Next Article in Journal
Identification of Alternatively-Activated Pathways between Primary Breast Cancer and Liver Metastatic Cancer Using Microarray Data
Next Article in Special Issue
Genistein Modulates Signaling Pathways and Targets Several Epigenetic Markers in HeLa Cells
Previous Article in Journal
Effects of Fat and Fatty Acids on the Formation of Autolysosomes in the Livers from Yellow Catfish Pelteobagrus Fulvidraco
Previous Article in Special Issue
Investigation of Precise Molecular Mechanistic Action of Tobacco-Associated Carcinogen ‘NNK’ Induced Carcinogenesis: A System Biology Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MicroRNA in Pancreatic Cancer: From Biology to Therapeutic Potential

1
Department of Internal Medicine, University of New Mexico School of Medicine, Albuquerque, NM 87131, USA
2
Department of Biotechnology, Kumaun University, Nainital, Uttarakhand 263001, India
3
Department of Internal Medicine, Loyola University Medical Center, Chicago, IL 60153, USA
4
Biosecurity and Public Health Group, Bioscience Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA
5
Department of Biomedical Engineering, Rutgers, The State University of New Jersey, Piscataway, NJ 08854, USA
*
Author to whom correspondence should be addressed.
Submission received: 31 July 2019 / Revised: 19 September 2019 / Accepted: 20 September 2019 / Published: 25 September 2019
(This article belongs to the Special Issue The Role of Genotoxicity in Infertility and Cancer Development)

Abstract

:
Pancreatic cancer is one of the most aggressive malignancies, accounting for more than 45,750 deaths annually in the U.S. alone. The aggressive nature and late diagnosis of pancreatic cancer, coupled with the limitations of existing chemotherapy, present the pressing need for the development of novel therapeutic strategies. Recent reports have demonstrated a critical role of microRNAs (miRNAs) in the initiation, progression, and metastasis of cancer. Furthermore, aberrant expressions of miRNAs have often been associated with the cause and consequence of pancreatic cancer, emphasizing the possible use of miRNAs in the effective management of pancreatic cancer patients. In this review, we provide a brief overview of miRNA biogenesis and its role in fundamental cellular process and miRNA studies in pancreatic cancer patients and animal models. Subsequent sections narrate the role of miRNA in, (i) cell cycle and proliferation; (ii) apoptosis; (iii) invasions and metastasis; and (iv) various cellular signaling pathways. We also describe the role of miRNA’s in pancreatic cancer; (i) diagnosis; (ii) prognosis and (iii) therapeutic intervention. Conclusion section describes the gist of review with future directions.

1. Introduction

Pancreatic cancer impinges profoundly on mankind and, is one of the most aggressive and fatal malignancies. Pancreatic cancer is the third leading cause of cancer-related deaths in the USA and is anticipated to take the second spot by 2020 [1,2]. In 2019, 45,750 pancreatic cancer deaths and 56,770 new cases were estimated in the USA alone [2]. Patients are often asymptomatic and thus diagnosed at a very advanced stage. Owing to the aggressive nature of this neoplasm, approximately 50% of patients already present with metastatic disease at the time of initial diagnosis [3]. Pancreas comprises of both exocrine and endocrine cells. The cancer of exocrine cells in the ducts of the pancreas which is known as pancreatic ductal adenocarcinoma (PDAC), which accounts for 90% of pancreatic cancer cases. Neoplasms arising from the endocrine cells are less common and are termed as “neuroendocrine tumors” [4]. PDAC is highly invasive and spreads to lymph nodes, liver, peritoneal cavity, lungs or the intestine, however, it rarely metastasizes to the brain or bones. In spite of a large number of on-going experimental treatment strategies for PDAC patients, the 5-year survival rate is very low. Once diagnosed and treatment has been initiated, approximately only 3–6% of patients survive up to 5 years [5,6]. The poor prognosis of PDAC patients mainly attributed to the late-stage identification and diagnosis, in combination with the invasive and aggressive nature of the neoplasm along with the resistance of pancreatic cancer cells to chemo and radiation therapies. Gemcitabine, the first-line treatment drug has moderate efficacy. Several modifications to its structures and attempts of combination therapy so far yielded only marginal improvement in gemcitabine drug efficacy [7]. Other drugs used in combination with gemcitabine such as fluorouracil, oxaliplatin, and irinotecan have yielded promising results, however, the patient survival and prognosis still remain grave [8,9]. Limitations of the existing therapeutic strategies emphasize the urgent need for improved understanding of the molecular pathways pertaining to early diagnosis, pathogenesis, and progression, in order to develop a novel more effective therapeutic regimen.
Despite enormous progress in the field of cancer research and treatment, the biomarkers for initiation and progression of pancreatic cancer are still not well-defined. Micro RNA (miRNA) might help to establish a better diagnostic tool for pancreatic cancer. Altered and aberrant expression of miRNA is a hallmark of several human malignancies [10]. Recent reports have revealed the quintessential role of miRNA in initiation, proliferation and progression, metastasis and chemo-resistance of cancer [11]. Recently, a miRNA based therapeutic formulation has been worked out for pancreatic cancer [12].
In this review, we provide a brief account of the biogenesis of miRNA and its effects on cellular functions and miRNA related studies in pancreatic cancer patients and animal models. We also review existing literature on the role of miRNA in cell cycle and cell proliferation, apoptosis, invasion, metastasis, and cellular signaling pathways. We also explore the use of miRNA as diagnostic, prognostic and therapeutic targets in pancreatic cancer.

2. Biogenesis of microRNA and Its Functions

miRNAs are 19–24 nucleotide bases of noncoding RNA that play a vital regulatory role in post-transcriptional gene expression [13,14]. miRNA was first discovered by Victor Ambros et al. in 1993 while working with Caenorhabditis elegans [15]. Since then, thousands of miRNAs have been identified in prokaryotic and eukaryotic organisms In humans, 2588 miRNA have been recognized with their complete biological identity, including sequence, location within the genome, and transcript annotation [16,17]. Biogenesis of miRNA requires the transcription of the miRNA gene (microRNAs can also be present in intronic regions of coding genes in eukaryotes) by RNA polymerase II to generate a longer precursor primary miRNA transcript (pri-miRNA) of variable length (100 to >1000 bps). The pri-miRNA is further processed in the nucleus by the ribonuclease complex Drosha-DGCR8 to form a hairpin-shaped intermediate pre-miRNA of 70–100 nucleotides [18,19]. To carry out the next processing step, pre-miRNA is exported to the cytoplasm with the help of transporter proteins Exportin 5 and Ran-GTP6 [20]. Inside the cytoplasm another endoribonuclease RNase III enzyme, DICER cleaves the terminal loop of the pre-miRNA to form a mature miRNA (double-stranded, 19–24 nucleotide) [21]. The mature miRNA is unwound in the cytoplasm such that one strand is degraded, and another mature strand is incorporated into the highly specialized Argonaut protein heteromultimer of the RNA induced silencing complex (RISC) [22,23]. miRNAs bind with their complementary sequence target mRNA, eventually downregulating it by either aiding cleaving of the target or by translational repression. miRNAs play a significant role in posttranscriptional gene regulation such that approximately 60% of all coding genes are estimated to be regulated by miRNA [24,25]. A single miRNA can regulate several mRNAs and vice versa, one mRNA can be a target of numerous miRNA. miRNAs play indispensable roles in several fundamental cellular processes like cell proliferation and differentiation, metabolism and apoptosis, signaling and hematopoiesis [26,27,28]. An overview of miRNA biogenesis and function is presented in Figure 1 [29].

3. microRNA in Pancreatic Cancer

In the last decade, the aberrant expression of miRNA has been associated with cancer [30]. Multiple reports have demonstrated that considerable dysregulation in the expression of miRNA is a cause and consequence in numerous human tumors [31,32]. Aberrant expression of miRNA is often coupled with complete anarchy in cell regulation that includes, uncontrolled cell proliferation, tumor suppressor avoidance, cell death modulation, tumor invasion, angiogenesis, and metastasis [30,33,34]. Comparative miRNA expression profiling studies using biopsy samples from pancreas tissue of healthy individuals, PDAC and pancreatitis patients have clearly indicated the differential expression of various miRNA in cancer cells as compared to normal cells, emphasizing the potential role of miRNAs in diagnostics, prognostics and anticancer therapies [35,36]. Poy et al. conducted the first miRNA expression profiling study in mouse pancreas, followed by various other studies with a different type of samples [37]. Hong and Parking revealed that a total of 158 miRNAs were found to be differentially expressed in patient’s PDAC tissue compared to adjacent normal pancreatic tissue. A total of 51 up-regulated miRNAs that include miR-196, miR-200a, miR-27a, and miR-21 together with 107 significantly downregulated miRNAs, such as miR-200, miR-96 and miR-217 [38]. Another study conducted by Schultz et al. on formalin-fixed paraffin-embedded tissue from pancreatic ductal adenocarcinoma and normal pancreas revealed the increased expression of 43 miRNAs whereas, 41 miRNAs were found to be downregulated [39]. In patients with resectable PDAC 22 differentially expressed miRNAs were unraveled with most significant differential expression observed among gene products; miR-136, miR-196, miR-492, miR-64, and miR-622 [40]. Papaconstantinou et al. conducted differential miRNAs expression studies by employing a large number of samples from pancreatic cancer patients and normal pancreatic tissues and reported the up-regulation of miR-222, miR-21, miR-210, miR-221 and miR-155 while, the expression of miR-146, miR-245, miR-122, and miR-31 were considerably declined in pancreatic cancer tissue [41]. Consistently, other studies have also reported a clear link between the aberrant expression of miR-21 and miR-155 with the progression of cancer [42]. Irrespective of the source of samples and profiling technique used, the up-regulation of miR-155, miR-221, miR-21 and down-regulation of miR-34 and miR-145 were constantly observed among most of the miRNA differential expression studies. In pancreatic cancer patients, apart from the aberrant expression of miRNAs in pancreatic cells and tissues, miRNA dysregulation has been also observed in the systemic circulation. For example, several studies have reported enhanced levels of miR-18a, miR-21, miR-22, miR-24, miR-25, miR-99a, miR-155, miR-185, miR-191, miR-196a, miR-642b and miR-885-5p in the blood of the patients affected with pancreatic cancer [43]. miR-2 has also been observed to be upregulated in serum [44]. Like PDAC, neuroendocrine tumors of the pancreas also had aberrant expression of miRNAs. One study found upregulated miR-155, miR-146a, miR-142-5p and miR-142-3p in cancerous tissues compared with normal healthy endocrine cells of the pancreas [45]. Importantly, miRNA expression profiling studies have not only assisted in the discrimination between healthy and cancerous tissues but have also been found to distinguish between various types of pancreatic pathologies, such as acute, benign or chronic. Bloomston et al. demonstrated that the up-regulation of 21 miRNAs and downregulation of 4 miRNAs could differentiate between benign and cancerous pancreatic tissues [46]. A list of miRNA expression profile in PADC is depicted in Figure 2.
All these studies show the important role of miRNAs in various biological processes beginning from the appearance of cancerous growth until its metastasis to a distant organ. A keen insight about the specific role of miRNAs during various processes could assist in a better understanding of malignancies at a molecular level paving the way for the development of novel diagnostic and therapeutic innovations. A brief summary of differentially expressed miRNAs with their targets and implications tabulated in Table 1.

4. Role of miRNAs in Cell Cycle and Cell Proliferation

Cell cycle and cell proliferation are tightly regulated by various checkpoints, tumor suppressor genes, oncogenes along with other controlling mechanisms. Recent studies have uncovered the regulatory role of miRNAs in cell cycle and proliferation by modulating the expression of oncogenes and tumor suppressor genes [6,47]. The overexpression or upregulation of oncogenic miRNAs negatively affects the expression of tumor suppressor genes, resulting in the augmentation of cell proliferation. PTEN, a tumor suppressor gene is negatively regulated by the overexpression of miR-21. The product of the PTEN gene hampers the proliferation of tumor cells and also controls the numbers of cell divisions in normal cells. The overexpression of miR-21 reduces the tumor-suppressive functions of PTEN by binding to the complementary region in the 3′ UTR region of PTEN mRNA transcript [48]. miRNA also control the various cyclin-dependent kinases (CDK) and cyclin complexes that control cell cycle progression. The translation of CDK N1B gene, a CDK inhibitor is influenced by overexpression of miR-221 in pancreatic cancer [49,50]. The p27 protein controls cell cycle progression to G1 phase, by preventing the activation of either the CDK2/Cyclin E or CDK4/CyclinD complex [48,51]. miR-424-5p is another oncogenic miRNA, whose overexpression in pancreatic cancer has been reported to downregulate the SOCS6 protein and result in increased ERK pathway activity [52] that enhance cell proliferation and migration. The up-regulation of oncogenic miR-27a negatively modulates the expression of tumor suppressor Spry 2 [53].
Apart from the overexpression of oncogenic miRNAs, the stumpy expression of several tumor suppressor miRNAs has also been observed in pancreatic cancer. These miRNAs regulate the expression of proto-oncogenes. The declined expression of miR-124 in the tissues of pancreatic cancer patients affects cell cycle regulation and cell motility. miR-124 negatively regulates the expression of the Rac1 oncogene of the MKK4-JNK-C-JUN pathway and inhibits cancer cell proliferation [54]. Rac1 is a GTPase protein, with multiple functions in cell cycle regulation, cell adhesion, motility, cytoskeletal re-organizations, epithelial differentiation and protein kinase activation [52,55]. The down-regulation of miR-124 is linked with poor survival among pancreatic cancer patients.
The down-regulation of miR-203 induces cell progression to the G1 phase, resulting in enhanced cell proliferation [51]. Xu et al. recently demonstrated that the up-regulation of miR-203, significantly impedes cell proliferation, induces apoptosis, and arrests cell cycle [56]. Down-regulation of another miRNA, miR-150 in pancreatic cancer, which further downregulates the MUC4 oncoprotein as it has been shown by Srivastava et al. MUC4 down-regulation has been shown to suppress the growth and malignancy of pancreatic cancer cells [6,47].
Along with these, there are numerous other miRNAs that regulate cell proliferation. Aberrant expression of miR-143, let-7-d, and miR-126 alter the expression of the KRAS oncogene and is correlated with abnormal cellular proliferation [51,57,58]. Cyclin E2, which is associated with the transition from G1 to S phase is directly regulated by miR-26a and is also a direct target of miR-223 [59,60]. miR-148a regulates the expression of CDC25B. The overexpression of miR-148a negatively affects the malignancy potential of pancreatic cancer cells [61]. All these studies demonstrate the important role of miRNAs in cell proliferation and growth in pancreatic cancer. A list of miRNAs that are dysregulated during initiation and progression stages of PDAC is depicted in Figure 3.

5. Role of miRNAs in Downregulation of Apoptosis in Pancreatic Cancer

Apoptosis or programmed cell death is a key regulatory mechanism for the maintenance of tissue homeostasis. DNA damage induces either DNA repair or apoptosis in a normal cell. However, in pancreatic cancer cells, the apoptotic pathway is likely to be perturbed, which can lead to immortality and acquisition of resistance to anticancer therapeutic agents, especially those targeting apoptotic pathways. Induction or inhibition of apoptosis is regulated by several miRNAs. miR-34a most likely regulates apoptosis by targeting Bcl-2 and Notch signaling modulators. The overexpression of miR-34a in pancreatic cancer cells, substantially downregulates the expression of Bcl-2, Notch 1 and Notch 2. The expression of miR-34a is dramatically downregulated or possibly completely abrogated in pancreatic cancer in comparison to normal tissues leading to a significant downregulation of Bsl-2, Notch 1, Notch 2 [57,62]. The overexpression of miR-155 in pancreatic cancer suppresses the pro-apoptotic gene p53 (TP53INP1). The miR-155 negatively regulates the expression of TP53INP1, resulting in the inhibition of apoptosis. Overexpression of TP53INP1 induces apoptosis and inhibits cancerous growth [63]. Aberrant expression of miR-203 also influences apoptosis. In normal cells, miR-203 regulates the expression of survivin, a baculoviral inhibitor of apoptosis. Down-expression of miR-203 in pancreatic cancer cells results in the upregulation of survivin, which in turn inhibits apoptosis [56,64]. There are several other miRNAs, which promote apoptosis such as miR-23a, miR-150, and miR-603. miR-23a regulates APAF1, which in turn induces caspase-9 with cytochrome c, which triggers apoptosis [65,66]. miR-630 and miR-150, both down-regulate the expression of IGF-1R, an insulin-like growth factor which inhibits apoptosis [67]. Recent reports also suggest that inhibition of miR-196a and overexpression of miR-24 promote apoptosis and inhibit invasion and proliferation. miR-196a along the with miR-214 targets ING4 and ING5 which in turn interacts with TP53 [68,69]. While, miR-24 directly regulates the expression of B1M which, comprise of a Bcl2 homology- 3 domains and is linked with the anti-apoptotic protein of BCL-2 superfamily [70].

6. Role of miRNAs in Pancreatic Cancer Invasion and Metastasis

Metastasis is a multi-step process and is one of the main causes of disease severity and mortality among cancer patients. It begins with the detachment of cancerous cells from the site of origin, followed by invasion into the bloodstream or lymphatic system through the basement membrane. Invasion requires epithelial to mesenchymal transition (EMT) of tumor cells, eventually, the tumor cells colonize distant sites. Thus, epithelial to mesenchymal transition (EMT) is a critical process. Various studies have reported several miRNAs regulating EMT, specifically the members of the miR-200 family (miR-141, miR-429, miR-200a, miR-200b, and miR-200c) [51,71,72,73]. Similarly, miR-203 and miR-208 can also trigger EMT [74,75]. Apart from EMT, numerous miRNAs have been identified for their role in invasion and metastasis. Up-regulation of miR-10a down-regulates homeobox transcription factors (HOX B1, HOX B2, and HOX A1) and induces metastasis in pancreatic cancer [76,77]. HOX factors are suppressors of transcription and metastasis. Huang et al. investigated the differential level of miRNA expression in metastatic and non-metastatic pancreatic cancer cell lines and observed the up-regulation of miR-100 in all metastatic cell lines [78]. Down-regulation of miR-34b has been reported in metastasis. miR-34b negatively regulates the expression of Smad 3, a promoter of metastasis [79]. Hence, low expression of miR-34b favors metastasis by up-regulating Smad 3 [79]. In 2010, Mees et al. identified the various regulatory miRNAs of a metastatic suppressor gene, EP 300 [80]. They observed a substantial up-regulation of miR-194, miR-429, miR-200b and miR-200c in metastatic adenocarcinoma cells compared to normal healthy control cells. EP 300 is negatively regulated by these miRNAs [80]. Hu et al. showed that the expression of miR-143 significantly declined in pancreatic cancer cells especially during invasion and metastasis. miR-143 negatively regulates the expression of GEF1, GEF2, K-RAS, MMP-2, and MMP-1 [57]. The down-regulation of miR-126 in PDAC results in the aberrant expression of one of its targets, ADAM9 [81]. Restoration of miR-126 has an inhibitory effect on the invasive potential of cancer cells and it reverses EMT. Like miR-126, expression of miR-146a also declines in metastatic pancreatic cancer tissue. Restoration of miR-146a significantly inhibits the invasive potential of pancreatic cancer cells by negatively regulating IRAK-1, EGFR, and MTA-2 [82]. He et al. suggested the role of miR-218 in metastatic pancreatic cancer using microarray analysis; they observed the gradual decline in the expression of miR-218 during progression from normal epithelium to pancreatic cancer to lymph node metastasis [83,84]. Up-regulation of miR-21 is correlated with liver metastasis and poor survival in the pancreatic cancer patient. Suppression of miR-21 results in the upregulation of PTEN and PDCD4, indicating that they may be the target of miR-21 [86]. Recently Yuan et al. have investigated the potential role of yet another, miR-4295, in the proliferation and invasion of pancreatic cancer cells. Further, it was suggested that a potential target of miR-4295 could be GPC5 (Glypican 5), as the suppression of miR-4295 resulted in the upregulation of GPC5, together with inhibited invasion via the stimulation of Wnt/β-catenin signaling pathway [85]. A Schematic illustration of the role of several tumor suppressor miRNAs at different stages in PDAC is highlighted in Figure 4.

7. Role of miRNAs in Signaling Pathways

7.1. KRAS Signaling Pathway

KRAS is a small membrane-bound guanosine triphosphate protein (21 kDa), belonging to the RAS family of GTPases that govern numerous signaling pathways, such as PI3K-AKT, RAF, N1-GEF, T--, MAPK and MAP2K, by altering GTP-bound (active) and GDP-bound (inactive) states in response to stimuli. Mutations in KRAS have been observed in almost 90% of PDAC patients in early stages as well as in advanced metastatic stages [88]. The mutated KRAS promotes the hyperactivation of downstream signaling pathways regulating cell survival and proliferation. Hyperactivation induces proliferation, invasion, and metastasis in pancreatic cancer. Recent studies have shown that numerous miRNAs negatively target the KRAS pathway by inhibiting the translation of, or by degrading the KRAS mRNA transcripts [89]. Here, we review the role of a few miRNAs involved in KRAS regulation in PDAC.
miR-96 binds to the KRAS at the 3′ UTR region and regulates cell proliferation. In pancreatic cancer, miR-96 is almost completely downregulated. miR-96 downregulates the expression of the HERG1 oncogene [90]. Restoration of miR-96 causes tumor suppression by negatively affecting cell proliferation, migration, and invasion [91]. Significant down-regulation has also been observed for miR-126. It also targets KRAS translation by binding to the 3′UTR of KRAS [58]. Mutated KRASG12D mediates the down-regulation of miR-143/145 via RREB (RAS- responsive element-binding protein) in pancreatic cancer cells. Mutated KRASG12D also promotes the expression of RREB1, which in response, negatively regulates the expression of miR-143/145 by binding to its promoter region. Oncogenic KRASG12D, RREB, and miR-143/145 operate synchronously via a feedback mechanism [92]. A reporter assay displayed, substantial down-regulation of miR-217 in PDAC cell lines. They also revealed that restoration of miR-217 downregulates the expression of KRAS and subsequently, there is a decrease in the phosphorylation of AKT, inhibiting cell survival and proliferation [93]. Lastly, downregulation or loss of Let 7 family miRNAs in pancreatic cancer cells leads to the overexpression of KRAS, as these miRNAs negatively regulate the expression of KRAS by binding to its 3′ UTR region [94].

7.2. AKT Signaling Pathway

The AKT/PI3K signal transduction pathway is related to cell survival and proliferation [95]. Several miRNAs have been reported to modulate the expression of genes involved in the AKT/PI3K signaling pathways. miR-21, miR-181a, and miR-221 target PTEN, a tumor suppressor gene which is found to be mutated in many human malignancies [48,66,96,97]. PTEN controls cell proliferation by negatively regulating the PI3K-AKT pathway [98]. Up-regulation of miR-21 and miR-221 in pancreatic cancer is associated with increased cell proliferation and metastasis. Inhibition of miR-221 in pancreatic cancer cells results in the up-regulation in the expression of tumor suppressor p27, p57, PUMA and PTEN [48]. Unlike miR-181a which induces migration of pancreatic cancer cells miR-200c and miR-375 act as a suppressor of tumor growth [97]. miR-375 suppresses the malignancy in PDAC cells by targeting the mRNA transcript of PDK1, (Pyruvate dehydrogenase lipoamide kinase isozyme 1) a kinase in the AKT pathway, downstream of PI3K [99,100]. Aberrant expression of miR-200c in PDAC is correlated with MUC4 and EMT [101,102]. MUC4 has been reported to stimulate AKT, which in turn activates downstream enzymes such as N-cadherin [103,104].

7.3. JAK/STAT Signaling Pathway

The Janus kinase/signal transducer and activator of transcription (JAK/STAT) is a critical pathway that promotes cell proliferation, cell migration, differentiation, and apoptosis. The JAK/STAT pathway is stimulated by various cytokines and growth factors [105]. miR-216a negatively regulates the expression of JAK 2 and restoration of miR-216a in PDAC cells reduces tumor growth possibly by inhibition of the translation of JAK 2 downstream oncogenes, including survivin, and induces apoptosis [106,107]. Similarly, miR-130b negatively regulates the expression of STAT3 by binding to its 3′UTR region. The significant downregulation of miR-130b in PDAC cells is associated with tumor progression and poor prognosis. Restoration of miR-130b inhibits the proliferation of cancer cells [108]. In contrast, miR-155 is an oncogenic miRNA and downregulates the expression of SOCS1, a tumor suppressor of the JAK/STAT pathway. miR-155 stimulate STAT3 resulting in the invasion and migration in cancer cells [109].

7.4. WNT/β-Catenin Signaling Pathway

This pathway is crucial for fundamental cellular processes such as cell proliferation, fate determination, and differentiation in embryonic development. In adults, it maintains homeostasis and in cancerous cells, it plays a crucial role in proliferation, invasion, and metastasis. Binding of extracellular WNT to a receptor complex comprising of the frizzled (FZD) transmembrane receptor and low-density lipoprotein receptor-related protein 5 and 6 (LRP) [110] results in the translocation of β-Catenin into the nucleus and activation of transcription of target genes [111]. Various miRNAs target these regulators of the WNT/β-Catenin pathway. FRATL, LRP-6, FZD4, and FZD5 are upstream regulators of WNT signaling and are the direct target of miR-29c. Several studies have demonstrated the down-regulation of miR-29c in PDAC cells. Transforming growth factor-β (TGF- β) inhibits the expression of miR-29c, resulting in the aberrant, constitutive activation of WNT [112]. A microarray study demonstrated that miR-23a and miR-24 expression regulate the expression of TMEM92, HNF1-β, and FZD5 directly or indirectly [113].

7.5. TGF-β Signaling Pathway

The TGF-β signaling is involved in basic biological processes such as cellular growth and differentiation, and apoptosis and cellular homeostasis, in adults and developing embryo [114]. Binding of a specific ligand to the type II TGF-β receptor, recruit type I TGF-β receptors resulting in the hetero-dimerization and trans-phosphorylation and activates SMADs, In humans, there are eight distinct SMADs protein, amongst these five SMADs, are modulated by TGF-β receptor (SMAD-1, SMAD-2, SMAD-3, SMAD-5, and SMAD-8) and known as R-SMADs [115]. Activation of TGF-β/SMAD-4 signaling pathway induces the overexpression of miR-155, which in turn induces invasion and metastasis [116]. Zhang et al. have shown that the expression of SMAD 4 is directly regulated by miR-199a suggesting a regulatory role of miR-199a in TGF-β/SMAD-4 signaling [117]. Diminished expression or inactivation of SMAD-4, a tumor suppressor gene has been observed in 55% of pancreatic cancer patients. In 35% of pancreatic cancer patients, it occurs due to a deletion in both alleles whereas, in 20% of pancreatic cancer individuals, an intragenic mutation could be the reason [118].
Recent studies have shown a vital role of miRNAs in regulating the various functions of dendritic cells, which eventually affects the immune response in cancer patients. A study conducted on dendritic cells in pancreatic cancer patients showed abnormal expression of miR-146a. Aberrant expression of miR-146a is associated with the downregulation of SMAD 4, which in turn negatively affects the differentiation and antigen presentation function in dendritic cells [119]. Up-regulation of miR-421 and miR-483-3p, inhibits the expression of tumor suppressor SMAD 4, resulting in the progression of pancreatic cancer [120].
In normal healthy cells, TGF-β suppresses cancerous growth by negatively regulating C-Myc, a transcription factor. Moreover, it promotes the expression of cell cycle inhibitors including p15 and p21 [121,122]. In cancer cells, aberrant functioning of the TGF-β pathway impairs its regulatory role in cell cycle arrest. Several studies have shown a wide array of miRNAs that regulate the downstream modulators/regulators of TGF-β pathway. The maturation of miR-21 with the help of a complex comprising of DROSHA and RNA helicase p68 is induced by TGF-β signaling. The up-regulation of miR-21 promotes tumor invasion and metastasis by negatively regulating the expression of tumor suppressor genes such as PTEN and PDCD4 [123]. According to a study conducted by Gasper et al., inhibition of TGF-β receptor by small inhibitors such as SB431542, LY-2157299, and SD-208, suppressed cancer progression and metastasis in mouse models [124]. Significant down-regulation of tumor-suppressing miR-494 has been observed in PDAC cells when compared to healthy normal cells. miR-494 negatively regulates the expression of FOXM1, a transcriptional activator. Diminished expression of miR-494, boost the expression of FOXM1 and nuclear translocation of β-catenin [125]. These studies suggest a critical interplay between WNT, and SMAD proteins of the TGF-β signaling pathways with various regulatory miRNAs.

8. Role of miRNAs in Pancreatic Cancer Diagnosis

One of the explanations for the poor prognosis and survival in pancreatic cancer patients is the late diagnosis of the disease, as it remains asymptomatic during the early stages. Thus, there is a pressing need for the identification and development of highly specific biomarkers for early diagnosis. At present, tissues biopsies are commonly used for the diagnosis. Carbohydrate antigen 19.9 is a blood antigen, approved as a biomarker for pancreatic cancer. However, its low specificity and suboptimal sensitivity limit its diagnostic application [126]. The crucial role of miRNAs during the progression of pancreatic cancer coupled with the high stability of miRNAs at extreme pH and temperature offers great promise in the development of a miRNA-based diagnostic system [127]. Several studies have built upon this initiative by demonstrating the specific changes in the expression profile of miRNAs during different stages such as invasion, proliferation, and metastasis [46,128,129]. The advent of recent technology advances and their application, such as Next-Generation-Sequencing, customized oligonucleotide-based microarray analysis, quantitative fluorescence probe-based PCR, in situ hybridization and Northern blotting using locked nucleic acid (LNA) modified probes, have all provided a number of specific miRNA expression profiles that could be used for diagnosis [130,131,132,133,134,135,136,137]. A microarray analysis study using blood samples comparison between pancreatic cancer patients and healthy individuals clearly showed a distinct miRNA expression profile (miR-22, miR-642b, and miR-885-5p) that identifies early pancreatic cancer [138]. The overexpression of seven miRNAs that include miR-21, miR-23a, miR-31, miR-100, miR-143, miR-155, miR-2214 and down-regulation of three miRNAs i.e., miR-148a, miR-375 and miR-217 differentiate a healthy control from pancreatic cancer patients [139]. In another study, overexpression of miR-196 in tissues samples coupled with low expression of miR-217 could identify a normal pancreas from PDAC [140]. Wang et al. demonstrated, that aberrant expression of miR-21, miR-155, miR-196a and miR-210 in plasma of PDAC patients, can easily distinguish from healthy controls [141]. High level of miR-210 in the serum of pancreatic cancer patients as compared to the normal individual may be exploited as a diagnostic marker [142]. Similarly, the level of miR-192 and miR-18a in the serum and plasma of pancreatic cancer patients could identify a healthy individual from a pancreatic cancer patient [43,143]. The high concentration of miR-155 and miR-210 in the serum of pancreatic cancer patient has been reported in various studies [42,142]. Due to the increased serum concentration of miR-155 in the early stages, it could be used for the diagnosis of early pancreatic neoplasia [42]. In an extensive study, Yu et al. investigated the miRNA expression profile in PanIN lesions. They analyzed 700 miRNAs and observed the aberrant expression of 35 miRNAs. It includes the overexpression of 24 miRNAs and reduced expression of 6 miRNAs in PanIN-3 lesions. Among all these 35 miRNAs, miR-196b surfaces as a potential biomarker in identifying PanIN-3 lesions [128]. In a similar study, Li et al. evaluated 735 miRNAs in the serum of pancreatic cancer patient and analysis lead to emergences of miR-1290 as a promising biomarker [144]. miR-1290 displays higher sensitivity (81%) and specificity (80%) when compared to healthy control groups [144]. In contrast to these approaches, a combination of various biomarkers brings discernible enhancement in the sensitivity and specificity. The combination of CA 19.9 with miR-16 and miR-196a can precisely distinguish between a pancreatic cancer patient and healthy controls [145]. Similarly, when the expression profile of miR-27a-3p was coupled with CA 19.9, pancreatic cancer patient and healthy controls could be differentiated with a sensitivity of 85.3% and specificity of 81.61% [51,146]. Recently, Lu et al. have investigated the expression profile of four miRNAs (miR-7, miR-9, miR-122, and miR-14) using qRT-PCR among severe acute pancreatitis patients. They observed that the significant surge in the serum levels of these four miRNAs as compared to healthy controls [114].

9. Role of miRNAs in Prognosis

Highly specific prognostic biomarkers could assist in the clinical management of pancreatic cancer patients (Figure 4) [87]. Recent studies have shown the utility of miRNAs in predicting prognosis, especially in terms of survival. The expression profile of miR-142-5p could be used as a biomarker for gemcitabine responsiveness. Down-regulation of miR-142-5p in PDAC patient samples correlates with gemcitabine resistance, while the up-regulation is associated with enhanced patient survival [77]. Similarly, the overexpression of miR-320c is observed in gemcitabine-resistant cells [147]. In contrast, low expression of miR-200c is associated with low median survival [102]. In a comprehensive study, Greither et al. examined the connection between the upregulation of four miRNAs (miR-222, miR-203, miR-155, miR-210) and survival in a study conducted with 56 PDAC patients. An elevated level of these four miRNAs was correlated with increased mortality by 6.2 fold in comparison to low levels of these miRNAs [148]. Roldo et al. investigated the role of six miRNAs i.e., miR-105, miR-127, miR-187, miR-309-3p, miR-45 and miR-518-a-2 as prognostic biomarkers and demonstrated that the expression profile of these miRNAs may be used to predict long-term survivors among node-positive patients [30]. Several studies have revealed the importance of miR-21 as a prognostic biomarker. Analysis of samples collected from PDAC patients shows that the upregulation of miR-21 is linked with a low rate of survival in node-negative PDAC patients and metastasis to lymph nodes as well as with poor gemcitabine response in the patient [149,150,151].

10. Role of miRNAs as Potential Therapeutic Targets

Given the decisive role of miRNAs in various stages of pancreatic cancer, starting from its initiation to progression to metastasis; miRNAs are potential targets for the therapeutic interventions [152]. miRNA-based anticancer therapies mainly involve introduction or restoration of tumor suppressor miRNAs as “miRNA mimics” and targets an oncogenic miRNA with the help of miRNA antagonists. A miRNA mimic is a chemically modified double-stranded miRNA that is used to mimic the function of a miRNA which is downregulated or diminished due to pancreatic cancer [153]. Kent et al. demonstrated that viral-mediated transduction of tumor suppressor miRNA, miR-143/145, inhibited the tumorous growth in pancreatic cancer cells [92]. Adenovirus-mediated delivery of miR-143 also exhibited an inhibitory effect in pancreatic cancer cells by blocking the metastasis [57]. Srivastava et al. reported that the restoration of miR-150 could substantially hinder the malignant potencies and growth of pancreatic cancer cells [47]. Likewise, replenishment of miR-34 demonstrated promising results by not only inhibiting pancreatic cancer cell growth but also enhancing their sensitivity to chemo and radiation therapies [154].
A separate approach comprises of miRNA antagonists. These are single-stranded antisense oligonucleotides corresponding to target miRNA, synthesized chemically with specific modification to provide high stability, binding affinity, and protection from nucleases [155,156]. They are complementary to the guide strand of target miRNAs and inhibit its activity by binding to seed region or by interfering in the miRNA biogenesis [157,158]. Along with these antagonist miRNAs, small-molecule inhibitors are also positively employed to target miRNAs in vitro [159,160]. Targeting oncogenic miR-21 with specific small molecule antagonist have yielded promising results and shown the inhibition of growth and cell proliferation in PDAC cells [48]. Similarly, silencing of miR-10a effectively inhibits metastasis in pancreatic cancer cells and primary human tumors [76]. Repression of miR-212 and miR-132 using antagonist miRNAs also inhibited the cancerous growth [161]. Passadouro et al. have demonstrated that the co-delivery of a human serum albumin-1-palmitoyl-2-oleoyl-sn-glycerol-3-ethylphosphocholine: cholesterol with an anti-miR oligonucleotide has efficiently suppressed the up-regulation of miRNAs of pancreatic cancer (miR-10, miR-21, miR-221 and miR-222) [162,163].

11. Conclusions

The association of miRNAs with cancer and using these miRNAs as therapeutic targets provides new hope for the effective diagnosis and treatment of fatal pancreatic cancer. Novel data about the biogenesis, regulation, and functions of miRNAs is increasing at an exponential rate. The cardinal regulatory role of miRNAs in key biological processes such as cell survival and proliferation, invasion and metastasis, apoptosis, and drug responsiveness, highlights their potential as diagnostic, prognostic and therapeutic targets. Also, little is known about inter-cellular communication mediated by miRNAs in pancreatic cancer. Moreover, a single miRNA target several genes and a single gene may be targeted by multiple miRNAs resulting in the additive action of miRNAs in the course of the disease. The aberrant expression of miRNAs is strongly correlated with pancreatic cancer, and in at least some cases appears to play a causative role in the establishment and progression of the disease. Restoration of down-regulated tumor suppressor miRNAs and inhibition of up-regulated oncogenic miRNAs holds the enormous prospect for the development of novel effective anticancer therapies or diagnostics. While there has been tremendous progress in the field with ample ongoing progress, there remain some limitations that need to be addressed before employing miRNAs as a therapeutic target. One of the most serious limitations is the potential for off-target effects and levels of toxicity of these chemically modified miRNAs on the healthy cells of the patients. The specific delivery of therapeutics into the pancreatic cancer cells is another consideration. Nonetheless, the recently developed intravenous injections and nanoparticle associated delivery system have shown great promise. Thus far, the progress in the field of miRNAs is impressive, but there is a long way to go. Pancreatic cancer remains a leading cause of mortality but the new body of work on the function of miRNA in cancer provides hope for miRNA-based strategies for early diagnosis, improved prognosis, and better management of clinical systems.

Author Contributions

Conceptualization and design, M.R. and G.P.; writing—original draft preparation, M.R., K.K., Y.G., A.K., P.S.G.C., O.K., S.K. and G.P.

Funding

This work received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Biancur, D.E.; Kimmelman, A.C. The plasticity of pancreatic cancer metabolism in tumor progression and therapeutic resistance. Biochim. Biophys. Acta Rev. Cancer 2018, 1870, 67–75. [Google Scholar] [CrossRef] [PubMed]
  2. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2019. CA Cancer J. Clin. 2019, 69, 7–34. [Google Scholar] [CrossRef] [PubMed]
  3. Sener, S.F.; Fremgen, A.; Menck, H.R.; Winchester, D.P. Pancreatic cancer: A report of treatment and survival trends for 100,313 patients diagnosed from 1985–1995, using the National Cancer Database. J. Am. Coll. Surg. 1999, 189, 1–7. [Google Scholar] [CrossRef]
  4. Hidalgo, M.; Cascinu, S.; Kleeff, J.; Labianca, R.; Lohr, J.M.; Neoptolemos, J.; Real, F.X.; Van Laethem, J.L.; Heinemann, V. Addressing the challenges of pancreatic cancer: Future directions for improving outcomes. Pancreatology 2015, 15, 8–18. [Google Scholar] [CrossRef] [PubMed]
  5. Arora, S.; Bhardwaj, A.; Singh, S.; Srivastava, S.K.; McClellan, S.; Nirodi, C.S.; Piazza, G.A.; Grizzle, W.E.; Owen, L.B.; Singh, A.P. An undesired effect of chemotherapy: Gemcitabine promotes pancreatic cancer cell invasiveness through reactive oxygen species-dependent, nuclear factor kappaB- and hypoxia-inducible factor 1alpha-mediated up-regulation of CXCR4. J. Biol. Chem. 2013, 288, 21197–21207. [Google Scholar] [CrossRef] [PubMed]
  6. Srivastava, S.K.; Arora, S.; Singh, S.; Bhardwaj, A.; Averett, C.; Singh, A.P. MicroRNAs in pancreatic malignancy: Progress and promises. Cancer Lett. 2014, 347, 167–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Moore, M.J.; Goldstein, D.; Hamm, J.; Figer, A.; Hecht, J.R.; Gallinger, S.; Au, H.J.; Murawa, P.; Walde, D.; Wolff, R.A.; et al. Erlotinib plus gemcitabine compared with gemcitabine alone in patients with advanced pancreatic cancer: A phase III trial of the National Cancer Institute of Canada Clinical Trials Group. J. Clin. Oncol. 2007, 25, 1960–1966. [Google Scholar] [CrossRef] [PubMed]
  8. Conroy, T.; Desseigne, F.; Ychou, M.; Ducreux, M.; Bouche, O.; Guimbaud, R.; Becouarn, Y.; Montoto-Grillot, C.; Gourgou-Bourgade, S.; Adenis, A.; et al. Randomized phase III trial comparing FOLFIRINOX (F: 5FU/leucovorin [LV], irinotecan [I], and oxaliplatin [O]) versus gemcitabine (G) as first-line treatment for metastatic pancreatic adenocarcinoma (MPA): Preplanned interim analysis results of the PRODIGE 4/ACCORD 11 trial. J. Clin. Oncol. 2010, 28, 4010. [Google Scholar]
  9. Falasca, M.; Kim, M.; Casari, I. Pancreatic cancer: Current research and future directions. Biochim. Biophys. Acta 2016, 1865, 123–132. [Google Scholar] [CrossRef]
  10. Slack, F.J.; Weidhaas, J.B. MicroRNA in cancer prognosis. N. Engl. J. Med. 2008, 359, 2720–2722. [Google Scholar] [CrossRef]
  11. Iorio, M.V.; Croce, C.M. Causes and consequences of microRNA dysregulation. Cancer J. 2012, 18, 215–222. [Google Scholar] [CrossRef] [PubMed]
  12. Gilles, M.-E.; Hao, L.; Huang, L.; Rupaimoole, R.; Lopez-Casas, P.P.; Pulver, E.; Jeong, J.C.; Muthuswamy, S.K.; Hidalgo, M.; Bhatia, S.N.; et al. Personalized RNA medicine for pancreatic cancer. Clin. Cancer Res. 2018, 24, 1734–1747. [Google Scholar] [CrossRef] [PubMed]
  13. Friedman, R.C.; Farh, K.K.; Burge, C.B.; Bartel, D.P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009, 19, 92–105. [Google Scholar] [CrossRef]
  14. Lewis, B.P.; Burge, C.B.; Bartel, D.P. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 2005, 120, 15–20. [Google Scholar] [CrossRef] [PubMed]
  15. Lee, R.C.; Feinbaum, R.L.; Ambros, V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 1993, 75, 843–854. [Google Scholar] [CrossRef]
  16. Kozomara, A.; Griffiths-Jones, S. miRBase: Annotating high confidence microRNAs using deep sequencing data. Nucleic Acids Res. 2014, 42, D68–D73. [Google Scholar] [CrossRef] [PubMed]
  17. Malczewska, A.; Kidd, M.; Matar, S.; Kos-Kudla, B.; Modlin, I.M. A Comprehensive Assessment of the Role of miRNAs as Biomarkers in Gastroenteropancreatic Neuroendocrine Tumors. Neuroendocrinology 2018, 107, 73–90. [Google Scholar] [CrossRef] [PubMed]
  18. Church, V.A.; Pressman, S.; Isaji, M.; Truscott, M.; Cizmecioglu, N.T.; Buratowski, S.; Frolov, M.V.; Carthew, R.W. Microprocessor Recruitment to Elongating RNA Polymerase II Is Required for Differential Expression of MicroRNAs. Cell Rep. 2017, 20, 3123–3134. [Google Scholar] [CrossRef] [Green Version]
  19. Garzon, R.; Marcucci, G.; Croce, C.M. Targeting microRNAs in cancer: Rationale, strategies and challenges. Nat. Rev. Drug Discov. 2010, 9, 775–789. [Google Scholar] [CrossRef]
  20. Bohnsack, M.T.; Czaplinski, K.; Gorlich, D. Exportin 5 is a RanGTP-dependent dsRNA-binding protein that mediates nuclear export of pre-miRNAs. RNA 2004, 10, 185–191. [Google Scholar] [CrossRef] [Green Version]
  21. Valencia-Sanchez, M.A.; Liu, J.; Hannon, G.J.; Parker, R. Control of translation and mRNA degradation by miRNAs and siRNAs. Genes Dev. 2006, 20, 515–524. [Google Scholar] [CrossRef] [PubMed]
  22. Hammond, S.M.; Bernstein, E.; Beach, D.; Hannon, G.J. An RNA-directed nuclease mediates post-transcriptional gene silencing in Drosophila cells. Nature 2000, 404, 293–296. [Google Scholar] [CrossRef] [PubMed]
  23. Chendrimada, T.P.; Gregory, R.I.; Kumaraswamy, E.; Norman, J.; Cooch, N.; Nishikura, K.; Shiekhattar, R. TRBP recruits the Dicer complex to Ago2 for microRNA processing and gene silencing. Nature 2005, 436, 740–744. [Google Scholar] [CrossRef] [PubMed]
  24. Krek, A.; Grun, D.; Poy, M.N.; Wolf, R.; Rosenberg, L.; Epstein, E.J.; MacMenamin, P.; da Piedade, I.; Gunsalus, K.C.; Stoffel, M.; et al. Combinatorial microRNA target predictions. Nat. Genet. 2005, 37, 495–500. [Google Scholar] [CrossRef] [PubMed]
  25. Betel, D.; Wilson, M.; Gabow, A.; Marks, D.S.; Sander, C. The microRNA.org resource: Targets and expression. Nucleic Acids Res. 2008, 36, D149–D153. [Google Scholar] [CrossRef]
  26. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef]
  27. Bartel, D.P. Metazoan MicroRNAs. Cell 2018, 173, 20–51. [Google Scholar] [CrossRef]
  28. Esquela-Kerscher, A.; Slack, F.J. Oncomirs—microRNAs with a role in cancer. Nat. Rev. Cancer 2006, 6, 259–269. [Google Scholar] [CrossRef]
  29. Ranganna, K.; Mathew, O.P.; Milton, S.G.; Hayes, B.E. MicroRNAome of vascular smooth muscle cells: Potential for MicroRNA-based vascular therapies. In Current Trends in Atherogenesis; InTechOpen: London, UK, 2013; p. 152. [Google Scholar]
  30. Roldo, C.; Missiaglia, E.; Hagan, J.P.; Falconi, M.; Capelli, P.; Bersani, S.; Calin, G.A.; Volinia, S.; Liu, C.G.; Scarpa, A.; et al. MicroRNA expression abnormalities in pancreatic endocrine and acinar tumors are associated with distinctive pathologic features and clinical behavior. J. Clin. Oncol. 2006, 24, 4677–4684. [Google Scholar] [CrossRef]
  31. Drakaki, A.; Iliopoulos, D. MicroRNA-gene signaling pathways in pancreatic cancer. Biomed. J. 2013, 36, 200–208. [Google Scholar]
  32. Macha, M.A.; Seshacharyulu, P.; Krishn, S.R.; Pai, P.; Rachagani, S.; Jain, M.; Batra, S.K. MicroRNAs (miRNAs) as biomarker(s) for prognosis and diagnosis of gastrointestinal (GI) cancers. Curr. Pharm. Des. 2014, 20, 5287–5297. [Google Scholar] [CrossRef] [PubMed]
  33. Calin, G.A.; Croce, C.M. MicroRNA signatures in human cancers. Nat. Rev. Cancer 2006, 6, 857–866. [Google Scholar] [CrossRef] [PubMed]
  34. Garzon, R.; Fabbri, M.; Cimmino, A.; Calin, G.A.; Croce, C.M. MicroRNA expression and function in cancer. Trends Mol. Med. 2006, 12, 580–587. [Google Scholar] [CrossRef] [PubMed]
  35. Khan, S.A.; Kumar, D.; Jaggi, M.; Chauhan, S.C. Targeting microRNAs in pancreatic cancer: Microplayers in the big game. Cancer Res. 2013, 73, 6541–6547. [Google Scholar] [CrossRef] [PubMed]
  36. Zen, K.; Zhang, C.Y. Circulating microRNAs: A novel class of biomarkers to diagnose and monitor human cancers. Med. Res. Rev. 2012, 32, 326–348. [Google Scholar] [CrossRef] [PubMed]
  37. Poy, M.N.; Eliasson, L.; Krutzfeldt, J.; Kuwajima, S.; Ma, X.; Macdonald, P.E.; Pfeffer, S.; Tuschl, T.; Rajewsky, N.; Rorsman, P.; et al. A pancreatic islet-specific microRNA regulates insulin secretion. Nature 2004, 432, 226–230. [Google Scholar] [CrossRef] [PubMed]
  38. Hong, T.H.; Park, I.Y. MicroRNA expression profiling of diagnostic needle aspirates from surgical pancreatic cancer specimens. Ann. Surg. Treat. Res. 2014, 87, 290–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Schultz, N.A.; Werner, J.; Willenbrock, H.; Roslind, A.; Giese, N.; Horn, T.; Wojdemann, M.; Johansen, J.S. MicroRNA expression profiles associated with pancreatic adenocarcinoma and ampullary adenocarcinoma. Mod. Pathol. 2012, 25, 1609–1622. [Google Scholar] [CrossRef] [Green Version]
  40. Calatayud, D.; Dehlendorff, C.; Boisen, M.K.; Hasselby, J.P.; Schultz, N.A.; Werner, J.; Immervoll, H.; Molven, A.; Hansen, C.P.; Johansen, J.S. Tissue MicroRNA profiles as diagnostic and prognostic biomarkers in patients with resectable pancreatic ductal adenocarcinoma and periampullary cancers. Biomark. Res. 2017, 5, 8. [Google Scholar] [CrossRef]
  41. Papaconstantinou, I.G.; Manta, A.; Gazouli, M.; Lyberopoulou, A.; Lykoudis, P.M.; Polymeneas, G.; Voros, D. Expression of microRNAs in patients with pancreatic cancer and its prognostic significance. Pancreas 2013, 42, 67–71. [Google Scholar] [CrossRef]
  42. Habbe, N.; Koorstra, J.B.; Mendell, J.T.; Offerhaus, G.J.; Ryu, J.K.; Feldmann, G.; Mullendore, M.E.; Goggins, M.G.; Hong, S.M.; Maitra, A. MicroRNA miR-155 is a biomarker of early pancreatic neoplasia. Cancer Biol. Ther. 2009, 8, 340–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Morimura, R.; Komatsu, S.; Ichikawa, D.; Takeshita, H.; Tsujiura, M.; Nagata, H.; Konishi, H.; Shiozaki, A.; Ikoma, H.; Okamoto, K.; et al. Novel diagnostic value of circulating miR-18a in plasma of patients with pancreatic cancer. Br. J. Cancer 2011, 105, 1733–1740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ali, S.; Almhanna, K.; Chen, W.; Philip, P.A.; Sarkar, F.H. Differentially expressed miRNAs in the plasma may provide a molecular signature for aggressive pancreatic cancer. Am. J. Transl. Res. 2010, 3, 28–47. [Google Scholar]
  45. Olson, P.; Lu, J.; Zhang, H.; Shai, A.; Chun, M.G.; Wang, Y.; Libutti, S.K.; Nakakura, E.K.; Golub, T.R.; Hanahan, D. MicroRNA dynamics in the stages of tumorigenesis correlate with hallmark capabilities of cancer. Genes Dev. 2009, 23, 2152–2165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Bloomston, M.; Frankel, W.L.; Petrocca, F.; Volinia, S.; Alder, H.; Hagan, J.P.; Liu, C.G.; Bhatt, D.; Taccioli, C.; Croce, C.M. MicroRNA expression patterns to differentiate pancreatic adenocarcinoma from normal pancreas and chronic pancreatitis. JAMA 2007, 297, 1901–1908. [Google Scholar] [CrossRef] [PubMed]
  47. Srivastava, S.K.; Bhardwaj, A.; Singh, S.; Arora, S.; Wang, B.; Grizzle, W.E.; Singh, A.P. MicroRNA-150 directly targets MUC4 and suppresses growth and malignant behavior of pancreatic cancer cells. Carcinogenesis 2011, 32, 1832–1839. [Google Scholar] [CrossRef] [PubMed]
  48. Park, J.K.; Lee, E.J.; Esau, C.; Schmittgen, T.D. Antisense inhibition of microRNA-21 or -221 arrests cell cycle, induces apoptosis, and sensitizes the effects of gemcitabine in pancreatic adenocarcinoma. Pancreas 2009, 38, e190–e199. [Google Scholar] [CrossRef]
  49. Basu, A.; Alder, H.; Khiyami, A.; Leahy, P.; Croce, C.M.; Haldar, S. MicroRNA-375 and MicroRNA-221: Potential Noncoding RNAs Associated with Antiproliferative Activity of Benzyl Isothiocyanate in Pancreatic Cancer. Genes Cancer 2011, 2, 108–119. [Google Scholar] [CrossRef] [Green Version]
  50. Su, A.; He, S.; Tian, B.; Hu, W.; Zhang, Z. MicroRNA-221 mediates the effects of PDGF-BB on migration, proliferation, and the epithelial-mesenchymal transition in pancreatic cancer cells. PLoS ONE 2013, 8, e71309. [Google Scholar] [CrossRef]
  51. Halkova, T.; Cuperkova, R.; Minarik, M.; Benesova, L. MicroRNAs in Pancreatic Cancer: Involvement in Carcinogenesis and Potential Use for Diagnosis and Prognosis. Gastroenterol. Res. Pract. 2015, 2015, 892903. [Google Scholar] [CrossRef]
  52. Wu, K.; Hu, G.; He, X.; Zhou, P.; Li, J.; He, B.; Sun, W. MicroRNA-424-5p suppresses the expression of SOCS6 in pancreatic cancer. Pathol. Oncol. Res. 2013, 19, 739–748. [Google Scholar] [CrossRef] [PubMed]
  53. Ma, Y.; Yu, S.; Zhao, W.; Lu, Z.; Chen, J. miR-27a regulates the growth, colony formation and migration of pancreatic cancer cells by targeting Sprouty2. Cancer Lett. 2010, 298, 150–158. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, P.; Chen, L.; Zhang, J.; Chen, H.; Fan, J.; Wang, K.; Luo, J.; Chen, Z.; Meng, Z.; Liu, L. Methylation-mediated silencing of the miR-124 genes facilitates pancreatic cancer progression and metastasis by targeting Rac1. Oncogene 2014, 33, 514–524. [Google Scholar] [CrossRef] [PubMed]
  55. Ridley, A.J. Rho GTPases and actin dynamics in membrane protrusions and vesicle trafficking. Trends Cell Biol. 2006, 16, 522–529. [Google Scholar] [CrossRef] [PubMed]
  56. Xu, D.; Wang, Q.; An, Y.; Xu, L. MiR203 regulates the proliferation, apoptosis and cell cycle progression of pancreatic cancer cells by targeting Survivin. Mol. Med. Rep. 2013, 8, 379–384. [Google Scholar] [CrossRef] [PubMed]
  57. Hu, Y.; Ou, Y.; Wu, K.; Chen, Y.; Sun, W. miR-143 inhibits the metastasis of pancreatic cancer and an associated signaling pathway. Tumor Biol. 2012, 33, 1863–1870. [Google Scholar] [CrossRef]
  58. Jiao, L.R.; Frampton, A.E.; Jacob, J.; Pellegrino, L.; Krell, J.; Giamas, G.; Tsim, N.; Vlavianos, P.; Cohen, P.; Ahmad, R.; et al. MicroRNAs targeting oncogenes are down-regulated in pancreatic malignant transformation from benign tumors. PLoS ONE 2012, 7, e32068. [Google Scholar] [CrossRef] [PubMed]
  59. Deng, J.; He, M.; Chen, L.; Chen, C.; Zheng, J.; Cai, Z. The loss of miR-26a-mediated post-transcriptional regulation of cyclin E2 in pancreatic cancer cell proliferation and decreased patient survival. PLoS ONE 2013, 8, e76450. [Google Scholar] [CrossRef]
  60. Ma, J.; Fang, B.; Zeng, F.; Ma, C.; Pang, H.; Cheng, L.; Shi, Y.; Wang, H.; Yin, B.; Xia, J.; et al. Down-regulation of miR-223 reverses epithelial-mesenchymal transition in gemcitabine-resistant pancreatic cancer cells. Oncotarget 2015, 6, 1740–1749. [Google Scholar] [CrossRef] [Green Version]
  61. Liffers, S.T.; Munding, J.B.; Vogt, M.; Kuhlmann, J.D.; Verdoodt, B.; Nambiar, S.; Maghnouj, A.; Mirmohammadsadegh, A.; Hahn, S.A.; Tannapfel, A. MicroRNA-148a is down-regulated in human pancreatic ductal adenocarcinomas and regulates cell survival by targeting CDC25B. Lab. Investig. 2011, 91, 1472–1479. [Google Scholar] [CrossRef] [Green Version]
  62. Tang, Y.; Tang, Y.; Cheng, Y.S. miR-34a inhibits pancreatic cancer progression through Snail1-mediated epithelial-mesenchymal transition and the Notch signaling pathway. Sci Rep. 2017, 7, 38232. [Google Scholar] [CrossRef]
  63. Gironella, M.; Seux, M.; Xie, M.J.; Cano, C.; Tomasini, R.; Gommeaux, J.; Garcia, S.; Nowak, J.; Yeung, M.L.; Jeang, K.T.; et al. Tumor protein 53-induced nuclear protein 1 expression is repressed by miR-155, and its restoration inhibits pancreatic tumor development. Proc. Natl. Acad. Sci. USA 2007, 104, 16170–16175. [Google Scholar] [CrossRef] [Green Version]
  64. Wang, W.; Liu, W.; Sun, H.; Chen, D.; Yao, X.; Zhao, J. miR-203 inhibits proliferation of HCC cells by targeting survivin. Cell Biochem. Funct. 2013, 31, 82–85. [Google Scholar]
  65. Liu, N.; Sun, Y.Y.; Zhang, X.W.; Chen, S.; Wang, Y.; Zhang, Z.X.; Song, S.W.; Qiu, G.B.; Fu, W.N. Oncogenic miR-23a in Pancreatic Ductal Adenocarcinogenesis Via Inhibiting APAF1. Dig. Sci. 2015, 60, 2000–2008. [Google Scholar] [CrossRef] [PubMed]
  66. Yonemori, K.; Kurahara, H.; Maemura, K.; Natsugoe, S. MicroRNA in pancreatic cancer. J. Hum. Genet. 2017, 62, 33–40. [Google Scholar] [CrossRef] [PubMed]
  67. Farhana, L.; Dawson, M.I.; Murshed, F.; Das, J.K.; Rishi, A.K.; Fontana, J.A. Upregulation of miR-150* and miR-630 induces apoptosis in pancreatic cancer cells by targeting IGF-1R. PLoS ONE 2013, 8, e61015. [Google Scholar] [CrossRef] [PubMed]
  68. Zhang, X.J.; Ye, H.; Zeng, C.W.; He, B.; Zhang, H.; Chen, Y.Q. Dysregulation of miR-15a and miR-214 in human pancreatic cancer. J. Hematol. Oncol. 2010, 3, 46. [Google Scholar] [CrossRef]
  69. Liu, M.; Du, Y.; Gao, J.; Liu, J.; Kong, X.; Gong, Y.; Li, Z.; Wu, H.; Chen, H. Aberrant expression miR-196a is associated with abnormal apoptosis, invasion, and proliferation of pancreatic cancer cells. Pancreas 2013, 42, 1169–1181. [Google Scholar] [CrossRef]
  70. Liu, R.; Zhang, H.; Wang, X.; Zhou, L.; Li, H.; Deng, T.; Qu, Y.; Duan, J.; Bai, M.; Ge, S.; et al. The miR-24-Bim pathway promotes tumor growth and angiogenesis in pancreatic carcinoma. Oncotarget 2015, 6, 43831–43842. [Google Scholar] [CrossRef] [PubMed]
  71. Burk, U.; Schubert, J.; Wellner, U.; Schmalhofer, O.; Vincan, E.; Spaderna, S.; Brabletz, T. A reciprocal repression between ZEB1 and members of the miR-200 family promotes EMT and invasion in cancer cells. EMBO Rep. 2008, 9, 582–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Bracken, C.P.; Gregory, P.A.; Kolesnikoff, N.; Bert, A.G.; Wang, J.; Shannon, M.F.; Goodall, G.J. A double-negative feedback loop between ZEB1-SIP1 and the microRNA-200 family regulates epithelial-mesenchymal transition. Cancer Res. 2008, 68, 7846–7854. [Google Scholar] [CrossRef] [PubMed]
  73. Paterson, E.L.; Kolesnikoff, N.; Gregory, P.A.; Bert, A.G.; Khew-Goodall, Y.; Goodall, G.J. The microRNA-200 family regulates epithelial to mesenchymal transition. Sci. World J. 2008, 8, 901–904. [Google Scholar] [CrossRef] [PubMed]
  74. Miao, L.; Xiong, X.; Lin, Y.; Cheng, Y.; Lu, J.; Zhang, J.; Cheng, N. miR-203 inhibits tumor cell migration and invasion via caveolin-1 in pancreatic cancer cells. Oncol. Lett. 2014, 7, 658–662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Liu, A.; Shao, C.; Jin, G.; Liu, R.; Hao, J.; Song, B.; Ouyang, L.; Hu, X. miR-208-induced epithelial to mesenchymal transition of pancreatic cancer cells promotes cell metastasis and invasion. Cell Biochem. Biophys. 2014, 69, 341–346. [Google Scholar] [CrossRef] [PubMed]
  76. Weiss, F.U.; Marques, I.J.; Woltering, J.M.; Vlecken, D.H.; Aghdassi, A.; Partecke, L.I.; Heidecke, C.D.; Lerch, M.M.; Bagowski, C.P. Retinoic acid receptor antagonists inhibit miR-10a expression and block metastatic behavior of pancreatic cancer. Gastroenterology 2009, 137, 2136–2145.e7. [Google Scholar] [CrossRef] [PubMed]
  77. Ohuchida, K.; Mizumoto, K.; Lin, C.; Yamaguchi, H.; Ohtsuka, T.; Sato, N.; Toma, H.; Nakamura, M.; Nagai, E.; Hashizume, M.; et al. MicroRNA-10a is overexpressed in human pancreatic cancer and involved in its invasiveness partially via suppression of the HOXA1 gene. Ann. Surg. Oncol. 2012, 19, 2394–2402. [Google Scholar] [CrossRef] [PubMed]
  78. Huang, J.S.; Egger, M.E.; Grizzle, W.E.; McNally, L.R. MicroRNA-100 regulates IGF1-receptor expression in metastatic pancreatic cancer cells. Biotech. Histochem. 2013, 88, 397–402. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, C.; Cheng, H.; Shi, S.; Cui, X.; Yang, J.; Chen, L.; Cen, P.; Cai, X.; Lu, Y.; Wu, C.; et al. MicroRNA-34b inhibits pancreatic cancer metastasis through repressing Smad3. Curr. Mol. Med. 2013, 13, 467–478. [Google Scholar] [CrossRef]
  80. Mees, S.T.; Mardin, W.A.; Wendel, C.; Baeumer, N.; Willscher, E.; Senninger, N.; Schleicher, C.; Colombo-Benkmann, M.; Haier, J. EP300—A miRNA-regulated metastasis suppressor gene in ductal adenocarcinomas of the pancreas. Int. J. Cancer 2010, 126, 114–124. [Google Scholar] [CrossRef]
  81. Frampton, A.E.; Krell, J.; Jacob, J.; Stebbing, J.; Castellano, L.; Jiao, L.R. Loss of miR-126 is crucial to pancreatic cancer progression. Expert Rev. Anticancer Ther. 2012, 12, 881–884. [Google Scholar] [CrossRef]
  82. Li, Y.; Vandenboom, T.G., 2nd; Wang, Z.; Kong, D.; Ali, S.; Philip, P.A.; Sarkar, F.H. miR-146a suppresses invasion of pancreatic cancer cells. Cancer Res. 2010, 70, 1486–1495. [Google Scholar] [CrossRef] [PubMed]
  83. He, H.; Di, Y.; Liang, M.; Yang, F.; Yao, L.; Hao, S.; Li, J.; Jiang, Y.; Jin, C.; Fu, D. The microRNA-218 and ROBO-1 signaling axis correlates with the lymphatic metastasis of pancreatic cancer. Oncol. Rep. 2013, 30, 651–658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. He, H.; Hao, S.J.; Yao, L.; Yang, F.; Di, Y.; Li, J.; Jiang, Y.J.; Jin, C.; Fu, D.L. MicroRNA-218 inhibits cell invasion and migration of pancreatic cancer via regulating ROBO1. Cancer Biol. Ther. 2014, 15, 1333–1339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Yuan, Q.; Zhang, Y.; Li, J.; Cao, G.; Yang, W. High expression of microRNA-4295 contributes to cell proliferation and invasion of pancreatic ductal adenocarcinoma by the down-regulation of Glypican-5. Biochem. Biophys. Res. Commun. 2018, 497, 73–79. [Google Scholar] [CrossRef] [PubMed]
  86. Wei, X.; Wang, W.; Wang, L.; Zhang, Y.; Zhang, X.; Chen, M.; Wang, F.; Yu, J.; Ma, Y.; Sun, G. MicroRNA-21 induces 5-fluorouracil resistance in human pancreatic cancer cells by regulating PTEN and PDCD4. Cancer Med. 2016, 5, 693–702. [Google Scholar] [CrossRef] [PubMed]
  87. Chitkara, D.; Mittal, A.; Mahato, R.I. miRNAs in pancreatic cancer: Therapeutic potential, delivery challenges and strategies. Adv. Drug Deliv. Rev. 2015, 81, 34–52. [Google Scholar] [CrossRef] [PubMed]
  88. di Magliano, M.P.; Logsdon, C.D. Roles for KRAS in pancreatic tumor development and progression. Gastroenterology 2013, 144, 1220–1229. [Google Scholar] [CrossRef]
  89. Collins, M.A.; Bednar, F.; Zhang, Y.; Brisset, J.C.; Galban, S.; Galban, C.J.; Rakshit, S.; Flannagan, K.S.; Adsay, N.V.; Pasca di Magliano, M. Oncogenic Kras is required for both the initiation and maintenance of pancreatic cancer in mice. J. Clin. Investig. 2012, 122, 639–653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Feng, J.; Yu, J.; Pan, X.; Li, Z.; Chen, Z.; Zhang, W.; Wang, B.; Yang, L.; Xu, H.; Zhang, G.; et al. HERG1 functions as an oncogene in pancreatic cancer and is downregulated by miR-96. Oncotarget 2014, 5, 5832–5844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Yu, S.; Lu, Z.; Liu, C.; Meng, Y.; Ma, Y.; Zhao, W.; Liu, J.; Yu, J.; Chen, J. miRNA-96 suppresses KRAS and functions as a tumor suppressor gene in pancreatic cancer. Cancer Res. 2010, 70, 6015–6025. [Google Scholar] [CrossRef] [PubMed]
  92. Kent, O.A.; Chivukula, R.R.; Mullendore, M.; Wentzel, E.A.; Feldmann, G.; Lee, K.H.; Liu, S.; Leach, S.D.; Maitra, A.; Mendell, J.T. Repression of the miR-143/145 cluster by oncogenic Ras initiates a tumor-promoting feed-forward pathway. Genes Dev. 2010, 24, 2754–2759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Zhao, W.G.; Yu, S.N.; Lu, Z.H.; Ma, Y.H.; Gu, Y.M.; Chen, J. The miR-217 microRNA functions as a potential tumor suppressor in pancreatic ductal adenocarcinoma by targeting KRAS. Carcinogenesis 2010, 31, 1726–1733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Johnson, S.M.; Grosshans, H.; Shingara, J.; Byrom, M.; Jarvis, R.; Cheng, A.; Labourier, E.; Reinert, K.L.; Brown, D.; Slack, F.J. RAS is regulated by the let-7 microRNA family. Cell 2005, 120, 635–647. [Google Scholar] [CrossRef] [PubMed]
  95. Ferro, R.; Falasca, M. Emerging role of the KRAS-PDK1 axis in pancreatic cancer. World J. Gastroenterol. 2014, 20, 10752–10757. [Google Scholar] [CrossRef] [PubMed]
  96. Sarkar, S.; Dubaybo, H.; Ali, S.; Goncalves, P.; Kollepara, S.L.; Sethi, S.; Philip, P.A.; Li, Y. Down-regulation of miR-221 inhibits proliferation of pancreatic cancer cells through up-regulation of PTEN, p27(kip1), p57(kip2), and PUMA. Am. J. Cancer Res. 2013, 3, 465–477. [Google Scholar] [PubMed]
  97. Liu, J.; Xu, D.; Wang, Q.; Zheng, D.; Jiang, X.; Xu, L. LPS induced miR-181a promotes pancreatic cancer cell migration via targeting PTEN and MAP2K4. Dig. Sci. 2014, 59, 1452–1460. [Google Scholar] [CrossRef] [PubMed]
  98. Song, M.S.; Salmena, L.; Pandolfi, P.P. The functions and regulation of the PTEN tumour suppressor. Nat. Rev. Mol. Cell Biol. 2012, 13, 283–296. [Google Scholar] [CrossRef]
  99. Zhou, J.; Song, S.; He, S.; Zhu, X.; Zhang, Y.; Yi, B.; Zhang, B.; Qin, G.; Li, D. MicroRNA-375 targets PDK1 in pancreatic carcinoma and suppresses cell growth through the Akt signaling pathway. Int. J. Mol. Med. 2014, 33, 950–956. [Google Scholar] [CrossRef]
  100. Song, S.D.; Zhou, J.; Zhou, J.; Zhao, H.; Cen, J.N.; Li, D.C. MicroRNA-375 targets the 3-phosphoinositide-dependent protein kinase-1 gene in pancreatic carcinoma. Oncol. Lett. 2013, 6, 953–959. [Google Scholar] [CrossRef] [Green Version]
  101. Li, Y.; VandenBoom, T.G., 2nd; Kong, D.; Wang, Z.; Ali, S.; Philip, P.A.; Sarkar, F.H. Up-regulation of miR-200 and let-7 by natural agents leads to the reversal of epithelial-to-mesenchymal transition in gemcitabine-resistant pancreatic cancer cells. Cancer Res. 2009, 69, 6704–6712. [Google Scholar] [CrossRef]
  102. Yu, J.; Ohuchida, K.; Mizumoto, K.; Sato, N.; Kayashima, T.; Fujita, H.; Nakata, K.; Tanaka, M. MicroRNA, hsa-miR-200c, is an independent prognostic factor in pancreatic cancer and its upregulation inhibits pancreatic cancer invasion but increases cell proliferation. Mol. Cancer 2010, 9, 169. [Google Scholar] [CrossRef] [PubMed]
  103. Radhakrishnan, P.; Mohr, A.M.; Grandgenett, P.M.; Steele, M.M.; Batra, S.K.; Hollingsworth, M.A. MicroRNA-200c modulates the expression of MUC4 and MUC16 by directly targeting their coding sequences in human pancreatic cancer. PLoS ONE 2013, 8, e73356. [Google Scholar] [CrossRef] [PubMed]
  104. Rachagani, S.; Macha, M.A.; Ponnusamy, M.P.; Haridas, D.; Kaur, S.; Jain, M.; Batra, S.K. MUC4 potentiates invasion and metastasis of pancreatic cancer cells through stabilization of fibroblast growth factor receptor 1. Carcinogenesis 2012, 33, 1953–1964. [Google Scholar] [CrossRef] [PubMed]
  105. Rawlings, J.S.; Rosler, K.M.; Harrison, D.A. The JAK/STAT signaling pathway. J. Cell Sci. 2004, 117, 1281–1283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Hou, B.H.; Jian, Z.X.; Cui, P.; Li, S.J.; Tian, R.Q.; Ou, J.R. miR-216a may inhibit pancreatic tumor growth by targeting JAK2. FEBS Lett. 2015, 589, 2224–2232. [Google Scholar] [CrossRef] [Green Version]
  107. Wang, S.; Chen, X.; Tang, M. MicroRNA-216a inhibits pancreatic cancer by directly targeting Janus kinase 2. Oncol. Rep. 2014, 32, 2824–2830. [Google Scholar] [CrossRef] [Green Version]
  108. Zhao, G.; Zhang, J.G.; Shi, Y.; Qin, Q.; Liu, Y.; Wang, B.; Tian, K.; Deng, S.C.; Li, X.; Zhu, S.; et al. MiR-130b is a prognostic marker and inhibits cell proliferation and invasion in pancreatic cancer through targeting STAT3. PLoS ONE 2013, 8, e73803. [Google Scholar]
  109. Huang, C.; Li, H.; Wu, W.; Jiang, T.; Qiu, Z. Regulation of miR-155 affects pancreatic cancer cell invasiveness and migration by modulating the STAT3 signaling pathway through SOCS1. Oncol. Rep. 2013, 30, 1223–1230. [Google Scholar] [CrossRef]
  110. Morris, J.P., IV; Wang, S.C.; Hebrok, M. KRAS, Hedgehog, Wnt and the twisted developmental biology of pancreatic ductal adenocarcinoma. Nat. Rev. Cancer 2010, 10, 683–695. [Google Scholar] [CrossRef]
  111. Reya, T.; Clevers, H. Wnt signalling in stem cells and cancer. Nature 2005, 434, 843–850. [Google Scholar] [CrossRef]
  112. Jiang, J.; Yu, C.; Chen, M.; Zhang, H.; Tian, S.; Sun, C. Reduction of miR-29c enhances pancreatic cancer cell migration and stem cell-like phenotype. Oncotarget 2015, 6, 2767–2778. [Google Scholar] [CrossRef] [PubMed]
  113. Listing, H.; Mardin, W.A.; Wohlfromm, S.; Mees, S.T.; Haier, J. MiR-23a/-24-induced gene silencing results in mesothelial cell integration of pancreatic cancer. Br. J. Cancer 2015, 112, 131–139. [Google Scholar] [CrossRef] [PubMed]
  114. Whitman, M. Smads and early developmental signaling by the TGFbeta superfamily. Genes Dev. 1998, 12, 2445–2462. [Google Scholar] [CrossRef] [PubMed]
  115. Massague, J.; Seoane, J.; Wotton, D. Smad transcription factors. Genes Dev. 2005, 19, 2783–2810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Kong, W.; Yang, H.; He, L.; Zhao, J.J.; Coppola, D.; Dalton, W.S.; Cheng, J.Q. MicroRNA-155 is regulated by the transforming growth factor beta/Smad pathway and contributes to epithelial cell plasticity by targeting RhoA. Mol. Cell. Biol. 2008, 28, 6773–6784. [Google Scholar] [CrossRef]
  117. Zhang, Y.; Fan, K.J.; Sun, Q.; Chen, A.Z.; Shen, W.L.; Zhao, Z.H.; Zheng, X.F.; Yang, X. Functional screening for miRNAs targeting Smad4 identified miR-199a as a negative regulator of TGF-beta signalling pathway. Nucleic Acids Res. 2012, 40, 9286–9297. [Google Scholar] [CrossRef] [PubMed]
  118. Hruban, R.H.; Offerhaus, G.J.; Kern, S.E.; Goggins, M.; Wilentz, R.E.; Yeo, C.J. Tumor-suppressor genes in pancreatic cancer. J. Hepato-Biliary-Pancreat. Surg. 1998, 5, 383–391. [Google Scholar] [CrossRef]
  119. Du, J.; Wang, J.; Tan, G.; Cai, Z.; Zhang, L.; Tang, B.; Wang, Z. Aberrant elevated microRNA-146a in dendritic cells (DC) induced by human pancreatic cancer cell line BxPC-3-conditioned medium inhibits DC maturation and activation. Med. Oncol. 2012, 29, 2814–2823. [Google Scholar] [CrossRef]
  120. Hao, J.; Zhang, S.; Zhou, Y.; Hu, X.; Shao, C. MicroRNA 483-3p suppresses the expression of DPC4/Smad4 in pancreatic cancer. FEBS Lett. 2011, 585, 207–213. [Google Scholar] [CrossRef]
  121. Adhikary, S.; Eilers, M. Transcriptional regulation and transformation by Myc proteins. Nat. Rev. Mol. Cell Biol. 2005, 6, 635–645. [Google Scholar] [CrossRef]
  122. Donovan, J.; Slingerland, J. Transforming growth factor-beta and breast cancer: Cell cycle arrest by transforming growth factor-beta and its disruption in cancer. Breast Cancer Res. 2000, 2, 116–124. [Google Scholar] [CrossRef] [PubMed]
  123. Davis, B.N.; Hilyard, A.C.; Lagna, G.; Hata, A. SMAD proteins control DROSHA-mediated microRNA maturation. Nature 2008, 454, 56–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Gaspar, N.J.; Li, L.; Kapoun, A.M.; Medicherla, S.; Reddy, M.; Li, G.; O’Young, G.; Quon, D.; Henson, M.; Damm, D.L.; et al. Inhibition of transforming growth factor beta signaling reduces pancreatic adenocarcinoma growth and invasiveness. Mol. Pharmacol. 2007, 72, 152–161. [Google Scholar] [CrossRef] [PubMed]
  125. Li, L.; Li, Z.; Kong, X.; Xie, D.; Jia, Z.; Jiang, W.; Cui, J.; Du, Y.; Wei, D.; Huang, S.; et al. Down-regulation of microRNA-494 via loss of SMAD4 increases FOXM1 and beta-catenin signaling in pancreatic ductal adenocarcinoma cells. Gastroenterology 2014, 147, 485–497.e18. [Google Scholar] [CrossRef] [PubMed]
  126. Humphris, J.L.; Chang, D.K.; Johns, A.L.; Scarlett, C.J.; Pajic, M.; Jones, M.D.; Colvin, E.K.; Nagrial, A.; Chin, V.T.; Chantrill, L.A.; et al. The prognostic and predictive value of serum CA19.9 in pancreatic cancer. Ann. Oncol. 2012, 23, 1713–1722. [Google Scholar] [CrossRef]
  127. Mitchell, P.S.; Parkin, R.K.; Kroh, E.M.; Fritz, B.R.; Wyman, S.K.; Pogosova-Agadjanyan, E.L.; Peterson, A.; Noteboom, J.; O’Briant, K.C.; Allen, A.; et al. Circulating microRNAs as stable blood-based markers for cancer detection. Proc. Natl. Acad. Sci. USA 2008, 105, 10513–10518. [Google Scholar] [CrossRef] [Green Version]
  128. Yu, J.; Li, A.; Hong, S.M.; Hruban, R.H.; Goggins, M. MicroRNA alterations of pancreatic intraepithelial neoplasias. Clin. Cancer Res. 2012, 18, 981–992. [Google Scholar] [CrossRef]
  129. Li, Y.; Sarkar, F.H. MicroRNA targeted therapeutic approach for pancreatic cancer. Int. J. Biol. Sci. 2016, 12, 326. [Google Scholar] [CrossRef]
  130. Morin, R.D.; O’Connor, M.D.; Griffith, M.; Kuchenbauer, F.; Delaney, A.; Prabhu, A.L.; Zhao, Y.; McDonald, H.; Zeng, T.; Hirst, M.; et al. Application of massively parallel sequencing to microRNA profiling and discovery in human embryonic stem cells. Genome Res. 2008, 18, 610–621. [Google Scholar] [CrossRef] [Green Version]
  131. Liang, R.Q.; Li, W.; Li, Y.; Tan, C.Y.; Li, J.X.; Jin, Y.X.; Ruan, K.C. An oligonucleotide microarray for microRNA expression analysis based on labeling RNA with quantum dot and nanogold probe. Nucleic Acids Res. 2005, 33, e17. [Google Scholar] [CrossRef]
  132. Thomson, J.M.; Parker, J.; Perou, C.M.; Hammond, S.M. A custom microarray platform for analysis of microRNA gene expression. Nat. Methods 2004, 1, 47–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Valoczi, A.; Hornyik, C.; Varga, N.; Burgyan, J.; Kauppinen, S.; Havelda, Z. Sensitive and specific detection of microRNAs by northern blot analysis using LNA-modified oligonucleotide probes. Nucleic Acids Res. 2004, 32, e175. [Google Scholar] [CrossRef] [PubMed]
  134. Duncan, D.D.; Eshoo, M.; Esau, C.; Freier, S.M.; Lollo, B.A. Absolute quantitation of microRNAs with a PCR-based assay. Anal. Biochem. 2006, 359, 268–270. [Google Scholar] [CrossRef] [PubMed]
  135. Raymond, C.K.; Roberts, B.S.; Garrett-Engele, P.; Lim, L.P.; Johnson, J.M. Simple, quantitative primer-extension PCR assay for direct monitoring of microRNAs and short-interfering RNAs. RNA 2005, 11, 1737–1744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Deo, M.; Yu, J.Y.; Chung, K.H.; Tippens, M.; Turner, D.L. Detection of mammalian microRNA expression by in situ hybridization with RNA oligonucleotides. Dev. Dyn. 2006, 235, 2538–2548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Kloosterman, W.P.; Wienholds, E.; de Bruijn, E.; Kauppinen, S.; Plasterk, R.H. In situ detection of miRNAs in animal embryos using LNA-modified oligonucleotide probes. Nat. Methods 2006, 3, 27–29. [Google Scholar] [CrossRef] [PubMed]
  138. Ganepola, G.A.; Rutledge, J.R.; Suman, P.; Yiengpruksawan, A.; Chang, D.H. Novel blood-based microRNA biomarker panel for early diagnosis of pancreatic cancer. World J. Gastrointest. Oncol. 2014, 6, 22–33. [Google Scholar] [CrossRef] [PubMed]
  139. Frampton, A.E.; Giovannetti, E.; Jamieson, N.B.; Krell, J.; Gall, T.M.; Stebbing, J.; Jiao, L.R.; Castellano, L. A microRNA meta-signature for pancreatic ductal adenocarcinoma. Expert Rev. Mol. Diagn. 2014, 14, 267–271. [Google Scholar] [CrossRef] [PubMed]
  140. Szafranska, A.E.; Davison, T.S.; John, J.; Cannon, T.; Sipos, B.; Maghnouj, A.; Labourier, E.; Hahn, S.A. MicroRNA expression alterations are linked to tumorigenesis and non-neoplastic processes in pancreatic ductal adenocarcinoma. Oncogene 2007, 26, 4442–4452. [Google Scholar] [CrossRef] [Green Version]
  141. Wang, J.; Chen, J.; Chang, P.; LeBlanc, A.; Li, D.; Abbruzzesse, J.L.; Frazier, M.L.; Killary, A.M.; Sen, S. MicroRNAs in plasma of pancreatic ductal adenocarcinoma patients as novel blood-based biomarkers of disease. Cancer Prev. Res. 2009, 2, 807–813. [Google Scholar] [CrossRef]
  142. Ho, A.S.; Huang, X.; Cao, H.; Christman-Skieller, C.; Bennewith, K.; Le, Q.T.; Koong, A.C. Circulating miR-210 as a Novel Hypoxia Marker in Pancreatic Cancer. Transl. Oncol. 2010, 3, 109–113. [Google Scholar] [CrossRef] [Green Version]
  143. Zhao, C.; Zhang, J.; Zhang, S.; Yu, D.; Chen, Y.; Liu, Q.; Shi, M.; Ni, C.; Zhu, M. Diagnostic and biological significance of microRNA-192 in pancreatic ductal adenocarcinoma. Oncol. Rep. 2013, 30, 276–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Li, A.; Yu, J.; Kim, H.; Wolfgang, C.L.; Canto, M.I.; Hruban, R.H.; Goggins, M. MicroRNA array analysis finds elevated serum miR-1290 accurately distinguishes patients with low-stage pancreatic cancer from healthy and disease controls. Clin. Cancer Res. 2013, 19, 3600–3610. [Google Scholar] [CrossRef] [PubMed]
  145. Liu, J.; Gao, J.; Du, Y.; Li, Z.; Ren, Y.; Gu, J.; Wang, X.; Gong, Y.; Wang, W.; Kong, X. Combination of plasma microRNAs with serum CA19-9 for early detection of pancreatic cancer. Int. J. Cancer 2012, 131, 683–691. [Google Scholar] [CrossRef] [PubMed]
  146. Wang, W.S.; Liu, L.X.; Li, G.P.; Chen, Y.; Li, C.Y.; Jin, D.Y.; Wang, X.L. Combined serum CA19-9 and miR-27a-3p in peripheral blood mononuclear cells to diagnose pancreatic cancer. Cancer Prev. Res. 2013, 6, 331–338. [Google Scholar] [CrossRef]
  147. Iwagami, Y.; Eguchi, H.; Nagano, H.; Akita, H.; Hama, N.; Wada, H.; Kawamoto, K.; Kobayashi, S.; Tomokuni, A.; Tomimaru, Y.; et al. miR-320c regulates gemcitabine-resistance in pancreatic cancer via SMARCC1. Br. J. Cancer 2013, 109, 502–511. [Google Scholar] [CrossRef] [Green Version]
  148. Greither, T.; Grochola, L.F.; Udelnow, A.; Lautenschlager, C.; Wurl, P.; Taubert, H. Elevated expression of microRNAs 155, 203, 210 and 222 in pancreatic tumors is associated with poorer survival. Int. J. Cancer 2010, 126, 73–80. [Google Scholar] [CrossRef]
  149. Dillhoff, M.; Liu, J.; Frankel, W.; Croce, C.; Bloomston, M. MicroRNA-21 is overexpressed in pancreatic cancer and a potential predictor of survival. J. Gastrointest. Surg. 2008, 12, 2171–2176. [Google Scholar] [CrossRef]
  150. Giovannetti, E.; Funel, N.; Peters, G.J.; Del Chiaro, M.; Erozenci, L.A.; Vasile, E.; Leon, L.G.; Pollina, L.E.; Groen, A.; Falcone, A.; et al. MicroRNA-21 in pancreatic cancer: Correlation with clinical outcome and pharmacologic aspects underlying its role in the modulation of gemcitabine activity. Cancer Res. 2010, 70, 4528–4538. [Google Scholar] [CrossRef]
  151. Kadera, B.E.; Li, L.; Toste, P.A.; Wu, N.; Adams, C.; Dawson, D.W.; Donahue, T.R. MicroRNA-21 in pancreatic ductal adenocarcinoma tumor-associated fibroblasts promotes metastasis. PLoS ONE 2013, 8, e71978. [Google Scholar] [CrossRef]
  152. Bader, A.G. miR-34—A microRNA replacement therapy is headed to the clinic. Front. Genet. 2012, 3, 120. [Google Scholar] [CrossRef] [PubMed]
  153. Pai, P.; Rachagani, S.; Are, C.; Batra, S.K. Prospects of miRNA-based therapy for pancreatic cancer. Curr. Drug Targets 2013, 14, 1101–1109. [Google Scholar] [CrossRef] [PubMed]
  154. Ji, Q.; Hao, X.; Zhang, M.; Tang, W.; Yang, M.; Li, L.; Xiang, D.; Desano, J.T.; Bommer, G.T.; Fan, D.; et al. MicroRNA miR-34 inhibits human pancreatic cancer tumor-initiating cells. PLoS ONE 2009, 4, e6816. [Google Scholar] [CrossRef]
  155. Davis, S.; Propp, S.; Freier, S.M.; Jones, L.E.; Serra, M.J.; Kinberger, G.; Bhat, B.; Swayze, E.E.; Bennett, C.F.; Esau, C. Potent inhibition of microRNA in vivo without degradation. Nucleic Acids Res. 2009, 37, 70–77. [Google Scholar] [CrossRef] [PubMed]
  156. 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]
  157. Esau, C.C. Inhibition of microRNA with antisense oligonucleotides. Methods 2008, 44, 55–60. [Google Scholar] [CrossRef] [PubMed]
  158. Krutzfeldt, J.; Rajewsky, N.; Braich, R.; Rajeev, K.G.; Tuschl, T.; Manoharan, M.; Stoffel, M. Silencing of microRNAs in vivo with “antagomirs”. Nature 2005, 438, 685–689. [Google Scholar] [CrossRef] [PubMed]
  159. Gumireddy, K.; Young, D.D.; Xiong, X.; Hogenesch, J.B.; Huang, Q.; Deiters, A. Small-molecule inhibitors of microrna miR-21 function. Angew. Chem. Int. Ed. Engl. 2008, 47, 7482–7484. [Google Scholar] [CrossRef]
  160. Young, D.D.; Connelly, C.M.; Grohmann, C.; Deiters, A. Small molecule modifiers of microRNA miR-122 function for the treatment of hepatitis C virus infection and hepatocellular carcinoma. J. Am. Chem. Soc. 2010, 132, 7976–7981. [Google Scholar] [CrossRef]
  161. Park, J.-K.; Henry, J.C.; Jiang, J.; Esau, C.; Gusev, Y.; Lerner, M.R.; Postier, R.G.; Brackett, D.J.; Schmittgen, T.D. miR-132 and miR-212 are increased in pancreatic cancer and target the retinoblastoma tumor suppressor. Biochem. Biophys. Res. Commun. 2011, 406, 518–523. [Google Scholar] [CrossRef] [Green Version]
  162. Passadouro, M.; Pedroso de Lima, M.C.; Faneca, H. MicroRNA modulation combined with sunitinib as a novel therapeutic strategy for pancreatic cancer. Int. J. Nanomed. 2014, 9, 3203–3217. [Google Scholar]
  163. Rachagani, S.; Macha, M.A.; Heimann, N.; Seshacharyulu, P.; Haridas, D.; Chugh, S.; Batra, S.K. Clinical implications of miRNAs in the pathogenesis, diagnosis and therapy of pancreatic cancer. Adv. Drug Deliv. Rev. 2015, 81, 16–33. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biogenesis of miRNA and its functions. The miRNA synthesis starts in the nucleus where pri-miRNA transcript 100 to >1000 nucleotides is generated by RNA polymerase II. Subsequently, pri-miRNA is cleaved by Drosha/DGCR8 to form ~70–100 nucleotides long hairpin loop pre-miRNA. Pre-miRNA is then transported from nucleus to cytoplasm through Exportin 5 and Ran-GTP6 wherein it is further processed by RNase activity of Dicer to 19–24 nucleotides double-stranded mature miRNA duplex. The miRNA duplex then loads onto Ago in the RISC complex and undergoes strand separation. The guide strand of the miRNA mediates gene silencing by degrading the target mRNA or interfering with the translational process. The passenger strand gets degraded. Adopted from [29] © 2013 Ranganna K, Mathew OM, Milton SG, Hayes BE under a CC-BY 3.0 license.
Figure 1. Biogenesis of miRNA and its functions. The miRNA synthesis starts in the nucleus where pri-miRNA transcript 100 to >1000 nucleotides is generated by RNA polymerase II. Subsequently, pri-miRNA is cleaved by Drosha/DGCR8 to form ~70–100 nucleotides long hairpin loop pre-miRNA. Pre-miRNA is then transported from nucleus to cytoplasm through Exportin 5 and Ran-GTP6 wherein it is further processed by RNase activity of Dicer to 19–24 nucleotides double-stranded mature miRNA duplex. The miRNA duplex then loads onto Ago in the RISC complex and undergoes strand separation. The guide strand of the miRNA mediates gene silencing by degrading the target mRNA or interfering with the translational process. The passenger strand gets degraded. Adopted from [29] © 2013 Ranganna K, Mathew OM, Milton SG, Hayes BE under a CC-BY 3.0 license.
Genes 10 00752 g001
Figure 2. The major structures of the pancreatic gland indicating the origin of PDAC. A cataloug of commonly downregulated miRNAs (in green), upregulated miRNAs (in red) and differentially regulated miRNAs (in blue) in PDAC are listed in different colors. Differentially regulated miRNAs refers to either up or down-regulated miRNAs compared to healthy pancreatic tissue.
Figure 2. The major structures of the pancreatic gland indicating the origin of PDAC. A cataloug of commonly downregulated miRNAs (in green), upregulated miRNAs (in red) and differentially regulated miRNAs (in blue) in PDAC are listed in different colors. Differentially regulated miRNAs refers to either up or down-regulated miRNAs compared to healthy pancreatic tissue.
Genes 10 00752 g002
Figure 3. Schematic illustration of different stages of pancreatic ductal adenocarcinoma (PDAC). The list of miRNAs that are upregulated (in red color) and downregulated (in green color) during initiation and progression stages of PDAC are listed.
Figure 3. Schematic illustration of different stages of pancreatic ductal adenocarcinoma (PDAC). The list of miRNAs that are upregulated (in red color) and downregulated (in green color) during initiation and progression stages of PDAC are listed.
Genes 10 00752 g003
Figure 4. Schematic illustration of the role of tumor suppressor miRNAs at different stages of pancreatic ductal adenocarcinoma (PDAC). The figure depicts the interrelationship between pancreatic cancer cells (PCCs), pancreatic stellate cells (PSCs) and cancer stem cells (CSCs). The epithelial pancreatic cells undergo epithelial-mesenchymal transition (EMT) and become mesenchymal cells which bear high propensity to metastasize. EMT could be reversed back to mesenchymal-to-epithelial transition (MET) by the intervention of certain miRNAs. This figure is adopted and modified from [87], © 2015 Chitkara D, Mittal A, Mahato RI. under a CC BY-NC-SA 4.0 license.
Figure 4. Schematic illustration of the role of tumor suppressor miRNAs at different stages of pancreatic ductal adenocarcinoma (PDAC). The figure depicts the interrelationship between pancreatic cancer cells (PCCs), pancreatic stellate cells (PSCs) and cancer stem cells (CSCs). The epithelial pancreatic cells undergo epithelial-mesenchymal transition (EMT) and become mesenchymal cells which bear high propensity to metastasize. EMT could be reversed back to mesenchymal-to-epithelial transition (MET) by the intervention of certain miRNAs. This figure is adopted and modified from [87], © 2015 Chitkara D, Mittal A, Mahato RI. under a CC BY-NC-SA 4.0 license.
Genes 10 00752 g004
Table 1. Summary of differentially expressed microRNAs with their targets and implications in pancreatic cancer.
Table 1. Summary of differentially expressed microRNAs with their targets and implications in pancreatic cancer.
miRNA/sSample Type/Site of ActionRegulationTarget (-ve)/(+ve)ImplicationRef.
miR-196, miR-200a, miR-27a, and miR-21PDAC tissue/adjacent normal pancreatic tissueUp 51↑[38]
miR-200, miR-96, and miR-217Down107↓
miR-136, miR-196, miR-492, miR-64, and miR-622PDAC vs. healthy pancreatic duct tissueDifferential 43↑[39]
41↓
miR-21 and miR-155PDAC vs. healthy pancreatic duct tissueDifferential Cancer progression↑[40]
miR-222, miR-21, miR-210, miR-221 and miR-155PDAC vs. healthy pancreatic duct large sample sizeUp 43↑[41]
miR-146, miR-245, miR-122, and miR-31Down41↓[41]
miR-155, miR-221, miR-21PDAC vs. healthy pancreatic duct tissueUp Common to most studies[38,39,41,42]
miR-34 and miR-145PDAC vs. healthy pancreatic duct tissueDown
miR-18a, miR-21, miR-22, miR-24, miR-25, miR-99a, miR-155, miR-185, miR-191, miR-196a, miR-642b and miR-885-5pBlood of PDAC patientsUp Chronic inflamation↑[43,44]
miR-2Serum of PDAC patientsUp
miR-155, miR-146a, miR-142-5p and miR-142-3pNeuroendocrine tumors/adjacent normal pancreatic tissueUp [45]
miR-150Various cancer samplesDownMUC4 (-ve)Tumor growth↑
Malignancy↑
[6,47]
miR-21Various cancer samplesUpPTEN (-ve)Oncogenic/poor survival patients[48]
miR-221Various cancer samplesUpCDK N1B (-ve)Oncogenic[49,50]
miR-424-5pVarious cancer samplesUpSOCS6 (-ve)Oncogenic[52]
miR-27aVarious cancer samplesUpSpry 2 (-ve)Oncogenic[53]
miR-124Various cancer samplesDownRac1 (-ve)Oncogenic/poor survival patients[54]
miR-143Metastatic pancreatic cancer; microarray analysisDownGEF1, GEF2, K-RAS, MMP-2, and MMP-1 (-ve)Invasive potential↑
EMT↑
[57]
miR-143, let-7-d, and miR-126Various cancer samples KRAS oncogeneAbnormal cellular proliferation↑[51,57,58]
miR-26a, miR-223Various cancer samples Cyclin E2Abnormal cellular proliferation↑[59,60]
miR-148aVarious cancer samplesUpCDC25B (-ve)Malignancy↓[61]
miR-34aVarious cancer samplesDownBcl-2, Notch 1 and Notch 2 (-ve)Tumor growth↑
Malignancy↑
[57,62]
miR-155Various cancer samplesUpp53 (TP53INP1) (-ve)Abnormal cellular proliferation↓[63]
miR-203Various cancer samplesDownSurvivin (-ve)Rampant tumor growth↑
EMT↑
[64]
miR-23a Various cancer samplesDownAPAF1caspase-9 with cytochrome c
Apoptosis↑
[65,66]
miR-630, miR-150Various cancer samplesUpIGF-1R (-ve)Apoptosis↑[67]
miR-196a, miR-214Various cancer samplesDownING4 and ING5TP53↓
Tumor growth↑
[68,69]
miR-24Various cancer samplesUpB1M (-ve)Tumor growth↓
Malignancy↓
[70]
miR-200 family (miR-141, miR-429, miR-200a, miR-200b, and miR-200c)Various metastatic cancer samplesUp/Down Metastasis↕[51,71,72,73]
miR-208Metastatic cell linesUP Metastasis↑[74,75]
miR-10aMetastatic cell linesDownhomeobox transcription factors (HOX B1, HOX B2, and HOX A1) (-ve) [76,77]
miR-100Metastatic cell linesUp [78]
miR-34bMetastatic adenocarcinoma cellsDownSmad 3 (-ve)EMT↑[79]
miR-194, miR-429, miR-200b and miR-200cMetastatic adenocarcinoma cellsUpEP 300 (-ve)EMT↑[80]
miR-126PDAC progressive samples with metastasisDownADAM9 (-ve)Invasive potential↑
EMT↑
[81]
miR-146aMetastatic pancreatic cancer; microarray analysisDownIRAK-1, EGFR, and MTA-2 (-ve)Invasive potential↑[82]
miR-218Metastatic pancreatic cancer; microarray analysisDown Progression and metastasis[83,84]
miR-4295Pancreatic cancer cellsDownGPC5↑proliferation and invasion↑[85]

Share and Cite

MDPI and ACS Style

Rawat, M.; Kadian, K.; Gupta, Y.; Kumar, A.; Chain, P.S.G.; Kovbasnjuk, O.; Kumar, S.; Parasher, G. MicroRNA in Pancreatic Cancer: From Biology to Therapeutic Potential. Genes 2019, 10, 752. https://0-doi-org.brum.beds.ac.uk/10.3390/genes10100752

AMA Style

Rawat M, Kadian K, Gupta Y, Kumar A, Chain PSG, Kovbasnjuk O, Kumar S, Parasher G. MicroRNA in Pancreatic Cancer: From Biology to Therapeutic Potential. Genes. 2019; 10(10):752. https://0-doi-org.brum.beds.ac.uk/10.3390/genes10100752

Chicago/Turabian Style

Rawat, Manmeet, Kavita Kadian, Yash Gupta, Anand Kumar, Patrick S.G. Chain, Olga Kovbasnjuk, Suneel Kumar, and Gulshan Parasher. 2019. "MicroRNA in Pancreatic Cancer: From Biology to Therapeutic Potential" Genes 10, no. 10: 752. https://0-doi-org.brum.beds.ac.uk/10.3390/genes10100752

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