Open access peer-reviewed chapter

Advantages of Noncoding RNAs in Molecular Diagnosis

Written By

Tomomi Fujii, Tomoko Uchiyama and Maiko Takeda

Submitted: 02 May 2022 Reviewed: 24 May 2022 Published: 22 June 2022

DOI: 10.5772/intechopen.105525

From the Edited Volume

Recent Advances in Noncoding RNAs

Edited by Lütfi Tutar

Chapter metrics overview

101 Chapter Downloads

View Full Metrics

Abstract

Noncoding RNAs contribute to physiological processes by regulating many intracellular molecules participating in the life-supporting mechanisms of development, differentiation, and regeneration as well as by disrupting various signaling mechanisms such as disease development and progression and tumor growth. Because microRNAs (miRNAs) target and regulate the functions of key proteins, it is very useful to identify specific miRNAs that contribute to cellular functions and to clarify the roles of their target molecules as diagnostic and therapeutic strategies for cancer prognosis and treatment. In this section, the roles of miRNAs in various cancers and the processes leading to the identification of their target molecules are described, and the latest diagnostic strategies using miRNAs are discussed with specific examples.

Keywords

  • microRNA
  • cancer
  • molecular diagnosis

1. Introduction

Noncoding RNAs (ncRNAs) comprise the majority of gene transcripts in eukaryotes. ncRNAs play little biological role, but microRNAs (miRNAs), first reported in 1993, have important roles in gene expression [1, 2]. One of the most well-known miRNAs, let-7, a heterochronic gene in C. elegans, is considered important for developmental differentiation and regulates the larva-to-adult transition during larval development [3]. As a summary of the biosynthetic process of miRNAs, double-stranded RNA, the source of RNA interference, is sequentially processed into single-stranded RNA fragments of 21–23 nucleotides (nt) by the action of various enzymes [4, 5]. Let-7 encodes a small 21-nt ncRNA that binds to the 3′-untranslated region (UTR) of lin genes, resulting in RNA–RNA interactions and negative regulation of mRNA [6]. More than 30,000 miRNAs have been identified from more than 250 species, ranging from prokaryotes to eukaryotes, invertebrates to vertebrates, and even viruses [7, 8]. More than 5000 of these miRNAs have been identified in humans [miRBase: The microRNA database; cited 2018 Aug; Available from: http://www.mirbase.org].

miRNAs function as posttranscriptional regulators (negative regulators) of various gene expressions by binding to complementary sequences, primarily in the 3′-UTR of their target genes. miRNAs can target a large number of genes, and the mRNAs targeted by miRNAs mRNAs targeted by miRNAs account for more than 60% of all genes. Most miRNAs are encoded within the introns of coding mRNAs, but some can also be found within the 3′-UTR sequences of noncoding or coding mRNAs but can also be found in the 3′-UTR sequence of noncoding or coding mRNAs [9]. When miRNAs bind to target mRNAs, these are destabilized by deadenylation factors becoming more susceptible to degradation. As a result, mRNA is degraded; however, sometimes the mRNA is not degraded and translation is regulated, a mechanism that is not yet unclear. Alterations in the genetic structure of miRNA-associated regions are known to occur in many pathologies, especially cancer. Such chromosomal aberrations include amplifications (e.g. miR-26a in gliomas), deletions (e.g. miR-15a/16–1 in chronic lymphocytic leukemia), mutations (e.g. miR-125a in breast cancer), translocations [e.g. miR-125b in acute myeloid leukemia with t(2;11)(p21;q23)], single-nucleotide polymorphisms (e.g. miR-608 in colorectal adenocarcinoma), and heterozygous deletions (such as the 14q32 cluster in acute lymphoblastic leukemia) [10, 11, 12, 13, 14, 15]. These changes can occur not only in the sequence of the mature miRNA itself but also in the promoter region/primary transcript (pri-miRNA) sequence or in the miRNA binding site of the target gene, and as somatic or germline mutations [16, 17, 18]. A number of transcription factors, as well as vital protein-coding genes, influence the expression levels of miRNAs, and this regulatory action is assumed to be particularly important for tissue specificity and developmental stage control. Conversely, loss of Dicer, DGCR8, Drosha, and Argonaute2(Ago2), components of pathways involved in the miRNA biosynthesis mechanism, is known to result in the inability to sustain life and early pregnancy death due to severe developmental defects. [19, 20, 21, 22]. Loss of reproductive function due to the deletion of components involved in miRNA biosynthesis, and developmental defects in the heart, lungs, and neuromuscular tissue have been reported [23, 24, 25, 26, 27, 28, 29, 30]. Dicer mutations have also been reported to cause tumor development and progression [31, 32]. Dysfunction of the miRNA biogenesis pathway causes the disruption of important life-support mechanisms associated with various diseases, among which cancer development and progression are strongly influenced. In particular, abnormal expression of Dicer and Drosha is known to lead to poor prognosis in various cancers [33, 34, 35, 36, 37, 38, 39]. Although there is a vast amount of evidence that miRNAs play fundamental roles in the mechanisms of many diseases, including cancer, unfortunately, they have not yet been fully exploited in diagnostic and therapeutic settings. It is therefore necessary to determine how miRNAs can potentially be utilized in clinical practice in the future. miRNAs likely have the greatest and most direct potential as novel biomarkers of diagnosis and prognosis, as well as predictors of therapeutic efficacy. It is also clear that miRNAs regulate or disrupt the expression of numerous target mRNAs, resulting in maintenance of vital activity or disease, which cannot be resolved by a single miRNA. Therefore, gene expression analysis by miRNA expression profiling in disease must accurately discriminate between disease diagnosis and tumor stage of development [40].

A more accurate diagnosis of tumors can be obtained by morphological histopathology using formalin-fixed paraffin-embedded (FFPE) tissues and profiling of expressed proteins via immunohistochemical staining. Genetic analysis using FFPE is also emerging as being important for therapeutic selection. However, the degradation of nucleic acids is a problem for FFPE, and nucleic acids extracted from aged tissues are degraded [41]. A particularly attractive property of miRNAs is their stability against chemical and enzymatic degradation. This implies that they can be purified from routinely prepared FFPE biopsy materials and measured robustly [42]. As a result, retrospective miRNA expression studies are underway, using FFPE as a conserved and valuable resource. The stability and clinical utility of miRNAs are due to their presence in extracellular biological fluids, including blood, as well as in tissues and cells. Tumor-associated miRNAs are found at higher levels in the serum and plasma of cancer patients than in healthy controls. Therefore, the use of miRNAs as potential noninvasive biomarkers of disease has attracted much attention, and miRNAs have been detected in many biological fluids, including plasma, serum, tears, urine, cerebrospinal fluid, breast milk, and saliva [43, 44]. The widespread use of biomarker testing, which is noninvasive and more reliable than the conventional histopathological diagnosis of cancer that invasive to the patient, could pave the way for screening programs, improve the early detection rate of disease, and enhance cancer prevention. Despite the growing functional importance of miRNAs as diagnostic and therapeutic strategies for cancer, there are still many knowledge gaps despite the potential to develop miRNA markers in the future.

Advertisement

2. Role of miRNAs as oncogenes and tumor suppressor genes

The role of miRNAs in regulating the genes involved in cancer growth and progression may be to suppress cancer progression primarily through suppressing target gene expression or to inhibit the function of genes that suppress cancer progression, therefore positively affecting cancer progression. The results obtained by detailed profiling of miRNA expression in cancer according to organ or cancer differentiation level revealed that the expression pattern of miRNAs varies depending on the organ, histological type, and differentiation level of cancer, indicating that the role of miRNAs is highly segmented [40]. Many miRNAs function in a suppressive manner in cancer. The expression of the aforementioned let-7 miRNA is downregulated in lung cancer, targeting and suppressing the expression of Ras, an important oncogene. This indicates that let-7 may function as a cancer suppressor gene [45]. The expression of miR-15 and miR-16 is absent or decreased in chronic lymphocytic leukemia, suggesting they act in a tumor-suppressive manner by targeting B cell lymphoma-2, an antiapoptotic factor [46]. Human miR-373 binds to the E-cadherin (CDH1) promoter and induces gene expression [47]. In addition to gene silencing through base pairing with DNA and mRNA target sequences, miRNAs are known to inhibit the function of regulatory proteins (decoy activity). miR-328 binds to poly C-binding protein 2 (PCBP2), also known as heterogeneous ribonucleoprotein E2. This binding does not involve the seed region of the miRNA, but it inhibits its interaction with the target mRNA. In chronic myeloid leukemia, downregulation of miR-328 inhibits myeloid differentiation via PCBP2, leading to tumor progression [48]. It is also important to understand the molecular mechanisms of miRNAs to define the mechanism of their suppressive properties against cancer. miRNAs primarily target the 3′-UTR of mRNAs and downregulate the cytoplasmic expression of protein-coding genes. As evidence for this, miRNAs are usually localized in the nucleus [49]. In addition to the 3′-UTR, other genetic regions at the DNA and RNA levels, that is, the 5′-UTR, promoter, and coding regions, as well as proteins can be targeted (Figure 1) [50, 51, 52]. By binding to various functional regions of genes, miRNAs can upregulate or downregulate gene expression, interact with other ncRNAs and RNA transcripts, and modulate biological networks [53, 54]. It has become clear that the actions of these miRNAs on their target molecules lead to the regulation of gene expression [55]. It has been shown that miRNAs not only bind to key sites in other RNAs but also to promoter regions at the DNA level, thereby affecting transcriptional activity. As a specific example, in terms of localization sites, miRNAs have been shown to function both in the cytoplasm and nucleus [56]. Investigation of the subcellular localization of miR-29b revealed it was mostly localized to the nucleus. The characteristic hexanucleotide terminal motif of miR-29b functions as a metastable nuclear localization element that induces the nuclear localization of the miRNA or small-interference RNA to which it binds [49]. It has been shown that miRNAs bind to adenylate-uridylate-rich elements and upregulate translation in cells that have undergone cell cycle arrest. This suggests that miRNAs can affect not only cell differentiation and proliferation stages in a broad sense but also the cell cycle [57].

Figure 1.

Predicted binding sites for miRNAs in genes. miRNAs have binding sites not only in the 3′UTR but also in every site of the gene.

Advertisement

3. Expression dynamics of miRNAs in cancer and expectations for clinical application

Cancer is a malignant disease in which a population of cells with genetically abnormal protein expression proliferate autonomously, destroying host tissues as they grow, metastasize, and settle in other organs via vascular flow. As a result, cancer causes the dysfunction of normal tissues and ultimately leads to organ failure. There is evidence that carcinogenesis is a multistep process in which malignant cells accumulate epigenetic and genetic changes that gradually transform normal cells into malignant cells. Triggers for carcinogenesis include tobacco, chemical exogenous stimuli such as chemicals and dust from air pollution, and accumulation of genetic damage due to the aging of individual cells [58]. The proliferation of cancer cells progresses by causing structural changes in genes that promote cancer cell proliferation (oncogenes), genes that prevent their proliferation (tumor suppressor genes), and abnormal gene expression [59, 60, 61, 62]. Malignant cells lose cell discipline in morphology and differentiation and acquire autonomous proliferation, antiapoptosis, and invasiveness properties. miRNAs have been analyzed in various human malignant tumors and can be involved in the pathophysiology of benign and malignant tumors in various ways. Although there are differences in miRNA expression profiles between normal and tumor tissues, many of these miRNAs are only indirectly altered by genetic changes that occur during carcinogenesis, epigenomic dynamics, and physiological changes in cell biology and do not directly trigger tumor development. Differentially expressed miRNAs between tumor and normal tissues have been identified in lymphomas, breast cancer, lung cancer, papillary thyroid cancer, glioblastoma, hepatocellular carcinoma, pancreatic tumor, pituitary adenoma, cervical cancer, brain tumors, prostate cancer (PCa), kidney and bladder cancer, and colon cancer [63, 64, 65, 66, 67, 68, 69, 70, 71, 72, 73, 74, 75, 76]. The involvement of miRNAs in the carcinogenic process may involve changes in the components of miRNA biosynthesis pathway, and overall suppression of miRNA maturation by mutations in DROSHA, DGCR8, or DICER has been shown to affect the transcription and translation of multiple mRNAs, ultimately promoting cell transformation and tumorigenesis [77, 78, 79]. The fact that conditional loss of DICER or DGCR8 inhibits bone and cartilage growth and differentiation in mice and in vitro has been demonstrated, suggesting that components involved in the miRNA biosynthesis mechanism are also key molecules in carcinogenesis and cancer progression [80, 81]. The expression of Dicer, Exportin-5, TRKRA, TARBP2, DGCR8, and Argonaut-2 changes dynamically with changes in miRNA expression during the oncogenic phase in pancreatic cancer, suggesting that these miRNA biosynthesis-related molecules can cause cellular changes, including carcinogenesis [82]. Decreased expression of some components that play an important role in miRNA biogenesis, such as DICER, is associated with cancer prognosis and sensitivity to molecularly targeted drugs and has been shown to correlate with short survival in non-small cell lung cancer while increasing sensitivity to gefitinib [36, 83]. The methylation of DICER and DROSHA is also a potential biomarker for lung cancer [84]. Some miRNA precursors, known as mirtrons, have the unique feature of entering the miRNA biosynthesis pathway without undergoing Drosha-mediated cleavage while mimicking the structural features of pre-miRNAs. They are susceptible to epigenetic regulation and have been shown to exert epigenetic effects through interactions between various miRNAs and host genes in urothelial cell carcinoma [85, 86, 87]. Breast cancer susceptibility gene 1 (BRCA1) promotes the processing of primary miRNA transcripts and increases the expression of both precursors and mature forms of let-7a-1, miR-16-1, miR-145, and miR-34a [88]. BRCA1 interacts with Smad3, p53, and DHX9 RNA helicases by directly binding to DROSHA and DDX5 in the DROSHA microprocessor complex that regulates miRNA maturation and recognizes RNA secondary structures through its DNA-binding domain, which interacts with pri-miRNA via the DNA-binding domain.

Thus, from their discovery to the present, miRNAs have brought great promise and success in cancer diagnosis, prognosis, and treatment both in clinical practice and experimental medicine. Powerful technologies, such as miRNA arrays, short RNA deep sequencing, specific quantitative polymerase chain reaction (PCR) of miRNAs, and antisense technologies, have revealed the functions of numerous miRNAs. It is expected that they will continue to have a major impact on clinical oncology in terms of cancer diagnosis and prognostic treatment. Because miRNAs are important factors that define cellular characteristics, direct detection of miRNAs and miRNA target molecules from algorithms based on the vast amount of data obtained by profiling analysis could be used as a valuable tool for cancer diagnosis. miRNA expression profiling has been considered more efficient and informative than conventional mRNA profiling in classifying tumors with respect to their tissue of origin and differentiation [89, 90, 91, 92, 93]. These findings indicate that in clinical practice, miRNAs are useful biomarkers for tracking the tissue of origin of tumors [94, 95, 96, 97]. miRNA expression profiles can be an important source of information for cancer progression prediction and prognosis. It has been shown that high expression of miR-21 and miR-155 in lung cancer and colorectal cancer (CRC) predicts efficient recurrence, poor prognosis, and reduced survival [98, 99, 100, 101]. Functional analysis of miRNAs through experimental processes has been found to be valuable for therapeutic strategies based on the regulation of miRNA activity. As a therapeutic strategy against miRNAs acting as oncogenes (oncomirs), anti-miRNAs are expected to exert therapeutic effects by blocking oncomirs through the action of oligonucleotides that are complementary to miRNAs. For them to function stably, they must be chemically modified to enhance their stability and deliverability to the target organs. As is the case with many drugs used in cancer therapy, issues related to the delivery of miRNA-containing compounds, their stability in vivo, and biological defense measures against toxicity are important issues that must be resolved as miRNAs progress from experimental to clinical trials. However, because miRNAs have multiple targets, inhibition of target molecules by miRNAs may cause serious side effects. These issues are expected to be resolved and applied clinically in the future.

Advertisement

4. Functions of miRNAs in various organs

Given their properties, miRNAs have many target molecules and each performs different functions during cell differentiation and proliferation. Therefore, miRNAs are expected to perform characteristic functions in various organs. Here, the functions of miRNAs in typical cancers are introduced.

4.1 Breast cancer

miRNA profiling studies in breast cancer have been conducted from various perspectives. The first miRNA shown to be highly expressed in metastatic breast cancer was miR-10b (using mouse and human cells), with clinical correlation in primary breast cancer [102]. Subsequent studies confirmed the elevated levels of miR-10b, miR-34a, and miR-155 in patients with metastatic breast cancer [103]. Furthermore, miR-10b and miR-373 were recently shown to be upregulated in lymph node-positive breast cancer [104]. In addition, miRNAs from previous miRNA profiling studies have been shown to play potential roles as tumor suppressor genes or oncogenes in breast cancer [65, 105, 106]. The decreased expression of many genes involved in cancer development and progression has become evident, and miRNAs and their target molecules that are downregulated in breast cancer tissues and cell lines have become the focus of research. Breast cancer is a cancerous tissue with high histopathological heterogeneity. Identifying miRNAs and their target molecules that are specifically expressed and function in this context and have diagnostic value is therefore difficult. miRNA profiling should facilitate the accurate classification of tumor subtypes and provide useful additional information to complement current classification methods and guide therapeutic decisions. Studies have nested to predict miRNAs in accordance with previous subtype classifications. Recent analyses of miRNAs have identified predicted miRNAs corresponding to hormone receptor and HER2 subtypes, including ER (miR-342, miR-299, miR-217, miR-190, miR-135b, and miR-218), PR (miR-520g, miR-377, miR- 527-518a, miR-520f-520c), and HER2 (miR-520d, miR-181c, miR-302c, miR-376b, and miR-30e). These expression kinetics classified cases with 100% accuracy compared to immunohistochemistry results [107]. Further classification is expected to increase miRNA profiling accuracy and the development of new diagnostic markers and therapeutic strategies based on miRNA target molecules.

4.2 Prostate cancer

Prostate cancer (PCa) is one of the most common causes of cancer-related death in men. Although most PCa progress very slowly and remain within the primary tumor, some have an aggressive behavior that can spread to other organs via the vascular system. Prostate-specific antigen (PSA) and prostate acid phosphatase (PAP) are useful tumor markers for prostate cancer. However, PAP has low specificity and PSA, despite its superior specificity, is inadequate for predicting the tumor types in which it is not elevated and histological differentiation and progression. Many studies have indicated that miRNAs are involved in the development of PCa [108, 109, 110, 111, 112]. Aberrant expression of miRNAs has been detected in PCa cell lines, in experimental carcinogenesis using cancer xenografts, and in clinical samples obtained from patients [113, 114, 115, 116, 117, 118]. These changes suggest that miRNAs play an important role in the pathogenesis of PCa.

Many miRNAs are expressed as oncogenes. For example, miR-375, miR-106a, and miR-194 have been reported to play important roles in tumor formation, proliferation, and progression [119, 120, 121]. When the expression of tumor suppressor miRNAs decreases, the expression of oncogenes increases, promoting cancer growth and progression. Experimental introduction of precursors of these miRNAs into cells and overexpression of these miRNAs to reproduce normal expression patterns suppressed the growth, invasion, clonogenesis, and metastatic potential of PCa cells. The most well-known miRNAs are miR-149-5p, miR-7-5p, miR-122, miR-124-3p, miR-188, and miR-129 [122, 123, 124, 125, 126, 127]. Downregulation of these miRNAs is associated with increased cell proliferation, migration, and chemotherapy resistance, primarily through the regulation of cell growth and proliferation via the mitogen-activated protein kinase pathway.

Clinically practical markers in PCa include serum PSA measurement and digital rectal examination; however, their diagnostic value is limited and does not extend to prognostic prediction based on cancer progression or assessment of differentiation related to tumor growth potential. Human glandular kallikrein 2, urokinase plasminogen activator and its receptor, transforming growth factor-β1, and interleukin 6 and its receptor are biomarkers with high sensitivity and specificity for the early diagnosis of PCa. miRNAs are also clinically promising diagnostic markers for this cancer. miRNAs can be used as diagnostic and disease-monitoring markers for the presence of minimal residual disease and early detection of recurrence. Furthermore, the establishment of a diagnostic algorithm combining multiple oncomir and tumor suppressor miRNAs is expected to enable a more stepwise diagnosis of PCa. In this cancer, histopathological diagnosis by biopsy is highly invasive and often fails to detect microscopic cancers owing to the extremely small amount of visible tissue. To compensate for the shortcomings of local biopsy pathology diagnosis, the availability of profiling tests with circulating miRNAs has the potential to predict the onset of cancer and the degree of differentiation through its combination algorithm. From the time of diagnosis and initiation of treatment, the detection of circulating miRNAs could contribute to the early detection of hormone refractory PCa. The role of cancer stem cells in PCa has been elucidated in previous studies. The CD44-positive PCa stem cell population has tumorigenic and metastatic potential, which has been mentioned in conjunction with other useful markers for cancer stem cells [128]. Functional analysis of miRNAs has shown that miR-9-5p, miR-34a, miR-373, miR-708, and miR-520c directly target CD44 suppressing its expression [129, 130, 131, 132, 133]. The fact that the expression levels of these miRNAs are decreased in PCa suggests an increased potential for tumor invasion and metastasis. Profiling of miRNAs in PCa holds promise as a more precise diagnostic tool before performing prostate biopsy.

4.3 CRC

CRC is a common cancer worldwide and is the second most common cause of cancer-related deaths in the Western world. Although the 5-year survival rate for early-stage cancer exceeds 90%, it decreases to approximately 70% for advanced cancer and to approximately 10% when distant metastases are present. Although it is certainly a carcinoma for which early diagnosis is very useful, in reality early detection is less than 40% due to insufficient noninvasive screening. Although direct diagnosis by endoscopy is most effective in CRC, it often relies on a fecal occult blood test because of its invasive nature. However, the fecal occult blood reaction can be positive even for inflammatory or nonneoplastic polyps and lacks specificity in terms of cancer diagnosis. If CRC can be found to have miRNA expression characteristics, analysis of miRNAs using stool and the detection of circulating miRNAs may be useful in detecting marginally present tumor cells. Therefore, we searched for miRNAs that were significantly elevated in CRC using stool and plasma and found a number of miRNAs, including miR-17, miR-18a, miR-19a, miR-21, miR-92a, miR-200c, miR-221, and miR-106a as diagnostic markers [134, 135, 136, 137, 138]. miR-141 is a miRNA that functions in an inhibitory manner against colon cancer cells in terms of cell proliferation, invasiveness, and migration [139, 140, 141, 142]. As a circulating miRNA, miR-141, together with miR-200 and miR-143, could be an independent prognostic marker for advanced CRC [143, 144].

CRC progresses from adenoma to colorectal adenocarcinoma as precancerous lesions. Investigation of miRNA expression in staged histological types has revealed that the expression of several miRNAs is altered in precancerous lesions, adenomas, and CRC tissues compared to that in normal colorectal tissue. This suggests that some miRNAs are involved in the process of CRC and may be important biomarkers in the transition of tumor tissue from benign to malignant. In addition, miRNAs are now being extracted from stored FFPE tissue specimens and retrospectively profiled for miRNA expression using next-generation sequencing (NGS) to understand the origin of CRC development and metastatic disease in cases with multiple lesions. The most studied miRNAs with respect to clinical prognosis are miR-21, miR-143, and miR-145. Although miR-21 suppresses the expression of various tumor suppressor genes, it also affects cell proliferation, apoptosis, invasion, and tumor progression. Increased miR-21 expression in CRC tumor tissue correlates with increased metastatic potential of the primary tumor and is associated with shorter disease-free and overall survival. Owing to the antiapoptotic effect of miR-21, its overexpression was also found to affect therapeutic efficacy by reducing the efficacy of 5-fluorouracil (5-FU) chemotherapy [145, 146, 147, 148, 149, 150]. Decreased expression of miR-143 and miR-145 in CRC tumor tissue has been observed in several studies, suggesting that they behave as tumor suppressors [132, 151, 152, 153, 154]. Their decreased expression has been associated with increased tumor growth and angiogenesis, and it is clinically associated with shorter disease-free survival. The expression of miR-143 directly correlates with the sensitivity of the colon cancer cell line HCT116 to 5-FU treatment. miR-215 is involved in the cell proliferative capacity of CRC and is sensitive to chemotherapy and molecular-targeted therapy. miR-215 is also involved in the prognostic prediction of CRC [155, 156, 157, 158, 159, 160].

Although biomarker analysis using stool as a method to directly capture tumor cells in CRC is open for consideration, the establishment of a methodology for good-quality nucleic acid extraction that enables good analysis and sample quality maintenance is required to obtain reliable data due to the heterogeneity of sample processing. In contrast, for the prediction of treatment potential and prognosis for advanced cancer, miRNA profiling can be performed by nucleic acid extraction from FFPE of excised tumor tissue, and sufficient technology is being established for accuracy and reliability.

Advertisement

5. Analysis of circulating miRNAs for their usefulness as biomarkers

In addition to serum and plasma, various body fluids, such as saliva, cerebrospinal fluid, and body cavity fluid, are candidates as analyzable materials for the detection of circulating miRNAs. However, in samples of these fluids from cancer patients, not only tumor cells but also blood cells, macrophages, exosomes, and microparticles are present; where the miRNAs are present, there is also the issue of how many of the targeted tumor-derived miRNAs are included or how reliable the analysis is based on which miRNAs are included. Nevertheless, regardless of the members, each of whom may possess miRNAs, if the changes caused by the coexistence of cancer can be captured for cancer patients, the usefulness of profiling will be well maintained. It is well known that analysis of exosome-derived miRNAs can detect more cancer-relevant miRNAs than analysis using untreated serum. An interesting aspect of the quantitative analysis of miRNAs in exosomes and microparticles is that they may be involved in cell–cell interactions, stimulate intracellular signaling, and modulate metabolic functions and homeostasis in a small number of stem cells. Quantitative analysis of circulating miRNAs is most useful for miRNAs quantitative real-time PCR (qRT-PCR), because it detects miRNAs from a very small amount of nucleic acids. Although NGS is superior for expression profiling, it requires a sufficient amount of nucleic acids and is likely to lack reliability due to detection errors or low sensitivity. Therefore, a single-miRNA assay using the TaqMan method should be performed on miRNAs with presumed abnormal expression at the transcriptional level, as detected from microarray analysis or NGS analysis based on FFPE, followed by qRT-PCR validation. Capturing abnormal miRNA expression in cancer is a highly useful diagnostic strategy but, because capturing circulating miRNAs involves sampling a physiological diverse environment, we may be quantifying miRNAs derived not only from cancer but also from blood or other organs. In addition, miRNAs derived not only from cancer but also from coexisting diseases may be captured; therefore, it is not always the case that miRNAs of cancer cells are captured. In other words, the possibility of observing the dynamics of miRNAs in a broad sense, including in the cancer microenvironment, must be fully understood. Nevertheless, the analysis of circulating miRNAs is highly useful as a potential candidate for blood-based biomarkers that can be tested repeatedly. Circulating miRNAs are promising diagnostic biomarkers owing to their high information content and disease-specific regulation. miRNAs inhibit the translation of various proteins involved in disease-related signaling pathways, including malignant and benign diseases, and alterations in physiological states, thereby inhibiting specific signaling pathways. Therefore, the analysis of the complex profiling of multiple miRNAs involved in complex pathways is expected to provide more specific diagnostic algorithms.

Biomarker development involves several important issues. The use of various analytical technology platforms, serum or plasma samples, extraction of miRNAs from whole serum/plasma or exosomes/microvesicles separated by specific disease-related markers, and sample collection, storage, and processing, all need to be addressed. Depending on the time point of blood sample collection (presurgery, during surgery, and posttreatment) and patient treatment (surgery, chemotherapy, immunotherapy, and/or drug therapy), the expression profile of circulating miRNAs may change. It is also important to establish endogenous miRNA control and assess its consistency with the patient’s clinicopathological data to assess prognosis.

Advertisement

6. Conclusions

miRNA expression profiling for cancer diagnosis can be subdivided into miRNA expression patterns based on signaling common to cancer, such as cell proliferation, invasion, migration, and antiapoptotic effects, and miRNA expression patterns according to various organ or tumor differentiation levels and tissue types. In addition, the expression patterns of miRNAs in cancers of various organs are different according to cancer differentiation levels and tissue types. In addition, the degree of differentiation and tissue classification are often heterogeneous among cancers of various organs, and the estimation of miRNAs based on the differentiation of each organ or developmental stage and the miRNAs that may be targeted may be important determinants for organ identification in so-called primary unknown cancers. A vast number of miRNA expression analyses in various cancers are expected to be reported in the near future. As more cancer-specific miRNAs become known, they will be systematically classified and miRNA expression profiling consistent with conventional immunohistochemical staining of cancer tissues is expected to have practical application.

Advertisement

Acknowledgments

We thank Aya Sugimoto for her excellent research skills and support of our work.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Lee RC, Feinbaum RL, Ambros V, The C. Elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75(5):843-854
  2. 2. Wightman B, Ha I, Ruvkun G. Posttranscriptional regulation of the heterochronic gene lin-14 by lin-4 mediates temporal pattern formation in C. elegans. Cell. 1993;75(5):855-862
  3. 3. Reinhart BJ, Slack FJ, Basson M, Pasquinelli AE, Bettinger JC, Rougvie AE, et al. The 21-nucleotide let-7 RNA regulates developmental timing in Caenorhabditis elegans. Nature. 2000;403(6772):901-906
  4. 4. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391(6669):806-811
  5. 5. Zamore PD, Tuschl T, Sharp PA, Bartel DP. RNAi: Double-stranded RNA directs the ATP-dependent cleavage of mRNA at 21 to 23 nucleotide intervals. Cell. 2000;101(1):25-33
  6. 6. Pasquinelli AE, Reinhart BJ, Slack F, Martindale MQ , Kuroda MI, Maller B, et al. Conservation of the sequence and temporal expression of let-7 heterochronic regulatory RNA. Nature. 2000;408(6808):86-89
  7. 7. Fan R, Xiao C, Wan X, Cha W, Miao Y, Zhou Y, et al. Small molecules with big roles in microRNA chemical biology and microRNA-targeted therapeutics. RNA Biology. 2019;16(6):707-718
  8. 8. Griffiths-Jones S, Grocock RJ, van Dongen S, Bateman A, Enright AJ. miRBase: microRNA sequences, targets and gene nomenclature. Nucleic Acids Research. 2006;34, (Database issue):D140-D144
  9. 9. Rodriguez A, Griffiths-Jones S, Ashurst JL, Bradley A. Identification of mammalian microRNA host genes and transcription units. Genome Research. 2004;14(10A):1902-1910
  10. 10. Agueli C, Cammarata G, Salemi D, Dagnino L, Nicoletti R, La Rosa M, et al. 14q32/miRNA clusters loss of heterozygosity in acute lymphoblastic leukemia is associated with up-regulation of BCL11a. American Journal of Hematology. 2010;85(8):575-578
  11. 11. Bousquet M, Quelen C, Rosati R, Mansat-De Mas V, La Starza R, Bastard C, et al. Myeloid cell differentiation arrest by miR-125b-1 in myelodysplastic syndrome and acute myeloid leukemia with the t(2;11)(p21;q23) translocation. The Journal of Experimental Medicine. 2008;205(11):2499-2506
  12. 12. Calin GA, Dumitru CD, Shimizu M, Bichi R, Zupo S, Noch E, et al. Frequent deletions and down-regulation of micro- RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia. Proceedings of the National Academy of Sciences of the United States of America. 2002;99(24):15524-15529
  13. 13. Huse JT, Brennan C, Hambardzumyan D, Wee B, Pena J, Rouhanifard SH, et al. The PTEN-regulating microRNA miR-26a is amplified in high-grade glioma and facilitates gliomagenesis in vivo. Genes & Development. 2009;23(11):1327-1337
  14. 14. Li W, Duan R, Kooy F, Sherman SL, Zhou W, Jin P. Germline mutation of microRNA-125a is associated with breast cancer. Journal of Medical Genetics. 2009;46(5):358-360
  15. 15. Lin M, Gu J, Eng C, Ellis LM, Hildebrandt MA, Lin J, et al. Genetic polymorphisms in MicroRNA-related genes as predictors of clinical outcomes in colorectal adenocarcinoma patients. Clinical Cancer Research. 2012;18(14):3982-3991
  16. 16. Abelson JF, Kwan KY, O'Roak BJ, Baek DY, Stillman AA, Morgan TM, et al. Sequence variants in SLITRK1 are associated with Tourette's syndrome. Science. 2005;310(5746):317-320
  17. 17. Calin GA, Ferracin M, Cimmino A, Di Leva G, Shimizu M, Wojcik SE, et al. A MicroRNA signature associated with prognosis and progression in chronic lymphocytic leukemia. The New England Journal of Medicine. 2005;353(17):1793-1801
  18. 18. Calin GA, Sevignani C, Dumitru CD, Hyslop T, Noch E, Yendamuri S, et al. Human microRNA genes are frequently located at fragile sites and genomic regions involved in cancers. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(9):2999-3004
  19. 19. Bernstein E, Kim SY, Carmell MA, Murchison EP, Alcorn H, Li MZ, et al. Dicer is essential for mouse development. Nature Genetics. 2003;35(3):215-217
  20. 20. Kao SH, Cheng WC, Wang YT, Wu HT, Yeh HY, Chen YJ, et al. Regulation of miRNA biogenesis and histone modification by K63-Polyubiquitinated DDX17 controls Cancer stem-like features. Cancer Research. 2019;79(10):2549-2563
  21. 21. Morita S, Horii T, Kimura M, Goto Y, Ochiya T, Hatada I. One Argonaute family member, Eif2c2 (Ago2), is essential for development and appears not to be involved in DNA methylation. Genomics. 2007;89(6):687-696
  22. 22. Wang Y, Medvid R, Melton C, Jaenisch R, Blelloch R. DGCR8 is essential for microRNA biogenesis and silencing of embryonic stem cell self-renewal. Nature Genetics. 2007;39(3):380-385
  23. 23. Chapnik E, Sasson V, Blelloch R, Hornstein E. Dgcr8 controls neural crest cells survival in cardiovascular development. Developmental Biology. 2012;362(1):50-56
  24. 24. Chen Z, Wu J, Yang C, Fan P, Balazs L, Jiao Y, et al. DiGeorge syndrome critical region 8 (DGCR8) protein-mediated microRNA biogenesis is essential for vascular smooth muscle cell development in mice. The Journal of Biological Chemistry. 2012;287(23):19018-19028
  25. 25. da Costa Martins PA, Bourajjaj M, Gladka M, Kortland M, van Oort RJ, Pinto YM, et al. Conditional dicer gene deletion in the postnatal myocardium provokes spontaneous cardiac remodeling. Circulation. 2008;118(15):1567-1576
  26. 26. Gonzalez G, Behringer RR. Dicer is required for female reproductive tract development and fertility in the mouse. Molecular Reproduction and Development. 2009;76(7):678-688
  27. 27. Harris KS, Zhang Z, McManus MT, Harfe BD, Sun X. Dicer function is essential for lung epithelium morphogenesis. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(7):2208-2213
  28. 28. O’Donnell KA, Wentzel EA, Zeller KI, Dang CV, Mendell JT. c-Myc-regulated microRNAs modulate E2F1 expression. Nature. 2005;435(7043):839-843
  29. 29. Shin D, Shin JY, McManus MT, Ptacek LJ, Fu YH. Dicer ablation in oligodendrocytes provokes neuronal impairment in mice. Annals of Neurology. 2009;66(6):843-857
  30. 30. Tao J, Wu H, Lin Q , Wei W, Lu XH, Cantle JP, et al. Deletion of astroglial dicer causes non-cell-autonomous neuronal dysfunction and degeneration. The Journal of Neuroscience. 2011;31(22):8306-8319
  31. 31. Kumar MS, Lu J, Mercer KL, Golub TR, Jacks T. Impaired microRNA processing enhances cellular transformation and tumorigenesis. Nature Genetics. 2007;39(5):673-677
  32. 32. Yang Z, Jin P, Xu S, Zhang T, Yang X, Li X, et al. Dicer reprograms stromal fibroblasts to a pro-inflammatory and tumor-promoting phenotype in ovarian cancer. Cancer Letters. 2018;415:20-29
  33. 33. Chiosea S, Jelezcova E, Chandran U, Acquafondata M, McHale T, Sobol RW, et al. Up-regulation of dicer, a component of the MicroRNA machinery, in prostate adenocarcinoma. The American Journal of Pathology. 2006;169(5):1812-1820
  34. 34. Faber C, Horst D, Hlubek F, Kirchner T. Overexpression of dicer predicts poor survival in colorectal cancer. European Journal of Cancer. 2011;47(9):1414-1419
  35. 35. Guo X, Liao Q , Chen P, Li X, Xiong W, Ma J, et al. The microRNA-processing enzymes: Drosha and dicer can predict prognosis of nasopharyngeal carcinoma. Journal of Cancer Research and Clinical Oncology. 2012;138(1):49-56
  36. 36. Karube Y, Tanaka H, Osada H, Tomida S, Tatematsu Y, Yanagisawa K, et al. Reduced expression of dicer associated with poor prognosis in lung cancer patients. Cancer Science. 2005;96(2):111-115
  37. 37. Khoshnaw SM, Rakha EA, Abdel-Fatah TM, Nolan CC, Hodi Z, Macmillan DR, et al. Loss of dicer expression is associated with breast cancer progression and recurrence. Breast Cancer Research and Treatment. 2012;135(2):403-413
  38. 38. Lin RJ, Lin YC, Chen J, Kuo HH, Chen YY, Diccianni MB, et al. microRNA signature and expression of dicer and Drosha can predict prognosis and delineate risk groups in neuroblastoma. Cancer Research. 2010;70(20):7841-7850
  39. 39. Merritt WM, Lin YG, Han LY, Kamat AA, Spannuth WA, Schmandt R, et al. Dicer, Drosha, and outcomes in patients with ovarian cancer. The New England Journal of Medicine. 2008;359(25):2641-2650
  40. 40. Lu J, Getz G, Miska EA, Alvarez-Saavedra E, Lamb J, Peck D, et al. MicroRNA expression profiles classify human cancers. Nature. 2005;435(7043):834-838
  41. 41. Fujii T, Uchiyama T, Matsuoka M, Myojin T, Sugimoto S, Nitta Y, et al. Evaluation of DNA and RNA quality from archival formalin-fixed paraffin-embedded tissue for next-generation sequencing - retrospective study in Japanese single institution. Pathology International. 2020;70(9):602-611
  42. 42. Lawrie CH, Soneji S, Marafioti T, Cooper CD, Palazzo S, Paterson JC, et al. MicroRNA expression distinguishes between germinal center B cell-like and activated B cell-like subtypes of diffuse large B cell lymphoma. International Journal of Cancer. 2007;121(5):1156-1161
  43. 43. Lawrie CH, Gal S, Dunlop HM, Pushkaran B, Liggins AP, Pulford K, et al. Detection of elevated levels of tumour-associated microRNAs in serum of patients with diffuse large B-cell lymphoma. British Journal of Haematology. 2008;141(5):672-675
  44. 44. Weber JA, Baxter DH, Zhang S, Huang DY, Huang KH, Lee MJ, et al. The microRNA spectrum in 12 body fluids. Clinical Chemistry. 2010;56(11):1733-1741
  45. 45. Johnson SM, Grosshans H, Shingara J, Byrom M, Jarvis R, Cheng A, et al. RAS is regulated by the let-7 microRNA family. Cell. 2005;120(5):635-647
  46. 46. Calin GA, Cimmino A, Fabbri M, Ferracin M, Wojcik SE, Shimizu M, et al. MiR-15a and miR-16-1 cluster functions in human leukemia. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(13):5166-5171
  47. 47. Place RF, Li LC, Pookot D, Noonan EJ, Dahiya R. MicroRNA-373 induces expression of genes with complementary promoter sequences. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(5):1608-1613
  48. 48. Eiring AM, Harb JG, Neviani P, Garton C, Oaks JJ, Spizzo R, et al. miR-328 functions as an RNA decoy to modulate hnRNP E2 regulation of mRNA translation in leukemic blasts. Cell. 2010;140(5):652-665
  49. 49. Hwang HW, Wentzel EA, Mendell JT. A hexanucleotide element directs microRNA nuclear import. Science. 2007;315(5808):97-100
  50. 50. Forman JJ, Legesse-Miller A, Coller HA. A search for conserved sequences in coding regions reveals that the let-7 microRNA targets dicer within its coding sequence. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(39):14879-14884
  51. 51. Lytle JR, Yario TA, Steitz JA. Target mRNAs are repressed as efficiently by microRNA-binding sites in the 5' UTR as in the 3' UTR. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(23):9667-9672
  52. 52. Tay Y, Zhang J, Thomson AM, Lim B, Rigoutsos I. MicroRNAs to Nanog, Oct4 and Sox2 coding regions modulate embryonic stem cell differentiation. Nature. 2008;455(7216):1124-1128
  53. 53. Salmena L, Poliseno L, Tay Y, Kats L, Pandolfi PP. A ceRNA hypothesis: The Rosetta stone of a hidden RNA language? Cell. 2011;146(3):353-358
  54. 54. Vasudevan S, Tong Y, Steitz JA. Switching from repression to activation: microRNAs can up-regulate translation. Science. 2007;318(5858):1931-1934
  55. 55. Fukao A, Aoyama T, Fujiwara T. The molecular mechanism of translational control via the communication between the microRNA pathway and RNA-binding proteins. RNA Biology. 2015;12(9):922-926
  56. 56. Stavast CJ, Erkeland SJ. The non-canonical aspects of MicroRNAs: Many roads to gene regulation. Cell. 2019;8:11
  57. 57. Vasudevan S, Tong Y, Steitz JA. Cell-cycle control of microRNA-mediated translation regulation. Cell Cycle. 2008;7(11):1545-1549
  58. 58. Fujii T, Shimada K, Nakai T, Ohbayashi C. MicroRNAs in smoking-related carcinogenesis: Biomarkers, functions, and therapy. Journal of Clinical Medicine. 2018;7:5
  59. 59. Saikia M, Paul S, Chakraborty S. Role of microRNA in forming breast carcinoma. Life Sciences. 2020;259:118256
  60. 60. Shenouda SK, Alahari SK. MicroRNA function in cancer: Oncogene or a tumor suppressor? Cancer Metastasis Reviews. 2009;28(3-4):369-378
  61. 61. Svoronos AA, Engelman DM, Slack FJ. OncomiR or tumor suppressor? The duplicity of MicroRNAs in Cancer. Cancer Research. 2016;76(13):3666-3670
  62. 62. Volinia S, Calin GA, Liu CG, Ambs S, Cimmino A, Petrocca F, et al. A microRNA expression signature of human solid tumors defines cancer gene targets. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(7):2257-2261
  63. 63. Bedewy AML, Elmaghraby SM, Shehata AA, Kandil NS. Prognostic value of miRNA-155 expression in B-cell non-Hodgkin lymphoma. Turkish Journal of Haematology. 2017;34(3):207-212
  64. 64. Gierlikowski W, Broniarek K, Cheda L, Rogulski Z, Kotlarek-Lysakowska M. MiR-181a-5p regulates NIS expression in papillary thyroid carcinoma. International Journal of Molecular Sciences. 2021;22(11):6067
  65. 65. Iorio MV, Ferracin M, Liu CG, Veronese A, Spizzo R, Sabbioni S, et al. MicroRNA gene expression deregulation in human breast cancer. Cancer Research. 2005;65(16):7065-7070
  66. 66. Komoll RM, Hu Q , Olarewaju O, von Dohlen L, Yuan Q , Xie Y, et al. MicroRNA-342-3p is a potent tumour suppressor in hepatocellular carcinoma. Journal of Hepatology. 2021;74(1):122-134
  67. 67. Kong X, Xu R, Wang W, Zeng M, Li Y, Lin M, et al. CircularLRRC7 is a potential tumor suppressor associated with miR-1281 and PDXP expression in glioblastoma. Frontiers in Molecular Biosciences. 2021;8:743417
  68. 68. Liang G, Meng W, Huang X, Zhu W, Yin C, Wang C, et al. miR-196b-5p-mediated downregulation of TSPAN12 and GATA6 promotes tumor progression in non-small cell lung cancer. Proceedings of the National Academy of Sciences of the United States of America. 2020;117(8):4347-4357
  69. 69. Pellatt DF, Stevens JR, Wolff RK, Mullany LE, Herrick JS, Samowitz W, et al. Expression profiles of miRNA subsets distinguish human colorectal carcinoma and Normal colonic mucosa. Clinical and Translational Gastroenterology. 2016;7:e152
  70. 70. Shen P, Yang T, Chen Q , Yuan H, Wu P, Cai B, et al. CircNEIL3 regulatory loop promotes pancreatic ductal adenocarcinoma progression via miRNA sponging and A-to-I RNA-editing. Molecular Cancer. 2021;20(1):51
  71. 71. Sommerova L, Frankova H, Anton M, Jandakova E, Vojtesek B, Hrstka R. Expression and functional characterization of miR-34c in cervical Cancer. Klinická Onkologie. 2018;31(Suppl. 2):82-87
  72. 72. Wang ZY, Zhang W, Yang JJ, Song DK, Wei JX. Expression of miRNA-630 in bladder urothelial carcinoma and its clinical significance. Journal of Huazhong University of Science and Technology. Medical Sciences. 2016;36(5):705-709
  73. 73. Yang J, Lai CC, Xian ZM, Wang WQ , Xu BL. Preliminary results indicate increased expression of miR-184 in patients with renal carcinoma. European Review for Medical and Pharmacological Sciences. 2019;23(16):6878-6887
  74. 74. Yi D, Wang R, Shi X, Xu L, Yilihamu Y, Sang J. METTL14 promotes the migration and invasion of breast cancer cells by modulating N6methyladenosine and hsamiR146a5p expression. Oncology Reports. 2020;43(5):1375-1386
  75. 75. Zhang Q , Wang Y, Zhou Y, Zhang Q , Xu C. Potential biomarkers of miRNA in non-functional pituitary adenomas. World Journal of Surgical Oncology. 2021;19(1):270
  76. 76. Zhang R, Li F, Wang Y, Yao M, Chi C. Prognostic value of microRNA-20b expression level in patients with prostate cancer. Histology and Histopathology. 2020;35(8):827-831
  77. 77. Gurtner A, Falcone E, Garibaldi F, Piaggio G. Dysregulation of microRNA biogenesis in cancer: The impact of mutant p53 on Drosha complex activity. Journal of Experimental & Clinical Cancer Research. 2016;35:45
  78. 78. Babu KR, Tay Y. The Yin-Yang regulation of reactive oxygen species and MicroRNAs in Cancer. International Journal of Molecular Sciences. 2019;20(21):5335
  79. 79. Krell J, Stebbing J, Carissimi C, Dabrowska AF, de Giorgio A, Frampton AE, et al. TP53 regulates miRNA association with AGO2 to remodel the miRNA-mRNA interaction network. Genome Research. 2016;26(3):331-341
  80. 80. Kobayashi T, Papaioannou G, Mirzamohammadi F, Kozhemyakina E, Zhang M, Blelloch R, et al. Early postnatal ablation of the microRNA-processing enzyme, Drosha, causes chondrocyte death and impairs the structural integrity of the articular cartilage. Osteoarthritis and Cartilage. 2015;23(7):1214-1220
  81. 81. Sugatani T, Hruska KA. Impaired micro-RNA pathways diminish osteoclast differentiation and function. The Journal of Biological Chemistry. 2009;284(7):4667-4678
  82. 82. Rachagani S, Macha MA, Menning MS, Dey P, Pai P, Smith LM, et al. Changes in microRNA (miRNA) expression during pancreatic cancer development and progression in a genetically engineered KrasG12D;Pdx1-Cre mouse (KC) model. Oncotarget. 2015;6(37):40295-40309
  83. 83. Chen JC, Su YH, Chiu CF, Chang YW, Yu YH, Tseng CF, et al. Suppression of dicer increases sensitivity to gefitinib in human lung cancer cells. Annals of Surgical Oncology. 2014;21(Suppl. 4):S555-S563
  84. 84. Szczyrek M, Grenda A, Kuznar-Kaminska B, Krawczyk P, Sawicki M, Batura-Gabryel H, et al. Methylation of DROSHA and DICER as a biomarker for the detection of lung Cancer. Cancers (Basel). 2021;13(23):6139
  85. 85. Butkyte S, Ciupas L, Jakubauskiene E, Vilys L, Mocevicius P, Kanopka A, et al. Splicing-dependent expression of microRNAs of mirtron origin in human digestive and excretory system cancer cells. Clinical Epigenetics. 2016;8:33
  86. 86. Dudziec E, Miah S, Choudhry HM, Owen HC, Blizard S, Glover M, et al. Hypermethylation of CpG islands and shores around specific microRNAs and mirtrons is associated with the phenotype and presence of bladder cancer. Clinical Cancer Research. 2011;17(6):1287-1296
  87. 87. Sibley CR, Seow Y, Saayman S, Dijkstra KK, El Andaloussi S, Weinberg MS, et al. The biogenesis and characterization of mammalian microRNAs of mirtron origin. Nucleic Acids Research. 2012;40(1):438-448
  88. 88. Kawai S, Amano A. BRCA1 regulates microRNA biogenesis via the DROSHA microprocessor complex. The Journal of Cell Biology. 2012;197(2):201-208
  89. 89. Hayes J, Peruzzi PP, Lawler S. MicroRNAs in cancer: Biomarkers, functions and therapy. Trends in Molecular Medicine. 2014;20(8):460-469
  90. 90. Hussen BM, Hidayat HJ, Salihi A, Sabir DK, Taheri M, Ghafouri-Fard S. MicroRNA: A signature for cancer progression. Biomedicine & Pharmacotherapy. 2021;138:111528
  91. 91. Iorio MV, Croce CM. MicroRNA dysregulation in cancer: Diagnostics, monitoring and therapeutics. A comprehensive review. EMBO Mol Med. 2012;4(3):143-159
  92. 92. Lee YS, Dutta A. MicroRNAs in cancer. Annual Review of Pathology. 2009;4:199-227
  93. 93. Munker R, Calin GA. MicroRNA profiling in cancer. Clinical Science (London, England). 2011;121(4):141-158
  94. 94. Barker EV, Cervigne NK, Reis PP, Goswami RS, Xu W, Weinreb I, et al. microRNA evaluation of unknown primary lesions in the head and neck. Molecular Cancer. 2009;8:127
  95. 95. Ferracin M, Pedriali M, Veronese A, Zagatti B, Gafa R, Magri E, et al. MicroRNA profiling for the identification of cancers with unknown primary tissue-of-origin. The Journal of Pathology. 2011;225(1):43-53
  96. 96. Laprovitera N, Riefolo M, Porcellini E, Durante G, Garajova I, Vasuri F, et al. MicroRNA expression profiling with a droplet digital PCR assay enables molecular diagnosis and prognosis of cancers of unknown primary. Molecular Oncology. 2021;15(10):2732-2751
  97. 97. Varadhachary GR, Spector Y, Abbruzzese JL, Rosenwald S, Wang H, Aharonov R, et al. Prospective gene signature study using microRNA to identify the tissue of origin in patients with carcinoma of unknown primary. Clinical Cancer Research. 2011;17(12):4063-4070
  98. 98. Lan J, Sun L, Xu F, Liu L, Hu F, Song D, et al. M2 macrophage-derived exosomes promote cell migration and invasion in Colon Cancer. Cancer Research. 2019;79(1):146-158
  99. 99. Wang Y, Li J, Tong L, Zhang J, Zhai A, Xu K, et al. The prognostic value of miR-21 and miR-155 in non-small-cell lung cancer: A meta-analysis. Japanese Journal of Clinical Oncology. 2013;43(8):813-820
  100. 100. Xue X, Liu Y, Wang Y, Meng M, Wang K, Zang X, et al. MiR-21 and MiR-155 promote non-small cell lung cancer progression by downregulating SOCS1, SOCS6, and PTEN. Oncotarget. 2016;7(51):84508-84519
  101. 101. Yang M, Shen H, Qiu C, Ni Y, Wang L, Dong W, et al. High expression of miR-21 and miR-155 predicts recurrence and unfavourable survival in non-small cell lung cancer. European Journal of Cancer. 2013;49(3):604-615
  102. 102. Ma L, Teruya-Feldstein J, Weinberg RA. Tumour invasion and metastasis initiated by microRNA-10b in breast cancer. Nature. 2007;449(7163):682-688
  103. 103. Roth C, Rack B, Muller V, Janni W, Pantel K, Schwarzenbach H. Circulating microRNAs as blood-based markers for patients with primary and metastatic breast cancer. Breast Cancer Research. 2010;12(6):R90
  104. 104. Chen W, Cai F, Zhang B, Barekati Z, Zhong XY. The level of circulating miRNA-10b and miRNA-373 in detecting lymph node metastasis of breast cancer: Potential biomarkers. Tumour Biology. 2013;34(1):455-462
  105. 105. Bertoli G, Cava C, Castiglioni I. MicroRNAs: New biomarkers for diagnosis, prognosis, therapy prediction and therapeutic tools for breast Cancer. Theranostics. 2015;5(10):1122-1143
  106. 106. Rahman MM, Brane AC, Tollefsbol TO. MicroRNAs and epigenetics strategies to reverse breast Cancer. Cell. 2019;8:10
  107. 107. Lowery AJ, Miller N, Devaney A, McNeill RE, Davoren PA, Lemetre C, et al. MicroRNA signatures predict oestrogen receptor, progesterone receptor and HER2/neu receptor status in breast cancer. Breast Cancer Research. 2009;11(3):R27
  108. 108. Chen QY, Des Marais T, Costa M. Deregulation of SATB2 in carcinogenesis with emphasis on miRNA-mediated control. Carcinogenesis. 2019;40(3):393-402
  109. 109. Culig Z. miRNA as regulators of prostate carcinogenesis and endocrine and Chemoresistance. Current Cancer Drug Targets. 2021;21(4):283-288
  110. 110. Deng JH, Deng Q , Kuo CH, Delaney SW, Ying SY. MiRNA targets of prostate cancer. Methods in Molecular Biology. 2013;936:357-369
  111. 111. Liu Y, Ao X, Ding W, Ponnusamy M, Wu W, Hao X, et al. Critical role of FOXO3a in carcinogenesis. Molecular Cancer. 2018;17(1):104
  112. 112. Vlaeminck-Guillem V. Extracellular vesicles in prostate Cancer carcinogenesis, diagnosis, and management. Frontiers in Oncology. 2018;8:222
  113. 113. Che Y, Shi X, Shi Y, Jiang X, Ai Q , Shi Y, et al. Exosomes derived from miR-143-overexpressing MSCs inhibit cell migration and invasion in human prostate Cancer by downregulating TFF3. Mol Ther Nucleic Acids. 2019;18:232-244
  114. 114. Du C, Lv C, Feng Y, Yu S. Activation of the KDM5A/miRNA-495/YTHDF2/m6A-MOB3B axis facilitates prostate cancer progression. Journal of Experimental & Clinical Cancer Research. 2020;39(1):223
  115. 115. Fujii T, Shimada K, Tatsumi Y, Fujimoto K, Konishi N. Syndecan-1 responsive microRNA-126 and 149 regulate cell proliferation in prostate cancer. Biochemical and Biophysical Research Communications. 2015;456(1):183-189
  116. 116. Fujii T, Shimada K, Tatsumi Y, Tanaka N, Fujimoto K, Konishi N. Syndecan-1 up-regulates microRNA-331-3p and mediates epithelial-to-mesenchymal transition in prostate cancer. Molecular Carcinogenesis. 2016;55(9):1378-1386
  117. 117. Meng XF, Liu AD, Li SL. SNHG1 promotes proliferation, invasion and EMT of prostate cancer cells through miR-195-5p. European Review for Medical and Pharmacological Sciences. 2020;24(19):9880-9888
  118. 118. Wu M, Huang Y, Chen T, Wang W, Yang S, Ye Z, et al. LncRNA MEG3 inhibits the progression of prostate cancer by modulating miR-9-5p/QKI-5 axis. Journal of Cellular and Molecular Medicine. 2019;23(1):29-38
  119. 119. Fernandes RC, Toubia J, Townley S, Hanson AR, Dredge BK, Pillman KA, et al. Post-transcriptional gene regulation by MicroRNA-194 promotes neuroendocrine Transdifferentiation in prostate Cancer. Cell Reports. 2021;34(1):108585
  120. 120. Lu J, Mu X, Yin Q , Hu K. miR-106a contributes to prostate carcinoma progression through PTEN. Oncology Letters. 2019;17(1):1327-1332
  121. 121. Szczyrba J, Nolte E, Wach S, Kremmer E, Stohr R, Hartmann A, et al. Downregulation of Sec23A protein by miRNA-375 in prostate carcinoma. Molecular Cancer Research. 2011;9(6):791-800
  122. 122. Liu H, Hou T, Ju W, Xing Y, Zhang X, Yang J. MicroRNA122 downregulates Rhoassociated protein kinase 2 expression and inhibits the proliferation of prostate carcinoma cells. Molecular Medicine Reports. 2019;19(5):3882-3888
  123. 123. Ma J, Wei H, Li X, Qu X. Hsa-miR-149-5p suppresses prostate carcinoma malignancy by suppressing RGS17. Cancer Management and Research. 2021;13:2773-2783
  124. 124. Xiao H. MiR-7-5p suppresses tumor metastasis of non-small cell lung cancer by targeting NOVA2. Cellular & Molecular Biology Letters. 2019;24:60
  125. 125. Xu S, Yi XM, Zhang ZY, Ge JP, Zhou WQ. miR-129 predicts prognosis and inhibits cell growth in human prostate carcinoma. Molecular Medicine Reports. 2016;14(6):5025-5032
  126. 126. Zhang Y, Zhang W, Xu A, Tian Y, Liang C, Wang Z. MicroRNA-188 inhibits proliferation migration and invasion of prostate carcinoma by targeting at MARCKS. American Journal of Translational Research. 2019;11(8):5019-5028
  127. 127. Zhang Z. MiR-124-3p suppresses prostatic carcinoma by targeting PTGS2 through the AKT/NF-kappaB pathway. Molecular Biotechnology. 2021;63(7):621-630
  128. 128. Shimada K, Anai S, Fujii T, Tanaka N, Fujimoto K, Konishi N. Syndecan-1 (CD138) contributes to prostate cancer progression by stabilizing tumour-initiating cells. The Journal of Pathology. 2013;231(4):495-504
  129. 129. Iczkowski KA. Cell adhesion molecule CD44: Its functional roles in prostate cancer. American Journal of Translational Research. 2010;3(1):1-7
  130. 130. Liu C, Kelnar K, Liu B, Chen X, Calhoun-Davis T, Li H, et al. The microRNA miR-34a inhibits prostate cancer stem cells and metastasis by directly repressing CD44. Nature Medicine. 2011;17(2):211-215
  131. 131. Saini S, Majid S, Shahryari V, Arora S, Yamamura S, Chang I, et al. miRNA-708 control of CD44(+) prostate cancer-initiating cells. Cancer Research. 2012;72(14):3618-3630
  132. 132. Wang X, Cai J, Zhao L, Zhang D, Xu G, Hu J, et al. NUMB suppression by miR-9-5P enhances CD44(+) prostate cancer stem cell growth and metastasis. Scientific Reports. 2021;11(1):11210
  133. 133. Yan X, Tang B, Chen B, Shan Y, Yang H, Reproducibility Project: Cancer, B. Replication study: The microRNA miR-34a inhibits prostate cancer stem cells and metastasis by directly repressing CD44. eLife. 2019;8:e43511
  134. 134. Koga Y, Yamazaki N, Yamamoto Y, Yamamoto S, Saito N, Kakugawa Y, et al. Fecal miR-106a is a useful marker for colorectal cancer patients with false-negative results in immunochemical fecal occult blood test. Cancer Epidemiology, Biomarkers & Prevention. 2013;22(10):1844-1852
  135. 135. Koga Y, Yasunaga M, Takahashi A, Kuroda J, Moriya Y, Akasu T, et al. MicroRNA expression profiling of exfoliated colonocytes isolated from feces for colorectal cancer screening. Cancer Prevention Research (Philadelphia, Pa.). 2010;3(11):1435-1442
  136. 136. Link A, Balaguer F, Shen Y, Nagasaka T, Lozano JJ, Boland CR, et al. Fecal MicroRNAs as novel biomarkers for colon cancer screening. Cancer Epidemiology, Biomarkers & Prevention. 2010;19(7):1766-1774
  137. 137. Ng EK, Chong WW, Jin H, Lam EK, Shin VY, Yu J, et al. Differential expression of microRNAs in plasma of patients with colorectal cancer: A potential marker for colorectal cancer screening. Gut. 2009;58(10):1375-1381
  138. 138. Wu CW, Ng SS, Dong YJ, Ng SC, Leung WW, Lee CW, et al. Detection of miR-92a and miR-21 in stool samples as potential screening biomarkers for colorectal cancer and polyps. Gut. 2012;61(5):739-745
  139. 139. Liang Z, Li X, Liu S, Li C, Wang X, Xing J. MiR-141-3p inhibits cell proliferation, migration and invasion by targeting TRAF5 in colorectal cancer. Biochemical and Biophysical Research Communications. 2019;514(3):699-705
  140. 140. Long ZH, Bai ZG, Song JN, Zheng Z, Li J, Zhang J, et al. miR-141 inhibits proliferation and migration of colorectal Cancer SW480 cells. Anticancer Research. 2017;37(8):4345-4352
  141. 141. Tong SJ, Zhang XY, Guo HF, Yang J, Qi YP, Lu S. Study on effects of miR-141-3p in proliferation, migration, invasion and apoptosis of colon cancer cells by inhibiting Bcl2. Clinical & Translational Oncology. 2021;23(12):2526-2535
  142. 142. Xing Y, Jing H, Zhang Y, Suo J, Qian M. MicroRNA-141-3p affected proliferation, chemosensitivity, migration and invasion of colorectal cancer cells by targeting EGFR. The International Journal of Biochemistry & Cell Biology. 2020;118:105643
  143. 143. Ardila HJ, Sanabria-Salas MC, Meneses X, Rios R, Huertas-Salgado A, Serrano ML. Circulating miR-141-3p, miR-143-3p and miR-200c-3p are differentially expressed in colorectal cancer and advanced adenomas. Mol Clin Oncol. 2019;11(2):201-207
  144. 144. Ding M, Zhang T, Li S, Zhang Y, Qiu Y, Zhang B. Correlation analysis between liver metastasis and serum levels of miR200 and miR141 in patients with colorectal cancer. Molecular Medicine Reports. 2017;16(5):7791-7795
  145. 145. Ding T, Cui P, Zhou Y, Chen C, Zhao J, Wang H, et al. Antisense oligonucleotides against miR-21 inhibit the growth and metastasis of colorectal carcinoma via the DUSP8 pathway. Mol Ther Nucleic Acids. 2018;13:244-255
  146. 146. Igder S, Mohammadiasl J, Mokarram P. Altered miR-21, miRNA-148a expression in relation to KRAS mutation status as Indicator of adenoma-carcinoma transitional pattern in colorectal adenoma and carcinoma lesions. Biochemical Genetics. 2019;57(6):767-780
  147. 147. Li G, Wang Q , Li Z, Shen Y. Serum miR-21 and miR-210 as promising non-invasive biomarkers for the diagnosis and prognosis of colorectal cancer. Revista Española de Enfermedades Digestivas. 2020;112(11):832-837
  148. 148. Liang G, Zhu Y, Ali DJ, Tian T, Xu H, Si K, et al. Engineered exosomes for targeted co-delivery of miR-21 inhibitor and chemotherapeutics to reverse drug resistance in colon cancer. J Nanobiotechnology. 2020;18(1):10
  149. 149. Liu Q , Yang W, Luo Y, Hu S, Zhu L. Correlation between miR-21 and miR-145 and the incidence and prognosis of colorectal cancer. Journal of BUON. 2018;23(1):29-35
  150. 150. Shan L, Ji Q , Cheng G, Xia J, Liu D, Wu C, et al. Diagnostic value of circulating miR-21 for colorectal cancer: A meta-analysis. Cancer Biomarkers. 2015;15(1):47-56
  151. 151. Gomes SE, Simoes AE, Pereira DM, Castro RE, Rodrigues CM, Borralho PM. miR-143 or miR-145 overexpression increases cetuximab-mediated antibody-dependent cellular cytotoxicity in human colon cancer cells. Oncotarget. 2016;7(8):9368-9387
  152. 152. Kent OA, Fox-Talbot K, Halushka MK. RREB1 repressed miR-143/145 modulates KRAS signaling through downregulation of multiple targets. Oncogene. 2013;32(20):2576-2585
  153. 153. Li C, Yan G, Yin L, Liu T, Li C, Wang L. Prognostic roles of microRNA 143 and microRNA 145 in colorectal cancer: A meta-analysis. The International Journal of Biological Markers. 2019;34(1):6-14
  154. 154. Li JM, Zhao RH, Li ST, Xie CX, Jiang HH, Ding WJ, et al. Down-regulation of fecal miR-143 and miR-145 as potential markers for colorectal cancer. Saudi Medical Journal. 2012;33(1):24-29
  155. 155. Caritg O, Navarro A, Moreno I, Martinez-Rodenas F, Cordeiro A, Munoz C, et al. Identifying high-risk stage II Colon Cancer patients: A three-MicroRNA-based score as a prognostic biomarker. Clinical Colorectal Cancer. 2016;15(4):e175-e182
  156. 156. Karaayvaz M, Pal T, Song B, Zhang C, Georgakopoulos P, Mehmood S, et al. Prognostic significance of miR-215 in colon cancer. Clinical Colorectal Cancer. 2011;10(4):340-347
  157. 157. Milanesi E, Dobre M, Bucuroiu AI, Herlea V, Manuc TE, Salvi A, et al. miRNAs-based molecular signature for KRAS mutated and wild type colorectal Cancer: An explorative study. Journal of Immunology Research. 2020;2020:4927120
  158. 158. Slattery ML, Herrick JS, Mullany LE, Valeri N, Stevens J, Caan BJ, et al. An evaluation and replication of miRNAs with disease stage and colorectal cancer-specific mortality. International Journal of Cancer. 2015;137(2):428-438
  159. 159. Svoboda M, Sana J, Fabian P, Kocakova I, Gombosova J, Nekvindova J, et al. MicroRNA expression profile associated with response to neoadjuvant chemoradiotherapy in locally advanced rectal cancer patients. Radiation Oncology. 2012;7:195
  160. 160. Zhang JX, Song W, Chen ZH, Wei JH, Liao YJ, Lei J, et al. Prognostic and predictive value of a microRNA signature in stage II colon cancer: A microRNA expression analysis. The Lancet Oncology. 2013;14(13):1295-1306

Written By

Tomomi Fujii, Tomoko Uchiyama and Maiko Takeda

Submitted: 02 May 2022 Reviewed: 24 May 2022 Published: 22 June 2022