Next Article in Journal
MITF Promotes Cell Growth, Migration and Invasion in Clear Cell Renal Cell Carcinoma by Activating the RhoA/YAP Signal Pathway
Next Article in Special Issue
miRNA Clusters with Up-Regulated Expression in Colorectal Cancer
Previous Article in Journal
Management of T-Cell Lymphoma: In Quest of the Holy Grail
Previous Article in Special Issue
Low MicroRNA-19b Expression Shows a Promising Clinical Impact in Locally Advanced Rectal Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Clinical Assessment of MicroRNA Diagnostic, Prognostic, and Theranostic Value in Colorectal Cancer

by
Hussein Al-Akhrass
1,† and
Niki Christou
2,3,4,*,†
1
Turku Bioscience, Tykistökatu 6, FI-20520 Turku, Finland
2
Digestive Surgery Department, University Hospital Limoges, Avenue Martin Luther King, CEDEX 1, 87042 Limoges, France
3
EA3842 CAPTuR Laboratory “Cell Activation Control, Tumor Progression and Therapeutic Resistance”, Faculty of Medicine, 2 rue du Docteur Marcland, CEDEX 1, 87025 Limoges, France
4
Colorectal Surgery Department, University Hospitals Birmingham NHS Foundation Trust, Birmingham B152TH, UK
*
Author to whom correspondence should be addressed.
Both authors share the First co-author.
Cancers 2021, 13(12), 2916; https://doi.org/10.3390/cancers13122916
Submission received: 28 April 2021 / Revised: 29 May 2021 / Accepted: 9 June 2021 / Published: 11 June 2021
(This article belongs to the Special Issue microRNAs in Colorectal Cancer)

Abstract

:

Simple Summary

MiRNAs are of great interest within colorectal cancers in diagnosis, prognosis, and within the field of personalized treatments; they are present within different biological fluids such as blood and can lead to specific information for daily clinical use. Herein, we review the current literature focusing on miRNAs as potential diagnostic and prognostic biomarkers in patients treated for colorectal cancers. Detection and analysis of miRNA expression are cost-effective and lead to high sensitivity and specificity rates. However, it is now necessary to highlight the most sensitive and specific miRNAs for each goal, either diagnostic, prognostic, or theranostic, thanks to multicentric prospective studies.

Abstract

MiRNAs have recently become a subject of great interest within cancers and especially colorectal cancers in diagnosis, prognosis, and therapy decisions; herein we review the current literature focusing on miRNAs in colorectal cancers, and we discuss future challenges to use this tool on a daily clinical basis. In liquid biopsies, miRNAs seem easily accessible and can give important information toward each step of the management of colorectal cancers. However, it is now necessary to highlight the most sensitive and specific miRNAs for each goal thanks to multicentric prospective studies. Conclusions: by their diversity and the feasibility of their use, miRNAs are getting part of the armamentarium of healthcare management of colorectal cancers.

1. Introduction

Colon and rectal cancers (CRCs) are respectively the 5th and 8th cancers in terms of incidence worldwide, and they represent the fourth most cancer-related deaths [1]. The most frequent histological type is Liberkhünien adenocarcinomas. The carcinogenesis of such tumors can be explained by several mutational and signaling network alterations. For instance, we can quote different alterations of oncogenes or tumor suppressor genes linked to signaling pathway changes such as within the Wnt (portmanteau created from the names Wingless and Int-1) [2] signaling pathway but also receptor tyrosine kinase pathways leading to cellular proliferation, DNA repair, cell cycle arrest, and apoptosis pathway inactivation [3].
Currently, miRNAs have been an ever-increasing topic of interest. MiRNAs belong to the RNA interference family originally discovered in 1998 by Andrew Fire and Craig Mello [4] (2006 Nobel Prize laureates in Physiology or Medicine). Their analysis is cost-effective and encompasses high levels of both sensitivity and specificity. They are short non-coding strands of RNA comprising around 19 to 22 nucleotides [5]. Unsurprisingly, miRNAs play a fundamental role in the regulation of physiological processes such as embryogenesis [6] as well as several human pathologies such as cancer, auto-immune, and cardiovascular diseases [7,8,9,10]. Until now, around 2600 mature miRNAs in the human genome have been discovered and are accessible through publicly available databases such as miRBase v.22 [11]. MiRNAs exhibit a broad regulatory potential since more than 60% of protein-coding mRNAs contain highly conserved miRNA-binding sequences [12]. Generally, miRNAs inhibit gene expression through binding to the untranslated region within the 3’ end of messenger RNAs (mRNAs) [13].
Different publications have shown miRNA release into fluids such as stools [14] and blood [15]. Extracellular miRNAs can be transported to target cells through several transport pathways. We can quote extracellular vesicles, such as exosomes, or the mechanism by interaction with proteins like argonaute 2 (Ago2) as described in human plasma analysis [15], high density lipoprotein (HDL), and nucleophosmin 1 (NPM 1). These associations are essential for them in order to avoid any digestion from RNase [16]. It is worth noting that the carrier of one miRNA may not be the only one. Furthermore, different carriers can exist for one miRNA and depend on the tissue specificity and cell type that are going to be targeted by the miRNA [17].
Here, we briefly introduce the biosynthesis of miRNAs, and we review the current clinical achievements related to their potential roles as diagnostic and prognostic biomarkers in CRCs.

2. Canonical miRNA Biosynthesis

The canonical miRNA biosynthesis pathway begins with the transcription of a primary miRNA (pri-miRNA) forming a hairpin structure. Pri-miRNA is subsequently processed into an miRNA precursor (pre-miRNA) by the nuclear microprocessor complex formed by the RNA-binding protein DiGeorge Critical Region 8 (DGCR8) and the ribonuclease 3 enzyme Drosha [18]. This step results in the cleavage of pri-miRNA at the base of the hairpin structure. Exportin-5 transports the pre-miRNA to the cytoplasm where it is processed by the RNase III endonuclease Dicer that eliminates the loop sequence, thus resulting in a mature miRNA duplex [19,20]. The directionality of each strand determines its name. Mature miRNAs 5p and 3p arise from the 5′ and 3′ ends, respectively. Ultimately, the miRNA duplex is unwound, and the remaining single-stranded miRNA is loaded into the Argonaute complex in an ATP-dependent manner, forming a central component of the RNA-induced silencing complex (RISC) [21,22]. It is worth noting that multiple Drosha/DGCR8- and Dicer-independent biosynthesis pathways can generate non-canonical miRNAs (reviewed in [23]).

3. RNAs Regulate Fundamental Processes of CRC Growth

In this section, we highlight the history of miRNA research in relation to CRCs and we review key fundamental roles of miRNAs in regulating CRC signaling. Without being exhaustive, we provide a focus on signaling pathways frequently deregulated in CRCs and underpinning essential processes for CRC progression such as the epithelial to mesenchymal transition (EMT), cancer angiogenesis, and uncontrolled cell division.
First of all, it is important to mention that colon and rectal cancer arise from different genetic and epigenetic alterations within normal colonic and rectal tissues. At the molecular scale, the development of an adenoma followed by apparition of a CRC corresponds to the progressive accumulation of gene mutations within the nucleus of colonic epithelial cells: activation of oncogenes and inactivation of tumor suppressor genes.
There are two main pathways of colorectal carcinogenesis that result from genetic instability: one, the most common, is detected at the chromosomal level (chromosomal instability, “CIN”); the second is at the nucleotide level (instability of the loci of microsatellites). These two distinct pathways result in morphologically similar lesions (adenomas) but with a different progression to cancer, more important in the path of micro satellite instability.
Regarding “CIN” tumors, they present a characteristic set of mutations on the tumor suppressor gene APC as well as on the proto-oncogene Kirsten RAS (KRAS), activating the initiation and progression signaling pathways of CRC [24]. In 2008, Nagel et al. showed that miR-135a&b was able to target the 3′ untranslated region of APC. This led to the abolishment of its expression, with induction of downstream Wnt pathway activity [25]. Wang et al., in 2016, proved that miR-384 repressed KRAS expression by directly targeting its 3’untranslated region [26].
The second pathway of colorectal carcinogenesis is the molecular mechanism concerning 15% of sporadic CRCs and is typically observed in Lynch syndrome (HNPCC: 3% of cases). These cancers are called replication errors (RER+) or microsatellite instabilities (MSI). Cancer cells have normal DNA content (diploidy) and have no chromosomal loss but have abnormalities in DNA mismatch repair (MMR) genes. The components of this repair system include: ATPases hMSH2, hMSH6, hMSH3, hMLH1, hPMS2, hPMS1, and hMLH3 [27]. These genes code for proteins whose role is to detect and repair DNA replication errors that occur during mitosis. Mutation or methylation of the promoter region of MMR genes induces a deficiency in this repair system and mutations will accumulate, preferably at the level of microsatellites, regions of the genome particularly prone to replication errors. The cell acquires a “hypermutator” phenotype predisposing to the occurrence of mutations in certain oncogenes (pro-apoptotic BAX genes) or tumor suppressor genes (genes encoding the type II receptor of TGF β). The chronology of RER+ cancer mutations is poorly understood: the mutation of the type II TGF β receptor appears to be the earliest. In 2010, Earle et al. demonstrated an association between miRNA expression in CRCs with MSI subgroups, underlining post-transcriptional gene regulation by miRNAs [28].
The more recent work of Slattery et al. in 2016 tried to determine if there was a specific molecular pattern linked to specific expressions of particular miRNAs [29]. A total of 1893 CRC samples from patients based in the United States were analyzed. They found that MSI tumors have the majority of significantly different miRNAs expressions. Indeed, few different expressions of miRNAs by TP53, KRAS, CpG island methylator phenotype, and BRAF molecular phenotype for either colon or rectal carcinoma were highlighted. On the contrary, 94 miRNAs were differentially expressed between MSI and MSS tumors for colon carcinomas and 41 miRNAs for rectal carcinomas.
In 2003, Michael et al. highlighted the first implication of miRNAs in CRCs. This study showed a downregulation of miR-143 and miR-145 in CRC specimens. Interestingly, the authors did not observe altered levels of miRNA precursors, indicating a post-transcriptional regulation of their expression in CRC cells. The inhibited expression of miR-143 and miR-145 correlates with enhanced cancer cell proliferation [30]. MiR-145 inhibits the expression of multiple targets essential for sustained CRC growth such as insulin receptor substrate 1 (IRS-1), the proto-oncogene c-Myc, the proto-oncogene Yamaguchi sarcoma viral oncogen homolog 1 (YES-1), and signal transducers and activators of transcription 1 (STAT1) [31,32,33]. MiR-143 expression was found to negatively correlate with CRC metastasis [34]. MiR-143 inhibits key targets essential for CRC growth such as the GTPase KRAS, the DNA (cytosine-5)-methyltransferase 3A (DNMT3A), the receptor tyrosine-protein kinase erbB-3, and the mitogen-activated protein kinase 7 (MAPK7) [31,32]. In addition, other miRNAs regulators of CRC growth have been described. For example, overexpression of miR-21 was reported in CRC tissue (the clinical value of miR-21 will be discussed below) [35]. Also, miRNAs from the cluster miR-17-92 were found to be overexpressed in a broad spectrum of human cancers including CRCs. Using an engineered mouse model overexpressing this miRNA cluster, Huabin et al. demonstrated that these miRNAs inhibit CRC progression by targeting key inducers of cancer angiogenesis such as the vascular endothelial growth factor A (VEGFA) [36]. Other miRNAs, such as members of the let-7 family of miRNAs, are downregulated in CRCs [37] (the clinical value of let-7 will be reviewed below).
In 2006, Cummins et al. initially characterized the CRC miRNome. In the study, authors identified 133 miRNAs specific to CRC tissue compared to non-tumor tissue. MiR-143 and miR-145 were identified as parts of this CRC-specific miRNA signature [38]. In 2012, The Cancer Genome Atlas Network performed a comprehensive molecular characterization of CRCs [39]. The study conducted genome-scale analyses, including an miRNA expression profile, using 276 CRC samples that were compared to healthy mucosa. The expression of multiple miRNAs was found to be altered in CRCs as well as to correlate with CRC aggressiveness.
MiRNAs are thought to contribute to the initiation of CRCs. MiRNA-135a and miRNA-135b [25] are overexpressed in CRCs and inhibit the expression of the tumor suppressor adenomatous polyposis coli (APC), a key component in a cytoplasmic complex that inhibits the Wnt/β-catenin signaling pathway [40]. This pathway is aberrantly activated in CRCs and contributes to cell transformation upon APC mutation or loss of function [40]. Therefore, a role of miR-135a and miR-135b in the early steps of CRC oncogenesis could be speculated.
The transcription factor p53 is the most studied transcription factor and tumor suppressor. Generally, p53 activation is triggered as a response to DNA damage. Activation of p53 blocks cell cycle progression, thus allowing mechanisms of DNA repair to take place [41]. Deregulation of p53 expression was reported in ~60% of CRC samples [41]. A miRNA-related p53-mediated mechanism was reported in CRCs. Chang et al. discovered that p53 directly activates the transcription of miRNA-34a. MiRNA-34a is a tumor suppressor that triggers apoptosis and mimics the effects of p53 by regulating cell cycle progression, DNA repair, and angiogenesis [42].
The transcription factors zinc finger E-box binding homeobox 1 (ZEB1) and ZEB2 promote the epithelial to mesenchymal transition (EMT), a key process in the onset of metastasis. A mechanism of mutual inhibition linking ZEB1/2 to miRNA-200 was reported. ZEB1/2 inhibit miRNA-200 expression and miRNA-200 inhibits ZEB1/2. Therefore, an inhibited expression of miR-200 induces an accumulation of ZEB1/2 and correlates with enhanced CRC aggressiveness [43,44,45].
We performed a literature search within Pubmed between 2011 and 2021 with the keywords “miRNAs and colorectal cancer pathways” or “miRNAs and colorectal cancer signaling”. After finding the most relevant works aimed at demonstrating the role of some miRNAs targeting specific “signaling pathways or proteins” involved in the carcinogenesis of CRCs, we organized each publication below in Table 1.

4. MicroRNAs as a Diagnosis Biomarker in CRCs

The last decade witnessed key achievements in understanding the origin and functions of miRNAs. These studies pointed out the potential interest of miRNAs in clinical practice for patients in healthy or diseased conditions. Thus, they are linked to biological processes such as cell differentiation, development of the nervous system, muscle development, the cell cycle, cell signaling, immune response, angiogenesis, apoptosis, and autophagy [54]. On the other hand, we can highlight them within different pathological processes. Consequently, various fields of implementation of miRNAs within medicine emerged notably in terms of CRC biomarkers. Indeed, different studies have suggested that they could be appealing biomarkers for cancer through vesicular trafficking communication [55], in cardiology to diagnose heart failure [56], or more recently in neurology for the diagnosis and prognosis of neurodegenerative pathologies [57].
Performing a literature search with the keywords “miRNA and colorectal cancer and diagnostic biomarker” between 2011 and 2021 within Pubmed encompassing only clinical trials and randomized control trials in English, we were able to highlight 23 articles. After screening them, five had to be considered for review focusing on miRNAs as a diagnosis biomarker in CRCs.
They are summarized in the table below (Table 2).
First, miRNAs can help detect CRCs at an early stage and have become part of the screening process. As blood search within stools is currently widely used for CRC screening (fecal immunotest called “FIT”) [63], Wu et al. evaluated miRNA levels within stool samples [58]. The authors showed that miRNA-92a stool levels for patients without cancers were lower than those with CCRs. Furthermore, miRNA-92a was also sensitive and specific for the detection between polyps (higher levels) and patients without CRCs. These results can be explained by the function of miRNA-92a, which is a member of the cluster miRNA-17-92 located at chromosome 13q13 [64]. More precisely, it generally promotes cell proliferation, leading to tumor progression, suppresses apoptosis of cancer cells, and is able to induce tumor angiogenesis. MiRNA-92a was found to be upregulated in CRCs. It interacts with its target PTEN (phosphatase and tensin homolog) via the PTEN/PI3K (phosphoinositide 3-kinase)/Akt (protein kinase B) signaling pathway, triggering a decrease of E-cadherin expression, a biomarker of epithelial-to-mesenchymal transition (EMT) [65].
Precancerous lesions such as adenomas that are diagnosed in colonoscopies can also be detected based on the expression of miRNAs in serum as demonstrated by Zheng et al. [62]. In the study, the authors highlighted miRNA-191-5p and U6 as the most stable pair of reference genes (controls) in colorectal adenocarcinoma, colorectal adenoma, and healthy controls for serum microRNA qPCR analysis. Thus, using these controls, they demonstrated that serum miRNA-92a-3p was significantly higher in colorectal adenocarcinoma patients than colorectal adenoma patients and healthy controls (p < 0.001). Moreover, the difference in miRNA-92a levels was significant between colorectal adenoma patients and healthy controls (p < 0.001).
Secondly, miRNAs are great candidates for the diagnosis of CRCs such as different studies highlighted it irrespective of the sample type (serum or plasma) [59,60,61]. Both sensitivity and specificity reached high rates between 70% and 95%. It is also worth noting that in the last few years miRNAs have been sought within exosome-enriched fractions within serum [66]. Indeed, these nanovesicles are secreted by either normal or cancerous cells leading to the transport of different molecules such as proteins or RNAs [67]. As RNAs in exososmal cargo are drastically different between healthy and cancer-stricken patients, they can be indicators of the presence of transformed cells. The study of Ogata-Kawata et al. has shown that miRNA-23a, miRNA-1246, and miR-21 present a diagnostic accuracy for CRCs. The area under the curve (AUC) in receiver operating characteristic (ROC) curves were around 0.953, 0.948, and 0.798, respectively [68].
A recent meta-analysis [69] encompassing 35 studies published from 2009 to 2019 and including 3258 CRC patients and 2683 healthy controls analyzed serum-derived miRNA panels. More precisely, it focused on multiple miRNA serum-derived panels (and not single-miRNA assays), demonstrating them as promising markers of early diagnosis of CRCs. However, it is important to stress that there is a high degree of heterogeneity between studies: most studies integrated subjects from the Asian continent, some combined carcinoembryonic antigen (CEA) measures to miRNAs (which may improve diagnostic efficiency of early CRC detection).

5. MicroRNAs as a Prognosis Biomarker in CRCs (CRC Risk)

We searched within Pubmed “miRNA and colorectal cancer and prognosis biomarker” for articles between 2011 and 2021. Only clinical trials and randomized control trials in English were integrated. We found 20 research articles, which were carefully screened, resulting in 12 remaining studies for review focusing on miRNAs as prognostic and predictive biomarkers in CRCs (Table 3).
Regarding the prognosis, while the study of Schou et al. demonstrated higher rates of miR-345 in whole blood linked to shorter overall survival (OS) [73], the work of Hansel et al. [75] reported further findings. They showed that high tumor expression of miRNA-126 was significantly related to a longer progression-free survival (PFS). MiRNAs have thus different actions and links within the molecular signaling pathways of the initiation and maintenance of carcinogenesis. Consequently, their modifications’ rates show indirectly their effects on these different pathways and especially in terms of survival: for instance, miR-21 targets the genes involved in the PI3K pathway, which is downstream from the EGFR pathway regulating tumor growth, angiogenesis, and metastasis [81]. Thus, miR-21 reflects the KRAS signaling, and high levels are found in KRAS wild-type patients predicting increased response to anti-EGFR therapy (cetuximab) [82]. Furthermore, as described by the meta-analysis of Peng et al. [83], higher expression of miRNA-21 in patients with CRCs correlated with significantly shorter disease-free survival.
Due to the ever-increasing interest in exosomes, miRNAs derived from exosome-enriched fractions are used to define the prognosis of CRCs [84]. Beyond that, they are also able to distinguish chemoresistance and predict recrudescence. For example, Yagi et al. in 2019 found that plasma exosomal microRNA-125b may have been a monitoring biomarker of resistance to mFOLFOX6-based chemotherapy in 55 patients with advanced and recurrent CRCs.
Targeted therapy has dramatically improved patients’ outcomes. However, drug resistance mechanisms leading to tumor recurrences remain a major clinical challenge [85]. Consequently, different studies have demonstrated the role of miRNAs as predictor markers of response to antiepidermal growth factor receptor (EGFR) monoclonal antibodies (cetuximab and panitumumab) [71,73,77,79], antiangiogenic therapy [74], chemotherapies (irinotecan, 5-fluorouracile) [78], and immunotherapy [76]. The main chemotherapy drugs used to treat CRCs are 5-fluorouracile and oxaliplatin. A study by Liu et al. [78] demonstrated the value of miR-21 in predicting response to 5-FU and oxaliplatin treatments. This was subsequently demonstrated by in vitro studies with CRC cell lines HT-29 [86]. MiRNAs can lead to high specificity when used in the development of personalized treatments, as was shown in the study of Sclafani et al. [80], which underlined such use with neoadjuvant treatment in the particular entity of rectal cancers.
In metastatic CRCs, clinical studies revealed a correlation between miRNA expression profiles and therapeutic response to EGFR- and vascular endothelial growth factor (VEGF)-targeted therapies.
Clinical studies of miRNAs in anti-EGFR therapy-treated CRC patients assessed only the miRNA expression in tumor tissue. In RAS wild-type metastatic CRC patients receiving anti-EGFR monoclonal antibody cetuximab and fluoropyrimidine-based therapy, upregulation of miR-31-5p and miR-31-3p correlated with worse progression-free survival. In the same setting of drug combinations and RAS wild-type metastatic CRC patients, decreased expression of miR-592 and miR-181a also correlated with worse PFS [87,88,89]. Furthermore, elevated expression of miRNA-140-5p and decreased expression of miR-1224-5p and miR-181a correlated with worse overall survival (OS) in these patients [87,89]. In KRAS-mutated and BRAF wild-type metastatic CRC patients receiving a cetuximab- and irinotecan-based third-line therapy, increased expression of miR-let-7a correlated with improved OS and PFS. This correlation was enhanced in patients with conserved miR-let-7a-binding sequence in KRAS-coding mRNA, suggesting that patients with high tumor expression of miR-let-7a may benefit from an anti-EGFR therapy irrespective of KRAS mutational status [90].
An original study revealed that in metastatic CRC patients receiving an anti-VEGF monoclonal antibody bevacizumab and capecitabine plus oxaliplatin therapy, overexpression of miR-664-3p and downregulation of miR-455-5p correlated with improved OS and PFS [91]. Subsequently, other clinical studies showed that elevated cancer tissue and plasmatic levels of miR-126 prior to bevacizumab therapy associated with improved therapeutic response [74,92]. Other circulating miRNA signatures were associated with improved response to bevacizumab. For example, an enriched basal plasmatic expression of miR-92b-3p, miR-3156-5p, miR-10a-5p, and miR-125a-5p predicted improved PFS in bevacizumab-treated metastatic CRC patients [93]. Moreover, elevated basal expression of miR-20b-3p, miR29b-3p, and miR-155-5p correlated with improved OS and PFS. However, a further increase of miR-155-5p expression at one month of bevacizumab treatment correlated with worse prognosis, highlighting an original prognostic value of the dynamic expression of circulating miRNAs in metastatic CRCs [94].

6. MicroRNAs as Therapeutics in CRCs

MiRNAs regulate the progression of various human diseases such as cancer, immune disorders, cardiovascular diseases, Alzheimer’s disease, and rheumatoid arthritis [95]. Recent studies highlighted different therapeutic miRNAs in several diseases. For instance, the liver-expressed miRNA-122, using the locked nucleic acid (LNA)–modified antisense oligonucleotide miravirsen, has been used in a phase II clinical trial as a potential treatment for hepatitis C viral infection (HCV) [96]. In this trial, 36 patients with chronic HCV genotype 1 infection were enrolled and randomized to three cohorts: 9 miravirsen-treated and 3 placebo-treated subjects in each cohort. They demonstrated that a four-week miravirsen treatment provides long-lasting suppression of the virus, with a good tolerance for patients with chronic HCV infection.
In cancer, miRNAs regulate several cellular processes such as cell proliferation, invasion, and metastasis. There are two potential therapeutic ways to restore the physiological miRNA expression in tumor cells: (1) by inhibiting miRNA activity in case of oncogenic miRNA overexpression; (2) by rescuing an miRNA activity in the situation of a downregulated tumor suppressor miRNA.
For the inhibition of miRNAs, different tools can be used such as locked nucleic acids (LNA), antisense anti-miR oligonucleotides (AMO), miRNA sponges, and miRNA antagomirs [97].
-
Locked nucleic acids (LNA), or “bridged nucleic acids (BNAs)”, are modified RNAs with a 2’sugar modification and act as anti-miRNAs. Due to their structure, they have high stability and affinity [98],
-
Antisense anti-miRNA oligonucleotides (AMO) are oligonucleotides with a complementary sequence used to neutralize miRNAs and their (dys)functions [99],
-
MiRNA sponges, or “sponge-miR-mask technology”, are used to prevent several miRNAs from binding. However, these masks have low specificity regarding gene blockage [100],
-
MiRNA antagomirs, or “anti-miRs “or “blockmirs”, are oligonucleotides designed to block molecules from binding to a specific site on miRNAs [101].
Conversely, for the restoration of miRNA activity, viral vectors expressing miRNAs or synthetic miRNA mimics [102] may be employed. However, while in specific cancers, such as mesothelioma, treatments based on miRNAs have been established [103,104], no clinical studies demonstrating miRNAs as effective therapeutics have been highlighted within the international literature on CRCs.
In 2016, Christopher et al. underlined the whole process from the discovery of a potential miRNA signature of a disease to clinical trials validating miRNA therapeutic development [105]. As a consequence, four steps have to be followed:
-
Identification of miRNA signatures of a specific disease.
-
Validation of miRNA signatures within gain- and loss-of-function studies in vitro and in vivo.
-
Pharmacologic analysis of in vivo delivery studies with pharmacodynamic and pharmacokinetic analysis.
-
Efficacy and toxicity analysis at a large scale within human clinical trials (phase I to phase III) aimed at approval for therapeutic use.
In the published literature on CRCs, only in vitro and in vivo studies have been found.
Akao et al. examined the miR-143 and miR-145 expression levels in 63 CRC and 65 colon and rectum adenoma specimens with paired non-cancerous tissues [106]. Their expressions were shown to be downregulated in the early phase of carcinogenesis. Thus, with the clinical therapeutic goal, they chemically modified these miRNAs to increase their activity and stability. Injections of modified synthetic miR-143 were performed within nude-mice xenografted tumors derived from DLD-1 human CRC cells, directly first in the tumors and then in the second experiment into veins. Tumor sizes and weights decreased after these injections, demonstrating the tumor-suppressive effect of chemically modified synthetic miR-143 in vivo. Zhang et al. [107] demonstrated that miR-34a was downregulated in SW480 CRC cell lines. Transfection of the cells with miR-34a mimic showed, consequently, attenuation of both cell migration and invasion.
Previous in vitro studies have the stressed pro-apoptotic and anti-proliferative roles of miR-145 [108] and repression activity of miR-33a on the proto-oncogene Pim-1 action thus as a tumor suppressor inhibiting cell-cycle progression [109]. Consequently, Ibrahim et al. [110], in 2011, delivered these two miRNAs through a mouse model to validate their therapeutic interest. They used a specific method of delivery with a polyethylenimine (PEI)-mediated delivery of unmodified miRNAs. It resulted in reduced tumor proliferation and increased apoptosis through these chemically unmodified miRNAs complexed with PEI. More recently, Karimi et al. [111] restored miR-143 expression [34] by transfection of the pCMV-miR-143 vector into the SW480 CRC cells. Viability and migratory potential decreased after such a transfection. Hejazi et al. [112] combined a treatment with miR-193a mimics (miR-193a-5p known to be downregulated in CRCs and correlated with advanced stages [113]) and Taxol (antimicrotubule agent) in HT-29 CRC cell lines. Inhibition of migration and colony formations was observed, suggesting miR-193a replacement combined with Taxol chemotherapy as a potential treatment for CRCs.
More recently, in vitro studies have tried to tackle the topic of chemoresistance. Thus, Liu et al. [114], transfected oxaliplatin-resistant CRCs and normal intestinal fetal human cells with an miR-128-3p expression lentivirus. They conducted both in vitro and in vivo studies, highlighting an improvement in oxaliplatin response. Sun et al. [115], in 2019, infected DLD-1 and Caco-2 CRC cells with lentiviruses carrying miR-302a or targeted miR-302a (anti-miR-302a). When miR-302a was overexpressed, both migration and invasion were decreased with restoration of cetuximab response contrasting with cells expressing an increase in miR-302a downregulation.
Therefore, miRNAs appear to be attractive targets to treat CRCs, whatever the stage of the disease and considering its molecular specificities such as resistance to conventional chemotherapies use. A lot of patents from Europe and the USA have been issued miRNA treatments in the past the last decade. However, as indicated above, these treatments have focused on viral infections such as HCV (US Patent No. 7,307,067) and glioblastoma (United States Patent 8778903) [116]. We hope that in the next few years, following the interesting in vitro studies that we described earlier, a multiplication of clinical trials will be initiated. It will be easier as administration modes of miRNAs with effective pharmacodynamic and pharmacokinetic properties have improved [117]. Systemic delivery can be performed, but also intratumor injections with increased specificity and efficacy, with a minimization of side effects [118] (Table 4).
Indeed, thanks to both bioavailability and decreased toxicity, the intratumoral injection of miRNAs can be more effective than systemic delivery. One main advantage of local delivery is minimal nonspecific uptake by healthy organs avoiding toxicity and immunogenicity. The polyethylenimine (PEI)-mediated local application of miR-145 demonstrated antitumor effects in CRC mouse models [110]. Similarly, intratumoral injection of nanoparticles of siRNAs demonstrated efficiency in CRC nude-mice xenografts [119]. Nevertheless, local administration can be difficult at a human level and needs exploration, especially at early disease stages with the possible use of endoscopy. Moreover, it is important to underline that this strategy seems at first glance limited to the localized and accessible tumor. Thus, when CRCs are spreading, systemic delivery has to be assessed. This route delivery should protect miRNAs from early degradation in the blood circulation. In addition, it has to carry them to the target cells and facilitate cellular uptake without inducing any immunogenic response [120]. Thus, oligonucleotide editing [121] and systemic miRNA delivery systems such as LNAs, which are chemically modified RNA nucleotide analogs with high nuclease resistance [122], have been examined and are still being explored.

7. Conclusions

In last decade, miRNAs have shown increasing value as potential diagnostic and prognostic biomarkers in CRCs. Different studies both at fundamental and clinical levels are the basis of these advances. It can be highlighted that there is no one but different couples of miRNAs allowing these functions. However, analyses at a large scale are necessary to show with both sensitivity and specificity which couple has to be used in daily clinical care.
More precisely, miRNAs are major post-transcriptional regulators. Their deregulation participates in the process of colorectal carcinogenesis. To date, the use of miRNAs as diagnostic biomarkers for the screening of CRCs appears promising but requires further research. The deregulation of some miRNAs can identify colorectal tumors at risk of recurrence. Moreover, some deregulated miRNAs take part in chemoresistance mechanisms. The identification of miRNA biomarkers of resistance to anti-EGFR or anti-VEGF antibodies would allow for the optimization of the management of patients with metastatic CRCs. The therapeutic use of miRNAs is a very recent field of application. Different clinical trials have highlighted their impact in treating lung cancer, glioblastoma, or breast cancer but not yet colorectal cancers. Moreover, to be completely efficient, several optimizations are required.
Challenges are to be considered and taken into consideration toward specificity, stability, pharmacodynamics and pharmacokinetics, immune activation, and efficacity toxicity in both in vivo and in vitro studies. Another important challenge in the next few years will be to integrate miRNAs within the therapeutic arsenal of personalized-medicine strategies for treatment of CRCs.

Author Contributions

Conceptualization, H.A.-A. and N.C.; methodology, H.A.-A. and N.C.; software, H.A.-A. and N.C.; validation, H.A.-A. and N.C.; formal analysis, H.A.-A. and N.C.; investigation, H.A.-A. and N.C.; resources, H.A.-A. and N.C.; data curation, H.A.-A. and N.C.; writing—original draft preparation, H.A.-A. and N.C.; writing—review and editing, H.A.-A. and N.C.; visualization, H.A.-A. and N.C. 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. 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]
  2. Nusse, R.; Brown, A.; Papkoff, J.; Scambler, P.; Shackleford, G.; McMahon, A.; Moon, R.; Varmus, H. A new nomenclature for int-1 and related genes: The Wnt gene family. Cell 1991, 64, 231. [Google Scholar] [CrossRef]
  3. Koveitypour, Z.; Panahi, F.; Vakilian, M.; Peymani, M.; Forootan, F.S.; Esfahani, M.H.N.; Ghaedi, K. Signaling pathways involved in colorectal cancer progression. Cell Biosci. 2019, 9, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Fire, A.; Xu, S.; Montgomery, M.K.; Kostas, S.A.; Driver, S.E.; Mello, C.C. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 1998, 391, 806–811. [Google Scholar] [CrossRef] [PubMed]
  5. Starega-Roslan, J.; Krol, J.; Koscianska, E.; Kozlowski, P.; Szlachcic, W.; Sobczak, K.; Krzyzosiak, W.J. Structural basis of microRNA length variety. Nucleic Acids Res. 2010, 39, 257–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Bhinge, A.; Poschmann, J.; Namboori, S.C.; Tian, X.; Loh, S.J.H.; Traczyk, A.; Prabhakar, S.; Stanton, L.W. Mi R -135b is a direct PAX 6 target and specifies human neuroectoderm by inhibiting TGF -β/ BMP signaling. EMBO J. 2014, 33, 1271–1283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Forterre, A.; Komuro, H.; Aminova, S.; Harada, M. A Comprehensive Review of Cancer MicroRNA Therapeutic Delivery Strategies. Cancers 2020, 12, 1852. [Google Scholar] [CrossRef]
  8. Qu, Z.; Li, W.; Fu, B. MicroRNAs in Autoimmune Diseases. BioMed Res. Int. 2014, 2014, 1–8. [Google Scholar] [CrossRef] [Green Version]
  9. Zhou, S.-S.; Jin, J.-P.; Wang, J.-Q.; Zhang, Z.-G.; Freedman, J.H.; Zheng, Y.; Cai, L. miRNAS in cardiovascular diseases: Potential biomarkers, therapeutic targets and challenges. Acta Pharmacol. Sin. 2018, 39, 1073–1084. [Google Scholar] [CrossRef] [Green Version]
  10. Weiss, C.N.; Ito, K. A Macro View of MicroRNAs: The Discovery of MicroRNAs and Their Role in Hematopoiesis and Hematologic Disease. Int. Rev. Cell Mol. Biol. 2017, 334, 99–175. [Google Scholar] [CrossRef] [Green Version]
  11. Plotnikova, O.; Baranova, A.; Skoblov, M. Comprehensive Analysis of Human microRNA–mRNA Interactome. Front. Genet. 2019, 10, 933. [Google Scholar] [CrossRef]
  12. Friedman, R.; Farh, K.K.-H.; Burge, C.B.; Bartel, D.P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2008, 19, 92–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. 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] [PubMed] [Green Version]
  14. Link, A.; Balaguer, F.; Shen, Y.; Nagasaka, T.; Lozano, J.J.; Boland, C.R.; Goel, A. Fecal MicroRNAs as Novel Biomarkers for Colon Cancer Screening. Cancer Epidemiol. Biomark. Prev. 2010, 19, 1766–1774. [Google Scholar] [CrossRef] [Green Version]
  15. Arroyo, J.D.; Chevillet, J.; Kroh, E.M.; Ruf, I.K.; Pritchard, C.C.; Gibson, D.F.; Mitchell, P.; Bennett, C.F.; Pogosova-Agadjanyan, E.L.; Stirewalt, D.L.; et al. Argonaute2 complexes carry a population of circulating microRNAs independent of vesicles in human plasma. Proc. Natl. Acad. Sci. USA 2011, 108, 5003–5008. [Google Scholar] [CrossRef] [Green Version]
  16. Weickmann, J.L.; Glitz, D.G. Human ribonucleases. Quantitation of pancreatic-like enzymes in serum, urine, and organ preparations. J. Biol. Chem. 1982, 257, 8705–8710. [Google Scholar] [CrossRef]
  17. Cui, M.; Wang, H.; Yao, X.; Zhang, D.; Xie, Y.; Cui, R.; Zhang, X. Circulating MicroRNAs in Cancer: Potential and Challenge. Front. Genet. 2019, 10, 626. [Google Scholar] [CrossRef] [Green Version]
  18. Lee, Y.; Ahn, C.; Han, J.; Choi, H.; Kim, J.; Yim, J.; Lee, J.; Provost, P.; Rådmark, O.; Kim, S.; et al. The nuclear RNase III Drosha initiates microRNA processing. Nat. Cell Biol. 2003, 425, 415–419. [Google Scholar] [CrossRef]
  19. Lund, E.; Güttinger, S.; Calado, A.; Dahlberg, J.E.; Kutay, U. Nuclear Export of MicroRNA Precursors. Science 2004, 303, 95–98. [Google Scholar] [CrossRef] [Green Version]
  20. Bernstein, E.; Caudy, A.; Hammond, S.M.; Hannon, G.J. Role for a bidentate ribonuclease in the initiation step of RNA interference. Nat. Cell Biol. 2001, 409, 363–366. [Google Scholar] [CrossRef]
  21. Bartel, D.P. MicroRNAs: Target Recognition and Regulatory Functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Rand, T.A.; Petersen, S.; Du, F.; Wang, X. Argonaute2 Cleaves the Anti-Guide Strand of siRNA during RISC Activation. Cell 2005, 123, 621–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Abdelfattah, A.M.; Park, C.; Choi, M.Y. Update on non-canonical microRNAs. Biomol. Concepts 2014, 5, 275–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pino, M.S.; Chung, D.C. The Chromosomal Instability Pathway in Colon Cancer. Gastroenterology 2010, 138, 2059–2072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Nagel, R.; Le Sage, C.; Diosdado, B.; Van Der Waal, M.; Vrielink, J.A.O.; Bolijn, A.; Meijer, G.A.; Agami, R. Regulation of the Adenomatous Polyposis Coli Gene by the miR-135 Family in Colorectal Cancer. Cancer Res. 2008, 68, 5795–5802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Wang, Y.-X.; Chen, Y.-R.; Liu, S.-S.; Ye, Y.-P.; Jiao, H.-L.; Wang, S.-Y.; Xiao, Z.-Y.; Wei, W.-T.; Qiu, J.-F.; Liang, L.; et al. MiR-384 inhibits human colorectal cancer metastasis by targeting KRAS and CDC42. Oncotarget 2016, 7, 84826–84838. [Google Scholar] [CrossRef] [Green Version]
  27. Kunkel, T.A.; Erie, D.A. DNA MISMATCH REPAIR. Annu. Rev. Biochem. 2005, 74, 681–710. [Google Scholar] [CrossRef] [Green Version]
  28. Earle, J.S.; Luthra, R.; Romans, A.; Abraham, R.; Ensor, J.; Yao, H.; Hamilton, S.R. Association of MicroRNA Expression with Microsatellite Instability Status in Colorectal Adenocarcinoma. J. Mol. Diagn. 2010, 12, 433–440. [Google Scholar] [CrossRef]
  29. Slattery, M.L.; Herrick, J.S.; Mullany, L.E.; Wolff, E.; Hoffman, M.D.; Pellatt, D.F.; Stevens, J.R.; Wolff, R.K. Colorectal tumor molecular phenotype and miRNA: Expression profiles and prognosis. Mod. Pathol. 2016, 29, 915–927. [Google Scholar] [CrossRef] [Green Version]
  30. Michael, M.Z.; O’ Connor, S.M.; van Holst Pellekaan, N.G.; Young, G.P.; James, R.J. Reduced Ac-cumulation of Specific MicroRNAs in Colorectal Neoplasia. Mol. Cancer Res. 2003, 1, 882–891. [Google Scholar]
  31. Goel, A.; Boland, C.R. Epigenetics of Colorectal Cancer. Gastroenterology 2012, 143, 1442–1460.e1. [Google Scholar] [CrossRef] [Green Version]
  32. Schetter, A.J.; Harris, C.C. Alterations of MicroRNAs Contribute to Colon Carcinogenesis. Semin. Oncol. 2011, 38, 734–742. [Google Scholar] [CrossRef] [Green Version]
  33. Pekow, J.; Meckel, K.; Dougherty, U.; Butun, F.; Mustafi, R.; Lim, J.; Crofton, C.; Chen, X.; Joseph, L.; Bissonnette, M. Tumor suppressors miR-143 and miR-145 and predicted target proteins API5, ERK5, K-RAS, and IRS-1 are differentially expressed in proximal and distal colon. Am. J. Physiol. Liver Physiol. 2015, 308, G179–G187. [Google Scholar] [CrossRef] [Green Version]
  34. Bai, J.-W.; Xue, H.-Z.; Zhang, C. Down-regulation of microRNA-143 is associated with colorectal cancer progression. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 4682–4687. [Google Scholar] [PubMed]
  35. Zhao, J.; Zhang, Y.; Zhao, G. Emerging role of microRNA-21 in colorectal cancer. Cancer Biomark. 2015, 15, 219–226. [Google Scholar] [CrossRef]
  36. Ma, H.; Pan, J.-S.; Jin, L.-X.; Wu, J.; Ren, Y.-D.; Chen, P.; Xiao, C.; Han, J. MicroRNA-17~92 inhibits colorectal cancer progression by targeting angiogenesis. Cancer Lett. 2016, 376, 293–302. [Google Scholar] [CrossRef] [PubMed]
  37. Chang, C.-M.; Wong, H.S.-C.; Huang, C.-Y.; Hsu, W.-L.; Maio, Z.-F.; Chiu, S.-J.; Tsai, Y.-T.; Chen, B.-K.; Wan, Y.-J.Y.; Wang, J.-Y.; et al. Functional Effects of let-7g Expression in Colon Cancer Metastasis. Cancers 2019, 11, 489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Cummins, J.M.; He, Y.; Leary, R.J.; Pagliarini, R.; Diaz, L.A.; Sjoblom, T.; Barad, O.; Bentwich, Z.; Szafranska, A.E.; Labourier, E.; et al. The colorectal microRNAome. Proc. Natl. Acad. Sci. USA 2006, 103, 3687–3692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Cancer Genome Atlas Network. Comprehensive molecular characterization of human colon and rectal cancer. Nature 2012, 487, 330–337. [Google Scholar] [CrossRef] [Green Version]
  40. Parker, T.; Neufeld, K.L. APC controls Wnt-induced β-catenin destruction complex recruitment in human colonocytes. Sci. Rep. 2020, 10, 1–14. [Google Scholar] [CrossRef] [Green Version]
  41. Nakayama, M.; Oshima, M. Mutant p53 in colon cancer. J. Mol. Cell Biol. 2019, 11, 267–276. [Google Scholar] [CrossRef] [Green Version]
  42. Chang, T.-C.; Wentzel, E.A.; Kent, O.; Ramachandran, K.; Mullendore, M.; Lee, K.H.; Feldmann, G.; Yamakuchi, M.; Ferlito, M.; Lowenstein, C.J.; et al. Transactivation of miR-34a by p53 Broadly Influences Gene Expression and Promotes Apoptosis. Mol. Cell 2007, 26, 745–752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. 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. Nat. Cell Biol. 2008, 10, 593–601. [Google Scholar] [CrossRef] [PubMed]
  44. 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] [Green Version]
  45. Fessler, E.; Jansen, M.; Melo, F.D.S.E.; Zhao, L.; Prasetyanti, P.R.; Rodermond, H.; Kandimalla, R.; Linnekamp, J.F.; Franitza, M.; Van Hooff, S.R.; et al. A multidimensional network approach reveals microRNAs as determinants of the mesenchymal colorectal cancer subtype. Oncogene 2016, 35, 6026–6037. [Google Scholar] [CrossRef] [Green Version]
  46. Abdelmaksoud-Dammak, R.; Chamtouri, N.; Triki, M.; Saadallah-Kallel, A.; Ayadi, W.; Charfi, S.; Khabir, A.; Ayadi, L.; Sallemi-Boudawara, T.; Mokdad-Gargouri, R. Overexpression of miR-10b in colorectal cancer patients: Correlation with TWIST-1 and E-cadherin expression. Tumor Biol. 2017, 39, 1010428317695916. [Google Scholar] [CrossRef] [Green Version]
  47. Zhang, J.; Fei, B.; Wang, Q.; Song, M.; Yin, Y.; Zhang, B.; Ni, S.; Guo, W.; Bian, Z.; Quan, C.; et al. MicroRNA-638 inhibits cell proliferation, invasion and regulates cell cycle by targeting tetraspanin 1 in human colorectal carcinoma. Oncotarget 2014, 5, 12083–12096. [Google Scholar] [CrossRef] [Green Version]
  48. Falzone, L.; Scola, L.; Zanghì, A.; Biondi, A.; Di Cataldo, A.; Libra, M.; Candido, S. Integrated analysis of colorectal cancer microRNA datasets: Identification of microRNAs associated with tumor development. Aging 2018, 10, 1000–1014. [Google Scholar] [CrossRef]
  49. Zhang, N.; Lu, C.; Chen, L. miR-217 regulates tumor growth and apoptosis by targeting the MAPK signaling pathway in colorectal cancer. Oncol. Lett. 2016, 12, 4589–4597. [Google Scholar] [CrossRef] [Green Version]
  50. Fasihi, A.; Soltani, B.M.; Atashi, A.; Nasiri, S. Introduction of hsa-miR-103a and hsa-miR-1827 and hsa-miR-137 as new regulators of Wnt signaling pathway and their relation to colorectal carcinoma. J. Cell. Biochem. 2018, 119, 5104–5117. [Google Scholar] [CrossRef]
  51. Hao, H.; Xia, G.; Wang, C.; Zhong, F.; Liu, L.; Zhang, D. miR-106a suppresses tumor cells death in colorectal cancer through targeting ATG7. Med Mol. Morphol. 2016, 50, 76–85. [Google Scholar] [CrossRef]
  52. Jia, L.; Luo, S.; Ren, X.; Li, Y.; Hu, J.; Liu, B.; Zhao, L.; Shan, Y.; Zhou, H. miR-182 and miR-135b Mediate the Tumorigenesis and Invasiveness of Colorectal Cancer Cells via Targeting ST6GALNAC2 and PI3K/AKT Pathway. Dig. Dis. Sci. 2017, 62, 3447–3459. [Google Scholar] [CrossRef] [PubMed]
  53. Ji, D.; Chen, Z.; Li, M.; Zhan, T.; Yao, Y.; Zhang, Z.; Xi, J.; Yan, L.; Gu, J. MicroRNA-181a promotes tumor growth and liver metastasis in colorectal cancer by targeting the tumor suppressor WIF-1. Mol. Cancer 2014, 13, 86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Tufekci, K.U.; Meuwissen, R.L.J.; Genc, S. The Role of MicroRNAs in Biological Processes. Methods Mol. Biol. 2014, 1107, 15–31. [Google Scholar] [CrossRef]
  55. Angius, A.; Uva, P.; Pira, G.; Muroni, M.R.; Sotgiu, G.; Saderi, L.; Uleri, E.; Caocci, M.; Ibba, G.; Cesaraccio, M.R.; et al. Integrated Analysis of miRNA and mRNA Endorses a Twenty miRNAs Signature for Colorectal Carcinoma. Int. J. Mol. Sci. 2019, 20, 4067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Tijsen, A.J.; Creemers, E.E.; Moerland, P.D.; De Windt, L.J.; Van Der Wal, A.C.; Kok, W.E.; Pinto, Y.M. MiR423-5p As a Circulating Biomarker for Heart Failure. Circ. Res. 2010, 106, 1035–1039. [Google Scholar] [CrossRef] [PubMed]
  57. Wiedrick, J.T.; Phillips, J.I.; Lusardi, T.A.; McFarland, T.J.; Lind, B.; Sandau, U.S.; Harrington, C.A.; Lapidus, J.A.; Galasko, D.R.; Quinn, J.F.; et al. Validation of MicroRNA Biomarkers for Alzheimer’s Disease in Human Cerebrospinal Fluid. J. Alzheimer’s Dis. 2019, 67, 875–891. [Google Scholar] [CrossRef]
  58. Wu, C.W.; Ng, S.S.M.; Dong, Y.J.; Ng, S.C.; Leung, W.W.; Lee, C.W.; Ni Wong, Y.; Chan, F.K.; Yu, J.; Sung, J.J.Y. Detection of miR-92a and miR-21 in stool samples as potential screening biomarkers for colorectal cancer and polyps. Gut 2012, 61, 739–745. [Google Scholar] [CrossRef]
  59. Wang, J.; Huang, S.-K.; Zhao, M.; Yang, M.; Zhong, J.-L.; Gu, Y.-Y.; Peng, H.; Che, Y.-Q.; Huang, C.-Z. Identification of a Circulating MicroRNA Signature for Colorectal Cancer Detection. PLoS ONE 2014, 9, e87451. [Google Scholar] [CrossRef]
  60. Basati, G.; Razavi, A.E.; Abdi, S.; Mirzaei, A. Elevated level of microRNA-21 in the serum of patients with colorectal cancer. Med Oncol. 2014, 31, 1–5. [Google Scholar] [CrossRef]
  61. Du, M.; Liu, S.; Gu, D.; Wang, Q.; Zhu, L.; Kang, M.; Shi, D.; Chu, H.; Tong, N.; Chen, J.; et al. Clinical potential role of circulating microRNAs in early diagnosis of colorectal cancer patients. Carcinogenesis 2014, 35, 2723–2730. [Google Scholar] [CrossRef] [Green Version]
  62. Zheng, G.; Wang, H.; Zhang, X.; Yang, Y.; Wang, L.; Du, L.; Li, W.; Li, J.; Qu, A.; Liu, Y.; et al. Identification and Validation of Reference Genes for qPCR Detection of Serum microRNAs in Colorectal Adenocarcinoma Patients. PLoS ONE 2013, 8, e83025. [Google Scholar] [CrossRef]
  63. Navarro, M.; Nicolas, A.; Ferrandez, A.; Lanas, A. Colorectal cancer population screening programs worldwide in 2016: An update. World J. Gastroenterol. 2017, 23, 3632–3642. [Google Scholar] [CrossRef]
  64. Mogilyansky, E.; Rigoutsos, I. The miR-17/92 cluster: A comprehensive update on its genomics, genetics, functions and increasingly important and numerous roles in health and disease. Cell Death Differ. 2013, 20, 1603–1614. [Google Scholar] [CrossRef]
  65. Christou, N.; Perraud, A.; Blondy, S.; Jauberteau, M.-O.; Battu, S.; Mathonnet, M. E-cadherin: A potential biomarker of colorectal cancer prognosis. Oncol. Lett. 2017, 13, 4571–4576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Karimi, N.; Feizi, M.A.H.; Safaralizadeh, R.; Hashemzadeh, S.; Baradaran, B.; Shokouhi, B.; Teimourian, S. Serum overexpression of miR-301a and miR-23a in patients with colorectal cancer. J. Chin. Med Assoc. 2019, 82, 215–220. [Google Scholar] [CrossRef]
  67. Xiao, Y.; Zhong, J.; Zhong, B.; Huang, J.; Jiang, L.; Jiang, Y.; Yuan, J.; Sun, J.; Dai, L.; Yang, C.; et al. Exosomes as potential sources of biomarkers in colorectal cancer. Cancer Lett. 2020, 476, 13–22. [Google Scholar] [CrossRef] [PubMed]
  68. Ogata-Kawata, H.; Izumiya, M.; Kurioka, D.; Honma, Y.; Yamada, Y.; Furuta, K.; Gunji, T.; Ohta, H.; Okamoto, H.; Sonoda, H.; et al. Circulating Exosomal microRNAs as Biomarkers of Colon Cancer. PLoS ONE 2014, 9, e92921. [Google Scholar] [CrossRef]
  69. Zuo, Z.; Jiang, Y.; Zeng, S.; Li, Y.; Fan, J.; Guo, Y.; Tao, H. The value of microRNAs as the novel biomarkers for colorectal cancer diagnosis: A meta-analysis. Pathol. Res. Pr. 2020, 216, 153130. [Google Scholar] [CrossRef] [PubMed]
  70. Tanaka, H.; Hazama, S.; Iida, M.; Tsunedomi, R.; Takenouchi, H.; Nakajima, M.; Tokumitsu, Y.; Kanekiyo, S.; Shindo, Y.; Tomochika, S.; et al. miR-125b-1 and miR-378a are predictive biomarkers for the efficacy of vaccine treatment against colorectal cancer. Cancer Sci. 2017, 108, 2229–2238. [Google Scholar] [CrossRef] [Green Version]
  71. Anandappa, G.; Lampis, A.; Cunningham, D.; Khan, K.H.; Kouvelakis, K.; Vlachogiannis, G.; Hedayat, S.; Tunariu, N.; Rao, S.; Watkins, D.; et al. miR-31-3p Expression and Benefit from Anti-EGFR Inhibitors in Metastatic Colorectal Cancer Patients Enrolled in the Prospective Phase II PROSPECT-C Trial. Clin. Cancer Res. 2019, 25, 3830–3838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Chen, P.; Xi, Q.; Wang, Q.; Wei, P. Downregulation of microRNA-100 correlates with tumor progression and poor prognosis in colorectal cancer. Med. Oncol. 2014, 31, 1–6. [Google Scholar] [CrossRef] [PubMed]
  73. Schou, J.V.; Rossi, S.; Jensen, B.V.; Nielsen, D.L.; Pfeiffer, P.; Høgdall, E.; Yilmaz, M.; Tejpar, S.; Delorenzi, M.; Kruhøffer, M.; et al. miR-345 in Metastatic Colorectal Cancer: A Non-Invasive Biomarker for Clinical Outcome in Non-KRAS Mutant Patients Treated with 3rd Line Cetuximab and Irinotecan. PLoS ONE 2014, 9, e99886. [Google Scholar] [CrossRef]
  74. Hansen, T.; Carlsen, A.L.; Heegaard, N.H.H.; Sørensen, F.B.; Jakobsen, A. Changes in circulating microRNA-126 during treatment with chemotherapy and bevacizumab predicts treatment response in patients with metastatic colorectal cancer. Br. J. Cancer 2015, 112, 624–629. [Google Scholar] [CrossRef]
  75. Hansen, T.F.; Christensen, R.D.; Andersen, R.F.; Sørensen, F.B.; Johnsson, A.; Jakobsen, A. MicroRNA-126 and epidermal growth factor-like domain 7–an angiogenic couple of importance in metastatic colorectal cancer. Results from the Nordic ACT trial. Br. J. Cancer 2013, 109, 1243–1251. [Google Scholar] [CrossRef] [Green Version]
  76. Kijima, T.; Hazama, S.; Tsunedomi, R.; Tanaka, H.; Takenouchi, H.; Kanekiyo, S.; Inoue, Y.; Nakashima, M.; Iida, M.; Sakamoto, K.; et al. MicroRNA-6826 and −6875 in plasma are valuable non-invasive biomarkers that predict the efficacy of vaccine treatment against metastatic colorectal cancer. Oncol. Rep. 2017, 37, 23–30. [Google Scholar] [CrossRef] [Green Version]
  77. Kiss, I.; Mlcochova, J.; Bortlicek, Z.; Poprach, A.; Drabek, J.; Vychytilova-Faltejskova, P.; Svoboda, M.; Buchler, T.; Batko, S.; Ryska, A.; et al. Efficacy and Toxicity of Panitumumab After Progression on Cetuximab and Predictive Value of MiR-31-5p in Metastatic Wild-type KRAS Colorectal Cancer Patients. Anticancer. Res. 2016, 36, 4955–4960. [Google Scholar] [CrossRef] [Green Version]
  78. Liu, K.; Li, G.; Fan, C.; Zhou, X.; Wu, B.; Li, J. Increased Expression of MicroRNA-21 and Its Association with Chemotherapeutic Response in Human Colorectal Cancer. J. Int. Med. Res. 2011, 39, 2288–2295. [Google Scholar] [CrossRef] [Green Version]
  79. Laurent-Puig, P.; Grisoni, M.-L.; Heinemann, V.; Liebaert, F.; Neureiter, D.; Jung, A.; Montestruc, F.; Gaston-Mathe, Y.; Thiébaut, R.; Stintzing, S. Validation of miR-31-3p Expression to Predict Cetuximab Efficacy When Used as First-Line Treatment in RAS Wild-Type Metastatic Colorectal Cancer. Clin. Cancer Res. 2018, 25, 134–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Sclafani, F.; Chau, I.; Cunningham, D.; Peckitt, C.; Lampis, A.; Hahne, J.; Braconi, C.; Tabernero, J.; Glimelius, B.; Cervantes, A.; et al. Prognostic role of the LCS6 KRAS variant in locally advanced rectal cancer: Results of the EXPERT-C trial. Ann. Oncol. 2015, 26, 1936–1941. [Google Scholar] [CrossRef]
  81. Ciardiello, F.; Tortora, G. EGFR Antagonists in Cancer Treatment. N. Engl. J. Med. 2008, 358, 1160–1174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Miller-Phillips, L.; Heinemann, V.; Stahler, A.; Von Weikersthal, L.F.; Kaiser, F.; Al-Batran, S.-E.; Neureiter, D.; Kahl, C.; Kullmann, F.; Moehler, M.H.; et al. Association of microRNA-21 with efficacy of cetuximab in RAS wild-type patients in the FIRE-3 study (AIO KRK-0306) and microRNA-21’s influence on gene expression in the EGFR signaling pathway. J. Clin. Oncol. 2019, 37, 3593. [Google Scholar] [CrossRef]
  83. Peng, Q.; Zhang, X.; Min, M.; Zou, L.; Shen, P.; Zhu, Y. The clinical role of microRNA-21 as a promising biomarker in the diagnosis and prognosis of colorectal cancer: A systematic review and meta-analysis. Oncotarget 2017, 8, 44893–44909. [Google Scholar] [CrossRef] [Green Version]
  84. Yan, S.; Han, B.; Gao, S.; Wang, X.; Wang, Z.; Wang, F.; Zhang, J.; Xu, D.; Sun, B. Exosome-encapsulated microRNAs as circulating biomarkers for colorectal cancer. Oncotarget 2017, 8, 60149–60158. [Google Scholar] [CrossRef] [Green Version]
  85. Van Der Jeught, K.; Xu, H.-C.; Li, Y.-J.; Lu, X.-B.; Ji, G. Drug resistance and new therapies in colorectal cancer. World J. Gastroenterol. 2018, 24, 3834–3848. [Google Scholar] [CrossRef]
  86. Deng, J.; Lei, W.; Fu, J.-C.; Zhang, L.; Li, J.-H.; Xiong, J.-P. Targeting miR-21 enhances the sensitivity of human colon cancer HT-29 cells to chemoradiotherapy in vitro. Biochem. Biophys. Res. Commun. 2014, 443, 789–795. [Google Scholar] [CrossRef]
  87. Mosakhani, N.; Lahti, L.; Borze, I.; Karjalainen-Lindsberg, M.-L.; Sundström, J.; Ristamäki, R.; Österlund, P.; Knuutila, S.; Sarhadi, V.K. MicroRNA profiling predicts survival in anti-EGFR treated chemorefractory metastatic colorectal cancer patients with wild-type KRAS and BRAF. Cancer Genet. 2012, 205, 545–551. [Google Scholar] [CrossRef] [PubMed]
  88. Mlcochova, J.; Faltejskova-Vychytilova, P.; Ferracin, M.; Zagatti, B.; Radova, L.; Svoboda, M.; Nemecek, R.; John, S.; Kiss, I.; Vyzula, R.; et al. MicroRNA expression profiling identifies miR-31-5p/3p as associated with time to progression in wild-type RAS metastatic colorectal cancer treated with cetuximab. Oncotarget 2015, 6, 38695–38704. [Google Scholar] [CrossRef] [Green Version]
  89. Pichler, M.; Winter, E.; Ress, A.L.; Bauernhofer, T.; Gerger, A.; Kiesslich, T.; Lax, S.; Samonigg, H.; Hoefler, G. miR-181a is associated with poor clinical outcome in patients with colorectal cancer treated with EGFR inhibitor. J. Clin. Pathol. 2013, 67, 198–203. [Google Scholar] [CrossRef] [PubMed]
  90. Ruzzo, A.; Graziano, F.; Vincenzi, B.; Canestrari, E.; Perrone, G.; Galluccio, N.; Catalano, V.; Loupakis, F.; Rabitti, C.; Santini, D.; et al. High Let-7a MicroRNA Levels in KRAS -Mutated Colorectal Carcinomas May Rescue Anti-EGFR Therapy Effects in Patients with Chemotherapy-Refractory Metastatic Disease. Oncology 2012, 17, 823–829. [Google Scholar] [CrossRef] [Green Version]
  91. Boisen, M.K.; Dehlendorff, C.; Linnemann, D.; Nielsen, B.S.; Larsen, J.S.; Østerlind, K.; Nielsen, S.E.; Tarpgaard, L.S.; Qvortrup, C.; Pfeiffer, P.; et al. Tissue MicroRNAs as Predictors of Outcome in Patients with Metastatic Colorectal Cancer Treated with First Line Capecitabine and Oxaliplatin with or without Bevacizumab. PLoS ONE 2014, 9, e109430. [Google Scholar] [CrossRef]
  92. Fiala, O.; Pitule, P.; Hosek, P.; Liška, V.; Šorejs, O.; Bruha, J.; Vycital, O.; Buchler, T.; Poprach, A.; Topolcan, O.; et al. The association of miR-126-3p, miR-126-5p and miR-664-3p expression profiles with outcomes of patients with metastatic colorectal cancer treated with bevacizumab. Tumor Biol. 2017, 39, 1010428317709283. [Google Scholar] [CrossRef] [Green Version]
  93. Kiss, I.; Mlčochová, J.; Součková, K.; Fabian, P.; Poprach, A.; Halamkova, J.; Svoboda, M.; Vyzula, R.; Slaby, O. MicroRNAs as outcome predictors in patients with metastatic colorectal cancer treated with bevacizumab in combination with FOLFOX. Oncol. Lett. 2017, 14, 743–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Ulivi, P.; Canale, M.; Passardi, A.; Marisi, G.; Valgiusti, M.; Frassineti, G.L.; Calistri, D.; Amadori, D.; Scarpi, E. Circulating Plasma Levels of miR-20b, miR-29b and miR-155 as Predictors of Bevacizumab Efficacy in Patients with Metastatic Colorectal Cancer. Int. J. Mol. Sci. 2018, 19, 307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Ardekani, A.M.; Naeini, M.M. The Role of MicroRNAs in Human Diseases. Avicenna J. Med. Biotechnol. 2010, 2, 161–179. [Google Scholar]
  96. Lindow, M.; Kauppinen, S. Discovering the first microRNA-targeted drug. J. Cell Biol. 2012, 199, 407–412. [Google Scholar] [CrossRef]
  97. 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] [PubMed] [Green Version]
  98. Lundin, K.E.; Højland, T.; Hansen, B.R.; Persson, R.; Bramsen, J.B.; Kjems, J.; Koch, T.; Wengel, J.; Smith, C.E. Biological Activity and Biotechnological Aspects of Locked Nucleic Acids. In Advances in Genetics; Friedmann, T., Dunlap, J.C., Goodwin, S.F., Eds.; Academic Press: Cambridge, MA, USA, 2013; Volume 82, pp. 47–107. [Google Scholar] [CrossRef]
  99. Lima, J.F.; Cerqueira, L.; Figueiredo, C.; Oliveira, C.; Azevedo, N.F. Anti-miRNA oligonucleotides: A comprehensive guide for design. RNA Biol. 2018, 15, 338–352. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, Z. MicroRNA Interference Technologies; Springer Science and Business Media LLC: Berlin, Germany, 2009. [Google Scholar] [CrossRef]
  101. Krützfeldt, J.; Rajewsky, N.; Braich, R.; Rajeev, K.G.; Tuschl, T.; Manoharan, M.; Stoffel, M. Silencing of microRNAs in vivo with ‘antagomirs’. Nat. Cell Biol. 2005, 438, 685–689. [Google Scholar] [CrossRef]
  102. Wu, N.; Fesler, A.; Liu, H.; Ju, J. Development of novel miR-129 mimics with enhanced efficacy to eliminate chemoresistant colon cancer stem cells. Oncotarget 2017, 9, 8887–8897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. van Zandwijk, N.; Pavlakis, N.; Kao, S.C.; Linton, A.; Boyer, M.J.; Clarke, S.; Huynh, Y.; Chrzanowska, A.; Fulham, M.; Bailey, D.L.; et al. Safety and activity of microRNA-loaded minicells in patients with recurrent malignant pleural mesothelioma: A first-in-man, phase 1, open-label, dose-escalation study. Lancet Oncol. 2017, 18, 1386–1396. [Google Scholar] [CrossRef]
  104. Viteri, S.; Rosell, R. An innovative mesothelioma treatment based on miR-16 mimic loaded EGFR targeted minicells (TargomiRs). Transl. Lung Cancer Res. 2018, 7, S1–S4. [Google Scholar] [CrossRef] [Green Version]
  105. Bansal, P.; Christopher, A.F.; Kaur, R.P.; Kaur, G.; Kaur, A.; Gupta, V. MicroRNA therapeutics: Discovering novel targets and developing specific therapy. Perspect. Clin. Res. 2016, 7, 68–74. [Google Scholar] [CrossRef]
  106. Akao, Y.; Nakagawa, Y.; Hirata, I.; Iio, A.; Itoh, T.; Kojima, K.; Nakashima, R.; Kitade, Y.; Naoe, T. Role of anti-oncomirs miR-143 and -145 in human colorectal tumors. Cancer Gene Ther. 2010, 17, 398–408. [Google Scholar] [CrossRef]
  107. Zhang, X.; Ai, F.; Li, X.; Tian, L.; Wang, X.; Shen, S.; Liu, F. MicroRNA-34a suppresses colorectal cancer metastasis by regulating Notch signaling. Oncol. Lett. 2017, 14, 2325–2333. [Google Scholar] [CrossRef]
  108. Sachdeva, M.; Mo, Y.-Y. miR-145-mediated suppression of cell growth, invasion and metastasis. Am. J. Transl. Res. 2010, 2, 170–180. [Google Scholar] [PubMed]
  109. A Thomas, M.; Langegrunweller, K.; Weirauch, U.; Gutsch, D.; Aigner, A.; Grunweller, A.; Hartmann, R.K. The proto-oncogene Pim-1 is a target of miR-33a. Oncogene 2011, 31, 918–928. [Google Scholar] [CrossRef] [Green Version]
  110. Ibrahim, A.F.; Weirauch, U.; Thomas, M.; Grünweller, A.; Hartmann, R.K.; Aigner, A. MicroRNA Replacement Therapy for miR-145 and miR-33a Is Efficacious in a Model of Colon Carcinoma. Cancer Res. 2011, 71, 5214–5224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Karimi, L.; Zeinali, T.; Hosseinahli, N.; Mansoori, B.; Mohammadi, A.; Yousefi, M.; Asadi, M.; Sadreddini, S.; Baradaran, B.; Shanehbandi, D. miRNA-143 replacement therapy harnesses the proliferation and migration of colorectal cancer cells in vitro. J. Cell. Physiol. 2019, 234, 21359–21368. [Google Scholar] [CrossRef]
  112. Hejazi, M.; Baghbani, E.; Amini, M.; Rezaei, T.; Aghanejad, A.; Mosafer, J.; Mokhtarzadeh, A.; Baradaran, B. MicroRNA-193a and taxol combination: A new strategy for treatment of colorectal cancer. J. Cell. Biochem. 2020, 121, 1388–1399. [Google Scholar] [CrossRef] [PubMed]
  113. Zhang, P.; Ji, D.-B.; Han, H.-B.; Shi, Y.-F.; Du, C.-Z.; Gu, J. Downregulation of miR-193a-5p correlates with lymph node metastasis and poor prognosis in colorectal cancer. World J. Gastroenterol. 2014, 20, 12241–12248. [Google Scholar] [CrossRef]
  114. Liu, T.; Zhang, X.; Du, L.; Wang, Y.; Liu, X.; Tian, H.; Wang, L.; Li, P.; Zhao, Y.; Duan, W.; et al. Exosome-transmitted miR-128-3p increase chemosensitivity of oxaliplatin-resistant colorectal cancer. Mol. Cancer 2019, 18, 1–17. [Google Scholar] [CrossRef] [Green Version]
  115. Sun, L.; Fang, Y.; Wang, X.; Han, Y.; Du, F.; Li, C.; Hu, H.; Liu, H.; Liu, Q.; Wang, J.; et al. miR-302a Inhibits Metastasis and Cetuximab Resistance in Colorectal Cancer by Targeting NFIB and CD44. Theranostics 2019, 9, 8409–8425. [Google Scholar] [CrossRef] [PubMed]
  116. Ghosh, D.; Nandi, S.; Bhattacharjee, S. Combination therapy to checkmate Glioblastoma: Clinical challenges and advances. Clin. Transl. Med. 2018, 7, 33. [Google Scholar] [CrossRef] [Green Version]
  117. Chen, Y.; Gao, D.-Y.; Huang, L. In vivo delivery of miRNAs for cancer therapy: Challenges and strategies. Adv. Drug Deliv. Rev. 2015, 81, 128–141. [Google Scholar] [CrossRef] [Green Version]
  118. Mercatelli, N.; Coppola, V.; Bonci, D.; Miele, F.; Costantini, A.; Guadagnoli, M.; Bonanno, E.; Muto, G.; Frajese, G.V.; De Maria, R.; et al. The Inhibition of the Highly Expressed Mir-221 and Mir-222 Impairs the Growth of Prostate Carcinoma Xenografts in Mice. PLoS ONE 2008, 3, e4029. [Google Scholar] [CrossRef]
  119. Sureban, S.M.; May, R.; Mondalek, F.G.; Qu, D.; Ponnurangam, S.; Pantazis, P.; Anant, S.; Ramanujam, R.P.; Houchen, C.W. Nanoparticle-based delivery of siDCAMKL-1 increases microRNA-144 and inhibits colorectal cancer tumor growth via a Notch-1 dependent mechanism. J. Nanobiotechnol. 2011, 9, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Wang, X.; Yu, B.; Ren, W.; Mo, X.; Zhou, C.; He, H.; Jia, H.; Wang, L.; Jacob, S.T.; Lee, R.J.; et al. Enhanced hepatic delivery of siRNA and microRNA using oleic acid based lipid nanoparticle formulations. J. Control. Release 2013, 172, 690–698. [Google Scholar] [CrossRef]
  121. Davis, S.; Lollo, B.; Freier, S.; Esau, C. Improved targeting of miRNA with antisense oligonucleotides. Nucleic Acids Res. 2006, 34, 2294–2304. [Google Scholar] [CrossRef] [PubMed]
  122. Elmén, J.; Lindow, M.; Silahtaroglu, A.; Bak, M.; Christensen, M.; Lind-Thomsen, A.; Hedtjärn, M.; Hansen, J.B.; Hansen, H.F.; Straarup, E.M.; et al. Antagonism of microRNA-122 in mice by systemically administered LNA-antimiR leads to up-regulation of a large set of predicted target mRNAs in the liver. Nucleic Acids Res. 2007, 36, 1153–1162. [Google Scholar] [CrossRef] [Green Version]
  123. Hiraki, M.; Nishimura, J.; Takahashi, H.; Wu, X.; Takahashi, Y.; Miyo, M.; Nishida, N.; Uemura, M.; Hata, T.; Takemasa, I.; et al. Concurrent Targeting of KRAS and AKT by MiR-4689 Is a Novel Treatment Against Mutant KRAS Colorectal Cancer. Mol. Ther. Nucleic Acids 2015, 4, e231. [Google Scholar] [CrossRef] [PubMed]
  124. Inoue, A.; Mizushima, T.; Wu, X.; Okuzaki, D.; Kambara, N.; Ishikawa, S.; Wang, J.; Qian, Y.; Hirose, H.; Yokoyama, Y.; et al. A miR-29b Byproduct Sequence Exhibits Potent Tumor-Suppressive Activities via Inhibition of NF-κB Signaling in KRAS-Mutant Colon Cancer Cells. Mol. Cancer Ther. 2018, 17, 977–987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Akao, Y.; Iio, A.; Itoh, T.; Noguchi, S.; Itoh, Y.; Ohtsuki, Y.; Naoe, T. Microvesicle-mediated RNA Molecule Delivery System Using Monocytes/Macrophages. Mol. Ther. 2011, 19, 395–399. [Google Scholar] [CrossRef] [Green Version]
  126. Tazawa, H.; Tsuchiya, N.; Izumiya, M.; Nakagama, H. Tumor-suppressive miR-34a induces senescence-like growth arrest through modulation of the E2F pathway in human colon cancer cells. Proc. Natl. Acad. Sci. USA 2007, 104, 15472–15477. [Google Scholar] [CrossRef] [Green Version]
Table 1. MiRNAs and CRC pathways.
Table 1. MiRNAs and CRC pathways.
MiRNA(s)SampleSignaling Pathways Involved or Proteins InvolvedRole in OncogenesisReference
MiRNA-10bTissueE cadherin proteinActivates metastasis, vascular invasion, and tumor differentiationAbdelmaksoud-Dammak et al. [46]
MiRNA-638TissueTSPAN1 proteinInhibits cell proliferation, invasion, and arrests the cell cycle in G1 phaseZhang et al., 2014 [47]
MiRNA-17, 19, 20 and 9TissueTGFβ-signalingCell proliferationPellat et al., 2018 [29]
MiRNA-21-5pTissueHippo signaling, Wnt signaling
RAS signaling
PI3K-AKT signaling
TGF-β signaling
Mismatch repair signaling
Erb signaling
p53 signaling
Cell proliferation and cancer progressionFalzone et al., 2018
[48]
MiRNA-195-5pMAPK signaling
MiRNA-497-5pp53 signaling
FoxO signaling
mTOR signaling
MAPK signaling
MiRNA-217TissueMAPK signalingCell growth
Apoptosis
Zhang et al., 2016
[49]
MiRNA-103a
MiRNA-1827
MiRNA-137
TissueWnt signalingCell cycle
Apoptosis
Fasihi et al., 2018
[50]
MiRNA-106aTissueATG7 proteinAutophagyHao et al., 2017 [51]
MiRNA-182 and miRNA-135bTissuePI3K-AKT signalingMigration, adhesion, invasion, proliferation, and tumor angiogenesisJia et al., 2017
[52]
MiRNA-181aTissueE cadherin proteinCell motility, invasion, tumor growthJi et al., 2014 [53]
MiRNA-125a-3pTissuePI3K-AKT signalingProliferation, migration, invasion, and angiogenesisLiang et al., 2017
Table 2. MiRNAs as diagnostic biomarkers.
Table 2. MiRNAs as diagnostic biomarkers.
MiRNA(s)SampleAimLimit of Detection (LOD) [Unit]SensibilitySpecificityReference
MiRNA-92a(miR-21)StoolsEarly diagnosis (screening)Diagnosis435 [copies/ng]71.6% for CRC 56.1% for polyp73.3% for CRC and polypWu et al., 2012 [58]
Six- miRNA signature:
miRNA-21,
let-7g,
miRNA-31,
miRNA-92a,
miRNA-181b, miRNA-203
SerumDiagnosis9.595 [“of risk score function”]96.4%88.1%Wang et al., 2014 [59]
MiRNA-21SerumDiagnosis1.49 [fold change: ratio of the changes between miRNA value of CCR patients and miRNA value of healthy patients]77%78%Basati et al., 2014 [60]
MiRNA-21PlasmaDiagnosis0.00220
[expression relative, log 10 (2ΔCt)]
76.2%93.2%Du et al., 2014 [61]
MiRNA-191-5pSerumReference gene for diagnosisNot reportedNot reportedNot reportedZheng et al., 2013 [62]
Table 3. MiRNAs as prognostic biomarkers.
Table 3. MiRNAs as prognostic biomarkers.
MiRNA(s)SampleAimResultsReference
MiRNA-125b-1
MiRNA-378a
TissueTo predict efficacy of vaccine (5-peptide combination) against CRCs+
After peptide vaccines combined with oxaliplatin-containing chemotherapy were given:
miR-125b-1 in cancer cells (p = 0.040), and miR-378a in both cancer cells (p = 0.009) and stromal cells (p < 0.001) were negatively associated with OS
Tanaka et al., 2017 [70]
MiRNA-31-3pTissuePredictive biomarker of selection for anti-EGFR mAbs.+
Low miR-31-3p expression linked to better overall response rate
Anandappa et al., 2019 [71]
MiRNA-100TissuePrognostic+
Its downregulation showed poor overall survival (OS)
Chen et al., 2014 [72]
MiRNA-345Whole bloodPrognostic
Predictive about cetuximab and irinotecan response
+
−MiR-345, single prognostic biomarker for both OS and progression-free survival
(PFS)
−High miR-345 expression was associated with lack of response to treatment with cetuximab and irinotecan
Schou et al., 2014 [73]
Circulating microRNA-126 (cir-miRNA-126)PlasmaPredictive of anti-angiogenic treatment resistance+
Non-response to anti-angiogenic treatment was linked to increase of cir-miRNA-126
Hansen et al., 2015 [74]
MiRNA-126TissuePrognostic+
High tumor expression of miRNA-126 was significantly related to a longer PFS
Hansen et al., 2013 [75]
MiRNA-6826 and miRNA-6875PlasmaPredictive to vaccine treatment response+
Plasma miR-6826 and miR-6875 may be predictive biomarkers for a poor response to vaccine treatment
Kijima et al., 2017 [76]
MiRNA-31-5pTissuePredictive response to cetuximab and panitumumab in metastatic wild-type KRAS colorectal cancer patients in progression after cetuximab in combination with irinotecan-based chemotherapy (FOLFIRI or irinotecan alone) who received panitumumab monotherapy+/−
−Predictive for cetuximab response
−Non-predictive for panitumumab response
Kiss et al., 2016 [77]
MiRNA-21TissuePredictive response to neoadjuvant chemotherapy (FOLFOX4) for locally advanced CRC (staging cT3–4, any N, M0 or cT2, N1)+
Cut-off: 10.32 for differentiating pathological responders from non-responders, with a sensitivity of 80.0% and specificity of 88.2%
Liu et al., 2011 [78]
MiRNA-31-3pTissuePredictive of cetuximab therapy efficacy for patients with RAS WT mCRC.+
Low miR-31-3p expressers significantly benefited from cetuximab compared with bevacizumab for PFS, OS in multivariate analyses
Laurent-Puig et al., 2019 [79]
Single nucleotide polymorphism (rs61764370, T > G base substitution) in the let-7 complementary site 6 (LCS-6) of KRAS miRNATissueRs61764370 predictive of neoadjuvant chemoradiotherapy for locally advanced rectal cancer+
Randomized phase II trial of neoadjuvant CAPOX (capecitabine + oxaliplatin) followed by chemoradiotherapy, surgery, and adjuvant CAPOX plus or minus cetuximab in locally advanced rectal cancer:carriers of the G allele had a statistically significantly higher rate of complete response (CR) after neoadjuvant therapy and a trend for better 5-year PFS and OS rates. Both CR and survival outcomes were independent of cetuximab use. The negative prognostic effect associated with KRAS mutation appeared to be stronger in patients with the LCS-6 TT genotype compared to those with the LCS-6 TG genotype
Sclafani et al., 2015 [80]
+: aim of the study achieved; +/−: aim of the study partially achieved; −: aim of the study not achieved.
Table 4. Potential approaches for miRNA therapeutic delivery.
Table 4. Potential approaches for miRNA therapeutic delivery.
Delivery SystemMiRNA(s)MiRNA TypeMode of DeliveryCRC SubtypeTarget GeneReference
PolyethylenimineMiR-145
MiR-33a
Double-stranded RNA (dsRNA)Intraperitoneal
Intravenous
CRCc-Myc, ERK5Ibrahim et al., 2011 [110]
Carbonate apatiteMiRNA-4689Mature-miRIntravenousKRAS-mutant CRCKRAS, AKT1Hiraki et al., 2015 [123]
Carbonate apatiteMiRNA-29b-1-5pMimic-miRIntravenousKRAS-mutant CRCBCL-2, MCL1Inoue et al., 2018 [124]
ExosomeMiRNA-143Chemically modified RNA molecules (BP-miR) entrapped by MVs (microvesicules)IntravenousCRCNoneAkao et al., 2011 [125]
AtelocollagenMiRNA-34aPrecursor-miRNA (pre-miRNA or pre-miR)SubcutaneousCRCE2FTazawa et al., 2007 [126]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Al-Akhrass, H.; Christou, N. The Clinical Assessment of MicroRNA Diagnostic, Prognostic, and Theranostic Value in Colorectal Cancer. Cancers 2021, 13, 2916. https://doi.org/10.3390/cancers13122916

AMA Style

Al-Akhrass H, Christou N. The Clinical Assessment of MicroRNA Diagnostic, Prognostic, and Theranostic Value in Colorectal Cancer. Cancers. 2021; 13(12):2916. https://doi.org/10.3390/cancers13122916

Chicago/Turabian Style

Al-Akhrass, Hussein, and Niki Christou. 2021. "The Clinical Assessment of MicroRNA Diagnostic, Prognostic, and Theranostic Value in Colorectal Cancer" Cancers 13, no. 12: 2916. https://doi.org/10.3390/cancers13122916

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