Cancer without Her2 Neu Overexpression in Postmenopausal Women. J. Clin. Oncol. 2020, 38, e12558. [Google Scholar] [CrossRef]
  115. Ohzawa, H.; Miki, A.; Teratani, T.; Shiba, S.; Sakuma, Y.; Nishimura, W.; Noda, Y.; Fukushima, N.; Fujii, H.; Hozumi, Y.; et al. Usefulness of MiRNA Profiles for Predicting Pathological Responses to Neoadjuvant Chemotherapy in Patients with Human Epidermal Growth Factor Receptor 2-Positive Breast Cancer. Oncol. Lett. 2017, 13, 1731–1740. [Google Scholar] [CrossRef]
  116. Han, S.-H.; Kim, H.J.; Gwak, J.M.; Kim, M.; Chung, Y.R.; Park, S.Y. MicroRNA-222 Expression as a Predictive Marker for Tumor Progression in Hormone Receptor-Positive Breast Cancer. J. Breast Cancer 2017, 20, 35. [Google Scholar] [CrossRef]
  117. Luo, Y.; Wang, X.; Niu, W.; Wang, H.; Wen, Q.; Fan, S.; Zhao, R.; Li, Z.; Xiong, W.; Peng, S.; et al. Elevated MicroRNA-125b Levels Predict a Worse Prognosis in HER2-Positive Breast Cancer Patients. Oncol. Lett. 2017, 13, 867–874. [Google Scholar] [CrossRef]
  118. Wu, X.; Somlo, G.; Yu, Y.; Palomares, M.R.; Li, A.X.; Zhou, W.; Chow, A.; Yen, Y.; Rossi, J.J.; Gao, H.; et al. De Novo Sequencing of Circulating MiRNAs Identifies Novel Markers Predicting Clinical Outcome of Locally Advanced Breast Cancer. J. Transl. Med. 2012, 10, 42. [Google Scholar] [CrossRef]
  119. Souza, K.C.B.; Evangelista, A.F.; Leal, L.F.; Souza, C.P.; Vieira, R.A.; Causin, R.L.; Neuber, A.C.; Pessoa, D.P.; Passos, G.A.S.; Reis, R.M.V.; et al. Identification of Cell-Free Circulating MicroRNAs for the Detection of Early Breast Cancer and Molecular Subtyping. J. Oncol. 2019, 2019, 8393769. [Google Scholar] [CrossRef]
  120. Mattie, M.D.; Benz, C.C.; Bowers, J.; Sensinger, K.; Wong, L.; Scott, G.K.; Fedele, V.; Ginzinger, D.; Getts, R.; Haqq, C. Optimized High-Throughput MicroRNA Expression Profiling Provides Novel Biomarker Assessment of Clinical Prostate and Breast Cancer Biopsies. Mol. Cancer 2006, 5, 24. [Google Scholar] [CrossRef]
  121. Toyama, T.; Kondo, N.; Endo, Y.; Sugiura, H.; Yoshimoto, N.; Iwasa, M.; Takahashi, S.; Fujii, Y.; Yamashita, H. High Expression of MicroRNA-210 Is an Independent Factor Indicating a Poor Prognosis in Japanese Triple-Negative Breast Cancer Patients. Jpn. J. Clin. Oncol. 2012, 42, 256–263. [Google Scholar] [CrossRef]
  122. Kalniete, D.; Nakazawa-Miklaševiča, M.; Štrumfa, I.; Āboliņš, A.; Irmejs, A.; Gardovskis, J.; Miklaševičs, E. High Expression of MiR-214 Is Associated with a Worse Disease-Specific Survival of the Triple-Negative Breast Cancer Patients. Hered. Cancer Clin. Pract. 2015, 13, 7. [Google Scholar] [CrossRef]
  123. Yao, L.; Liu, Y.; Cao, Z.; Li, J.; Huang, Y.; Hu, X.; Shao, Z. MicroRNA-493 Is a Prognostic Factor in Triple-negative Breast Cancer. Cancer Sci. 2018, 109, 2294–2301. [Google Scholar] [CrossRef]
  124. Uva, P.; Cossu-Rocca, P.; Loi, F.; Pira, G.; Murgia, L.; Orrù, S.; Floris, M.; Muroni, M.R.; Sanges, F.; Carru, C.; et al. MiRNA-135b Contributes to Triple Negative Breast Cancer Molecular Heterogeneity: Different Expression Profile in Basal-like Versus Non-Basal-like Phenotypes. Int. J. Med. Sci. 2018, 15, 536–548. [Google Scholar] [CrossRef]
  125. Kolesnikov, N.N.; Veryaskina, Y.A.; Titov, S.E.; Rodionov, V.V.; Gening, T.P.; Abakumova, T.V.; Kometova, V.V.; Torosyan, M.K.; Zhimulev, I.F. Expression of Micrornas in Molecular Genetic Breast Cancer Subtypes. Cancer Treat. Res. Commun. 2019, 20, 100026. [Google Scholar] [CrossRef]
  126. Moi, L.; Braaten, T.; Al-Shibli, K.; Lund, E.; Busund, L.-T.R. Differential Expression of the MiR-17-92 Cluster and MiR-17 Family in Breast Cancer According to Tumor Type; Results from the Norwegian Women and Cancer (NOWAC) Study. J. Transl. Med. 2019, 17, 334. [Google Scholar] [CrossRef] [PubMed]
  127. Harris, E.E.R. Precision Medicine for Breast Cancer: The Paths to Truly Individualized Diagnosis and Treatment. Int. J. Breast Cancer. 2018, 2018, 4809183. [Google Scholar] [CrossRef]
  128. Hamam, R.; Hamam, D.; Alsaleh, K.A.; Kassem, M.; Zaher, W.; Alfayez, M.; Aldahmash, A.; Alajez, N.M. Circulating microRNAs in breast cancer: Novel diagnostic and prognostic biomarkers. Cell Death Dis. 2017, 8, e3045. [Google Scholar] [CrossRef]
  129. Mitchell, P.S.; Parkin, R.K.; Kroh, E.M.; Fritz, B.R.; Wyman, S.K.; Pogosova-Agadjanyan, E.L.; Peterson, A.; Noteboom, J.; O’Briant, K.C.; Allen, A.; et al. Circulating microRNAs as stable blood-based markers for cancer detection. Proc. Natl. Acad. Sci. USA 2008, 105, 10513–10518. [Google Scholar] [CrossRef]
  130. Zhao, H.; Shen, J.; Medico, L.; Wang, D.; Ambrosone, C.B.; Liu, S. A pilot study of circulating miRNAs as potential biomarkers of early stage breast cancer. PLoS ONE 2010, 5, e13735. [Google Scholar] [CrossRef]
  131. Bahrami, A.; Aledavood, A.; Anvari, K.; Hassanian, S.M.; Maftouh, M.; Yaghobzade, A.; Salarzaee, O.; ShahidSales, S.; Avan, A. The prognostic and therapeutic application of microRNAs in breast cancer: Tissue and circulating microRNAs. J. Cell. Physiol. 2018, 233, 774–786. [Google Scholar] [CrossRef]
  132. Terkelsen, T.; Russo, F.; Gromov, P.; Haakensen, V.D.; Brunak, S.; Gromova, I.; Krogh, A.; Papaleo, E. Secreted Breast Tumor Interstitial Fluid MicroRNAs and Their Target Genes Are Associated with Triple-Negative Breast Cancer, Tumor Grade, and Immune Infiltration. Breast Cancer Res. 2020, 22, 73. [Google Scholar] [CrossRef]
  133. Li, J.; He, D.; Bi, Y.; Liu, S. The Emerging Roles of Exosomal MiRNAs in Breast Cancer Progression and Potential Clinical Applications. Breast Cancer Targets Ther. 2023, 15, 825–840. [Google Scholar] [CrossRef] [PubMed]
  134. Cardinali, B.; Tasso, R.; Piccioli, P.; Ciferri, M.C.; Quarto, R.; Del Mastro, L. Circulating miRNAs in Breast Cancer Diagnosis and Prognosis. Cancers 2022, 14, 2317. [Google Scholar] [CrossRef] [PubMed]
  135. Gao, J.; Zhang, Q.; Xu, J.; Guo, L.; Li, X. Clinical Significance of Serum MiR-21 in Breast Cancer Compared with CA153 and CEA. Chin. J. Cancer Res. 2013, 25, 743–748. [Google Scholar] [CrossRef] [PubMed]
  136. Canatan, D.; Sönmez, Y.; Yılmaz, Ö.; Çim, A.; Coşkun, H.Ş.; Sezgin Göksu, S.; Ucar, S.; Aktekin, M.R. MicroRNAs as biomarkers for breast cancer. Acta Biomed. 2021, 92, e2021028. [Google Scholar] [CrossRef]
  137. Shimomura, A.; Shiino, S.; Kawauchi, J.; Takizawa, S.; Sakamoto, H.; Matsuzaki, J.; Ono, M.; Takeshita, F.; Niida, S.; Shimizu, C.; et al. Novel Combination of Serum MicroRNA for Detecting Breast Cancer in the Early Stage. Cancer Sci. 2016, 107, 326–334. [Google Scholar] [CrossRef] [PubMed]
  138. Heneghan, H.M.; Miller, N.; Lowery, A.J.; Sweeney, K.J.; Newell, J.; Kerin, M.J. Circulating microRNAs as novel minimally invasive biomarkers for breast cancer. Ann. Surg. 2010, 251, 499–505. [Google Scholar] [CrossRef] [PubMed]
  139. Schwarzenbach, H.; Milde-Langosch, K.; Steinbach, B.; Müller, V.; Pantel, K. Diagnostic Potential of PTEN-Targeting MiR-214 in the Blood of Breast Cancer Patients. Breast Cancer Res. Treat. 2012, 134, 933–941. [Google Scholar] [CrossRef]
  140. Chen, W.; Cai, F.; Zhang, B.; Barekati, Z.; Zhong, X.Y. The Level of Circulating MiRNA-10b and MiRNA-373 in Detecting Lymph Node Metastasis of Breast Cancer: Potential Biomarkers. Tumor Biol. 2013, 34, 455–462. [Google Scholar] [CrossRef]
  141. van Schooneveld, E.; Wouters, M.C.; Van der Auwera, I.; Peeters, D.J.; Wildiers, H.; Van Dam, P.A.; Vergote, I.; Vermeulen, P.B.; Dirix, L.Y.; Van Laere, S.J. Expression profiling of cancerous and normal breast tissues identifies microRNAs that are differentially expressed in serum from patients with (metastatic) breast cancer and healthy volunteers. Breast Cancer Res. 2012, 14, R34. [Google Scholar] [CrossRef]
  142. Jung, E.J.; Santarpia, L.; Kim, J.; Esteva, F.J.; Moretti, E.; Buzdar, A.U.; Di Leo, A.; Le, X.F.; Bast, R.C., Jr.; Park, S.T.; et al. Plasma microRNA 210 levels correlate with sensitivity to trastuzumab and tumor presence in breast cancer patients. Cancer 2012, 118, 2603–2614. [Google Scholar] [CrossRef]
  143. Roth, C.; Rack, B.; Müller, V.; Janni, W.; Pantel, K.; Schwarzenbach, H. Circulating microRNAs as blood-based markers for patients with primary and metastatic breast cancer. Breast Cancer Res. 2010, 12, R90. [Google Scholar] [CrossRef]
  144. Wu, Q.; Lu, Z.; Li, H.; Lu, J.; Guo, L.; Ge, Q. Next-Generation Sequencing of MicroRNAs for Breast Cancer Detection. J. Biomed. Biotechnol. 2011, 2011, 597145. [Google Scholar] [CrossRef]
  145. Wang, Y.; Yin, W.; Lin, Y.; Yin, K.; Zhou, L.; Du, Y.; Yan, T.; Lu, J. Downregulated Circulating MicroRNAs after Surgery: Potential Noninvasive Biomarkers for Diagnosis and Prognosis of Early Breast Cancer. Cell Death Discov. 2018, 4, 87. [Google Scholar] [CrossRef] [PubMed]
  146. Cuk, K.; Zucknick, M.; Heil, J.; Madhavan, D.; Schott, S.; Turchinovich, A.; Arlt, D.; Rath, M.; Sohn, C.; Benner, A.; et al. Circulating MicroRNAs in Plasma as Early Detection Markers for Breast Cancer. Int. J. Cancer 2013, 132, 1602–1612. [Google Scholar] [CrossRef]
  147. Ng, E.K.O.; Li, R.; Shin, V.Y.; Jin, H.C.; Leung, C.P.H.; Ma, E.S.K.; Pang, R.; Chua, D.; Chu, K.-M.; Law, W.L.; et al. Circulating MicroRNAs as Specific Biomarkers for Breast Cancer Detection. PLoS ONE 2013, 8, e53141. [Google Scholar] [CrossRef] [PubMed]
  148. He, B.; Zhao, Z.; Cai, Q.; Zhang, Y.; Zhang, P.; Shi, S.; Xie, H.; Peng, X.; Yin, W.; Tao, Y.; et al. MiRNA-Based Biomarkers, Therapies, and Resistance in Cancer. Int. J. Biol. Sci. 2020, 16, 2628–2647. [Google Scholar] [CrossRef] [PubMed]
  149. Fu, Z.; Wang, L.; Li, S.; Chen, F.; Au-Yeung, K.K.-W.; Shi, C. MicroRNA as an Important Target for Anticancer Drug Development. Front. Pharmacol. 2021, 12, 736323. [Google Scholar] [CrossRef] [PubMed]
  150. Hussen, B.M.; Rasul, M.F.; Abdullah, S.R.; Hidayat, H.J.; Faraj, G.S.H.; Ali, F.A.; Salihi, A.; Baniahmad, A.; Ghafouri-Fard, S.; Rahman, M.; et al. Targeting MiRNA by CRISPR/Cas in Cancer: Advantages and Challenges. Mil. Med. Res. 2023, 10, 32. [Google Scholar] [CrossRef] [PubMed]
  151. Singh, S.; Saini, H.; Sharma, A.; Gupta, S.; Huddar, V.G.; Tripathi, R. Breast Cancer: MiRNAs Monitoring Chemoresistance and Systemic Therapy. Front. Oncol. 2023, 13, 1155254. [Google Scholar] [CrossRef] [PubMed]
  152. Sell, M.C.; Ramlogan-Steel, C.A.; Steel, J.C.; Dhungel, B.P. MicroRNAs in Cancer Metastasis: Biological and Therapeutic Implications. Expert. Rev. Mol. Med. 2023, 25, e14. [Google Scholar] [CrossRef]
  153. Shah, M.Y.; Ferrajoli, A.; Sood, A.K.; Lopez-Berestein, G.; Calin, G.A. MicroRNA Therapeutics in Cancer—An Emerging Concept. eBioMedicine 2016, 12, 34–42. [Google Scholar] [CrossRef]
  154. Munoz, J.P.; Perez-Moreno, P.; Perez, Y.; Calaf, G.M. The Role of MicroRNAs in Breast Cancer and the Challenges of Their Clinical Application. Diagnostics 2023, 13, 3072. [Google Scholar] [CrossRef]
  155. Balacescu, O.; Visan, S.; Baldasici, O.; Balacescu, L.; Vlad, C.; Achimas-Cadariu, P. miRNA-Based Therapeutics in Oncology, Realities, and Challenges. In Antisense Therapy; IntechOpen: London, UK, 2019. [Google Scholar]
  156. Okumura, S.; Hirano, Y.; Komatsu, Y. Stable Duplex-Linked Antisense Targeting MiR-148a Inhibits Breast Cancer Cell Proliferation. Sci. Rep. 2021, 11, 11467. [Google Scholar] [CrossRef]
  157. Van Der Ree, M.H.; Van Der Meer, A.J.; De Bruijne, J.; Maan, R.; Van Vliet, A.; Welzel, T.M.; Zeuzem, S.; Lawitz, E.J. Long-term safety and efficacy of microRNA-targeted therapy in chronic hepatitis C patients. Antiviral Res. 2014, 111, 53–59. [Google Scholar] [CrossRef]
  158. Roberts, T.C.; Langer, R.; Wood, M.J.A. Advances in Oligonucleotide Drug Delivery. Nat. Rev. Drug Discov. 2020, 19, 673–694. [Google Scholar] [CrossRef]
  159. Querfeld, C.; Pacheco, T.; Foss, F.M.; Halwani, A.S.; Porcu, P.; Seto, A.G.; Ruckman, J.; Landry, M.L.; Jackson, A.L.; Pestano, L.A.; et al. Preliminary results of a phase 1 trial evaluating MRG-106, a synthetic microRNA antagonist (LNA antimiR) of microRNA-155, in patients with CTCL. Blood 2016, 128, 1829. [Google Scholar] [CrossRef]
  160. De Cola, A.; Lamolinara, A.; Lanuti, P.; Rossi, C.; Iezzi, M.; Marchisio, M.; Todaro, M.; De Laurenzi, V. MiR-205-5p Inhibition by Locked Nucleic Acids Impairs Metastatic Potential of Breast Cancer Cells. Cell Death Dis. 2018, 9, 821. [Google Scholar] [CrossRef]
  161. Haftmann, C.; Riedel, R.; Porstner, M.; Wittmann, J.; Chang, H.-D.; Radbruch, A.; Mashreghi, M.-F. Direct Uptake of Antagomirs and Efficient Knockdown of MiRNA in Primary B and T Lymphocytes. J. Immunol. Methods 2015, 426, 128–133. [Google Scholar] [CrossRef]
  162. Ebert, M.S.; Sharp, P.A. Emerging Roles for Natural MicroRNA Sponges. Curr. Biol. 2010, 20, R858–R861. [Google Scholar] [CrossRef]
  163. Barta, T.; Peskova, L.; Hampl, A. MiRNAsong: A Web-Based Tool for Generation and Testing of MiRNA Sponge Constructs in Silico. Sci. Rep. 2016, 6, 36625. [Google Scholar] [CrossRef]
  164. Liang, A.-L.; Zhang, T.-T.; Zhou, N.; Wu, C.Y.; Lin, M.-H.; Liu, Y.-J. MiRNA-10b Sponge: An Anti-Breast Cancer Study in Vitro. Oncol. Rep. 2016, 35, 1950–1958. [Google Scholar] [CrossRef] [PubMed]
  165. Mandal, C.C.; Ghosh-Choudhury, T.; Dey, N.; Choudhury, G.G.; Ghosh-Choudhury, N. miR-21 is targeted by omega-3 polyunsaturated fatty acid to regulate breast tumor CSF-1 expression. Carcinogenesis 2012, 33, 1897–1908. [Google Scholar] [CrossRef] [PubMed]
  166. Fan, R.; Xiao, C.; Wan, X.; Cha, W.; Miao, Y.; Zhou, Y.; Qin, C.; Cui, T.; Su, F.; Shan, X. Small Molecules with Big Roles in MicroRNA Chemical Biology and MicroRNA-Targeted Therapeutics. RNA Biol. 2019, 16, 707–718. [Google Scholar] [CrossRef]
  167. Sun, J.; Xu, M.; Ru, J.; James-Bott, A.; Xiong, D.; Wang, X.; Cribbs, A.P. Small Molecule-Mediated Targeting of MicroRNAs for Drug Discovery: Experiments, Computational Techniques, and Disease Implications. Eur. J. Med. Chem. 2023, 257, 115500. [Google Scholar] [CrossRef]
  168. Melo, S.; Villanueva, A.; Moutinho, C.; Davalos, V.; Spizzo, R.; Ivan, C.; Rossi, S.; Setien, F.; Casanovas, O.; Simo-Riudalbas, L.; et al. Small Molecule Enoxacin Is a Cancer-Specific Growth Inhibitor That Acts by Enhancing TAR RNA-Binding Protein 2-Mediated MicroRNA Processing. Proc. Natl. Acad. Sci. USA 2011, 108, 4394–4399. [Google Scholar] [CrossRef]
  169. Monroig-Bosque, P.d.C.; Shah, M.Y.; Fu, X.; Fuentes-Mattei, E.; Ling, H.; Ivan, C.; Nouraee, N.; Huang, B.; Chen, L.; Pileczki, V.; et al. OncomiR-10b Hijacks the Small Molecule Inhibitor Linifanib in Human Cancers. Sci. Rep. 2018, 8, 13106. [Google Scholar] [CrossRef] [PubMed]
  170. Hosseinahli, N.; Aghapour, M.; Duijf, P.H.G.; Baradaran, B. Treating Cancer with MicroRNA Replacement Therapy: A Literature Review. J. Cell Physiol. 2018, 233, 5574–5588. [Google Scholar] [CrossRef]
  171. Liang, Z.; Ahn, J.; Guo, D.; Votaw, J.R.; Shim, H. MicroRNA-302 Replacement Therapy Sensitizes Breast Cancer Cells to Ionizing Radiation. Pharm. Res. 2013, 30, 1008–1016. [Google Scholar] [CrossRef]
  172. Zhang, J.; Zhang, Z.; Wang, Q.; Xing, X.-J.; Zhao, Y. Overexpression of MicroRNA-365 Inhibits Breast Cancer Cell Growth and Chemo-Resistance through GALNT4. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 4710–4718. [Google Scholar]
  173. Hong, D.S.; Kang, Y.K.; Borad, M.; Sachdev, J.; Ejadi, S.; Lim, H.Y.; Brenner, A.J.; Park, K.; Lee, J.L.; Kim, T.Y.; et al. Phase 1 study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br. J. Cancer 2020, 122, 1630–1637. [Google Scholar] [CrossRef]
Figure 1. Biosynthesis of miRNA: Pre-miRNAs are further cleaved into an asymmetric duplex using the action of Dicer and accessory proteins. Transactivation-responsive RNA-binding protein (TRBP) and PACT in humans remove the loop sequence by forming a short-lived asymmetric duplex intermediate (miRNA: miRNA), with a duplex of about 22 nucleotides in length. This pre-cursor is cleaved to generate ~21–25-nucleotide mature miRNAs. The mature miRNA is loaded into the miRNA-induced silencing complex (miRISC), which binds to target mRNA, resulting in either the degradation of mRNA or the blockage of translation without mRNA degradation (adapted from [17,18]).
Figure 1. Biosynthesis of miRNA: Pre-miRNAs are further cleaved into an asymmetric duplex using the action of Dicer and accessory proteins. Transactivation-responsive RNA-binding protein (TRBP) and PACT in humans remove the loop sequence by forming a short-lived asymmetric duplex intermediate (miRNA: miRNA), with a duplex of about 22 nucleotides in length. This pre-cursor is cleaved to generate ~21–25-nucleotide mature miRNAs. The mature miRNA is loaded into the miRNA-induced silencing complex (miRISC), which binds to target mRNA, resulting in either the degradation of mRNA or the blockage of translation without mRNA degradation (adapted from [17,18]).
Biomedicines 12 00691 g001
Figure 2. Variety of miRNA inhibition therapies in breast cancer discussed in this review.
Figure 2. Variety of miRNA inhibition therapies in breast cancer discussed in this review.
Biomedicines 12 00691 g002
Figure 3. Potential approaches to miRNA replacement therapy for breast cancer using a variety of miRNA delivery systems.
Figure 3. Potential approaches to miRNA replacement therapy for breast cancer using a variety of miRNA delivery systems.
Biomedicines 12 00691 g003
Table 1. List of miRNAs and the function of their targets in breast cancer.
Table 1. List of miRNAs and the function of their targets in breast cancer.
MicroRNATargetFunctionReferences
Oncogenic miRNAs in breast cancer
miRNA-10bHOXD10Promotes cell migration, invasion, and metastasis[25]
miRNA-17/92 clusterCOL4A3, LAMA3,
TIMP2/3, ADORA1
Promotes lymph node metastasis, enhanced cell proliferation, colony formation, migration, and invasion in Triple Negative Breast Cancer (TNBC)[26,27]
miRNA-21PDCD4, PTEN, TPM1,
TIMP3
Promotes invasion, metastasis, and migration[28,29,30,31]
miRNA-24Nanog, Oct-3/4, BimL, F1H1, HIF-1, Snail, VEGFAHypoxia-inducible miRNA[32]
miRNA-122Pyruvate kinase and
citrate synthase
Promotes metastasis by reprogrammed glucose metabolism[33]
miRNA-135bLATS2, CDK2,
p-YAP
Promotes cell proliferation and S-G2/M cell cycle progression[34]
miRNA-155SOCS1, TP53INP1, FOXO3Promotes cell growth, proliferation, and survival[35,36,37]
miRNA-181a BimPromotes epithelial-to-mesenchymal transition (EMT), migration, and invasion[38]
miRNA-191-5pSOX4, caspase-3, caspase-7, p53Promotes apoptosis resistance and doxorubicin resistance[39]
miRNA-200bEzrin/Radixin/Moesin (ERM)Promotes metastasis and invasion[40]
miRNA-206 NK-1Promotes breast cancer cell invasion, migration, proliferation, and colony formation in vitro[41]
miRNA-210Pax-5Modulating EMT and hypoxia[42]
miRNA-331HER2, HOTAIR, E2F1,
DOHH
Promotes metastasis and invasion by elevation in plasma of metastatic breast cancer patients[43]
miRNA-373CD44Promotes cell migration, invasion, and metastasis[44]
miRNA-455-3pEI24Promotes proliferation, invasion, and migration[45]
miRNA-498BRCA1Promotes TNBC cell proliferation[46]
miRNA-520cCD44Promotes cell migration, invasion, and metastasis[44]
Tumor-suppressor miRNAs in breast cancer
miRNA-17-92 Mekk2, cyclin D1Promotes NK cell antitumoral activity and reduces metastasis, regulates G1 to S phase transition[47,48]
miRNA-7SETDB1Inhibits cell invasion and metastasis, decreases the BCSC population, and partially reverses EMT[49]
let-7d Cyclin D1Induces stem cells radiation sensitization[50]
miRNA-30 Ubc9, ITGB3Inhibits self-renewal of breast tumor-initiating cells (BT-ICs), trigger apoptosis[51]
miRNA-33bHMGA2, SALL4, Twist1Regulates cell stemness and metastasis[52]
miRNA-34a IMP3Regulates TNBC stem cell property[53]
miRNA-125bERBB2, EPO, EPOR, ENPEP, CK2-α, CCNJ, MEGF9Inhibits cell proliferation and differentiation,
migration and invasion
[54,55]
miRNA-137 FSTL1Suppresses TNBC stemness[56]
miRNA-143ERK5, MAP3K7, Cyclin D1Anti-proliferative[57]
miRNA-148aWNT1, MMP13Inhibits cell proliferation, migration and
invasion
[58,59]
miRNA-200aTFAMRegulates breast cancer cell growth and mtDNA copy number[60]
miRNA-203ΔNp63αForfeiture of self-renewing capacity associated with epithelial stem cells, suppresses proliferation and colony formation[61]
miRNA-205HMGB3, Notch-2Suppresses proliferation and invasion and inhibits EMT and stem cell properties[62,63]
miRNA-206Cyclin D2, Cx43Reduces migration, invasion, and metastasis[64,65]
miRNA-223HAX-1Re-sensitizes TNBC stem cells to tumor necrosis factor-related apoptosis[66]
miRNA-483-3pCyclin E1, p-NPAT,
NPAT, CDK2
Anti-proliferative and G1-S cell cycle arrest[67]
miRNA-497Cyclin E1Anti-proliferative and reduces migration[68]
miRNA-519d-3pLIMK1Suppresses growth and motility[69]
Table 2. Characteristics of miRNAs in subtypes of breast cancer.
Table 2. Characteristics of miRNAs in subtypes of breast cancer.
SubtypemiRNA SignatureFindingsAnalysis TypeReferences
Luminal-AmiRNA-99a/let-7c/
miRNA-125b
High expression in Luminal A compared to Luminal B tissueBreast cancer tissues-cluster analysis of small RNAseq[107]
Luminal-AmiRNA-221High expressionPlasma samples—analyzed using qRT-PCR[108]
Luminal-AmiRNA-16, miRNA-21,
miRNA-155, miRNA-195
Higher expression than healthy controlsSerum circulating miRNAs analyzed using qRT-PCR[109]
Luminal-AmiRNA-29a,
miRNA-181a and
miRNA-652
Reliably differentiate between cancers and controlsMicroarray analysis and confirmed using qRT-PCR[110]
Luminal-AmiRNA-206Higher expressionSerum samples and analyzed using qRT-PCR[111]
Luminal-BmiRNA-342Higher expressionEarly-stage breast cancer specimens—
microarray and qRT-PCR analysis
[112]
Luminal-BmiRNA-182-5p and miRNA-200b-3pSignificantly higher than in normal breast tissuesBreast cancer tissue: global focus miRNA PCR Panel and analyzed using qRT-PCR[113]
Luminal-BmiRNA-21,
miRNA-221,
miRNA-200a and
miRNA-196a
Overexpression in Luminal B breast cancer without overexpression of HER2/neuTumor tissues of patients with Luminal B breast cancer and analyzed using qRT-PCR[114]
Luminal-BmiRNA-210Upregulated in the analyses of all 40 patients’ samplesFFPE blocks: miRNA microarray and confirmed using qRT-PCR[115]
Luminal-BmiRNA-222Higher expression in Luminal-B/(HER2+) subtypes than in Luminal A and TNBC subtypesBreast cancer tissues and analyzed using qRT-PCR[116]
HER2-enrichedmiRNA-125bSignificantly increased expression in breast cancer tissues compared with that in the non-cancerous tissuesFFPE breast cancer tissues, luciferase activity and qRT-PCR[117]
HER2-
enriched
miRNA-375 and
miRNA-122
High levels of circulating miRNA-375 and low levels of miRNA-122 were associated with HER2 statusSerum specimens: Solexa deep sequencing and confirmed using qRT-PCR[118]
HER2-
enriched
miRNA-548ar-5p,
miRNA-584-3p, miRNA-615-3p, and miRNA-1283
Significantly differential expression in the HER2-enriched subtypeSerum samples: Multiplexed gene expression analysis[119]
HER2-
enriched
let-7f, let-7g, miRNA-107, miRNA-10b,
miRNA-126, miRNA-154 and miRNA-195
Inversely correlated with HER2 overexpressionPrimary breast cancer biopsies: microarray and confirmed using qRT-PCR[120]
HER2-
enriched
miRNA-520d,
miRNA-376b
Highly expressed and accurately predicted HER2 status in early-stage breast tumorsEarly-stage breast cancer specimens-
microarray and qPCR analysis
[112]
TNBC or
basal-like
miRNA-210Significantly higher expression compared to ER+/HER2 breast cancersBreast cancer tissue and analyzed using qRT-PCR[121]
TNBC or
basal-like
miRNA-214Higher expression FFPE tissues and confirmed using qRT-PCR[122]
TNBC or
basal-like
miRNA-493High expressionTissue microarrays [123]
TNBC or
basal-like
miRNA-135bHigh expression TaqMan low-density array [124]
TNBC or
basal-like
miRNA-20a, miRNA-221Higher expression compared to Luminal A and Luminal B/HER2 breast cancer subtypesBiopsies of tumor tissue; and confirmed using qRT-PCR[125]
TNBC or
basal-like
miRNA-17 familyOverexpressed in high grade and TNBC associated with aggressive behaviorFFPE- tissue: miRCURY LNA microarray and confirmed using qRT-PCR[126]
TNBC or
basal-like
miRNA-17-92Elevated in TNBC but reduced in ER-positive breast cancer and associated with poor outcome. The miRNA-17–92 expression enhanced cell growth and invasion of TNBC cellsBreast cancer cell lines: high-throughput mRNA sequencing and confirmed using qRT-PCR[27]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Subramanian, K.; Sinha, R. Functions of Differentially Regulated miRNAs in Breast Cancer Progression: Potential Markers for Early Detection and Candidates for Therapy. Biomedicines 2024, 12, 691. https://doi.org/10.3390/biomedicines12030691

AMA Style

Subramanian K, Sinha R. Functions of Differentially Regulated miRNAs in Breast Cancer Progression: Potential Markers for Early Detection and Candidates for Therapy. Biomedicines. 2024; 12(3):691. https://doi.org/10.3390/biomedicines12030691

Chicago/Turabian Style

Subramanian, Kumar, and Raghu Sinha. 2024. "Functions of Differentially Regulated miRNAs in Breast Cancer Progression: Potential Markers for Early Detection and Candidates for Therapy" Biomedicines 12, no. 3: 691. https://doi.org/10.3390/biomedicines12030691

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop