Next Article in Journal
Identification of Ureidocoumarin-Based Selective Discoidin Domain Receptor 1 (DDR1) Inhibitors via Drug Repurposing Approach, Biological Evaluation, and In Silico Studies
Previous Article in Journal
Trichostatin C Synergistically Interacts with DNMT Inhibitor to Induce Antineoplastic Effect via Inhibition of Axl in Bladder and Lung Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tumor Suppressor MicroRNAs in Clinical and Preclinical Trials for Neurological Disorders

by
Austin Lui
1,
Timothy Do
1,
Omar Alzayat
1,
Nina Yu
1,
Su Phyu
2,
Hillary Joy Santuya
1,
Benjamin Liang
1,
Vidur Kailash
1,
Dewey Liu
1,
Sabra S. Inslicht
3,4,
Kiarash Shahlaie
5 and
DaZhi Liu
1,6,*
1
Department of Neurology, University of California at Davis, Davis, CA 95616, USA
2
Department of Neurological Surgery, University of California at San Francisco, San Francisco, CA 94143, USA
3
Department of Psychiatry and Behavioral Sciences, University of California at San Francisco, San Francisco, CA 94143, USA
4
San Francisco VA Health Care System, San Francisco, CA 94121, USA
5
Department of Neurological Surgery, University of California at Davis, Davis, CA 95616, USA
6
Mirnova Therapeutics Inc., Davis, CA 95618, USA
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2024, 17(4), 426; https://doi.org/10.3390/ph17040426
Submission received: 17 February 2024 / Revised: 18 March 2024 / Accepted: 20 March 2024 / Published: 27 March 2024
(This article belongs to the Special Issue RNA Biomarkers and Drugs)

Abstract

:
Cancers and neurological disorders are two major types of diseases in humans. We developed the concept called the “Aberrant Cell Cycle Disease (ACCD)” due to the accumulating evidence that shows that two different diseases share the common mechanism of aberrant cell cycle re-entry. The aberrant cell cycle re-entry is manifested as kinase/oncoprotein activation and tumor suppressor (TS) inactivation, which are associated with both tumor growth in cancers and neuronal death in neurological disorders. Therefore, some cancer therapies (e.g., kinase/oncogene inhibition and TS elevation) can be leveraged for neurological treatments. MicroRNA (miR/miRNA) provides a new style of drug-target binding. For example, a single tumor suppressor miRNA (TS-miR/miRNA) can bind to and decrease tens of target kinases/oncogenes, producing much more robust efficacy to block cell cycle re-entry than inhibiting a single kinase/oncogene. In this review, we summarize the miRNAs that are altered in both cancers and neurological disorders, with an emphasis on miRNA drugs that have entered into clinical trials for neurological treatment.

1. Introduction

We developed a novel concept of “Aberrant Cell Cycle Disease” (ACCD), revealing that two major types of diseases (cancers and neurological diseases) share the same mechanism of “aberrant cell cycle re-entry”. The aberrant cell cycle is manifested as oncogenic kinase activation and/or tumor suppressor inhibition, which are hallmarks of both tumor growth in cancers and neuronal death in neurological disorders [1]. Based on the ACCD concept, the key cell cycle players can be expanded from cyclin and cyclin-dependent kinases (CDKs) to Src family kinase (SFK), Jun N-terminal kinase (JNK), extracellular signal-regulated kinase (ERK), and other many oncogenic kinases (Figure 1) [1].
It is well known that genetic, environmental, and other factors cause cancers, which is featured as the aberrant cell cycle (unlimited cell proliferation) of cancer cells [1]. As for mature neurons, they are post-mitotic cells and unable to pass the mitotic phase to regain the stable, resting G0 phase of the cell cycle. Considering tat the cell cycle is an irreversible process, mature neurons ultimately die once they re-enter the cell cycle aberrantly under certain conditions [1]. Therefore, we concluded that the aberrant cell cycle is a common mechanism of both cancers and neurological disorders [1], as blocking the cell cycle can kill cancer cells for cancer therapy and rescue mature neurons for neurological therapy.
There are in general two strategies to block the cell cycle, namely the inhibition of oncogenic factors (e.g., kinases) and the elevation of tumor suppressors. Focusing on the oncogenic kinases, we recently reviewed all 74 FDA-approved kinase inhibitors (mostly for cancer therapy) by the end of 2022 in preclinical and clinical trials for neurological treatment, with nearly 500 references supporting the possibility of repurposing kinase inhibitors for the treatment of neurological disorders, such as Alzheimer’s disease (AD), intracerebral hemorrhage (ICH), ischemic stroke (IS), traumatic brain injury (TBI), and other disorders [2].
However, most current drug-development efforts (including kinase inhibitors) are based on the “receptor theory”—the basic framework for drug development for more than a century—which focuses on small molecules, with each molecule targeting a single gene, protein, or enzyme. The drugs developed by the traditional “one-drug-on-one-target” binding style may not be effective for the treatment of neurological disorders in humans, because (1) neurological disorders are heterogeneous in general with different causes, severities, and genetic backgrounds of patients and (2) increasing evidence shows that a large number of oncogenic kinases are significantly increased in neurological disorders [2]. Indeed, the selected testing of hundreds of traditional compounds that are effective in animal models has led to little success in clinical trials for the treatment of neurological disorders in humans [3]. The combinatorial therapy “multi-drug-on-multi-target” may be effective, but determining the optimum combination and dynamic interaction has been a challenge.
MicroRNAs (miR/miRNA), a type of small non-coding RNA, are regarded as next-generation promising drug targets for various diseases, including cancers and neurological disorders [4,5,6,7,8]. This is mostly because miRNAs provide a new “one-drug-on-numerous-target” binding style—a single miRNA binds to the 3′ untranslated region (3′UTR) of hundreds of target genes [9,10], leading to post-translational suppression of those target genes [11,12,13,14,15]. For a tumor suppressor miRNA (TS-miR/TS-miRNA), it is capable of inhibiting tens of oncogenic kinases among hundreds of target genes (Figure 1). Therefore, elevating a TS-miR has the potential to inhibit numerous oncogenic kinases and may lead to much more robust efficacy compared to targeting a single kinase by a kinase inhibitor.
In this review, we summarize miRNAs that are altered in cancers and neurological disorders and highlight the TS-miRs that have been tested in experimental models and/or clinical trials for the treatment of neurological disorders to discuss the feasibility and applicability of leveraging TS-miRs for neurological treatment.

2. MicroRNA History and Biogenesis

MicroRNAs (miR/miRNA) are small non-coding RNAs, about ~22 nucleotides in length, that are capable of modulating gene expression. In 1993, the Ambros and Ruvkun groups reported that lin-4 gene coded a small 22-nucleotide RNA instead of a conventional protein-coding gene in their studies of the lin-4 gene in Caenorhabditis elegans [16,17,18,19,20,21]. Since lin-4 had a complementary sequence to the 3′ untranslated region (3’UTR) of another gene lin-14, and lin-14 was post-transcriptionally downregulated through binding the 3′ UTR of Lin-14 [16], they postulated that lin-4 could regulate lin-14 by a post-transcriptional mechanism [17]. This discovery revealed the existence of miRNAs while having limited significance outside of the C. elegans research community, until the discovery of the second miRNA, let-7, in 2000 [22]. Unlike lin-4, let-7 is conserved across many organisms, including humans, suggesting that these miRNAs should have a more general role in biology and the regulation of gene expression [23]. In addition to the discovery of the post-transcriptional miRNA-induced gene modulation, these findings unfolded a more complete picture of the components of the RNA-induced silencing complex (RISC) by which miRNA regulates gene expression, kicking off the new field of miRNA research [24,25,26,27,28,29,30].
As shown in Figure 2, the miRNA biogenesis starts with transcribing a gene into a primary miRNA (pri-miRNA), which is 5′-capped and 3′-polyadenylated, with the transcription being performed by RNA Polymerase II. The pri-miRNAs are then cleaved by a microprocessor complex, consisting of RNA-binding protein DGCR8 and type III RNase Drosha, into a ~85-nucleotide stem–loop structure known as precursor miRNA (pre-miRNA) [31]. Pre-miRNA is exported and transported by Exportin 5 (XPO5) from the nucleus to the cytoplasm, where the pre-miRNA is processed by another RNase III enzyme Dicer into a mature miRNA. This creates the miRNA-induced silencing complex (miRISC), which consists of glycine-tryptophan protein (GW182) and Argonaute 2 protein, forming a complex of miRNA and mRNA interaction. MiRNAs usually interact with the 3′ UTR of their target genes post-transcriptionally to suppress expression under certain conditions, one of the conditions being RISC, but interaction with other regions of the gene, including the 5′ UTR, coding sequence, and even gene promoters has also been reported [32,33]. In addition, miRNAs are able to activate gene expression or control the rate of translation/transcription of genes [34,35]. miRNAs have even been shown to be in extracellular fluids where they may potentially serve as biomarkers or mediate cell-to-cell communication [36,37,38]. The aberrant expression of miRNAs is associated with many human diseases, showing that miRNAs are critical for normal human development and a variety of biological processes, making them a promising target for medical therapeutics [24,39,40].

3. MicroRNA Biomarkers in Diseases

In this section, we summarize miRNA fingerprints in different types of samples for each sub-type of disease. As shown in the Supplementary Table S1, there were (1) a total of 1132 miRNAs that were found to be dysregulated (upregulated or downregulated) across 15 types of cancers (sarcomas, gastrointestinal cancers, nervous system cancers, skin cancers, hepatobiliary cancers, bone cancers, urological cancers, breast cancers, lung cancers, female genital neoplasms, male genital neoplasms, heart cancers, head and neck cancers, eye cancers, adpancreatic cancers) in different types of samples, including plasma, serum, blood, exosomes, tissues, and cell lines, and (2) a total of 779 miRNAs were found to be dysregulated (upregulated or downregulated) across 7 types of neurological disorders (AD, PD, ALS, HD, Stroke, TBI, and seizures) in different sample types, including CSF, serum, plasma, blood, central nervous system tissue, exosomes, cell lines, and immune cells (Supplementary Materials).

3.1. MicroRNA Dysregulation and Biomarkers in Cancers

The discovery of emerging biomarkers is critical for cancers that require early identification and therapy. miRNAs can be detected in tissue and extracellular samples not limited to feces and blood [41] and are reported to be relatively stable in a variety of biological fluids as well as formalin-fixed paraffin-embedded tissues [42,43]. Extracellular miRNAs have been widely reported as potential biomarkers for various malignancies [44], with dysregulated circulating miRNAs being reported particularly in the early stages of various cancers [45].
Abnormal miRNA expression (i.e., upregulation or downregulation) in malignant cells is often due to changes in genomic miRNA copy numbers and gene locations pertaining to amplification, deletion or translocation, impacting multiple pathological processes in a variety of cancers. The earliest discovery of miRNA gene location change is the loss of miR-15a/16-1 cluster gene at chromosome 13q14, which is chronically seen in B-cell chronic lymphocytic leukemia patients [46]. In lung cancer, the 5q33 region containing miR-143 and miR-145 can often be deleted, leading to the downregulation of the respective miRNAs in question. Additionally, the amplification of the miR-17–92 cluster gene was observed in both lung cancer and B-cell lymphoma, and the translocation of the cluster gene was also observed in T-cell acute lymphoblastic leukemia, leading to the overexpression of the respective miRNAs [47].
It was reported that miR-29c was significantly downregulated in the plasma of hepatocellular carcinoma (HCC) patients [48], and abnormal miRNA expression is implicated in the carcinogenesis pathway, including the decrease in the apoptosis of malignant cells and increased proliferation [49,50]. It was also reported that the downregulation of miR-29a-3p is implicated in tumorigenesis and the development of colorectal cancer (CRC) and other cancers by increasing cell proliferation, migration, and invasion [51,52]. In prostate cancer (PCa), mouse models have revealed that the loss of miR-15 and miR-16 and the simultaneous overexpression of miR-21 promote cancer progression [53]. By decreasing levels of Bcl2, Mcl1, CCND1, and WNT3A, miR-15a and miR-16 exert tumor suppressor functions [54]. The administration in vivo of antagonistic miRNA directed against miR-15a and miR-16 produces prostate hyperplasia and leads to the development of a tumor-prone phenotype [54]. Furthermore, tumor regression, growth arrest, and apoptosis induction have been observed in miR-15a- and miR-16-reconstituted LNCaP xenograft tumors [55]. Therefore, the abnormal miRNA expression levels can be used as prognostic and diagnostic biomarkers in certain types of cancers [56,57].

3.2. MicroRNA Dysregulation and Biomarkers in Acute Neurological Disorders:

The dysregulation of miRNA is implicated in the pathology of acute brain injuries. Acute brain injuries encompass several neurological disorders, including acute ischemic stroke (AIS), intracerebral hemorrhage (ICH), traumatic brain injury (TBI), seizures, and others. We and others have reported that miRNAs are significantly changed in the blood and/or brain at different timepoints in both humans and experimental animals after different types of acute brain injuries [58,59,60]. Our studies have demonstrated that TS-miRNAs (including miR-122, miR-125b, miR-140, and others) are decreased in the blood in both humans and/or rats after AIS, ICH, and TBI [58,61]. Aside from the patent of miR-125b and miR-122 for TBI therapy [62], we demonstrated that miR-122 mimic has a 6 h therapeutic window to improve multiple outcomes after AIS in rats [61,63]. Another particular miRNA, miR-124, was shown to be downregulated in human AIS patients within 24 h of AIS and paradoxically upregulated approximately 48 to 72 h after AIS [64]. Interestingly, miR-124 was shown to have both beneficial and detrimental effects in the context of ischemia [65]. One possible interpretation was that miR-124 may trigger anti-inflammatory responses and attenuate apoptosis [66] but at the same time exert detrimental effects on synaptic plasticity and axonal growth [65]. Separate studies reported that miR-98 was downregulated [67], while miR-210, miR-191, miR-15a, and miR-16-1 were upregulated after AIS in rodent models [68,69,70].
In a recent study examining miRNA levels in veterans with chronic mild TBI, there were changes in levels of 32 miRNAs in plasma and 45 miRNAs in extracellular vesicles compared to control patients [71]. For example, in extracellular vesicles, miR-21-5p, miR-30d-5p, and miR-423-3p levels were changed compared to control patients, and plasma levels of miR-103b-1-5p, miR-106a-5p, and miR-132-5p were changed compared to control patients [71]. In human blood miRNA profiling studies, one study showed the upregulation of miR-16, miR-92a, and miR-765 in patients with severe TBI [8], and another study showed the downregulation of miR-425–5p and miR-502 in mild TBI patients at an early time period [72]. miR-425–5p levels were also shown to be a strong predictor of outcomes 6 months after mild TBI [72]. miR-212-5p was downregulated from 6 through 72 h after TBI in mice [73]. It was reported that miR-124-3p was upregulated at 3 and 14 h post-TBI but downregulated at 24 h post-TBI in mice [74], while miR-873a-5p was upregulated in human brain tissue after TBI [75]. Functional studies have revealed that miR-124-3p targeted Rela, an inhibitory transcription factor of ApoE that promotes the breakdown of amyloid-β (Aβ) plaques; ultimately alleviated neurodegeneration; inhibited inflammation; and promoted neurite outgrowth after TBI in rodents [74,76].
In various epilepsy models, miR-128, miR-124, miR-155, and miR-23b-3p levels were downregulated, while miR-134, miR-126a, miR-137, miR-129-5p, miR-203, miR-199a-5p, miR-22, miR-132, miR-324-5p, miR-34a, and miR-184 were upregulated [59,77,78,79]. In human studies, it was reported that miR-124 was downregulated in patients with epilepsy [80], while miR181a and miR-146a were upregulated [80,81]. Functional studies showed that miR-124 inhibited NMDA receptors, decreased expression of CREB1, and decreased anti-epileptogenic activity [80,82]. miR-134 regulated dendrite spine density and altered epileptogenesis [83,84], while miR181a inhibited caspase-3 and was associated with seizure-induced neuronal apoptosis [81]. miR-128 suppressed the expression of different types of ion channels and signaling components of the ERK2 signaling pathway [78,85], and miR-155 modulated brain-derived neurotrophic factor (BDNF) levels [86].

3.3. MicroRNA Dysregulation and Biomarkers in Neurodegenerative Disorders

Neurodegenerative diseases occur secondarily to the loss of neurons in the central nervous system, and the incidence typically increases with age [87,88]. As life expectancy continues to increase, more and more elderly people suffer from neurodegenerative diseases [89], such as Alzheimer’s Disease (AD), Parkinson’s Disease (PD), Amyotrophic Lateral Sclerosis (ALS), Huntington’s Disease (HD), and others. Currently, there are limited treatment options for neurodegenerative disorders [90,91], despite some supportive therapies.
AD is the most common form of dementia in people over the age of 65, and the disease is characterized by progressive neuronal loss and inflammation, further impacting cognitive function. The AD pathology is due to the deposition of Aβ in the brain, which is a potent mitogenic factor resulting in abnormal cell cycle re-entry and death of neurons [92,93]. Since Aβ is derived from the cleavage of amyloid precursor protein (APP) by beta-site APP-cleaving enzyme 1 (BACE1) and the β-secretase complex [94], targeting APP has the potential to delay AD progress. A series of in vitro studies revealed that several miRNAs are capable of modulating APP expression, such as miR-106a and miR-520c [95]; miR-16 and miR-101 [96]; and miR-147, miR-655, miR-323-3p, miR-644, and miR-153 [97]. Moreover, multiple miRNAs can indirectly inhibit APP by downregulating BACE1 [98]. Two separate studies demonstrated that miR-124 was decreased in the brains of AD patients [99], and elevating miR-124 was able to downregulate BACE1 expression in experimental AD models [100]. The inverse relationship was also observed between BACE1 levels and several miRNAs (e.g., miR-29c, miR-298, miR-328, miR-195) in multiple experimental AD models [101,102]. Additionally, miRNA dysregulation has been found to play a role in neuroinflammation in AD. Specifically, miR-146a has been observed in AD studies to be a potential target to reduce astrocytic inflammation and stymie cognitive impairment associated with AD [103,104,105]. The proposed mechanisms by which miR-146a exerts these effects include the downregulation of tumor necrosis factor (TNF) receptor-associated factor 6 and NF-kB and the switching of microglial phenotypes to secrete lower levels of pro-inflammatory factors and increase phagocytic function [103,106]. Since these inflammatory cytokines are mitotic and associated with cell cycle re-entry, the anti-inflammatory approach has the potential to rescue neurons by blocking cell cycle re-entry. In preclinical investigations, miR-132 has shown promise as a potential therapeutic target for alleviating AD symptoms and progression [107,108]. Walgrave et al. recently reported miR-132 replacement in adult mice with AD-restored adult hippocampal neurogenesis and memory deficits [108], corroborating other studies that had similar findings [109,110]. Inhibitors for miR-181a, a TS-miRNA that is downregulated in cancers [111,112,113], were found to decrease Tau protein in the hippocampus of mice to rescue memory deficits from AD [114]. Inhibitors of miR-124, a TS-miRNA also found to be downregulated in cancers [115,116,117], also reduced Tau hyperphosphorylation and rescued learning and memory [118].
PD is the second most common neurodegenerative disorder, and its symptoms are often attributed to a loss of dopaminergic neurons of the substantia nigra. The two genes typically involved in PD include a-synuclein (SNCA) and leucine-rich repeat kinase2 (LRRK). Since higher expression of SNCA and the expression of the three mutant forms of SNCA contribute to the insoluble aggregates that comprise the main structure of Lewy Bodies [119], the downregulation of SNCA is a possible mechanism in treating PD. miR-7 and miR-153 have been shown to inhibit the expression of SNCA [120]. The dysregulation of LRRK2, which is highly expressed in the brain, including the hippocampus and striatum [121], is associated with the pathogenesis of PD. An experiment demonstrated that the transfection of miR-205 in neurons expressing a PD-related LRKK2 R1441G mutant had a role in preventing neurite outgrowth [122]. Other miRNA types that have been identified as altered in PD also include miR-29-c-3p, miR-124, mir-150, miR-626, and miR-4639-5p [123,124]. In PD, there is a progressive loss of dopamine (DA) and attenuated DA signaling, α-synuclein-filled Lewy bodies, and increased neuroinflammation [125]. While Levodopa, DA receptor agonists, and several inhibitors are mainstay medications for PD treatment, interesting miRNA studies have been conducted to investigate whether miRNA therapy can serve as curative measures. Mimics of miR-150 and miR-29c-3p, both of which are TS-miRNAs in various cancers [112,126,127], have been observed to reduce inflammatory cytokines and microglial inflammasome activation, respectively [128,129]. Hu et al. found that the administration of mimics of miR-425, a TS-miRNA that is downregulated in cancer [130], intracerebrally improved locomotive behaviors and reduced DA neurodegeneration and necroptosis [131].
The dysregulation of miRNA is shown to be associated with ALS, also known as Lou Gehrig’s Disease, which is a progressive and fatal disease that affects the neurons in the brain and spinal cord. The motor neuronal damage leads to weakness, paralysis, and eventually death within 5 years of symptom onset. Some studies have found a positive correlation between the expression of miR-338-3p and miR-143-3p in CSF and serum from ALS patients [132]. As illustrated in many studies, miR-338, miR-142, miR-183, and let-7d play a role in neurodegeneration and apoptosis, while miR-206, miR-133a, miR-133b, and miR-27a are specifically expressed in striated muscle proliferation and at the muscle level [132]. The upregulation of miR-206, miR-133a, miR-133b, and miR-27a and inflammatory miRNAs (miR-155, miR-146a, and miR-221) were discovered in ALS patients with an earlier age of onset and longer disease duration [132]. In the peripheral tissues of ALS patients, eight miRNAs (miR-338-3p, miR-451, miR-1275, miR-328, miR-638, miR-149, miR-665, and miR-583) were deregulated in ALS patients. Among these, miR-338-3p was found in the brain tissues of ALS patients [133].
The dysregulation of miRNA is also shown to be associated with HD, an intractable neurodegenerative disease caused by CAG repeat expansion in the huntingtin gene located on chromosome 4p16.3. Patients present with cognitive defects and motor control impairment due to the progressive loss of cortical and striatal neurons [134]. The huntingtin protein was demonstrated to bind with repressor element 1 silencing transcription factor (REST) in neurons. In control individuals, the huntingtin protein sequesters REST in the cytoplasm and prevents the repressor from binding to DNA. In contrast, HD patients have mutant huntingtin protein that does not inhibit REST, which then permits REST to relocate to the nucleus of HD neurons and downregulate many of its target genes. One of them is BDNF, which is critical for the survivability of neurons [135]. Johnson et al. were able to identify a set of neuron-specific REST-target miRNAs in the human genome: miR-9-1, 9-3, 29a, 29b-1, 124a-1, 124a-2, 124a-3, 132, 135b, 139, 203, 204, 212, 330, and 346. Among these, miR-29a, miR-124a, miR-132, and miR-330 were found to be decreased in animal models for HD; however, only the downregulation of miR-132 was confirmed in human samples [136]. The expression profile of miRNAs in the frontal cortex and striatum of HD patients was also analyzed using qRT-PCR, RNA sequencing, and microarray. In all three of these different techniques, miR-100, miR-151-3p, miR-16, miR-219-2-3p, miR27b, miR-451, and miR-92a were found to be overexpressed in HD brain tissue [137]. In contrast, miR-128, miR-139-3p, miR-222, miR-382, miR-433, and miR-483-3p levels were decreased in the brain [137]. Other studies also found that miR-22, miR-29c, miR-132, miR-138, miR-218, miR-344, and miR-674 are downregulated in rodent models with HD [138,139]. Studies targeting specific miRNAs in HD models, such as miR-132, a TS-miRNA decreased in various types of malignancies [140,141], have shown HD symptom relief and stunting of HD progression [139].

4. MiRNAs Consistently Upregulated or Downregulated in Both Cancers and Neurological Disorders

The concept of ACCD suggests that cancers and neurological disorders share a common mechanism of dysregulated cell cycles, leading to tumorigenesis and progression in cancers and neuronal cell death in neurological disorders. With kinase/oncoproteins and TS being an integral part of the cell cycle regulation, we hypothesize that there will be a common set of oncogenic miRNAs that are upregulated and a common set of tumor suppressive miRNAs that are downregulated in both cancers and neurological disorders. We then identify the overlap in upregulated miRNAs and downregulated miRNAs in cancers and neurological disorders. A total of 273 miRNAs were found to be consistently upregulated in cancers only, while a total of 239 miRNAs were found to be consistently upregulated in neurological disorders only. The following 31 miRNAs were consistently upregulated in both cancers and neurological disorders across all disease types, studies, and sample types: miR-18a-5p, miR-18b-5p, miR-19a-3p, miR-20a-5p, miR-22-5p, miR-105-5p, miR-126-5p, miR-181a-5p, miR-223-5p, miR-323b-3p, miR-372-3p, miR-373, miR-455-3p, miR-488, miR-549, miR-582-5p, miR-585, miR-595, miR-627-5p, miR-671-5p, miR-762, miR-877-5p, miR-941, miR-1225-5p, miR-1249, miR-3613-5p, miR-4317, miR-4435, miR-iR-4293, let-7e-5p, and mIR-17-92 cluster (Table 1a). By contrast, a total of 269 miRNAs were found to be consistently downregulated in cancers only, while a total of 182 miRNAs were found to be consistently downregulated in neurological disorders only. The following 17 miRNAs were consistently downregulated in both cancers and neurological disorders across all disease types, studies, and sample types: miR-28-5p, miR-148-3p, miR-199b-3p, miR-208b, miR-219-5p, miR-320b, miR-323a-3p, miR-330, miR-374b, miR-449b-5p, miR-493-5p, miR-509-3p, miR-511, miR-575, miR-1297, miR-4487, and let-7i-5p (Table 1b). Since the downregulated miRs typically function as tumor suppressors, elevating those TS-miRNAs has the potential to treat both cancers and neurological disorders.
In general, upregulated miRNAs in cancers may be responsible for the promotion of oncomiRs, which results in unlimited tumor cell division in cancers [142], whereas downregulated miRNAs in cancers often serve as TS, and loss of TS-miRNAs play critical roles in uncontrolled tumor growth [143]. It was reported that miR-15/16 triggered apoptosis via inhibiting oncogenes, such as cyclin D1, CDC2, JUN, ETS1, MCL1, and others [142]. MiR-15/16 also repressed the expression of ROR1, a receptor for Wnt5a that promotes cell proliferation in leukemia and anoncoembryonic protein expressed in chronic lymphocytic leukemia (CLL) cells, as miR-15/16 expression is inversely related to ROR1 levels [142]. It was also reported that miR-140 targeted BCL9 and BCL2 and prevented colorectal cancer progression, miR-148a targeted BCL2 in non-small cell lung cancer, and miR-340 decreased BCL2 and increased Bax to induce cell apoptosis in ovarian cancer [144].
TS-miRNAs can block cell proliferation by blocking the WNT/B-catenin pathway in cancers. In particular, miR-200 targeted B-catenin, and miR-19 inhibited myocyte enhancer factor 2D (MEF2D) to block the Wnt pathway, while miR-133a-5p targeted TCF (which enhanced the transcription of oncogenes), thereby stopping cell proliferation in gastric cancer [145]. TS-miRNAs can also inhibit the endothelial-to-mesenchymal transition (EMT), which is critical for cancer metastasis. The miR-200 family inhibits EMT by directing decreasing amounts of zinc-finger E-box-binding homeobox1 (ZEB1) and SIP [146]. ZEB1 itself is capable of providing negative feedback to regulate the expression of miR-200. Thus, the disruption of ZEB1 has been shown to prevent EMT. Another factor that is critical for EMT maintenance is TBF-B, which downregulates miR-200 via the reversible DNA methylation of miRNA-200 [147].
To our knowledge, no other studies have explicitly linked the function of miRNAs as TSs to neurological disorders, except for our patent that demonstrated that elevating TS-miR-125b improves outcomes after TBI in rats by downregulating multiple oncogenic kinases (e.g., Mknk2, SFK, others) [62]. Instead, most studies have linked functions related to TS-miRNAs, such as apoptosis, to neurological disorders. For example, one study reported that miR-411 attenuated inflammatory damage and apoptosis in spinal cord injury in animal models [148]. Other studies showed that miR-124 may trigger anti-inflammatory responses and attenuate apoptosis in AIS [66].

5. Clinical Trials of miR drugs for Neurological Disorders

Although a large number of experimental studies demonstrated the robust efficacy of miRNA drugs for the treatment of neurological disorders, only three miRNA drugs have been tested so far in clinical trials for neurological indications (Table 2). One of the latest ongoing U.S. clinical trials to prevent HD progression involves AMT-130, which is a gene therapy that utilizes adeno-associated viral vector serotype 5 (AAV5) to deliver artificial miRNA (hsa-pre-miR-451a scaffold) that targets HTT to ultimately lower HTT [149] (NCT04120493) and a similar clinical trial is ongoing in Europe (NCT05243017). Of note, miR-451a is a known TS-miRNA in multiple cancers [150,151,152]. Though the trial is not yet complete, this approach has demonstrated success in reducing HTT expression in other models [153,154]. There are a few other clinical trials examining the use of miR therapies in other neurological disorders, such as ALS and epilepsy. One phase 1/2 clinical trial starting in 2024 (NCT06100276) and projected to end in 2027 is examining the safety, tolerability, and efficacy of AMT-162, a gene therapy that utilizes adeno-associated viral vector serotype 5 (AAV9) to deliver miRNA, which silences the SOD1 gene, in patients with ALS with the SOD1 mutation. One phase 1/2b clinical trial that started in 2023 and is projected to end in 2027 (NCT06063850) is examining the safety, tolerability, and efficacy of AMT-260, a gene therapy utilizing AAV9 to deliver miRNA to silence the GRIK2 gene, in patients with unilateral refractory mesial temporal lobe epilepsy.
There are a few major challenges that prevent the clinical development of miRNA therapeutics. First, the development of effective and safe delivery methods of miRNAs presents a challenge [155]. It is crucial to design tissue-specific targeting, as this will reduce off-target effects and decrease toxicity while improving efficacy [155]. Since oligonucleotides tend to be hydrophilic and negatively charged, another challenge that arises includes crossing the cell membrane to reach its target binding sites [156]. Ways of improving the ability of miRNAs to cross the cell membrane include using lipid nanoparticles and polymer-based carriers and the modification of the miRNA [156]. Secondly, it may be difficult to identify the most ideal miRNAs to target [155]. This is due to the heterogeneity of miRNA expression since multiple conditions, such as hypoxia and inflammation, can influence miRNA expression [155]. However, new next-generation RNA sequencing may help mitigate this issue [155]. Moreover, several databases and computational pipelines are being used to design synthetic miRNAs. For example, Laganà et al. present miR-Synth, a computational resource for the design of synthetic miRNAs able to target multiple genes in multiple sites. The proposed strategy constitutes a valid alternative to the use of siRNA, allowing the employment of a single molecule for the inhibition of multiple to numerous targets. This may represent a great advantage in designing therapies for diseases caused by crucial cellular pathways altered by numerous dysregulated genes [157]. Understanding these mechanisms will allow us to develop targeted solutions at the transcriptional level and make breakthroughs in therapeutics.

6. Conclusions

In summary, miRNA therapeutics is at the infancy stage in drug development, with some TS-miRNAs, which were originally identified in cancers, being investigated in clinical and preclinical trials for neurological treatments. As an emerging therapeutic approach, TS-miRNAs have the potential to target multiple oncogenic kinases at a time, which makes them a promising candidate to rescue neurons for the treatment of neurological diseases that typically involve multiple harmful pathways of brain injury. The understanding of miRNA characterization and fundamental scientific knowledge have grown explosively over the past few years. As such, the potential for miRNAs in the application of personalized medicine is being recognized due to their involvement in the regulation of genes, ease of access as a biomarker, and ability to influence multiple targets as therapeutics. With the advent of new techniques, being able to overcome the barriers of delivery, specific targeting, and identifying ideal miRNA candidates will immensely help promising preclinical miRNA drugs (including TS-miRNAs for neurological treatment) reach clinical trials and eventually FDA approval.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph17040426/s1, Table S1: MicroRNA fingerprints in different types of samples for each sub-type of cancers and neurological disorders.

Author Contributions

A.L., T.D., O.A., N.Y., S.P., H.J.S., B.L., V.K. and D.L. (Dewey Liu) read the literature, drafted the manuscript, and made the tables and figures. S.S.I. and K.S. added clinical potential and revised the manuscript. A.L., S.S.I. and D.L. (DaZhi Liu) edited the manuscript. D.L. (DaZhi Liu) developed the concept. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge the support of NIH/NINDS grants (R01NS089901 and R01NS114061 to DaZhi Liu).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

DaZhi Liu is a Professor at UC Davis and the Founder of Mirnova Therapeutics, Inc. MirnovaTX is in the process of licensing the technology from UC Davis. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Liu, D.Z.; Ander, B.P. Cell cycle inhibition without disruption of neurogenesis is a strategy for treatment of aberrant cell cycle diseases: An update. Sci. World J. 2012, 2012, 491737. [Google Scholar] [CrossRef] [PubMed]
  2. Lui, A.; Vanleuven, J.; Perekopskiy, D.; Liu, D.; Xu, D.; Alzayat, O.; Elgokhy, T.; Do, T.; Gann, M.; Martin, R.; et al. FDA-Approved Kinase Inhibitors in Preclinical and Clinical Trials for Neurological Disorders. Pharmaceuticals 2022, 15, 1546. [Google Scholar] [CrossRef] [PubMed]
  3. Lui, A.; Alzayat, O.; Do, T.; Perekopskiy, D.; Gann, M.; Elgokhy, T.S.; Gao, J.; Liu, D. Multi-targeted anti-inflammatory drugs for the treatment of neurological disorders. Neural Regen. Res. 2023, 18, 805–806. [Google Scholar] [CrossRef] [PubMed]
  4. Srinivasan, S.; Selvan, S.T.; Archunan, G.; Gulyas, B.; Padmanabhan, P. MicroRNAs -the next generation therapeutic targets in human diseases. Theranostics 2013, 3, 930–942. [Google Scholar] [CrossRef] [PubMed]
  5. Schmidt, M.F. Drug target miRNAs: Chances and challenges. Trends Biotechnol. 2014, 32, 578–585. [Google Scholar] [CrossRef] [PubMed]
  6. Chakraborty, C.; Sharma, A.R.; Sharma, G.; Doss, C.G.P.; Lee, S.S. Therapeutic miRNA and siRNA: Moving from Bench to Clinic as Next Generation Medicine. Mol. Ther. Nucleic Acids 2017, 8, 132–143. [Google Scholar] [CrossRef] [PubMed]
  7. Martin, C.G.; Kim, H.; Yun, S.; Livingston, W.; Fetta, J.; Mysliwiec, V.; Baxter, T.; Gill, J.M. Circulating miRNA associated with posttraumatic stress disorder in a cohort of military combat veterans. Psychiatry Res. 2017, 251, 261–265. [Google Scholar] [CrossRef] [PubMed]
  8. Redell, J.B.; Moore, A.N.; Ward, N.H., 3rd; Hergenroeder, G.W.; Dash, P.K. Human traumatic brain injury alters plasma microRNA levels. J. Neurotrauma 2010, 27, 2147–2156. [Google Scholar] [CrossRef]
  9. Hannon, G.J. RNA interference. Nature 2002, 418, 244–251. [Google Scholar] [CrossRef]
  10. Reynolds, A.; Leake, D.; Boese, Q.; Scaringe, S.; Marshall, W.S.; Khvorova, A. Rational siRNA design for RNA interference. Nat. Biotechnol. 2004, 22, 326–330. [Google Scholar] [CrossRef]
  11. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [PubMed]
  12. Bartel, D.P. MicroRNAs: Target recognition and regulatory functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef]
  13. Jakymiw, A.; Pauley, K.M.; Li, S.; Ikeda, K.; Lian, S.; Eystathioy, T.; Satoh, M.; Fritzler, M.J.; Chan, E.K. The role of GW/P-bodies in RNA processing and silencing. J. Cell Sci. 2007, 120, 1317–1323. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, L.; Fan, J.; Belasco, J.G. MicroRNAs direct rapid deadenylation of mRNA. Proc. Natl. Acad. Sci. USA 2006, 103, 4034–4039. [Google Scholar] [CrossRef] [PubMed]
  15. Lim, L.P.; Lau, N.C.; Garrett-Engele, P.; Grimson, A.; Schelter, J.M.; Castle, J.; Bartel, D.P.; Linsley, P.S.; Johnson, J.M. Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 2005, 433, 769–773. [Google Scholar] [CrossRef] [PubMed]
  16. 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] [PubMed]
  17. Wightman, B.; Ha, I.; Ruvkun, G. Posttranscriptional regulation of the heterochronic gene lin-14 by lin-4 mediates temporal pattern formation in C. elegans. Cell 1993, 75, 855–862. [Google Scholar] [CrossRef] [PubMed]
  18. Horvitz, H.R.; Sulston, J.E. Isolation and genetic characterization of cell-lineage mutants of the nematode Caenorhabditis elegans. Genetics 1980, 96, 435–454. [Google Scholar] [CrossRef] [PubMed]
  19. Chalfie, M.; Horvitz, H.R.; Sulston, J.E. Mutations that lead to reiterations in the cell lineages of C. elegans. Cell 1981, 24, 59–69. [Google Scholar] [CrossRef]
  20. Lee, R.; Feinbaum, R.; Ambros, V. A short history of a short RNA. Cell 2004, 116, S89–S92. [Google Scholar] [CrossRef]
  21. Almeida, M.I.; Reis, R.M.; Calin, G.A. MicroRNA history: Discovery, recent applications, and next frontiers. Mutat. Res. 2011, 717, 1–8. [Google Scholar] [CrossRef] [PubMed]
  22. Reinhart, B.J.; Slack, F.J.; Basson, M.; Pasquinelli, A.E.; Bettinger, J.C.; Rougvie, A.E.; Horvitz, H.R.; Ruvkun, G. The 21-nucleotide let-7 RNA regulates developmental timing in Caenorhabditis elegans. Nature 2000, 403, 901–906. [Google Scholar] [CrossRef]
  23. Pasquinelli, A.E.; Reinhart, B.J.; Slack, F.; Martindale, M.Q.; Kuroda, M.I.; Maller, B.; Hayward, D.C.; Ball, E.E.; Degnan, B.; Müller, P.; et al. Conservation of the sequence and temporal expression of let-7 heterochronic regulatory RNA. Nature 2000, 408, 86–89. [Google Scholar] [CrossRef] [PubMed]
  24. 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]
  25. 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]
  26. Hamilton, A.J.; Baulcombe, D.C. A species of small antisense RNA in posttranscriptional gene silencing in plants. Science 1999, 286, 950–952. [Google Scholar] [CrossRef] [PubMed]
  27. Zamore, P.D.; Tuschl, T.; Sharp, P.A.; Bartel, D.P. RNAi: Double-stranded RNA directs the ATP-dependent cleavage of mRNA at 21 to 23 nucleotide intervals. Cell 2000, 101, 25–33. [Google Scholar] [CrossRef] [PubMed]
  28. 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]
  29. Hutvágner, G.; McLachlan, J.; Pasquinelli, A.E.; Bálint, E.; Tuschl, T.; Zamore, P.D. A cellular function for the RNA-interference enzyme Dicer in the maturation of the let-7 small temporal RNA. Science 2001, 293, 834–838. [Google Scholar] [CrossRef]
  30. Grishok, A.; Pasquinelli, A.E.; Conte, D.; Li, N.; Parrish, S.; Ha, I.; Baillie, D.L.; Fire, A.; Ruvkun, G.; Mello, C.C. Genes and mechanisms related to RNA interference regulate expression of the small temporal RNAs that control C. elegans developmental timing. Cell 2001, 106, 23–34. [Google Scholar] [CrossRef]
  31. Pong, S.K.; Gullerova, M. Noncanonical functions of microRNA pathway enzymes—Drosha, DGCR8, Dicer and Ago proteins. FEBS Lett. 2018, 592, 2973–2986. [Google Scholar] [CrossRef]
  32. Ha, M.; Kim, V.N. Regulation of microRNA biogenesis. Nat. Rev. Mol. Cell Biol. 2014, 15, 509–524. [Google Scholar] [CrossRef]
  33. Broughton, J.P.; Lovci, M.T.; Huang, J.L.; Yeo, G.W.; Pasquinelli, A.E. Pairing beyond the Seed Supports MicroRNA Targeting Specificity. Mol. Cell 2016, 64, 320–333. [Google Scholar] [CrossRef] [PubMed]
  34. Vasudevan, S. Posttranscriptional upregulation by microRNAs. Wiley Interdiscip. Rev. RNA 2012, 3, 311–330. [Google Scholar] [CrossRef]
  35. Makarova, J.A.; Shkurnikov, M.U.; Wicklein, D.; Lange, T.; Samatov, T.R.; Turchinovich, A.A.; Tonevitsky, A.G. Intracellular and extracellular microRNA: An update on localization and biological role. Prog. Histochem. Cytochem. 2016, 51, 33–49. [Google Scholar] [CrossRef]
  36. Hayes, J.; Peruzzi, P.P.; Lawler, S. MicroRNAs in cancer: Biomarkers, functions and therapy. Trends Mol. Med. 2014, 20, 460–469. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, J.; Chen, J.; Sen, S. MicroRNA as Biomarkers and Diagnostics. J. Cell. Physiol. 2016, 231, 25–30. [Google Scholar] [CrossRef]
  38. Huang, W. MicroRNAs: Biomarkers, Diagnostics, and Therapeutics. Methods Mol. Biol. 2017, 1617, 57–67. [Google Scholar] [CrossRef] [PubMed]
  39. Tüfekci, K.U.; Oner, M.G.; Meuwissen, R.L.; Genç, S. The role of microRNAs in human diseases. Methods Mol. Biol. 2014, 1107, 33–50. [Google Scholar] [CrossRef]
  40. Paul, P.; Chakraborty, A.; Sarkar, D.; Langthasa, M.; Rahman, M.; Bari, M.; Singha, R.S.; Malakar, A.K.; Chakraborty, S. Interplay between miRNAs and human diseases. J. Cell. Physiol. 2018, 233, 2007–2018. [Google Scholar] [CrossRef]
  41. Mori, M.A.; Ludwig, R.G.; Garcia-Martin, R.; Brandão, B.B.; Kahn, C.R. Extracellular miRNAs: From Biomarkers to Mediators of Physiology and Disease. Cell Metab. 2019, 30, 656–673. [Google Scholar] [CrossRef] [PubMed]
  42. Kakimoto, Y.; Tanaka, M.; Kamiguchi, H.; Ochiai, E.; Osawa, M. MicroRNA Stability in FFPE Tissue Samples: Dependence on GC Content. PLoS ONE 2016, 11, e0163125. [Google Scholar] [CrossRef]
  43. Glinge, C.; Clauss, S.; Boddum, K.; Jabbari, R.; Jabbari, J.; Risgaard, B.; Tomsits, P.; Hildebrand, B.; Kääb, S.; Wakili, R.; et al. Stability of Circulating Blood-Based MicroRNAs—Pre-Analytic Methodological Considerations. PLoS ONE 2017, 12, e0167969. [Google Scholar] [CrossRef] [PubMed]
  44. Condrat, C.E.; Thompson, D.C.; Barbu, M.G.; Bugnar, O.L.; Boboc, A.; Cretoiu, D.; Suciu, N.; Cretoiu, S.M.; Voinea, S.C. miRNAs as Biomarkers in Disease: Latest Findings Regarding Their Role in Diagnosis and Prognosis. Cells 2020, 9, 276. [Google Scholar] [CrossRef]
  45. Buhagiar, A.; Seria, E.; Borg, M.; Borg, J.; Ayers, D. Overview of microRNAs as liquid biopsy biomarkers for colorectal cancer sub-type profiling and chemoresistance. Cancer Drug Resist. 2021, 4, 934–945. [Google Scholar] [CrossRef]
  46. Hutter, K.; Rülicke, T.; Szabo, T.G.; Andersen, L.; Villunger, A.; Herzog, S. The miR-15a/16-1 and miR-15b/16-2 clusters regulate early B cell development by limiting IL-7 receptor expression. Front. Immunol. 2022, 13, 967914. [Google Scholar] [CrossRef]
  47. Peng, Y.; Croce, C.M. The role of MicroRNAs in human cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef]
  48. Ratnasari, N.; Lestari, P.; Renovaldi, D.; Raditya Ningsih, J.; Qoriansas, N.; Wardana, T.; Hakim, S.; Signa Aini Gumilas, N.; Indrarti, F.; Triwikatmani, C.; et al. Potential plasma biomarkers: miRNA-29c, miRNA-21, and miRNA-155 in clinical progression of Hepatocellular Carcinoma patients. PLoS ONE 2022, 17, e0263298. [Google Scholar] [CrossRef]
  49. Oura, K.; Morishita, A.; Masaki, T. Molecular and Functional Roles of MicroRNAs in the Progression of Hepatocellular Carcinoma-A Review. Int. J. Mol. Sci. 2020, 21, 8362. [Google Scholar] [CrossRef] [PubMed]
  50. Chu, R.; Mo, G.; Duan, Z.; Huang, M.; Chang, J.; Li, X.; Liu, P. miRNAs affect the development of hepatocellular carcinoma via dysregulation of their biogenesis and expression. Cell Commun. Signal. 2014, 12, 45. [Google Scholar] [CrossRef]
  51. Mo, W.Y.; Cao, S.Q. MiR-29a-3p: A potential biomarker and therapeutic target in colorectal cancer. Clin. Transl. Oncol. 2023, 25, 563–577. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, J.Y.; Zhang, Q.; Wang, D.D.; Yan, W.; Sha, H.H.; Zhao, J.H.; Yang, S.J.; Zhang, H.D.; Hou, J.C.; Xu, H.Z.; et al. MiR-29a: A potential therapeutic target and promising biomarker in tumors. Biosci. Rep. 2018, 38, BSR20171265. [Google Scholar] [CrossRef] [PubMed]
  53. Bonci, D.; De Maria, R. miR-15/miR-16 loss, miR-21 upregulation, or deregulation of their target genes predicts poor prognosis in prostate cancer patients. Mol. Cell Oncol. 2016, 3, e1109744. [Google Scholar] [CrossRef] [PubMed]
  54. Arrighetti, N.; Beretta, G.L. miRNAs as Therapeutic Tools and Biomarkers for Prostate Cancer. Pharmaceutics 2021, 13, 380. [Google Scholar] [CrossRef] [PubMed]
  55. Jin, W.; Chen, F.; Wang, K.; Song, Y.; Fei, X.; Wu, B. miR-15a/miR-16 cluster inhibits invasion of prostate cancer cells by suppressing TGF-β signaling pathway. Biomed. Pharmacother. 2018, 104, 637–644. [Google Scholar] [CrossRef] [PubMed]
  56. He, B.; Zhao, Z.; Cai, Q.; Zhang, Y.; Zhang, P.; Shi, S.; Xie, H.; Peng, X.; Yin, W.; Tao, Y.; et al. miRNA-based biomarkers, therapies, and resistance in Cancer. Int. J. Biol. Sci. 2020, 16, 2628–2647. [Google Scholar] [CrossRef] [PubMed]
  57. Parra-Medina, R.; Lopez-Kleine, L.; Ramirez-Clavijo, S.; Payan-Gomez, C. Identification of candidate miRNAs in early-onset and late-onset prostate cancer by network analysis. Sci. Rep. 2020, 10, 12345. [Google Scholar] [CrossRef] [PubMed]
  58. Jickling, G.C.; Ander, B.P.; Zhan, X.; Noblett, D.; Stamova, B.; Liu, D. microRNA expression in peripheral blood cells following acute ischemic stroke and their predicted gene targets. PLoS ONE 2014, 9, e99283. [Google Scholar] [CrossRef]
  59. Liu, D.Z.; Tian, Y.; Ander, B.P.; Xu, H.; Stamova, B.S.; Zhan, X.; Turner, R.J.; Jickling, G.; Sharp, F.R. Brain and blood microRNA expression profiling of ischemic stroke, intracerebral hemorrhage, and kainate seizures. J. Cereb. Blood Flow Metab. 2010, 30, 92–101. [Google Scholar] [CrossRef]
  60. Jeyaseelan, K.; Lim, K.Y.; Armugam, A. MicroRNA expression in the blood and brain of rats subjected to transient focal ischemia by middle cerebral artery occlusion. Stroke 2008, 39, 959–966. [Google Scholar] [CrossRef]
  61. Lv, B.; Cheng, X.; Sharp, F.R.; Ander, B.P.; Liu, D.Z. MicroRNA-122 Mimic Improves Stroke Outcomes and Indirectly Inhibits NOS2 After Middle Cerebral Artery Occlusion in Rats. Front. Neurosci. 2018, 12, 767. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, D.Z. Methods (of MicroRNA Therapeutics) for Treating Brain Injury. WO Patent 2019165267A1, 2019. Available online: https://patents.google.com/patent/WO2019165267A1/en (accessed on 29 August 2019).
  63. Liu, D.Z.; Jickling, G.C.; Ander, B.P.; Hull, H.; Zhan, X.; Cox, C.; Shroff, N.; Dykstra-Aiello, C.; Stamova, B.; Sharp, F.R. Elevating microRNA-122 in blood improves outcomes after temporary middle cerebral artery occlusion in rats. J. Cereb. Blood Flow Metab. 2016, 36, 1374–1383. [Google Scholar] [CrossRef] [PubMed]
  64. Sun, M.; Hou, X.; Ren, G.; Zhang, Y.; Cheng, H. Dynamic changes in miR-124 levels in patients with acute cerebral infarction. Int. J. Neurosci. 2019, 129, 649–653. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, X.; Feng, Z.; Du, L.; Huang, Y.; Ge, J.; Deng, Y.; Mei, Z. The Potential Role of MicroRNA-124 in Cerebral Ischemia Injury. Int. J. Mol. Sci. 2019, 21, 120. [Google Scholar] [CrossRef] [PubMed]
  66. Song, Y.; Li, Z.; He, T.; Qu, M.; Jiang, L.; Li, W.; Shi, X.; Pan, J.; Zhang, L.; Wang, Y.; et al. M2 microglia-derived exosomes protect the mouse brain from ischemia-reperfusion injury via exosomal miR-124. Theranostics 2019, 9, 2910–2923. [Google Scholar] [CrossRef]
  67. Bernstein, D.L.; Zuluaga-Ramirez, V.; Gajghate, S.; Reichenbach, N.L.; Polyak, B.; Persidsky, Y.; Rom, S. miR-98 reduces endothelial dysfunction by protecting blood-brain barrier (BBB) and improves neurological outcomes in mouse ischemia/reperfusion stroke model. J. Cereb. Blood Flow Metab. 2020, 40, 1953–1965. [Google Scholar] [CrossRef] [PubMed]
  68. Huang, L.; Ma, Q.; Li, Y.; Li, B.; Zhang, L. Inhibition of microRNA-210 suppresses pro-inflammatory response and reduces acute brain injury of ischemic stroke in mice. Exp. Neurol. 2018, 300, 41–50. [Google Scholar] [CrossRef] [PubMed]
  69. Sun, P.; Zhang, K.; Hassan, S.H.; Zhang, X.; Tang, X.; Pu, H.; Stetler, R.A.; Chen, J.; Yin, K.J. Endothelium-Targeted Deletion of microRNA-15a/16-1 Promotes Poststroke Angiogenesis and Improves Long-Term Neurological Recovery. Circ. Res. 2020, 126, 1040–1057. [Google Scholar] [CrossRef]
  70. Du, K.; Zhao, C.; Wang, L.; Wang, Y.; Zhang, K.Z.; Shen, X.Y.; Sun, H.X.; Gao, W.; Lu, X. MiR-191 inhibit angiogenesis after acute ischemic stroke targeting VEZF1. Aging 2019, 11, 2762–2786. [Google Scholar] [CrossRef]
  71. Ghai, V.; Fallen, S.; Baxter, D.; Scherler, K.; Kim, T.K.; Zhou, Y.; Meabon, J.S.; Logsdon, A.F.; Banks, W.A.; Schindler, A.G.; et al. Alterations in Plasma microRNA and Protein Levels in War Veterans with Chronic Mild Traumatic Brain Injury. J. Neurotrauma 2020, 37, 1418–1430. [Google Scholar] [CrossRef]
  72. Di Pietro, V.; Ragusa, M.; Davies, D.; Su, Z.; Hazeldine, J.; Lazzarino, G.; Hill, L.J.; Crombie, N.; Foster, M.; Purrello, M.; et al. MicroRNAs as Novel Biomarkers for the Diagnosis and Prognosis of Mild and Severe Traumatic Brain Injury. J. Neurotrauma 2017, 34, 1948–1956. [Google Scholar] [CrossRef] [PubMed]
  73. Xiao, X.; Jiang, Y.; Liang, W.; Wang, Y.; Cao, S.; Yan, H.; Gao, L.; Zhang, L. miR-212-5p attenuates ferroptotic neuronal death after traumatic brain injury by targeting Ptgs2. Mol. Brain 2019, 12, 78. [Google Scholar] [CrossRef] [PubMed]
  74. Ge, X.; Guo, M.; Hu, T.; Li, W.; Huang, S.; Yin, Z.; Li, Y.; Chen, F.; Zhu, L.; Kang, C.; et al. Increased Microglial Exosomal miR-124-3p Alleviates Neurodegeneration and Improves Cognitive Outcome after rmTBI. Mol. Ther. 2020, 28, 503–522. [Google Scholar] [CrossRef] [PubMed]
  75. Long, X.; Yao, X.; Jiang, Q.; Yang, Y.; He, X.; Tian, W.; Zhao, K.; Zhang, H. Astrocyte-derived exosomes enriched with miR-873a-5p inhibit neuroinflammation via microglia phenotype modulation after traumatic brain injury. J. Neuroinflammation 2020, 17, 89. [Google Scholar] [CrossRef] [PubMed]
  76. Huang, S.; Ge, X.; Yu, J.; Han, Z.; Yin, Z.; Li, Y.; Chen, F.; Wang, H.; Zhang, J.; Lei, P. Increased miR-124-3p in microglial exosomes following traumatic brain injury inhibits neuronal inflammation and contributes to neurite outgrowth via their transfer into neurons. FASEB J. 2018, 32, 512–528. [Google Scholar] [CrossRef] [PubMed]
  77. Brennan, G.P.; Henshall, D.C. microRNAs in the pathophysiology of epilepsy. Neurosci. Lett. 2018, 667, 47–52. [Google Scholar] [CrossRef] [PubMed]
  78. Tan, C.L.; Plotkin, J.L.; Venø, M.T.; von Schimmelmann, M.; Feinberg, P.; Mann, S.; Handler, A.; Kjems, J.; Surmeier, D.J.; O’Carroll, D.; et al. MicroRNA-128 governs neuronal excitability and motor behavior in mice. Science 2013, 342, 1254–1258. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, D.; Li, Z.; Zhang, Y.; Wang, G.; Wei, M.; Hu, Y.; Ma, S.; Jiang, Y.; Che, N.; Wang, X.; et al. Targeting of microRNA-199a-5p protects against pilocarpine-induced status epilepticus and seizure damage via SIRT1-p53 cascade. Epilepsia 2016, 57, 706–716. [Google Scholar] [CrossRef]
  80. Ma, Y. The Challenge of microRNA as a Biomarker of Epilepsy. Curr. Neuropharmacol. 2018, 16, 37–42. [Google Scholar] [CrossRef]
  81. Ren, L.; Zhu, R.; Li, X. Silencing miR-181a produces neuroprotection against hippocampus neuron cell apoptosis post-status epilepticus in a rat model and in children with temporal lobe epilepsy. Genet. Mol. Res. 2016, 15, 15017798. [Google Scholar] [CrossRef]
  82. Wang, W.; Wang, X.; Chen, L.; Zhang, Y.; Xu, Z.; Liu, J.; Jiang, G.; Li, J.; Zhang, X.; Wang, K.; et al. The microRNA miR-124 suppresses seizure activity and regulates CREB1 activity. Expert Rev. Mol. Med. 2016, 18, e4. [Google Scholar] [CrossRef] [PubMed]
  83. Jimenez-Mateos, E.M.; Engel, T.; Merino-Serrais, P.; Fernaud-Espinosa, I.; Rodriguez-Alvarez, N.; Reynolds, J.; Reschke, C.R.; Conroy, R.M.; McKiernan, R.C.; deFelipe, J.; et al. Antagomirs targeting microRNA-134 increase hippocampal pyramidal neuron spine volume in vivo and protect against pilocarpine-induced status epilepticus. Brain Struct. Funct. 2015, 220, 2387–2399. [Google Scholar] [CrossRef] [PubMed]
  84. Jimenez-Mateos, E.M.; Engel, T.; Merino-Serrais, P.; McKiernan, R.C.; Tanaka, K.; Mouri, G.; Sano, T.; O’Tuathaigh, C.; Waddington, J.L.; Prenter, S.; et al. Silencing microRNA-134 produces neuroprotective and prolonged seizure-suppressive effects. Nat. Med. 2012, 18, 1087–1094. [Google Scholar] [CrossRef] [PubMed]
  85. Yuan, Y.; Xiang, W.; Yanhui, L.; Ruofei, L.; Yunhe, M.; Jiewen, L.; Qing, M. Dysregulation of microRNA-128 expression in WHO grades 2 glioma is associated with glioma-associated epilepsy: Down-regulation of miR-128 induces glioma-associated seizure. Epilepsy Res. 2016, 127, 6–11. [Google Scholar] [CrossRef] [PubMed]
  86. Cai, Z.; Li, S.; Li, S.; Song, F.; Zhang, Z.; Qi, G.; Li, T.; Qiu, J.; Wan, J.; Sui, H.; et al. Antagonist Targeting microRNA-155 Protects against Lithium-Pilocarpine-Induced Status Epilepticus in C57BL/6 Mice by Activating Brain-Derived Neurotrophic Factor. Front. Pharmacol. 2016, 7, 129. [Google Scholar] [CrossRef]
  87. Wilson, D.M., 3rd; Cookson, M.R.; Van Den Bosch, L.; Zetterberg, H.; Holtzman, D.M.; Dewachter, I. Hallmarks of neurodegenerative diseases. Cell 2023, 186, 693–714. [Google Scholar] [CrossRef] [PubMed]
  88. Kritsilis, M.; Rizou, S.V.; Koutsoudaki, P.N.; Evangelou, K.; Gorgoulis, V.G.; Papadopoulos, D. Ageing, Cellular Senescence and Neurodegenerative Disease. Int. J. Mol. Sci. 2018, 19, 2937. [Google Scholar] [CrossRef]
  89. Johnson, I.P. Age-related neurodegenerative disease research needs aging models. Front. Aging Neurosci. 2015, 7, 168. [Google Scholar] [CrossRef]
  90. Tzeplaeff, L.; Wilfling, S.; Requardt, M.V.; Herdick, M. Current State and Future Directions in the Therapy of ALS. Cells 2023, 12, 1523. [Google Scholar] [CrossRef]
  91. Ferguson, M.W.; Kennedy, C.J.; Palpagama, T.H.; Waldvogel, H.J.; Faull, R.L.M.; Kwakowsky, A. Current and Possible Future Therapeutic Options for Huntington’s Disease. J. Cent. Nerv. Syst. Dis. 2022, 14, 11795735221092517. [Google Scholar] [CrossRef]
  92. Copani, A.; Condorelli, F.; Caruso, A.; Vancheri, C.; Sala, A.; Giuffrida Stella, A.M.; Canonico, P.L.; Nicoletti, F.; Sortino, M.A. Mitotic signaling by beta-amyloid causes neuronal death. FASEB J. 1999, 13, 2225–2234. [Google Scholar] [CrossRef] [PubMed]
  93. Williamson, R.; Scales, T.; Clark, B.R.; Gibb, G.; Reynolds, C.H.; Kellie, S.; Bird, I.N.; Varndell, I.M.; Sheppard, P.W.; Everall, I.; et al. Rapid tyrosine phosphorylation of neuronal proteins including tau and focal adhesion kinase in response to amyloid-beta peptide exposure: Involvement of Src family protein kinases. J. Neurosci. 2002, 22, 10–20. [Google Scholar] [CrossRef] [PubMed]
  94. Angelucci, F.; Cechova, K.; Valis, M.; Kuca, K.; Zhang, B.; Hort, J. MicroRNAs in Alzheimer’s Disease: Diagnostic Markers or Therapeutic Agents? Front. Pharmacol. 2019, 10, 665. [Google Scholar] [CrossRef] [PubMed]
  95. Hébert, S.S.; De Strooper, B. Alterations of the microRNA network cause neurodegenerative disease. Trends Neurosci. 2009, 32, 199–206. [Google Scholar] [CrossRef] [PubMed]
  96. Vilardo, E.; Barbato, C.; Ciotti, M.; Cogoni, C.; Ruberti, F. MicroRNA-101 regulates amyloid precursor protein expression in hippocampal neurons. J. Biol. Chem. 2010, 285, 18344–18351. [Google Scholar] [CrossRef] [PubMed]
  97. Delay, C.; Calon, F.; Mathews, P.; Hébert, S.S. Alzheimer-specific variants in the 3’UTR of Amyloid precursor protein affect microRNA function. Mol. Neurodegener. 2011, 6, 70. [Google Scholar] [CrossRef] [PubMed]
  98. Hebert, S.S.; Horre, K.; Nicolai, L.; Papadopoulou, A.S.; Mandemakers, W.; Silahtaroglu, A.N.; Kauppinen, S.; Delacourte, A.; De Strooper, B. Loss of microRNA cluster miR-29a/b-1 in sporadic Alzheimer’s disease correlates with increased BACE1/beta-secretase expression. Proc. Natl. Acad. Sci. USA 2008, 105, 6415–6420. [Google Scholar] [CrossRef]
  99. Smith, P.; Al Hashimi, A.; Girard, J.; Delay, C.; Hébert, S.S. In Vivo regulation of amyloid precursor protein neuronal splicing by microRNAs. J. Neurochem. 2011, 116, 240–247. [Google Scholar] [CrossRef] [PubMed]
  100. Fang, M.; Wang, J.; Zhang, X.; Geng, Y.; Hu, Z.; Rudd, J.A.; Ling, S.; Chen, W.; Han, S. The miR-124 regulates the expression of BACE1/β-secretase correlated with cell death in Alzheimer’s disease. Toxicol. Lett. 2012, 209, 94–105. [Google Scholar] [CrossRef]
  101. Zong, Y.; Wang, H.; Dong, W.; Quan, X.; Zhu, H.; Xu, Y.; Huang, L.; Ma, C.; Qin, C. miR-29c regulates BACE1 protein expression. Brain Res. 2011, 1395, 108–115. [Google Scholar] [CrossRef]
  102. Zhu, H.C.; Wang, L.M.; Wang, M.; Song, B.; Tan, S.; Teng, J.F.; Duan, D.X. MicroRNA-195 downregulates Alzheimer’s disease amyloid-beta production by targeting BACE1. Brain Res. Bull. 2012, 88, 596–601. [Google Scholar] [CrossRef]
  103. Nakano, M.; Kubota, K.; Hashizume, S.; Kobayashi, E.; Chikenji, T.S.; Saito, Y.; Fujimiya, M. An enriched environment prevents cognitive impairment in an Alzheimer’s disease model by enhancing the secretion of exosomal microRNA-146a from the choroid plexus. Brain Behav. Immun. Health 2020, 9, 100149. [Google Scholar] [CrossRef]
  104. Huang, P.; Xu, M.; He, Y.X. MicroRNA-146a regulates the expression of the Aβ1-42 protein in Alzheimer’s disease. Eur. Rev. Med. Pharmacol. Sci. 2021, 25, 2003–2012. [Google Scholar] [CrossRef]
  105. Ma, Y.; Ye, J.; Zhao, L.; Pan, D. MicroRNA-146a inhibition promotes total neurite outgrowth and suppresses cell apoptosis, inflammation, and STAT1/MYC pathway in PC12 and cortical neuron cellular Alzheimer’s disease models. Braz. J. Med. Biol. Res. 2021, 54, e9665. [Google Scholar] [CrossRef]
  106. Liang, C.; Zou, T.; Zhang, M.; Fan, W.; Zhang, T.; Jiang, Y.; Cai, Y.; Chen, F.; Chen, X.; Sun, Y.; et al. MicroRNA-146a switches microglial phenotypes to resist the pathological processes and cognitive degradation of Alzheimer’s disease. Theranostics 2021, 11, 4103–4121. [Google Scholar] [CrossRef]
  107. Fu, X.; Liu, J.; Xie, J.; Chen, G.; Zhang, H.; Meng, F.; Wu, M.; Li, Q.; Liu, Y.; Wang, W.; et al. Identification of potential therapeutic and diagnostic characteristics of Alzheimer disease by targeting the miR-132-3p/FOXO3a-PPM1F axis in APP/PS1 mice. Brain Res. 2022, 1790, 147983. [Google Scholar] [CrossRef]
  108. Walgrave, H.; Balusu, S.; Snoeck, S.; Vanden Eynden, E.; Craessaerts, K.; Thrupp, N.; Wolfs, L.; Horré, K.; Fourne, Y.; Ronisz, A.; et al. Restoring miR-132 expression rescues adult hippocampal neurogenesis and memory deficits in Alzheimer’s disease. Cell Stem Cell 2021, 28, 1805–1821.e8. [Google Scholar] [CrossRef]
  109. Qu, J.; Xiong, X.; Hujie, G.; Ren, J.; Yan, L.; Ma, L. MicroRNA-132-3p alleviates neuron apoptosis and impairments of learning and memory abilities in Alzheimer’s disease by downregulation of HNRNPU stabilized BACE1. Cell Cycle 2021, 20, 2309–2320. [Google Scholar] [CrossRef]
  110. Liang, Y.; Wang, L. Inflamma-MicroRNAs in Alzheimer’s Disease: From Disease Pathogenesis to Therapeutic Potentials. Front. Cell Neurosci. 2021, 15, 785433. [Google Scholar] [CrossRef]
  111. Zhang, X.; Ma, W.; Xue, W.; Wang, Y.; Chen, P.; Li, Q.; Li, Y.Y.; Hu, X.; Zhao, Y.; Zhou, H. miR-181a plays the tumor-suppressor role in chronic myeloid leukemia CD34+ cells partially via SERPINE1. Cell. Mol. Life Sci. 2023, 81, 10. [Google Scholar] [CrossRef]
  112. Viera, G.M.; Salomao, K.B.; de Sousa, G.R.; Baroni, M.; Delsin, L.E.A.; Pezuk, J.A.; Brassesco, M.S. miRNA signatures in childhood sarcomas and their clinical implications. Clin. Transl. Oncol. 2019, 21, 1583–1623. [Google Scholar] [CrossRef]
  113. Hermansen, S.K.; Kristensen, B.W. MicroRNA biomarkers in glioblastoma. J. Neurooncol. 2013, 114, 13–23. [Google Scholar] [CrossRef]
  114. Rodriguez-Ortiz, C.J.; Prieto, G.A.; Martini, A.C.; Forner, S.; Trujillo-Estrada, L.; LaFerla, F.M.; Baglietto-Vargas, D.; Cotman, C.W.; Kitazawa, M. miR-181a negatively modulates synaptic plasticity in hippocampal cultures and its inhibition rescues memory deficits in a mouse model of Alzheimer’s disease. Aging Cell 2020, 19, e13118. [Google Scholar] [CrossRef]
  115. Sanuki, R.; Yamamura, T. Tumor Suppressive Effects of miR-124 and Its Function in Neuronal Development. Int. J. Mol. Sci. 2021, 22, 5919. [Google Scholar] [CrossRef]
  116. Kohama, I.; Kosaka, N.; Chikuda, H.; Ochiya, T. An Insight into the Roles of MicroRNAs and Exosomes in Sarcoma. Cancers 2019, 11, 428. [Google Scholar] [CrossRef]
  117. Yaghoubi, N.; Zahedi Avval, F.; Khazaei, M.; Aghaee-Bakhtiari, S.H. MicroRNAs as potential investigative and predictive biomarkers in colorectal cancer. Cell Signal. 2021, 80, 109910. [Google Scholar] [CrossRef]
  118. Hou, T.Y.; Zhou, Y.; Zhu, L.S.; Wang, X.; Pang, P.; Wang, D.Q.; Liuyang, Z.Y.; Man, H.; Lu, Y.; Zhu, L.Q.; et al. Correcting abnormalities in miR-124/PTPN1 signaling rescues tau pathology in Alzheimer’s disease. J. Neurochem. 2020, 154, 441–457. [Google Scholar] [CrossRef]
  119. Saiki, S.; Sato, S.; Hattori, N. Molecular pathogenesis of Parkinson’s disease: Update. J. Neurol. Neurosurg. Psychiatry 2012, 83, 430–436. [Google Scholar] [CrossRef]
  120. Junn, E.; Lee, K.W.; Jeong, B.S.; Chan, T.W.; Im, J.Y.; Mouradian, M.M. Repression of alpha-synuclein expression and toxicity by microRNA-7. Proc. Natl. Acad. Sci. USA 2009, 106, 13052–13057. [Google Scholar] [CrossRef]
  121. Galter, D.; Westerlund, M.; Carmine, A.; Lindqvist, E.; Sydow, O.; Olson, L. LRRK2 expression linked to dopamine-innervated areas. Ann. Neurol. 2006, 59, 714–719. [Google Scholar] [CrossRef]
  122. Cho, H.J.; Liu, G.; Jin, S.M.; Parisiadou, L.; Xie, C.; Yu, J.; Sun, L.; Ma, B.; Ding, J.; Vancraenenbroeck, R.; et al. MicroRNA-205 regulates the expression of Parkinson’s disease-related leucine-rich repeat kinase 2 protein. Hum. Mol. Genet. 2013, 22, 608–620. [Google Scholar] [CrossRef]
  123. Chen, Y.; Gao, C.; Sun, Q.; Pan, H.; Huang, P.; Ding, J.; Chen, S. MicroRNA-4639 Is a Regulator of DJ-1 Expression and a Potential Early Diagnostic Marker for Parkinson’s Disease. Front. Aging Neurosci. 2017, 9, 232. [Google Scholar] [CrossRef]
  124. Saraiva, C.; Paiva, J.; Santos, T.; Ferreira, L.; Bernardino, L. MicroRNA-124 loaded nanoparticles enhance brain repair in Parkinson’s disease. J. Control. Release 2016, 235, 291–305. [Google Scholar] [CrossRef]
  125. Tufekci, K.U.; Meuwissen, R.; Genc, S.; Genc, K. Inflammation in Parkinson’s disease. Adv. Protein Chem. Struct. Biol. 2012, 88, 69–132. [Google Scholar] [CrossRef]
  126. Ling, Z.; Fan, G.; Yao, D.; Zhao, J.; Zhou, Y.; Feng, J.; Zhou, G.; Chen, Y. MicroRNA-150 functions as a tumor suppressor and sensitizes osteosarcoma to doxorubicin-induced apoptosis by targeting RUNX2. Exp. Ther. Med. 2020, 19, 481–488. [Google Scholar] [CrossRef]
  127. Lv, T.; Jiang, L.; Kong, L.; Yang, J. MicroRNA-29c-3p acts as a tumor suppressor gene and inhibits tumor progression in hepatocellular carcinoma by targeting TRIM31. Oncol. Rep. 2020, 43, 953–964. [Google Scholar] [CrossRef]
  128. Li, H.; Yu, L.; Li, M.; Chen, X.; Tian, Q.; Jiang, Y.; Li, N. MicroRNA-150 serves as a diagnostic biomarker and is involved in the inflammatory pathogenesis of Parkinson’s disease. Mol. Genet. Genom. Med. 2020, 8, e1189. [Google Scholar] [CrossRef]
  129. Wang, R.; Li, Q.; He, Y.; Yang, Y.; Ma, Q.; Li, C. miR-29c-3p inhibits microglial NLRP3 inflammasome activation by targeting NFAT5 in Parkinson’s disease. Genes Cells 2020, 25, 364–374. [Google Scholar] [CrossRef]
  130. Ghafouri-Fard, S.; Gholipour, M.; Taheri, M. MicroRNA Signature in Melanoma: Biomarkers and Therapeutic Targets. Front. Oncol. 2021, 11, 608987. [Google Scholar] [CrossRef]
  131. Hu, Y.B.; Zhang, Y.F.; Wang, H.; Ren, R.J.; Cui, H.L.; Huang, W.Y.; Cheng, Q.; Chen, H.Z.; Wang, G. miR-425 deficiency promotes necroptosis and dopaminergic neurodegeneration in Parkinson’s disease. Cell Death Dis. 2019, 10, 589. [Google Scholar] [CrossRef]
  132. Ricci, C.; Marzocchi, C.; Battistini, S. MicroRNAs as Biomarkers in Amyotrophic Lateral Sclerosis. Cells 2018, 7, 219. [Google Scholar] [CrossRef]
  133. Shioya, M.; Obayashi, S.; Tabunoki, H.; Arima, K.; Saito, Y.; Ishida, T.; Satoh, J. Aberrant microRNA expression in the brains of neurodegenerative diseases: miR-29a decreased in Alzheimer disease brains targets neurone navigator 3. Neuropathol. Appl. Neurobiol. 2010, 36, 320–330. [Google Scholar] [CrossRef]
  134. Cattaneo, E.; Zuccato, C.; Tartari, M. Normal huntingtin function: An alternative approach to Huntington’s disease. Nat. Rev. Neurosci. 2005, 6, 919–930. [Google Scholar] [CrossRef]
  135. Zuccato, C.; Tartari, M.; Crotti, A.; Goffredo, D.; Valenza, M.; Conti, L.; Cataudella, T.; Leavitt, B.R.; Hayden, M.R.; Timmusk, T.; et al. Huntingtin interacts with REST/NRSF to modulate the transcription of NRSE-controlled neuronal genes. Nat. Genet. 2003, 35, 76–83. [Google Scholar] [CrossRef]
  136. Johnson, R.; Zuccato, C.; Belyaev, N.D.; Guest, D.J.; Cattaneo, E.; Buckley, N.J. A microRNA-based gene dysregulation pathway in Huntington’s disease. Neurobiol. Dis. 2008, 29, 438–445. [Google Scholar] [CrossRef]
  137. Martí, E.; Pantano, L.; Bañez-Coronel, M.; Llorens, F.; Miñones-Moyano, E.; Porta, S.; Sumoy, L.; Ferrer, I.; Estivill, X. A myriad of miRNA variants in control and Huntington’s disease brain regions detected by massively parallel sequencing. Nucleic Acids Res. 2010, 38, 7219–7235. [Google Scholar] [CrossRef]
  138. Lee, S.T.; Chu, K.; Im, W.S.; Yoon, H.J.; Im, J.Y.; Park, J.E.; Park, K.H.; Jung, K.H.; Lee, S.K.; Kim, M.; et al. Altered microRNA regulation in Huntington’s disease models. Exp. Neurol. 2011, 227, 172–179. [Google Scholar] [CrossRef]
  139. Fukuoka, M.; Takahashi, M.; Fujita, H.; Chiyo, T.; Popiel, H.A.; Watanabe, S.; Furuya, H.; Murata, M.; Wada, K.; Okada, T.; et al. Supplemental Treatment for Huntington’s Disease with miR-132 that Is Deficient in Huntington’s Disease Brain. Mol. Ther. Nucleic Acids 2018, 11, 79–90. [Google Scholar] [CrossRef]
  140. Rafat, M.; Moraghebi, M.; Afsa, M.; Malekzadeh, K. The outstanding role of miR-132-3p in carcinogenesis of solid tumors. Hum. Cell 2021, 34, 1051–1065. [Google Scholar] [CrossRef]
  141. Li, Y.; Zu, L.; Wang, Y.; Wang, M.; Chen, P.; Zhou, Q. miR-132 inhibits lung cancer cell migration and invasion by targeting SOX4. J. Thorac. Dis. 2015, 7, 1563–1569. [Google Scholar] [CrossRef]
  142. Otmani, K.; Lewalle, P. Tumor Suppressor miRNA in Cancer Cells and the Tumor Microenvironment: Mechanism of Deregulation and Clinical Implications. Front. Oncol. 2021, 11, 708765. [Google Scholar] [CrossRef]
  143. Feng, J.; Hu, S.; Liu, K.; Sun, G.; Zhang, Y. The Role of MicroRNA in the Regulation of Tumor Epithelial-Mesenchymal Transition. Cells 2022, 11, 1981. [Google Scholar] [CrossRef]
  144. Elnaggar, G.N.; El-Hifnawi, N.M.; Ismail, A.; Yahia, M.; Elshimy, R.A.A. Micro RNA-148a Targets Bcl-2 in Patients with Non-Small Cell Lung Cancer. Asian Pac. J. Cancer Prev. 2021, 22, 1949–1955. [Google Scholar] [CrossRef]
  145. Peng, Y.; Zhang, X.; Feng, X.; Fan, X.; Jin, Z. The crosstalk between microRNAs and the Wnt/β-catenin signaling pathway in cancer. Oncotarget 2017, 8, 14089–14106. [Google Scholar] [CrossRef]
  146. Ashrafizadeh, M.; Ang, H.L.; Moghadam, E.R.; Mohammadi, S.; Zarrin, V.; Hushmandi, K.; Samarghandian, S.; Zarrabi, A.; Najafi, M.; Mohammadinejad, R.; et al. MicroRNAs and Their Influence on the ZEB Family: Mechanistic Aspects and Therapeutic Applications in Cancer Therapy. Biomolecules 2020, 10, 1040. [Google Scholar] [CrossRef]
  147. Davalos, V.; Moutinho, C.; Villanueva, A.; Boque, R.; Silva, P.; Carneiro, F.; Esteller, M. Dynamic epigenetic regulation of the microRNA-200 family mediates epithelial and mesenchymal transitions in human tumorigenesis. Oncogene 2012, 31, 2062–2074. [Google Scholar] [CrossRef]
  148. Sun, F.; Li, S.G.; Zhang, H.W.; Hua, F.W.; Sun, G.Z.; Huang, Z. MiRNA-411 attenuates inflammatory damage and apoptosis following spinal cord injury. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 491–498. [Google Scholar] [CrossRef]
  149. Rodrigues, F.B.; Wild, E.J. Huntington’s Disease Clinical Trials Corner: April 2020. J. Huntingtons Dis. 2020, 9, 185–197. [Google Scholar] [CrossRef]
  150. Zhao, S.; Li, J.; Zhang, G.; Wang, Q.; Wu, C.; Zhang, Q.; Wang, H.; Sun, P.; Xiang, R.; Yang, S. Exosomal miR-451a Functions as a Tumor Suppressor in Hepatocellular Carcinoma by Targeting LPIN1. Cell Physiol. Biochem. 2019, 53, 19–35. [Google Scholar] [CrossRef]
  151. Sun, H.; Jiang, P. MicroRNA-451a acts as tumor suppressor in cutaneous basal cell carcinoma. Mol. Genet. Genom. Med. 2018, 6, 1001–1009. [Google Scholar] [CrossRef]
  152. Tan, J.; Li, C.; Ren, L.; Zhu, X.; Hua, F.; Fu, Y. miR-451a suppresses papillary thyroid cancer cell proliferation and invasion and facilitates apoptosis through targeting DCBLD2 and AKT1. Mol. Cell Probes 2022, 66, 101863. [Google Scholar] [CrossRef]
  153. Keskin, S.; Brouwers, C.C.; Sogorb-Gonzalez, M.; Martier, R.; Depla, J.A.; Vallès, A.; van Deventer, S.J.; Konstantinova, P.; Evers, M.M. AAV5-miHTT Lowers Huntingtin mRNA and Protein without Off-Target Effects in Patient-Derived Neuronal Cultures and Astrocytes. Mol. Ther. Methods Clin. Dev. 2019, 15, 275–284. [Google Scholar] [CrossRef]
  154. Miniarikova, J.; Zimmer, V.; Martier, R.; Brouwers, C.C.; Pythoud, C.; Richetin, K.; Rey, M.; Lubelski, J.; Evers, M.M.; van Deventer, S.J.; et al. AAV5-miHTT gene therapy demonstrates suppression of mutant huntingtin aggregation and neuronal dysfunction in a rat model of Huntington’s disease. Gene Ther. 2017, 24, 630–639. [Google Scholar] [CrossRef]
  155. Rupaimoole, R.; Slack, F.J. MicroRNA therapeutics: Towards a new era for the management of cancer and other diseases. Nat. Rev. Drug Discov. 2017, 16, 203–222. [Google Scholar] [CrossRef]
  156. Winkle, M.; El-Daly, S.M.; Fabbri, M.; Calin, G.A. Noncoding RNA therapeutics—Challenges and potential solutions. Nat. Rev. Drug Discov. 2021, 20, 629–651. [Google Scholar] [CrossRef]
  157. Laganà, A.; Acunzo, M.; Romano, G.; Pulvirenti, A.; Veneziano, D.; Cascione, L.; Giugno, R.; Gasparini, P.; Shasha, D.; Ferro, A.; et al. miR-Synth: A computational resource for the design of multi-site multi-target synthetic miRNAs. Nucleic Acids Res. 2014, 42, 5416–5425. [Google Scholar] [CrossRef]
Figure 1. The schematic of Aberrant Cell Cycle Diseases (ACCDs). The molecules and related mitogenic pathways contributing to the aberrant cell cycle re-entry that is associated with not only tumorigenesis in cancers but also neuronal death in neurological diseases. The solid arrow lines indicate activation, while the dashed arrow lines indicate inhibition. The arrows do not necessarily indicate direct binding and/or activation of the downstream effectors; intermediate molecules may exist. Akt: protein kinase B; Ca2+: calcium; Cdk: cyclin-dependent kinase; ERK: extracellular signal-regulated kinase; FAK: focal adhesion kinase; GPCR: G protein-coupled receptor; GSK3β: glycogen synthase-3 beta; IP3: inositol trisphosphate; JAK: Janus kinase; JNK: c-Jun N-terminal kinases; MEK: mitogen-activated protein kinase kinase; miR-122: microRNA-122; miR-451a: microRNA-451a; MLK: mixed lineage kinases; mTOR: mammalian target of rapamycin; NF-kB: nuclear factor kappa B; PI3K: phosphatidylinositol 3-kinase; PLC: phospholipase C; Ras: rat sarcoma virus kinase; Rac1: ras-related C3 botulinum toxin substrate 1; Raf: rapidly accelerated fibrosarcoma; ROS: reactive oxygen species; SFKs: Src family kinases; STAT: signal transducer and activator of transcription.
Figure 1. The schematic of Aberrant Cell Cycle Diseases (ACCDs). The molecules and related mitogenic pathways contributing to the aberrant cell cycle re-entry that is associated with not only tumorigenesis in cancers but also neuronal death in neurological diseases. The solid arrow lines indicate activation, while the dashed arrow lines indicate inhibition. The arrows do not necessarily indicate direct binding and/or activation of the downstream effectors; intermediate molecules may exist. Akt: protein kinase B; Ca2+: calcium; Cdk: cyclin-dependent kinase; ERK: extracellular signal-regulated kinase; FAK: focal adhesion kinase; GPCR: G protein-coupled receptor; GSK3β: glycogen synthase-3 beta; IP3: inositol trisphosphate; JAK: Janus kinase; JNK: c-Jun N-terminal kinases; MEK: mitogen-activated protein kinase kinase; miR-122: microRNA-122; miR-451a: microRNA-451a; MLK: mixed lineage kinases; mTOR: mammalian target of rapamycin; NF-kB: nuclear factor kappa B; PI3K: phosphatidylinositol 3-kinase; PLC: phospholipase C; Ras: rat sarcoma virus kinase; Rac1: ras-related C3 botulinum toxin substrate 1; Raf: rapidly accelerated fibrosarcoma; ROS: reactive oxygen species; SFKs: Src family kinases; STAT: signal transducer and activator of transcription.
Pharmaceuticals 17 00426 g001
Figure 2. Biogenesis, processing, and function of miRNAs (AGO2: Argonaute 2 protein; GW182: glycine-tryptophan protein of 182 kDa; RISC: RNA-induced silencing complex; 3′UTR: 3′ untranslated region).
Figure 2. Biogenesis, processing, and function of miRNAs (AGO2: Argonaute 2 protein; GW182: glycine-tryptophan protein of 182 kDa; RISC: RNA-induced silencing complex; 3′UTR: 3′ untranslated region).
Pharmaceuticals 17 00426 g002
Table 1. (a) miRNAs upregulated in cancers and neurological disorders; (b) miRNAs downregulated in cancers and neurological disorders.
Table 1. (a) miRNAs upregulated in cancers and neurological disorders; (b) miRNAs downregulated in cancers and neurological disorders.
(a)
miRNAs Upregulated in Cancers OnlymiRNAs Upregulated in Both Cancers and Neurological DisordersmiRNAs Upregulated in Neurological Disorders Only
miR-BART7, miR-BART13, miR-2, miR-3, miR-3P, miR-9-1, miR-9-2, miR-9-3, miR-15b-5p, miR-16-1-3p, miR-16-2, miR-16-2-3p, miR-18, miR-19b, miR-19b-3p, miR-20, miR-20a-3p, miR-20b-3p, miR-21-3p, miR-24-2, miR-24-3p, miR-25a, miR-25-5p, miR-27, miR-27a, miR-27-3p, miR-28, miR-28-3p, miR-29a-3p, miR-29c-3p, miR-30b, miR-30c-2, miR-30e-3p, miR-31-5p, miR-32-5p, miR-33, miR-33b-3p, miR-45, miR-48, miR-59, miR-81a, miR-92a-3p, miR-92b, miR-93-5p, miR-95-3p, miR-96-3p, miR-96-5p, miR-98-5p, miR-106a-3633, miR-106b-25, miR-125b-1-3p, miR-126a, miR-129-1, miR-129-2, miR-130, miR-130a-3p, miR-131a, miR-132-5p, miR-135, miR-135a, miR-135b, miR-138-1-3p, miR-138-2-3p, miR-138b, miR-141b, miR-145b, miR-146-5p, miR-146b-3p, miR-146b-5p, miR-148b, miR-151, miR-151-3p, miR-151a, miR-151a-3p, miR-152-3p, miR-159, miR-159a, miR-181a, miR-181a-2-3p, miR-181b-5p, miR-185-5p, miR-187-5p, miR-189-5p, miR-190, miR-194-5p, miR-196, miR-196b, miR-197-5p, miR-199-5p, miR-199-s, miR-203a, miR-203b, miR-208a, miR-208a-3p, miR-210, miR-210-3p, miR-215-5p, miR-218-1, miR-221/222 cluster, miR-222-3p, miR-222-5p, miR-223-3p, miR-244, miR-275, miR-300, miR-301, miR-301b, miR-301b-3p, miR-302-367 cluster, miR-302-3p, miR-302b-3p, miR-320e, miR-325, miR-339-5p, miR-361, miR-365a, miR-365b, miR-367, miR-369-3p, miR-370-3p, miR-371, miR-371a-3p, miR-371a-5p, miR-373-3p, miR-373-5p, miR-374a, miR-378a-5p, miR-378d, miR-409, miR-411, miR-412, miR-412-3p, miR-421, miR-423-3p, miR-423-5p, miR-449, miR-450a, miR-450b-5p, miR-452-5p, miR-466, miR-483, miR-483-5p, miR-485-3p, miR-485-5, miR-493-3p, miR-498, miR-500, miR-500a, miR-502-3p, miR-503-5p, miR-505-3p, miR-506-514 cluster, miR-512-5p, miR-513-1, miR-513-2, miR-514, miR-516a-3p, miR-517, miR-517b-3p, miR-518a, miR-518c, miR-519a, miR-520c-3p, miR-520f, miR-520g, miR-520h, miR-548a, miR-548ah, miR-548ar, miR-548b, miR-551b-3p, miR-551b-5p, miR-556-5p, miR-574-5p, miR-582-3p, miR-584-5p, miR-589, miR-592, miR-615-3p, miR-616, miR-618, miR-626, miR-628, miR-628-5p, miR-629-5p, miR-633, miR-635, miR-640, miR-642a-5p, miR-642b-3p, miR-645, miR-647, miR-657, miR-6741-3p, miR-675-3p, miR-711, miR-764, miR-767, miR-769, miR-857, miR-877-3p, miR-886-3p, miR-886-5p, miR-933, miR-938, miR-939, miR-1185-2-3p, miR-1234, miR-1247, miR-1248, miR-1254, miR-1269, miR-1273g-3p, miR-1275, miR-1293, miR-1303, miR-1307, miR-1323, miR-1537, miR-1537-3p, miR-1825, miR-1908, miR-1908-5p, miR-3065, miR-3131, miR-3136, miR-3141, miR-3144-3p, miR-3147, miR-3151, miR-3153, miR-3162, miR-3176, miR-3177-3p, miR-3189, miR-3200-5p, miR-3201, miR-3613-3p, miR-3917, miR-3976, miR-4267, miR-4270, miR-4283, miR-4289, miR-4418, miR-4429, miR-4465, miR-4484, miR-4644, miR-4652, miR-4664-3p, miR-4665-5p, miR-4709, miR-4764-3p, miR-5001-5p, miR-5100, miR-5191-3p, miR-5694, miR-6796-3p, miR-6826, miR-6852, miR-6875, miR-7641, miR-7702, let-5p, let-7b, let-7b-3p, let-7c-5p, let-7d-5p, let-7f-2, let-7f-5p, let-7gmiR-18a-5p, miR-18b-5p, miR-19a-3p, miR-20a-5p, miR-22-5p, miR-105-5p, miR-126-5p, miR-181a-5p, miR-223-5p, miR-323b-3p, miR-372-3p, miR-373, miR-455-3p, miR-488, miR-549, miR-582-5p, miR-585, miR-595, miR-627-5p, miR-671-5p, miR-762, miR-877-5p, miR-941, miR-1225-5p, miR-1249, miR-3613-5p, miR-4317, miR-4435, miR-iR-4293, let-7e-5p, mIR-17-92 clustermiR-3p-57664, miR-7-1-3p, miR-7b-5p, miR-9-3p, miR-9a-3p, miR-10, miR-10a, miR-15a, miR-17-5p, miR-18b, miR-21a-5p, miR-25, miR-26a-2, miR-26b-3p, miR-27a-5p, miR-27b, miR-27b-5p, miR-30a-3p, miR-30a-5p, miR-30c-2-3p, miR-30d, miR-30e-5p, miR-32, miR-32-3p, miR-93a-5p, miR-99a, miR-99a-5p, miR-99b, miR-99b-5p, miR-101-3p, miR-101b-3p, miR-106-5p, miR-124a, miR-124-2, miR-124-3, miR-125a-5p, miR-125b-2, miR-128b, miR-130b, miR-133, miR-133a-1, miR-133a-2-3p, miR-134-5, miR-135b-3p, miR-136, miR-136-5p, miR-139, miR-142, miR-142a-5p, miR-145, miR-146b, miR-150, miR-152, miR-154-5p, miR-155-5p, miR-181a-3p, miR-181b-1-3p, miR-181c-3p, miR-184, miR-186-3p, miR-187, miR-191, miR-192-3p, miR-192-5p, miR-193b-5p, miR-194, miR-194-3p, miR-195, miR-195-3p, miR-196a, miR-199a, miR-199b, miR-200b, miR-200c, miR-203, miR-203-3p, miR-204-5p, miR-206, miR-208a-5p, miR-220c, miR-290, miR-211, miR-211-5p, miR-219a.2-3p, miR-219a-5p, miR-292-5p, miR-298, miR-300-3p, miR-302d-3p, miR-322, miR-323a-5p, miR-328-5p, miR-330-3p, miR-339, miR-339-3p, miR-340-3p, miR-345-3p, miR-362, miR-362-3p, miR-363-3p, miR-365a, miR-375, miR-376b-3p, miR-378c, miR-381, miR-382-3p, miR-424, miR-424-5p, miR-425, miR-433-3p, miR-434-3p, miR-451-5p, miR-454-3p, miR-455-5p, miR-486, miR-488-3p, miR-490-3p, miR-494, miR-495-3p, miR-499, miR-499a-3p, miR-499a-5p, miR-500a-3p, miR-505-5p, miR-509-3-5p, miR-513a-5p, miR-518f-3p, miR-519, miR-520d-3p, miR-520f-3p, miR-525-5p, miR-532, miR-532-3p, miR-545, miR-548at-5p, miR-548d, miR-550, miR-552, miR-553, miR-572, miR-579, miR-579-3p, miR-590-5p, miR-601, miR-602, miR-603, miR-611, miR-613, miR-615-5p, miR-617, miR-623, miR-627, miR-628-3p, miR-629-3p, miR-637, miR-638, miR-641, miR-656-3p, miR-659, miR-660, miR-660-5p, miR-663a, miR-665, miR-668, miR-671, miR-671-3p, miR-675, miR-682, miR-686, miR-762, miR-765, miR-874, miR-883a-3p, miR-883b-3p, miR-891, miR-920, miR-923 miR-942, miR-943, miR-1183, miR-1184, miR-1185-1-3p, miR-1202, miR-1224, miR-1246, miR-1249, miR-1255b, miR-1261, miR-1274a, miR-1274b, miR-1277-3p, miR-1285, miR-1285-5p, miR-1289, miR-1290, miR-1291, miR-1321, miR-1843a-5p, miR-2861, miR-3190-3p, miR-3195, miR-3196, miR-3615, miR-3646, miR-3928-5p, miR-3939, miR-4306, miR-4433b-3p, miR-4449, miR-4521, miR-4649-5p, miR-4669, miR-4674, miR-4747-3p, miR-5001-3p, miR-5010-3p, miR-6096-5p, miR-6716-3p, miR-6736-3p, miR-6740-3p, miR-6747-3p, miR-6750-5p, miR-6753-3p, miR-6754-3p, miR-6761-3p, miR-6762-3p, miR-6777-3p, miR-6778-3p, miR-6787-3p, miR-6836-3p, miR-6867-5p, miR-6875-3p, miR-8082, miR-B19527a, let-7-c-3p let-7e, let-7g-3p, miR-PC-5p-12969, miR-23-27-24 cluster
(b)
miRNAs Downregulated in Cancers OnlymiRNAs Downregulated in Both Cancers and Neurological DisordersmiRNAs Downregulated in Neurological Disorders Only
miR-1-3p, miR-6-3p, miR-10, miR-15-5p, miR-15/16 cluster, miR-15a-3p, miR-16-1, miR-16.1, miR-23a-3p, miR-24-1-5p, miR-26, miR-26a-5p, miR-27a-5p, miR-29b, miR-29b-1-5p, miR-29c-5p, miR-30, miR-30b-3p, miR-30b-5p, miR-30c-2-3p, miR-33a-5p, miR-34b-3p, miR-34b-5p, miR-92, miR-92a cluster, miR-93-3p, miR-99-3b, miR-99b-5p, miR-103b, miR-104-5p, miR-122a, miR-124-5p, miR-126b, miR-127-3p, miR-128-1, miR-129-3p, miR-129b-5p, miR-130b-3p, miR-130b-5p, miR-132-5p, miR-133, miR-133a, miR-133a-3p, miR-134-5p, miR-135-5p, miR-136, miR-136-5p, miR-137, miR-138-5p, miR-145, miR-147, miR-147b, miR-148a-3p, miR-149-5p, miR-152, miR-153, miR-153-3p, miR-181a-3p, miR-181d-5p, miR-188, miR-193-3p, miR-193-5p, miR-193a, miR-196b-5p, miR-199-3p, miR-216-5p, miR-216a, miR-216a-3p, miR-216b, miR-217, miR-218-5p, miR-219a-2-3p, miR-220, miR-220a, miR-220b, miR-269-3p, miR-296-3p, miR-299, miR-302, miR-302a, miR-302b, miR-302c, miR-320a-3p, miR-323, miR-323-3p, miR-323-5p, miR-328, miR-328-3p, miR-330-3p, miR-330-3p-3p, miR-330-5p, miR-333-3p, miR-335-5p, miR-336, miR-337, miR-339-3p, miR-342, miR-362, miR-363, miR-365b-3p, miR-365b-5p, miR-374c, miR-374c-5p, miR-376, miR-376a-3p, miR-381-3p, miR-382-5p, miR-383-5p, miR-384, miR-385, miR-397a, miR-397b, miR-397c, miR-411-5p, miR-424-3p, miR-429-5p, miR-432, miR-433-3p, miR-448, miR-449b, miR-449c, miR-451, miR-451a.1, miR-454-3p, miR-485-5p, miR-486-3p, miR-487b, miR-488-3p, miR-488-5p, miR-489, miR-489-3p, miR-490-3p, miR-491, miR-491-3p, miR-491-5p, miR-497-5p, miR-506-3p, miR-508-3p, miR-509-5p, miR-509-3-5p, miR-511-5p, miR-513, miR-513c-5p, miR-514a-3p, miR-516b, miR-519, miR-520a-3p, miR-524-5p, miR-526b, miR-526b-5p, miR-532-5p, miR-539, miR-544, miR-545, miR-548b-5p, miR-548c-3p, miR-551b, miR-564, miR-573, miR-574, miR-579-3p, miR-582, miR-584, miR-590-3p, miR-593-3p, miR-596, miR-599, miR-601, miR-605, miR-610, miR-612, miR-613, miR-622, miR-625, miR-625-5p, miR-627, miR-628-3p, miR-632, miR-634, miR-636, miR-637, miR-639, miR-650, miR-652-5p, miR-654-5p, miR-655, miR-656, miR-661, miR-665, miR-708, miR-708-5p, miR-744, miR-744-5p, miR-760, miR-768-3p, miR-770-5p, miR-802, miR-874, miR-877, miR-886-3p, miR-887, miR-891a, miR-892b, miR-936, miR-937, miR-937-5p, miR-942, miR-1179, miR-1180-3p, miR-1182, miR-1203, miR-1205, miR-1226, miR-1227, miR-1228-5p, miR-1231, miR-1250, miR-1257, miR-1260, miR-1266, miR-1271, miR-1274a, miR-1280, miR-1283, miR-1287, miR-1296, miR-1301-3p, miR-1302, miR-1321, miR-1587, miR-1913, miR-2682-3p, miR-2861, miR-3065-5p, miR-3156-3p, miR-3170, miR-3185, miR-3196, miR-3612, miR-3614, miR-3651, miR-3662, miR-3666, miR-3714, miR-3928, miR-3960, miR-4259, miR-4282, miR-4324, miR-4430, miR-4458, miR-4467, miR-4478, miR-4485-5p, miR-4488, miR-4501, miR-4633-5p, miR-4635, miR-4647, miR-4706, miR-4730, miR-4731, miR-5096, miR-6086, miR-6503, miR-6510, miR-6511b-5p, miR-6887-5p, let-7amiR-28-5p, miR-148-3p, miR-199b-3p, miR-208b, miR-219-5p, miR-320b, miR-323a-3p, miR-330, miR-374b, miR-449b-5p, miR-493-5p, miR-509-3p, miR-511, miR-575, miR-1297, miR-4487, let-7i-5pmiR-2-5p, miR-7, miR-7c-5p, miR-7g, miR-9a-5p, miR-15, miR-15b, miR-15b-3p, miR-15b-5p, miR-16-2-3p, miR-16-5p, miR-17-3p, miR-18a, miR-19b-5p, miR-22-5p, miR-23b, miR-23b-3p, miR-26a, miR-26a-1, miR-26a-2-5p, miR-26b-5p, miR-27a-3p, miR-29, miR-30c, miR-31-5p, miR-33b-5p, miR-92a-5p, miR-92b, miR-96, miR-96-5p, miR-98, miR-98-5p, miR-101, miR-103, miR-106a-5, miR-106b-3p, miR-112, miR-122, miR-125, miR-125a, miR-125a-3p, miR-125b-1-3p, miR-126-3p, miR-128-2-5p, miR-130a, miR-130-b-5p, miR-134-3p, miR-139-3p, miR-140, miR-146-5p, miR-146a-3p, miR-146b-5p, miR-148b, miR-148b-3p, miR-148b-5p, miR-149, miR-150-3p, miR-151-3p, miR-151b, miR-161, miR-181-3p, miR-181-5p, miR-182, miR-182-5p, miR-183, miR-183-5p, miR-186, miR-186-5p, miR-187-3p, miR-187-5p, miR-190, miR-192, miR-193a-5p, miR-193b, miR-199a-5p, miR-203a-3p, miR-203b-5p, miR-204, miR-208a, miR-208b-3p, miR-212, miR-212-3p, miR-219, miR-221-5p, miR-298-5p, miR-299-5p, miR-301, miR-320, miR-320a, miR-320c, miR-320e, miR-325-5p, miR-331-3p, miR-335-3p, miR-337-5p, miR-340, miR-340-5p, miR-342-3p, miR-351-3p, miR-352, miR-362-5p, miR-370-3p, miR-374a, miR-374b-5p, miR-379, miR-380, miR-423-3p, miR-423-5p, miR-425-p, miR-433, miR-483-3p, miR-485, miR-486-1, miR-487a, miR-487b-3p, miR-493, miR-500, miR-501-5p, miR-502, miR-502-3p, miR-502-5p, miR-519e, miR-543, miR-545-3p, miR-574-3p, miR-576-5p, miR-598, miR-598-3p, miR-605-5p, miR-626, miR-629, miR-652, miR-669c-5p, miR-744, miR-744-3p, miR-764-3p, miR-767, miR-769, miR-769-5p, miR-873-3p, miR-877-3p, miR-885-3p, miR-885-5p, miR-886-5p, miR-1185-2-3p, miR-1224-5p, miR-1233-5p, miR-1234-3p, miR-1237-5p, miR-1249-3p, miR-1259, miR-1294, miR-1299, miR-1303-3p, miR-1306-3p, miR-1825, miR-1909-3p, miR-1915-3p, miR-3552, miR-3613-3p, miR-3916, miR-3935, miR-4299, miR-4322, miR-4422, miR-4429, miR-4448, miR-4530, miR-4745-5p, miR-4772-3p, miR-6131, miR-6236, miR-6805-5p, miR-6781-5p, miR-7046-3p, miR-7660-3p, miR-7665-5p, miR-7674-5p, miR-8071, let-7a-5p, let-7c, let-7g
Table 2. Study details of clinical trials on miRNA therapeutics in neurological disorders.
Table 2. Study details of clinical trials on miRNA therapeutics in neurological disorders.
Clinical Trial NCT NumberAims of Clinical TrialPhaseStart and End DateStatusResults
NCT04120493Establish safety and proof-of-concept of AMT-130 in patients with early-stage HD.Interventional (Phase I/II)6 September 2019–June 2019 (estimated)RecruitingNone
NCT05243017Examine safety and efficacy of AMT-130 in patients with HD.Interventional (Phase I/II)7 October 2021–7 October 2027 (estimated)RecruitingNone
NCT06100276Examine safety, tolerability, and efficacy of AMT-162 in patients with ALS.Interventional (phase 1/II)1 Febuary 2024–27 Febuary 2032 (estimated)Not Recruiting YetNone
NCT06063850Examine safety, tolerability, and efficacy of AMT-260 in patients with unilateral refractory mesial temporal lobe epilepsy.Interventional (phase I/II)17 November 2023–30 November 2026 (estimated)RecruitingNone
HD = Huntington’s Disease, ALS = amyotrophic lateral sclerosis.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lui, A.; Do, T.; Alzayat, O.; Yu, N.; Phyu, S.; Santuya, H.J.; Liang, B.; Kailash, V.; Liu, D.; Inslicht, S.S.; et al. Tumor Suppressor MicroRNAs in Clinical and Preclinical Trials for Neurological Disorders. Pharmaceuticals 2024, 17, 426. https://doi.org/10.3390/ph17040426

AMA Style

Lui A, Do T, Alzayat O, Yu N, Phyu S, Santuya HJ, Liang B, Kailash V, Liu D, Inslicht SS, et al. Tumor Suppressor MicroRNAs in Clinical and Preclinical Trials for Neurological Disorders. Pharmaceuticals. 2024; 17(4):426. https://doi.org/10.3390/ph17040426

Chicago/Turabian Style

Lui, Austin, Timothy Do, Omar Alzayat, Nina Yu, Su Phyu, Hillary Joy Santuya, Benjamin Liang, Vidur Kailash, Dewey Liu, Sabra S. Inslicht, and et al. 2024. "Tumor Suppressor MicroRNAs in Clinical and Preclinical Trials for Neurological Disorders" Pharmaceuticals 17, no. 4: 426. https://doi.org/10.3390/ph17040426

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