Next Article in Journal
Regulation of Smooth Muscle Cell Proliferation by Mitochondrial Ca2+ in Type 2 Diabetes
Previous Article in Journal
Mitochondrial Redox Balance of Fibroblasts Exposed to Ti-6Al-4V Microplates Subjected to Different Types of Anodizing
Previous Article in Special Issue
Systemic Metabolism and Mitochondria in the Mechanism of Alzheimer’s Disease: Finding Potential Therapeutic Targets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of microRNAs in Epigenetic Regulation of Signaling Pathways in Neurological Pathologies

by
Pavel P. Tregub
1,2,3,*,
Irada Ibrahimli
1,
Anton S. Averchuk
3,
Alla B. Salmina
3,4,
Peter F. Litvitskiy
1,
Zaripat Sh. Manasova
1 and
Inga A. Popova
1
1
Department of Pathophysiology, I.M. Sechenov First Moscow State Medical University, 119991 Moscow, Russia
2
Scientific and Educational Resource Center “Innovative Technologies of Immunophenotyping, Digital Spatial Profiling and Ultrastructural Analysis”, RUDN University, 117198 Moscow, Russia
3
Research Center of Neurology, 125367 Moscow, Russia
4
Research Institute of Molecular Medicine and Pathobiochemistry, Prof. V. F. Voino-Yasenetsky Krasnoyarsk State Medical University, 660022 Krasnoyarsk, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(16), 12899; https://doi.org/10.3390/ijms241612899
Submission received: 30 July 2023 / Revised: 11 August 2023 / Accepted: 15 August 2023 / Published: 17 August 2023

Abstract

:
In recent times, there has been a significant increase in researchers’ interest in the functions of microRNAs and the role of these molecules in the pathogenesis of many multifactorial diseases. This is related to the diagnostic and prognostic potential of microRNA expression levels as well as the prospects of using it in personalized targeted therapy. This review of the literature analyzes existing scientific data on the involvement of microRNAs in the molecular and cellular mechanisms underlying the development of pathologies such as Alzheimer’s disease, cerebral ischemia and reperfusion injury, and dysfunction of the blood–brain barrier.

1. Introduction

Discovered 30 years ago, microRNAs (miRNAs) are small non-coding RNA molecules that are 18–25 nucleotides long (22 nucleotides on average) and involved in the transcriptional and post-transcriptional regulation of gene expression by RNA interference, which is of great interest to molecular biologists, geneticists, and biochemists [1,2]. These molecules are mainly present intracellularly, but there is also an extracellular (circulating) microRNA fraction [3]. To date, the existence and functions of more than 2500 human miRNAs are known [4]. The existing database of fundamental knowledge about the structure and functional significance of microRNAs has significantly increased researchers’ interest in studying the role of these molecules in oncological and hereditary pathologies [5]. Promising trends include the diagnostic and prognostic value of assessing microRNA expression levels and their use as targets for personalized targeted therapy [6].
The determination of the role of microRNAs in the epigenetic regulation of signaling pathways in the pathogenesis of multifactorial disorders such as Alzheimer’s disease, ischemia–reperfusion injury, and dysfunction of the blood–brain barrier deserves special attention [7,8,9]. The aim of this review is to summarize and analyze the results of contemporary scientific data on the involvement of microRNAs in the molecular and cellular mechanisms underlying the development of these neurological pathologies.

2. Role of miRNAs in Epigenetic Regulation of Genome Translation

Approximately 70% of the human genome is transcribed into non-coding RNAs that serve regulatory functions. Among them are ribosomal RNA, transfer RNA, small nuclear RNA, and small nucleolar RNA. Non-coding RNAs also include small interfering RNAs, microRNAs, and long non-coding RNAs, which influence the normal expression of target genes, making them a novel target for the development of targeted therapy.
Small interfering RNAs (siRNAs) regulate endonucleolytic cleavage of mRNA and participate in the epigenetic regulation of translation through the RNA interference mechanism, leading to a natural “knockdown” of the target gene [1,10]. MicroRNAs can suppress the expression of target genes and inhibit the translation process. They are capable of recognizing specific mRNA sequences, even with only 6–8 nucleotides at the 5’-end. Long non-coding RNAs control genome integrity, chromatin organization, gene expression, and signal transmission. It has been established and experimentally confirmed that the targeted action of non-coding RNAs on mRNA (in the form of complementary antisense oligonucleotides) can alter the expression levels of target genes [11].
MicroRNAs are efficient signaling regulators of genomic translation. They can bind to several different targets on mRNA, and multiple microRNA variants can be specific to one mRNA target [1]. Additionally, the involvement of microRNAs in processes of positive regulation of transcription and translation has been demonstrated [12]. Such an epigenetic mechanism is undoubtedly phylogenetically more recent and characteristic of higher eukaryotes [3].
The majority of genes encoding microRNAs are antisense and transcribed as independent units [13], but sometimes microRNAs are located in a sense orientation and are co-transcribed with the gene that encodes their mRNA target locus [14]. Nearly half of the known microRNAs are generated from polycistronic units and are encoded by genes located within the introns of non-coding genes [15]. Additionally, for 16% of microRNAs, RNA editing has been demonstrated, allowing for the generation of different microRNA variants from one gene [16].
Mature microRNAs become part of the active complex RISC/microRNP (RNA-induced silencing complex/microRNA–ribonucleoprotein complex) [17]. Proteins of the Argonaute (Ago) family and chaperones Hsp70/Hsp90 play a central role in assembling RISC [18,19]. Once RISC is assembled, Ago proteins are essential for mRNA interference, and they possess two conserved domains for binding to microRNAs: the PAZ domain (binding the 3’-end) and the PIWI domain (binding the 5’-end) [20]. The mature microRNA within the RISC complex is oriented to interact with the target mRNA region [21].
Suppression of gene expression can occur by degrading the mRNA or preventing its translation. If the microRNA is fully complementary to the target mRNA, Ago2 cleaves the mRNA and leads to its degradation [22]. In the case of partial complementarity, RNA interference is achieved through translation blocking (Figure 1).

3. Methods for Assessing the Pool and miRNA Expression

For a detailed investigation of the role of microRNAs in physiological and pathological processes, several methods for isolating microRNAs have been developed, but their stability often raises concerns [23,24]. MicroRNA molecules degrade much more easily than mRNA, which is due to their short length and the activity of RNases [13]. Consequently, all steps involving sample handling require materials, solutions, and tools that are thoroughly RNase-free, and the investigated samples may need good cooling or even additional stabilization using nanomaterials [25].
One of the standard and widely used methods for detecting microRNAs is Northern blotting [26,27]. It can be used to detect not only mature microRNAs but also their precursors. For quantitative assessment of microRNA expression (both absolute and relative), polymerase chain reaction (PCR) methods can be utilized: reverse transcription PCR followed by amplification and real-time fluorescent detection [28,29], isothermal loop-mediated amplification [30], and rolling circle amplification [31,32].
Biosensor [33] and microarray [34] methods are known for hybridizing microRNAs, enabling the simultaneous detection of a large number (hundreds or thousands) of microRNA targets. High-throughput sequencing (Sanger sequencing or next-generation sequencing) [35] is used to identify new microRNA sequences.
To study spatiotemporal expression and intracellular transport, it is essential to investigate microRNA expression in vivo, for which the in situ hybridization method is most suitable [36]. For specific detection of microRNAs in situ, locked nucleic acid (LNA) probes [37] or morpholinos [38] can be used. LNA, due to its locked conformation, exhibits high sensitivity and specificity for detecting microRNA molecules [37]. Moreover, through using LNA and morpholinos, the activity level of microRNAs can be experimentally inhibited [39]. Similar potential exists with 2′-O-methylated oligoribonucleotides and complementary oligonucleotides called antagomirs [40]. Additionally, microRNA maturation and their binding to mRNA sites can be arrested at various points using space-blocking oligonucleotides [41].
One of the modern and promising methods for assessing microRNA effects is the technology of multiplex detection based on microsphere arrays [36,42,43]. This technology enables high-throughput detection of both proteins and nucleic acid targets in various samples. Xu Y et al. [44] described the development of a set of fluorophore-labeled microspheres (known as rainbowarray microspheres). The authors also modified the liquid hybridization method for multiplex detection of microRNA targets and demonstrated the practicality of the technology by quantitatively measuring the expression of two types of microRNA during the differentiation of 3T3-L1 cells.
Methods for measuring the levels of circulating microRNAs in blood serum deserve special attention as they can serve as diagnostic biomarkers for various diseases [45,46,47,48,49,50,51]. The most commonly used methods for this purpose include quantitative reverse transcription polymerase chain reaction (qRT-PCR) [52], loop-mediated isothermal amplification [53,54], and other amplification techniques [55]. There are reports of one-step direct measurement of microRNA molecules in human blood serum using a modified beacon probe with LNA and fluorescence spectroscopy and microscopy [56] as well as reports of amplification with magnetic nanoparticles and ligation [57].
To enhance the efficiency of microRNA detection methods in blood serum, which hold significant diagnostic prospects, approaches utilizing nanomaterials [58] are becoming popular. For instance, superparamagnetic nanoparticles coated with polyethylene glycol and biotin conjugates [59] or intracellular gold nanoprobes [60] are being employed for this purpose.
Performing precise quantitative determination of microRNA pools is methodologically challenging and often associated with a high number of errors [55]. Therefore, bioinformatic methods are crucial for studying microRNA expression levels and investigating their effects at this level [61]. Comparison of data on mRNA targets and microRNAs (based on sequence information) is carried out in specialized databases [62,63]. Sophisticated bioinformatic methods and analytical tools have been developed for comparing data on mRNA targets and microRNAs [64,65,66,67].

4. The Role of MicroRNAs in Signaling Pathways of Alzheimer’s-Type Neurodegeneration

Disruption of epigenetic regulation and alterations in microRNA expression, which usually occur in conjunction, are important factors in the pathogenesis of many neurological disorders [8,9].
Several recent and extensive systematic reviews [68,69,70,71,72,73] have addressed the role of microRNAs in the pathogenesis of Alzheimer’s disease. These reviews explore numerous microRNA variants involved in the epigenetic regulation of the neurodegenerative process that serve as potential targets for diagnosis and targeted therapy.
Examples of such microRNAs include the molecules miR-200b, miR-135a, miR-10a-5p, miR-142a-5p, miR-146a-5p, miR-155-5p, miR-211-5p, miR-455-5p, miR-34a, miR-125b, miR-181c, miR-9, miR-191-5p, miR-181c, and miR-206, which are considered potential diagnostic markers for Alzheimer’s disease [74]. Additionally, attention should be paid to microRNA molecules that possess both diagnostic and therapeutic potential: miR-128, miR-574, miR-146-a, miR-181, miR-132, miR-188-5p, and miR-137 [74].
There are also studies describing modern bioinformatic approaches that use artificial intelligence and machine learning algorithms for identifying new biomarkers and improving the accuracy of molecular diagnosis of Alzheimer’s disease [75,76,77,78,79].
Some review papers present important data on other non-coding RNAs, such as circular RNAs and long non-coding RNAs, which participate in the pathogenesis of neurodegeneration and regulate the interplay between microRNAs/mRNAs/regulatory signaling pathways, which are mediated genetically [80,81]. These RNA molecules are often considered by researchers as more efficient targets for diagnostic and therapeutic strategies. Another important contemporary research direction in Alzheimer’s disease pathogenesis is the study of the regulatory functions of both microRNAs localized in specific organelles (such as mitochondria [82]) and exosomal microRNAs [83,84].
The assessment of neurovascular dysfunction is of particular interest as it plays a crucial role in the onset and progression of the neurodegenerative process associated with the accumulation of β-amyloid peptide in brain neurons and cerebral vessel walls [85,86]. Special attention is given to the dysfunction of the blood–brain barrier, reduced cerebral blood flow, and impaired vascular clearance of beta-amyloid from the brain into the glymphatic system and meningeal lymphatic vessels [87]. These disturbances may, in turn, be linked to the reprogramming of epigenetic regulation [88,89].
According to several authors, RNA interference mediated by microRNAs can initiate neurovascular events leading to Alzheimer’s-type neurodegeneration [90,91,92,93]. This provides a strong theoretical basis for the development of new directions in targeted genetic therapy for neurodegenerative diseases [94,95,96,97].

5. Regulation of Signaling Pathways in Ischemia and Reperfusion Injury of Nerve Cells Involving microRNAs

In in vitro models, it has been shown that the activation of miR-496 and miR-874-3p reduces the consequences of ischemic–reperfusion injury in neurons by negatively regulating BCL2L14 and BMF/BCL2L13, respectively [98,99]. Activation of miR-92b-3p expression inhibits apoptosis, mitochondrial dysfunction, and inflammation through the inhibition of TRAF3 [100]. MiR-182-5p and miR-193b-3p also exert neuroprotective effects in cerebral ischemia–reperfusion injury by negatively regulating the inflammatory response mediated by Toll-like receptor 4 and 5-lipoxygenase, respectively [101,102]. Additionally, miR-30c acts by inhibiting neuronal apoptosis in ischemia–reperfusion injury of the brain, thus suppressing the expression of SOX9/MAPK [103,104], while miR-449a downregulates the expression of amphiregulin (AREG) [105].
Similar effects are observed with the inhibition of miR-19a-3p, which reduces the extent and area of cerebral ischemia–reperfusion injury by regulating inflammation and apoptosis through increased expression of IGFBP3 both in vivo and in vitro [106]. It has also been discovered that inhibiting exosomal miR-200a-3p/141-3p, which originates from astrocytes and targets the gene SIRT1 and its associated signaling pathways, reduces the pathological consequences of cerebral ischemia–reperfusion injury in a mouse model [107]. MiR-370 has been shown to exacerbate neuronal reperfusion injury by impacting the expression of SIRT6 and the regulation mechanism of the Nrf2/ARE signaling pathway [108], while the exosomal form of this microRNA (370-3p) increases blood–brain barrier permeability during ischemia–reperfusion injury through the interference of MAPK1 [109].
Long non-coding RNA MEG3, by binding to miR-485, promotes the exacerbation of cerebral ischemia–reperfusion injury through potentiating pyroptosis via AIM2 [110]. Additionally, the knockdown of miR-155-5p acts by inhibiting the pyroptosis mechanism and inflammation through interference with DUSP14 [111,112]. It should be noted that Zhang L and colleagues [113], without emphasizing the orientation of miR-155, also demonstrated its activating influence on apoptosis and inflammatory processes in neural tissue.
In addition to the mentioned studies, there are several other works concerning the role of microRNAs in the pathogenesis of ischemia–reperfusion injury and the signaling pathways through which their effects are realized during RNA interference (Table 1, Figure 2). The majority of neuroprotective mechanisms affected by the overexpression or inhibition of microRNAs involve anti-apoptotic, anti-inflammatory, and antioxidant signaling pathways.
Most studies from various researchers have demonstrated unidirectional effectiveness regarding the neuroprotective role of miR-10b-3p [116,117], miR-124-3p [119,120], miR-141-3p [107,131], miR-153-3p [141,142], miR-182-5p [101,150], miR-20a-3p [162,163], miR-22-3p [171,172], miR-24-3p [175,176], miR-27a-3p [181,182,183], miR-342-5p [191,192], miR-455-3p [200,201], miR-488-3p [203,204], and miR-92b-3p [100,221] as well as the damaging role of miR-141-3p [107,131], miR-155-5p [111,112], and miR-182 [148,149]. Therefore, these specific microRNAs should be considered as priority targets for further translation into clinical practice.
Indeed, along with the discovery of unidirectional effects of microRNA expression, contradictory results concerning the same molecules are also encountered. For example, according to the findings presented by Jia T et al. [205], activation of miR-489-3p expression in in vivo and in vitro models reduces the pathological consequences of cerebral ischemia–reperfusion by inhibiting histone deacetylase 2 (HDAC2), thereby reducing apoptosis intensity and enhancing cell viability. In contrast, Song L et al. [206] obtained opposing results: in mice subjected to temporary middle cerebral artery occlusion, an intensification of oxidative stress and neuron apoptosis was observed under conditions of increased miR-489-3p levels, which inhibits Sirtuin1 (SIRT1).
In the study investigating the effects of miR-421-3p in cerebral ischemia–reperfusion [199], both in vivo and in vitro models showed that increased expression of miR-421-3p reduces the intensity of inflammation through the YTHDF1/NF-κB p65 signaling pathway. On the contrary, Xu J. et al. [198] found a reverse effect of this microRNA concerning the intensity of apoptosis and inflammation in models of ischemia–reperfusion nerve tissue damage, mediated through the myocyte enhancer factor 2C (MEF2C).
In the study by Wei XY et al. [170], positive effects of the long non-coding RNA (lncRNA) RPL34-AS1 were observed in patients with stroke as well as in cerebral ischemia in rats and in an OGD cell model. This molecule inhibits miR-223-3p, which targets the insulin-like growth factor 1 receptor (IGF1R). The authors attribute the positive effects of RPL34-AS1 to its influence on this specific mechanism. However, there are contradictory results regarding the effects of miR-223-3p in studies focusing on the circular RNA circPDS5B and its impact on angiogenesis [168] and regarding the positive impact of miR-223-3p expression on the development of the inflammatory response in the zone of ischemia–reperfusion and its surrounding area [169].
The evaluation of the effects of miR-19a and its sense form miR-19a-3p shows a negative role in the mechanism of cerebral ischemia–reperfusion injury played by excessive stimulation of apoptosis and inflammation [106,156,157]. Similar observations are confirmed for the structurally related miR-19b-3p, which also intensifies the neuroinflammatory process in the ischemia–reperfusion zone [157]. However, these observations are contradicted by data showing that knockdown of the long non-coding RNA H19 and overexpression of miR-19a-3p attenuated the severity of ischemia–reperfusion-induced oxidative stress and apoptosis in neurons, as reported by Gao N. et al. [158].
The study of the effects of the lncRNA Malat1 revealed a positive influence of miR-26b [179] and a negative impact of miR-142-3p expression [132] on apoptosis and cell proliferation during experimental brain hypoxia–ischemia. However, the results from Li J. et al. [133] show an opposite effect when miR-142-3p expression is enhanced: inhibiting FBXO3. It is important to note that this study has limitations as it was conducted only on an in vitro model.
In two studies investigating the role of miR-140-3p on in vitro models with OGD, opposite effects were demonstrated: on the PC12 cell line, an enhancement of miR-140-3p expression showed a weakening effect on apoptosis and oxidative stress [130], while on the N2a cell line, an overexpression of miR-140-3p potentiated apoptosis and oxidative stress [129]. Supporting the greater significance of the first study and the neuroprotective role of this molecule is the fact that the in vitro results obtained by Zhang Y. et al. were replicated in the study on patients with ischemic stroke.
Another contradiction in determining the role of miRNAs in the pathogenesis of ischemia–reperfusion brain injury is the data on miR-128-3p. They indicate its proapoptotic efficacy in in vitro and in vivo models by inhibiting the FOXO/Relaxin signaling pathway [122] as well as the neuroprotective efficacy (potentiation of differentiation and myelination processes) of this miRNA in exosomal localization in an in vivo experiment [123].
Moreover, the studies mentioned above demonstrate that miR-135a, miR-181c, and miR-211-5p, whose expression plays a positive role in protection against cerebral ischemia–reperfusion injury, also act as neuroprotective agents in Alzheimer’s disease [74]. Conversely, the overexpression of miR-155-5p, miR-200a-3p, and miR-206 leads to damage to neural tissue both in the context of ischemia–reperfusion and Alzheimer’s-like neurodegeneration [74].

6. Peculiarities of Epigenetic Regulation in BBB Dysfunction

Modern data underscore the important role of blood–brain barrier (BBB) dysfunction in the development of various neurological pathologies [222]. Disruption of the integrity and permeability of the BBB is a significant element in the pathogenesis of hypoxic–ischemic and infectious brain injuries [223,224,225]. There is also evidence showing specific features of BBB functioning in developmental brain disorders and in autoimmune and neurodegenerative pathologies [226,227,228,229].
Increased permeability of the blood–brain barrier (BBB) in conditions of chronic brain hypoxia–ischemia during the peri- and neonatal periods may be one of the contributing factors to the progression of neurodegenerative processes. This can occur due to the entry of barrier-breaking antigens into the peripheral bloodstream, leading to subsequent immune responses and dysfunction of proteolytic and neurotransmitter systems [230]. There is compelling experimental evidence of a disrupted tight junction structure and dysregulation of BBB transporter proteins in Alzheimer’s disease and Parkinson’s disease [228,231].
Epigenetic factors responsible for DNA methylation and histone remodeling are crucial for normal ontogenetic development of the brain and its barrier structures [88,232]. Among these factors are small non-coding RNA molecules [233]. They are susceptible to reprogramming under the influence of hypoxic stimuli, which can result in both damage and preconditioning effects [234,235]. If the stimulus is excessive and/or prolonged, it can lead to the development of hypoxic–ischemic brain injury and contribute to future neurological disorders.
It has been shown that blocking histone deacetylase in microglia culture exerts a protective effect on oligodendrocytes after experimental hypoxia–ischemia [236]. Moreover, the M1 fraction suppresses the activity of oligodendrocytes, while the M2 fraction enhances it and allows for reducing the expression of pro-inflammatory factors and increasing the expression of anti-inflammatory cytokines. Methylation and demethylation processes involve a large number of enzymes, including histone acetyltransferases, deacetylases, histone methyltransferases, and demethylases [237]. These processes contribute to the modulation of transcription in response to endogenous and exogenous factors such as hypoxia and inflammation [238]. In in vivo models, it has been demonstrated that experimental hypoxia–ischemia in the neonatal period leads to increased expression of caspase-3, decreased expression of synapsin, and inhibition of histone methylation H3K4me2/-me3 and H3K27me2/-me3, resulting in the induction of neuronal apoptosis in the hippocampus [239].
Hypoxia–ischemia disrupts the structure and organization of the components of the blood–brain barrier (BBB), leading to increased permeability of the cerebral endothelium [240]. Degradation of BBB structures has been shown to increase the risk of early stroke, while blocking BBB disruption provides protection to the brain parenchyma [241]. The degradation of tight junction proteins or cell adhesion proteins under the influence of metalloproteinases is directly regulated by microRNAs [242]. Specifically, miR-539, miR-132, miR-21, and miR-671-5p inhibit the expression of MMP-9 and protect endothelial cells during experimental ischemia, preventing increased BBB permeability [243,244,245,246]. Bai Y and colleagues [247] found that circular RNA DLGAP4 positively regulates endothelial–mesenchymal transition, associated with BBB integrity, through miR-143 during ischemic stroke. Interestingly, the positive effect in the form of apoptosis suppression upon the inhibition of miR-143 expression is also observed in the ischemia–reperfusion injury model [135]. Increased adhesion of immune cells to the damaged endothelium contributes to their infiltration into surrounding tissue. Certain microRNAs can help preserve BBB integrity by reducing immune cell adhesion and the expression of pro-inflammatory cytokines [235]. For example, miR-210 activates the expression of TNF-α, IL-1β, IL-6, and chemokine ligands CCL1 and CCL2, which are linked to the pro-inflammatory response in a mouse model of ischemic stroke [248].
It is worth noting that miR-210 is recognized as one of the key regulators of neonatal hypoxia–ischemia [249]; its overexpression, along with that of miR-130a, increases BBB permeability by inhibiting the expression of occludin and β-catenin [250,251]. Overexpression of miR-126-3p/-5p and miR-98 in the ischemic brain of mice suppresses the effects of pro-inflammatory cytokines and adhesion molecules, preserving the integrity of the cerebral endothelium and thus reducing the negative consequences of stroke [252,253,254]. A similar effect is observed with overexpression of miR-152-3p, which reduces the degree of neurological deficit and disruption of BBB integrity in experimental ischemia by inhibiting apoptosis of endothelial cells through negative modulation of the MAP3K2/JNK/c-Jun pathway [255].
On the contrary, miR-34a is activated in endothelial cells during episodes of acute hypoxia–ischemia, which negatively affects mitochondrial function in these cells by targeting cytochrome C [256]. Moreover, knockout of miR-34a reduces BBB permeability, attenuates disruptions of tight intercellular contacts, improves stroke outcomes [242], and improves cognitive functions in mice with neurodegeneration [74].
MiR-15a is also activated in the cerebral endothelium of mice after acute oxygen–glucose deprivation, leading to cell death. This process is inhibited when the δ-receptor is activated by the peroxisome proliferator-activated receptor (PPAR), which acts as a potential neuroprotective factor during ischemic stroke [257]. The main role in maintaining BBB permeability belongs to the endothelium and, to a lesser extent, pericytes. It has been shown that during hypoxic–ischemic injury, pericytes detach from the basal membrane, contributing to increased BBB permeability. This process is induced by miR-149-5p [258].
There is also evidence of the involvement of other microRNAs in regulating the structure and functions of the cerebral endothelium. For instance, the upregulation of miR-150-5p leads to the negative regulation of the Vezf1 protein synthesis, resulting in dysfunction of the endothelial cells in the cerebral cortex’s blood vessels [259].
Particular attention should be given to the molecule miR-132, the expression of which not only exerts a protective effect on the blood–brain barrier [244] but also reduces oxidative stress and autophagy during cerebral ischemia–reperfusion injury [124] and possesses neuroprotective properties in Alzheimer’s disease [74].

7. Conclusions

The materials reviewed in this study indicate the significant importance of epigenetic regulation, including the involvement of microRNAs, in the pathogenesis of blood–brain barrier dysfunction, neurodegenerative disorders, and ischemia–reperfusion injuries of the nervous tissue. This has generated increasing interest in advancing technologies for studying the properties of these molecules and evaluating the prospects for translating this knowledge into practical medicine for predicting neurological pathologies, their differential diagnosis, and the search for targets for targeted therapies.

Author Contributions

Conceptualization, P.P.T. and A.B.S.; methodology, P.P.T. and I.I.; validation, P.P.T., I.I. and I.A.P.; writing—original draft preparation, P.P.T. and A.S.A.; writing—review and editing, P.P.T., I.I. and Z.S.M.; visualization, P.P.T.; project administration, P.P.T. and P.F.L.; funding acquisition, P.P.T. and A.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This publication has been supported by RUDN University project № 210500-2-000: ‘Molecular mechanisms of mast cell participation in the formation of the immune and stromal landscape of a specific tissue microenvironment in normal and pathological conditions (oncogenesis, inflammation, fibrosis)’.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicting financial interests or commercial considerations related to the issues presented herein.

Abbreviations

AIM2Interferon-inducible protein AIM2 (absent in melanoma 2)
AREGAmphiregulin
BCL2L14Apoptosis facilitator Bcl-2-like protein 14 is a protein that, in humans, is encoded by the BCL2L14 gene
BCL2L13A protein which, in humans, is encoded by the BCL2L13 gene on chromosome 22
BBBBlood–brain barrier
CCL1Chemokine (C-C motif) ligand 1
DNADeoxyribonucleic acid
DLGAP4Disks large-associated protein 4 (DAP-4)
DUSP14Dual-specificity Phosphatase 14
FOXOForkhead box protein O1 (FOXO1), also known as forkhead in rhabdomyosarcoma (FKHR)
FBXO3F-box protein 3
GTPGuanosine triphosphate
HDAC2Histone deacetylase 2
HSP70The 70-kilodalton heat shock proteins
HSP90The 90-kilodalton heat shock proteins
Il-1βInterleukin-1 beta
Il-6Interleukin 6
IGF1RInsulin-like growth factor 1 receptor
IGFBP3Insulin-like growth factor-binding protein 3
LNALocked nucleic acid
MAPK 1Mitogen-activated protein kinase 1
MEG3Maternally expressed 3 imprinted long non-coding RNA gene
MEF2CMyocyte enhancer factor 2C
MMP9Matrix metallopeptidase 9
MAP3K2Mitogen-activated protein Kinase 2
mRNAMessenger RNA
Nrf2/ARENuclear factor erythroid 2-related factor 2
OGDOxygen–glucose deprivation
PCRPolymerase chain reaction
qRT-PCRQuantitative reverse transcription polymerase chain reaction
RNA Ribonucleic acid
RPL34-AS1Ribosomal protein L34
RISC/microRNPRNA-induced silencing complex/miR–ribonucleoprotein complex
RNaseRibonuclease
SOX9SRY (sex-determining region Y)-Box
SIRT1(Silent mating type information regulation 2 homolog) 1
SIRT6Stress-responsive protein deacetylase and mono-ADP ribosyl transferase enzyme encoded by the SIRT6 gene
siRNAsSmall interfering RNAs
TRAF3TNF receptor-associated factor
TRBPTransactivation response RNA binding protein
TNF-αTumor necrosis factor
YTHDF1YTH domain family; member 1 is a protein that, in humans, is encoded by the YTHDF1 gene

References

  1. Lu, T.X.; Rothenberg, M.E. MicroRNA. J. Allergy Clin. Immunol. 2018, 141, 1202–1207. [Google Scholar] [CrossRef]
  2. Morselli, M.; Dieci, G. Epigenetic regulation of human non-coding RNA gene transcription. Biochem. Soc. Trans. 2022, 50, 723–736. [Google Scholar] [CrossRef] [PubMed]
  3. Mohr, A.M.; Mott, J.L. Overview of microRNA biology. Semin. Liver Dis. 2015, 35, 3–11. [Google Scholar] [CrossRef]
  4. Nachtigall, P.G.; Bovolenta, L.A. Computational Detection of MicroRNA Targets. Methods Mol. Biol. 2022, 2257, 187–209. [Google Scholar] [PubMed]
  5. Ho, P.T.B.; Clark, I.M.; Le, L.T.T. MicroRNA-Based Diagnosis and Therapy. Int. J. Mol. Sci. 2022, 23, 7167. [Google Scholar] [CrossRef] [PubMed]
  6. 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] [PubMed]
  7. Marques, S.; Outeiro, T.F. Epigenetics in Parkinson’s and Alzheimer’s diseases. Epigenetics Dev. Dis. 2013, 61, 507–525. [Google Scholar]
  8. Neal, M.; Richardson, J.R. Epigenetic regulation of astrocyte function in neuroinflammation and neurodegeneration. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 432–443. [Google Scholar] [CrossRef]
  9. Ghosh, P.; Saadat, A. Neurodegeneration and epigenetics: A review. Neurologia 2023, 38, e62–e68. [Google Scholar] [CrossRef]
  10. Aigner, A. Applications of RNA interference: Current state and prospects for siRNA-based strategies in vivo. Appl. Microbiol. Biotechnol. 2007, 76, 9–21. [Google Scholar] [CrossRef]
  11. Aksenenko, M.B.; Komina, A.V.; Palkina, N.V.; Averchuk, A.S.; Rybnikov, Y.A.; Dyhno, Y.A.; Ruksha, T.G. Transcriptomic analysis of melanoma cells extracted from different sites of the primary tumor. Sib. J. Oncol. 2018, 17, 59–66. [Google Scholar] [CrossRef]
  12. Kurakula, H.; Vaishnavi, S.; Sharif, M.Y.; Ellipilli, S. Emergence of Small Interfering RNA-Based Gene Drugs for Various Diseases. ACS Omega 2023, 8, 20234–20250. [Google Scholar] [CrossRef] [PubMed]
  13. Lee, Y.; Kim, M.; Han, J.; Yeom, K.H.; Lee, S.; Baek, S.H.; Kim, V.N. MicroRNA genes are transcribed by RNA polymerase II. EMBO J. 2004, 23, 4051–4060. [Google Scholar] [CrossRef] [PubMed]
  14. Mraz, M.; Dolezalova, D.; Plevova, K.; Stano Kozubik, K.; Mayerova, V.; Cerna, K.; Musilova, K.; Tichy, B.; Pavlova, S.; Borsky, M.; et al. MicroRNA-650 expression is influenced by immunoglobulin gene rearrangement and affects the biology of chronic lymphocytic leukemia. Blood 2012, 119, 2110–2113. [Google Scholar] [CrossRef] [PubMed]
  15. Gatti da Silva, G.H.; Coltri, P.P. A Reporter Assay to Analyze Intronic microRNA Maturation in Mammalian Cells. J. Vis. Exp. 2022, 184, e63498. [Google Scholar]
  16. Wong, A.C.H.; Rasko, J.E.J. Splice and Dice: Intronic microRNAs, Splicing and Cancer. Biomedicines 2021, 9, 1268. [Google Scholar] [CrossRef]
  17. Wakiyama, M.; Ogami, K.; Iwaoka, R.; Aoki, K.; Hoshino, S.I. MicroRNP-mediated translational activation of nonadenylated mRNAs in a mammalian cell-free system. Genes Cells 2018, 23, 332–344. [Google Scholar] [CrossRef]
  18. Nakanishi, K. Anatomy of RISC: How do small RNAs and chaperones activate Argonaute proteins? Wiley Interdiscip. Rev. RNA 2016, 7, 637–660. [Google Scholar] [CrossRef]
  19. Iwakawa, H.O.; Tomari, Y. Life of RISC: Formation, action, and degradation of RNA-induced silencing complex. Mol. Cell 2022, 82, 30–43. [Google Scholar] [CrossRef]
  20. Jin, S.; Zhan, J.; Zhou, Y. Argonaute proteins: Structures and their endonuclease activity. Mol. Biol. Rep. 2021, 48, 4837–4849. [Google Scholar] [CrossRef]
  21. Kobayashi, H.; Tomari, Y. RISC assembly: Coordination between small RNAs and Argonaute proteins. Biochim. Biophys. Acta 2016, 1859, 71–81. [Google Scholar] [CrossRef]
  22. Bibel, B.; Elkayam, E.; Silletti, S.; Komives, E.A.; Joshua-Tor, L. Target binding triggers hierarchical phosphorylation of human Argonaute-2 to promote target release. Elife 2022, 11, e76908. [Google Scholar] [CrossRef]
  23. Warnement, C.M.; Cismowski, M.J.; Rogers, L.K. Optimizing miR-29 measurements in biobanked, heparinized samples. Life Sci. 2019, 238, 116894. [Google Scholar] [CrossRef] [PubMed]
  24. Ingle, R.G.; Fang, W.J. An Overview of the Stability and Delivery Challenges of Commercial Nucleic Acid Therapeutics. Pharmaceutics 2023, 15, 1158. [Google Scholar] [CrossRef] [PubMed]
  25. Low, S.S.; Ji, D.; Chai, W.S.; Liu, J.; Khoo, K.S.; Salmanpour, S.; Karimi, F.; Deepanraj, B.; Show, P.L. Recent Progress in Nanomaterials Modified Electrochemical Biosensors for the Detection of MicroRNA. Micromachines. Micromachines 2021, 12, 1409. [Google Scholar] [CrossRef] [PubMed]
  26. Torres, A.G.; Fabani, M.M.; Vigorito, E.; Gait, M.J. MicroRNA fate upon targeting with anti-miRNA oligonucleotides as revealed by an improved Northern-blot-based method for miRNA detection. RNA 2011, 17, 933–943. [Google Scholar] [CrossRef] [PubMed]
  27. Tian, T.; Wang, J.; Zhou, X. A review: microRNA detection methods. Org. Biomol. Chem. 2015, 13, 2226–2238. [Google Scholar] [CrossRef] [PubMed]
  28. Trinh, M.P.; Carballo, J.G.; Adkins, G.B.; Guo, K.; Zhong, W. Physical and chemical template-blocking strategies in the exponential amplification reaction of circulating microRNAs. Anal. Bioanal. Chem. 2020, 412, 2399–2412. [Google Scholar] [CrossRef]
  29. de Ferronato, G.A.; Cerezetti, M.B.; Bridi, A.; Prado, C.M.; Dos Santos, G.; Bastos, N.M.; da Rosa, P.M.S.; Ferst, J.G.; da Silveira, J.C. MicroRNA Profiling Using a PCR-Based Method. Methods Mol. Biol. 2023, 2595, 159–170. [Google Scholar]
  30. Komori, M.; Komiya, K.; Shirakawa, T.; Morikawa, T.J.; Yoshimura, T. Measurement of microRNA with isothermal DNA amplification on fully automated immunoassay analyzers. Anal. Bioanal. Chem. 2019, 411, 3789–3800. [Google Scholar] [CrossRef]
  31. Liu, H.; Li, L.; Duan, L. High specific and ultrasensitive isothermal detection of microRNA by padlock probe-based exponential rolling circle amplification. Anal. Chem. 2013, 85, 7941–7947. [Google Scholar] [CrossRef] [PubMed]
  32. Deng, R.; Tang, L.; Tian, Q.; Wang, Y.; Lin, L.; Li, J. Toehold-initiated rolling circle amplification for visualizing individual microRNAs in situ in single cells. Angew. Chem. 2014, 126, 2421–2425. [Google Scholar] [CrossRef]
  33. Aziz, N.B.; Mahmudunnabi, R.G.; Umer, M.; Sharma, S.; Rashid, M.A.; Alhamhoom, Y.; Shim, Y.B.; Salomon, C.; Shiddiky, M.J.A. MicroRNAs in ovarian cancer and recent advances in the development of microRNA-based biosensors. Analyst 2020, 145, 2038–2057. [Google Scholar] [CrossRef] [PubMed]
  34. Martino, S.; Tammaro, C.; Misso, G.; Falco, M.; Scrima, M.; Bocchetti, M.; Rea, I.; De Stefano, L.; Caraglia, M. microRNA Detection via Nanostructured Biochips for Early Cancer Diagnostics. Int. J. Mol. Sci. 2023, 24, 7762. [Google Scholar] [CrossRef] [PubMed]
  35. Withanage, M.H.H.; Liang, H.; Zeng, E. RNA-Seq Experiment and Data Analysis. Methods Mol. Biol. 2022, 2418, 405–424. [Google Scholar]
  36. Yaylak, B.; Akgül, B. Experimental MicroRNA Detection Methods. Methods Mol. Biol. 2022, 2257, 33–55. [Google Scholar]
  37. Takegawa-Araki, T.; Yasukawa, K.; Iwazaki, N.; Utsugi, M.; Furukawa, H.; Kumagai, S.; Sawamoto, H.; Obika, S. 2’-N-Alkylaminocarbonyl-2’-amino-LNA: Synthesis, duplex stability, nuclease resistance, and in vitro anti-microRNA activity. Bioorg. Med. Chem. 2023, 78, 117148. [Google Scholar] [CrossRef]
  38. Langner, H.K.; Jastrzebska, K.; Caruthers, M.H. Synthesis and Characterization of Thiophosphoramidate Morpholino Oligonucleotides and Chimeras. J. Am. Chem. Soc. 2020, 142, 16240–16253. [Google Scholar] [CrossRef]
  39. Javanmard, S.H.; Vaseghi, G.; Ghasemi, A.; Rafiee, L.; Ferns, G.A.; Esfahani, H.N.; Nedaeinia, R. Therapeutic inhibition of microRNA-21 (miR-21) using locked-nucleic acid (LNA)-anti-miR and its effects on the biological behaviors of melanoma cancer cells in preclinical studies. Cancer Cell Int. 2020, 20, 384. [Google Scholar] [CrossRef]
  40. Smith, C.I.E.; Zain, R. Therapeutic Oligonucleotides: State of the Art. Annu. Rev. Pharmacol. Toxicol. 2019, 59, 605–630. [Google Scholar] [CrossRef]
  41. Bhattarai, U.; Hsieh, W.C.; Yan, H.; Guo, Z.F.; Shaikh, A.Y.; Soltani, A.; Song, Y.; Ly, D.H.; Liang, F.S. Bifunctional small molecule-oligonucleotide hybrid as microRNA inhibitor. Bioorg. Med. Chem. 2020, 28, 115394. [Google Scholar] [CrossRef] [PubMed]
  42. Graham, H.; Chandler, D.J.; Dunbar, S.A. The genesis and evolution of bead-based multiplexing. Methods 2019, 158, 2–11. [Google Scholar] [CrossRef] [PubMed]
  43. Cai, S.; Pataillot-Meakin, T.; Shibakawa, A.; Ren, R.; Bevan, C.L.; Ladame, S.; Ivanov, A.P.; Edel, J.B. Single-molecule amplification-free multiplexed detection of circulating microRNA cancer biomarkers from serum. Nat. Commun. 2021, 12, 3515. [Google Scholar] [CrossRef] [PubMed]
  44. Xu, Y.; Zhang, P.; Yi, F.; Wei, J.; Du, Q. Rainbowarray Microsphere-Based Gene Detection Assay. SLAS Technol. 2020, 25, 258–266. [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. Ohtsuka, M.; Iwamoto, K.; Naito, A.; Imasato, M.; Hyuga, S.; Nakahara, Y.; Mikamori, M.; Furukawa, K.; Moon, J.; Asaoka, T.; et al. Circulating MicroRNAs in Gastrointestinal Cancer. Cancers 2021, 13, 3348. [Google Scholar] [CrossRef]
  47. Ionescu, R.F.; Cretoiu, S.M. MicroRNAs as monitoring markers for right-sided heart failure and congestive hepatopathy. J. Med. Life 2021, 14, 142–147. [Google Scholar] [CrossRef] [PubMed]
  48. Wu, Y.; Li, Q.; Zhang, R.; Dai, X.; Chen, W.; Xing, D. Circulating microRNAs: Biomarkers of disease. Clin. Chim. Acta 2021, 516, 46–54. [Google Scholar] [CrossRef]
  49. Pozniak, T.; Shcharbin, D.; Bryszewska, M. Circulating microRNAs in Medicine. Int. J. Mol. Sci. 2022, 23, 3996. [Google Scholar] [CrossRef]
  50. Metintaş, M.; Ak, G.; Özbayer, C.; Boğar, F.; Metintaş, S. Serum Expression Levels of Certain miRNAs in Predicting Diagnosis, Prognosis, and Response to Chemotherapy in Malignant Pleural Mesothelioma. Balkan Med. J. 2022, 39, 246–254. [Google Scholar] [CrossRef]
  51. Takizawa, S.; Matsuzaki, J.; Ochiya, T. Circulating microRNAs: Challenges with their use as liquid biopsy biomarkers. Cancer Biomark. 2022, 35, 1–9. [Google Scholar] [CrossRef] [PubMed]
  52. Franco, S.; Pluvinet, R.; Sanchez-Herrero, J.F.; Sumoy, L.; Martinez, M.A. Rapid and accurate quantification of isomiRs by RT-qPCR. Sci. Rep. 2022, 12, 17220. [Google Scholar] [CrossRef] [PubMed]
  53. Deng, R.; Zhang, K.; Li, J. Isothermal Amplification for MicroRNA Detection: From the Test Tube to the Cell. Acc. Chem. Res. 2017, 50, 1059–1068. [Google Scholar] [CrossRef] [PubMed]
  54. Gines, G.; Menezes, R.; Xiao, W.; Rondelez, Y.; Taly, V. Emerging isothermal amplification technologies for microRNA biosensing: Applications to liquid biopsies. Mol. Asp. Med. 2020, 72, 100832. [Google Scholar] [CrossRef] [PubMed]
  55. Ye, J.; Xu, M.; Tian, X.; Cai, S.; Zeng, S. Research advances in the detection of miRNA. J. Pharm. Anal. 2019, 9, 217–226. [Google Scholar] [CrossRef]
  56. Beta, M.; Krishnakumar, S.; Elchuri, S.V.; Salim, B.; Narayanan, J. A comparative fluorescent beacon-based method for serum microRNA quantification. Anal. Sci. 2015, 31, 231–235. [Google Scholar] [CrossRef]
  57. Chan, H.N.; Ho, S.L.; He, D.; Li, H.W. Direct and sensitive detection of circulating miRNA in human serum by ligase-mediated amplification. Talanta 2020, 206, 120217. [Google Scholar] [CrossRef]
  58. Masud, M.K.; Umer, M.; Hossain, M.S.A.; Yamauchi, Y.; Nguyen, N.T.; Shiddiky, M.J.A. Nanoarchitecture Frameworks for Electrochemical MiRNA Detection. Trends Biochem. Sci. 2019, 44, 433–452. [Google Scholar] [CrossRef]
  59. Vilímová, I.; Chourpa, I.; David, S.; Soucé, M.; Hervé-Aubert, K. Two-step formulation of magnetic nanoprobes for microRNA capture. RSC Adv. 2022, 12, 7179–7188. [Google Scholar] [CrossRef]
  60. Wu, Y.; Huang, J.; Yang, X.; Yang, Y.; Quan, K.; Xie, N.; Li, J.; Ma, C.; Wang, K. Gold Nanoparticle Loaded Split-DNAzyme Probe for Amplified miRNA Detection in Living Cells. Anal. Chem. 2017, 89, 8377–8383. [Google Scholar] [CrossRef]
  61. Remsburg, C.; Konrad, K.; Sampilo, N.F.; Song, J.L. Analysis of microRNA functions. Methods Cell Biol. 2019, 151, 323–334. [Google Scholar]
  62. Li, Y.; Zhang, Z. Computational Biology in microRNA. Wiley Interdiscip. Rev. RNA 2015, 6, 435–452. [Google Scholar] [CrossRef] [PubMed]
  63. de Amorim, T.S.; Pedro, D.L.F.; Paschoal, A.R. MicroRNA Databases and Tools. Methods Mol. Biol. 2022, 2257, 131–166. [Google Scholar]
  64. Tomasello, L.; Cluts, L.; Croce, C.M. Experimental Validation of MicroRNA Targets: Mutagenesis of Binding Regions. Methods Mol. Biol. 2019, 1970, 331–339. [Google Scholar] [PubMed]
  65. Yu, D.; Chen, S.; Li, D.; Knox, B.; Guo, L.; Ning, B. FREMSA: A Method That Provides Direct Evidence of the Interaction between microRNA and mRNA. Methods Mol. Biol. 2020, 2102, 557–566. [Google Scholar] [PubMed]
  66. Mortazavi, S.S.; Bahmanpour, Z.; Daneshmandpour, Y.; Roudbari, F.; Sheervalilou, R.; Kazeminasab, S.; Emamalizadeh, B. An updated overview and classification of bioinformatics tools for MicroRNA analysis, which one to choose? Comput. Biol. Med. 2021, 134, 104544. [Google Scholar] [CrossRef] [PubMed]
  67. Ragini Sakshi Mani, I.; Singh, V. Applications of bioinformatics in epigenetics. Prog. Mol. Biol. Transl. Sci. 2023, 198, 1–13. [Google Scholar]
  68. Takousis, P.; Sadlon, A.; Schulz, J.; Wohlers, I.; Dobricic, V.; Middleton, L.; Lill, C.M.; Perneczky, R.; Bertram, L. Differential expression of microRNAs in Alzheimer’s disease brain, blood, and cerebrospinal fluid. Alzheimer’s Dement. 2019, 15, 1468–1477. [Google Scholar] [CrossRef]
  69. Swarbrick, S.; Wragg, N.; Ghosh, S.; Stolzing, A. Systematic Review of miRNA as Biomarkers in Alzheimer’s Disease. Mol. Neurobiol. 2019, 56, 6156–6167. [Google Scholar] [CrossRef]
  70. Nunomura, A.; Perry, G. RNA and Oxidative Stress in Alzheimer’s Disease: Focus on microRNAs. Oxidative Med. Cell. Longev. 2020, 2020, 2638130. [Google Scholar] [CrossRef]
  71. Walgrave, H.; Zhou, L.; De Strooper, B.; Salta, E. The promise of microRNA-based therapies in Alzheimer’s disease: Challenges and perspectives. Mol. Neurodegener. 2021, 16, 76. [Google Scholar] [PubMed]
  72. Mayo, S.; Benito-León, J.; Peña-Bautista, C.; Baquero, M.; Cháfer-Pericás, C. Recent Evidence in Epigenomics and Proteomics Biomarkers for Early and Minimally Invasive Diagnosis of Alzheimer’s and Parkinson’s Diseases. Curr. Neuropharmacol. 2021, 19, 1273–1303. [Google Scholar] [CrossRef]
  73. Liu, S.; Fan, M.; Zheng, Q.; Hao, S.; Yang, L.; Xia, Q.; Qi, C.; Ge, J. MicroRNAs in Alzheimer’s disease: Potential diagnostic markers and therapeutic targets. Biomed. Pharmacother. 2022, 148, 112681. [Google Scholar] [CrossRef] [PubMed]
  74. Silvestro, S.; Bramanti, P.; Mazzon, E. Role of miRNAs in Alzheimer’s Disease and Possible Fields of Application. Int. J. Mol. Sci. 2019, 20, 3979. [Google Scholar] [CrossRef] [PubMed]
  75. Yuen, S.C.; Liang, X.; Zhu, H.; Jia, Y.; Leung, S.W. Prediction of differentially expressed microRNAs in blood as potential biomarkers for Alzheimer’s disease by meta-analysis and adaptive boosting ensemble learning. Alzheimer’s Res. Ther. 2021, 13, 126. [Google Scholar]
  76. Chang, C.H.; Lin, C.H.; Lane, H.Y. Machine Learning and Novel Biomarkers for the Diagnosis of Alzheimer’s Disease. Int. J. Mol. Sci. 2021, 22, 2761. [Google Scholar] [CrossRef]
  77. Xu, A.; Kouznetsova, V.L.; Tsigelny, I.F. Alzheimer’s Disease Diagnostics Using MiRNA Biomarkers and Machine Learning. J. Alzheimer’s Dis. 2022, 86, 841–859. [Google Scholar] [CrossRef]
  78. Shokhirev, M.N.; Johnson, A.A. An integrative machine-learning meta-analysis of high-throughput omics data identifies age-specific hallmarks of Alzheimer’s disease. Ageing Res. Rev. 2022, 81, 101721. [Google Scholar] [CrossRef]
  79. Chiricosta, L.; D’Angiolini, S.; Gugliandolo, A.; Mazzon, E. Artificial Intelligence Predictor for Alzheimer’s Disease Trained on Blood Transcriptome: The Role of Oxidative Stress. Int. J. Mol. Sci. 2022, 23, 5237. [Google Scholar] [CrossRef]
  80. Akhter, R. Circular RNA and Alzheimer’s Disease. Adv. Exp. Med. Biol. 2018, 1087, 239–243. [Google Scholar]
  81. Ma, N.; Tie, C.; Yu, B.; Zhang, W.; Wan, J. Identifying lncRNA-miRNA-mRNA networks to investigate Alzheimer’s disease pathogenesis and therapy strategy. Aging 2020, 12, 2897–2920. [Google Scholar] [CrossRef] [PubMed]
  82. Gowda, P.; Reddy, P.H.; Kumar, S. Deregulated mitochondrial microRNAs in Alzheimer’s disease: Focus on synapse and mitochondria. Ageing Res. Rev. 2022, 73, 101529. [Google Scholar] [PubMed]
  83. Su, L.; Li, R.; Zhang, Z.; Liu, J.; Du, J.; Wei, H. Identification of altered exosomal microRNAs and mRNAs in Alzheimer’s disease. Ageing Res. Rev. 2022, 73, 101497. [Google Scholar] [CrossRef]
  84. Garcia, G.; Pinto, S.; Ferreira, S.; Lopes, D.; Serrador, M.J.; Fernandes, A.; Vaz, A.R.; Mendonça, A.; Edenhofer, F.; Malm, T.; et al. Emerging Role of miR-21-5p in Neuron-Glia Dysregulation and Exosome Transfer Using Multiple Models of Alzheimer’s Disease. Cells 2022, 11, 3377. [Google Scholar] [CrossRef]
  85. Ryazanova, M.V.; Averchuk, A.S.; Novikova, S.V.; Salmina, A.B. Molecular mechanisms of angiogenesis: Brain is in the focus. Opera Med. Physiol. 2022, 9, 54–72. [Google Scholar]
  86. Tregub, P.P.; Averchuk, A.S.; Baranich, T.I.; Ryazanova, M.V.; Salmina, A.B. Physiological and Pathological Remodeling of Cerebral Microvessels. Int. J. Mol. Sci. 2022, 23, 12683. [Google Scholar] [CrossRef]
  87. Da Mesquita, S.; Louveau, A.; Vaccari, A.; Smirnov, I.; Cornelison, R.C.; Kingsmore, K.M.; Contarino, C.; Onengut-Gumuscu, S.; Farber, E.; Raper, D.; et al. Functional aspects of meningeal lymphatics in ageing and Alzheimer’s disease. Nature 2018, 560, 185–191. [Google Scholar] [CrossRef]
  88. Ihezie, S.A.; Mathew, I.E.; McBride, D.W.; Dienel, A.; Blackburn, S.L.; Thankamani Pandit, P.K. Epigenetics in blood-brain barrier disruption. Fluids Barriers CNS 2021, 18, 17. [Google Scholar] [PubMed]
  89. Neag, M.A.; Mitre, A.O.; Burlacu, C.C.; Inceu, A.I.; Mihu, C.; Melincovici, C.S.; Bichescu, M.; Buzoianu, A.D. miRNA Involvement in Cerebral Ischemia-Reperfusion Injury. Front. Neurosci. 2022, 16, 901360. [Google Scholar]
  90. Van den Hove, D.L.; Kompotis, K.; Lardenoije, R.; Kenis, G.; Mill, J.; Steinbusch, H.W.; Lesch, K.P.; Fitzsimons, C.P.; De Strooper, B.; Rutten, B.P. Epigenetically regulated microRNAs in Alzheimer’s disease. Neurobiol. Aging 2014, 35, 731–745. [Google Scholar] [CrossRef]
  91. Stoccoro, A.; Coppedè, F. Role of epigenetics in Alzheimer’s disease pathogenesis. Neurodegener. Dis. Manag. 2018, 8, 181–193. [Google Scholar] [CrossRef] [PubMed]
  92. Nikolac Perkovic, M.; Videtic Paska, A.; Konjevod, M.; Kouter, K.; Svob Strac, D.; Nedic Erjavec, G.; Pivac, N. Epigenetics of Alzheimer’s Disease. Biomolecules 2021, 11, 195. [Google Scholar] [CrossRef] [PubMed]
  93. Bryzgalov, L.O.; Korbolina, E.E.; Merkulova, T.I. Exploring the Genetic Predisposition to Epigenetic Changes in Alzheimer’s Disease. Int. J. Mol. Sci. 2023, 24, 7955. [Google Scholar] [CrossRef] [PubMed]
  94. Coneys, R.; Wood, I.C. Alzheimer’s disease: The potential of epigenetic treatments and current clinical candidates. Neurodegener. Dis. Manag. 2020, 10, 543–558. [Google Scholar] [CrossRef]
  95. Pandey, D.; Pal, T.; Sharma, A.; Flora, S. Potential Epigenetic Targets for Combating Alzheimer’s Disease. Mini Rev. Med. Chem. 2021, 21, 1527–1540. [Google Scholar] [CrossRef] [PubMed]
  96. Coppedè, F. Epigenetic regulation in Alzheimer’s disease: Is it a potential therapeutic target? Expert Opin. Ther. Targets 2021, 25, 283–298. [Google Scholar] [CrossRef] [PubMed]
  97. Hajjo, R.; Sabbah, D.A.; Abusara, O.H.; Al Bawab, A.Q. A Review of the Recent Advances in Alzheimer’s Disease Research and the Utilization of Network Biology Approaches for Prioritizing Diagnostics and Therapeutics. Diagnostics 2022, 12, 2975. [Google Scholar] [CrossRef]
  98. Yao, X.; Yao, R.; Yi, J.; Huang, F. Upregulation of miR-496 decreases cerebral ischemia/reperfusion injury by negatively regulating BCL2L14. Neurosci. Lett. 2019, 696, 197–205. [Google Scholar] [CrossRef]
  99. Jiang, D.; Sun, X.; Wang, S.; Man, H. Upregulation of miR-874-3p decreases cerebral ischemia/reperfusion injury by directly targeting BMF and BCL2L13. Biomed. Pharmacother. 2019, 117, 108941. [Google Scholar] [CrossRef]
  100. Liu, E.; Sun, H.; Wu, J.; Kuang, Y. MiR-92b-3p regulates oxygen and glucose deprivation-reperfusion-mediated apoptosis and inflammation by targeting TRAF3 in PC12 cells. Exp. Physiol. 2020, 105, 1792–1801. [Google Scholar] [CrossRef]
  101. Wang, J.; Xu, Z.; Chen, X.; Li, Y.; Chen, C.; Wang, C.; Zhu, J.; Wang, Z.; Chen, W.; Xiao, Z.; et al. MicroRNA-182-5p attenuates cerebral ischemia-reperfusion injury by targeting Toll-like receptor 4. Biochem. Biophys. Res. Commun. 2018, 505, 677–684. [Google Scholar] [CrossRef] [PubMed]
  102. Chen, Z.; Yang, J.; Zhong, J.; Luo, Y.; Du, W.; Hu, C.; Xia, H.; Li, Y.; Zhang, J.; Li, M.; et al. MicroRNA-193b-3p alleviates focal cerebral ischemia and reperfusion-induced injury in rats by inhibiting 5-lipoxygenase expression. Exp. Neurol. 2020, 327, 113223. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, M.; Zhu, Y.; Wei, M.; Liu, H. Neuroprotective effects of miR-30c on rats with cerebral ischemia/reperfusion injury by targeting SOX9. Pathol. Res. Pract. 2020, 216, 153271. [Google Scholar] [CrossRef] [PubMed]
  104. Zhang, H.; Li, M.; Liang, J.; Li, M.; Sun, X. Long Non-coding RNA PVT1 Inhibits miR-30c-5p to Upregulate Rock2 to Modulate Cerebral Ischemia/Reperfusion Injury Through MAPK Signaling Pathway Activation. Mol. Neurobiol. 2021, 58, 6032–6048. [Google Scholar] [CrossRef]
  105. Yu, Y.; Zhang, X.; Han, Z.; Zhao, W.; Zhang, L. Expression and regulation of miR-449a and AREG in cerebral ischemic injury. Metab. Brain Dis. 2019, 34, 821–832. [Google Scholar] [CrossRef]
  106. Chai, Z.; Gong, J.; Zheng, P.; Zheng, J. Inhibition of miR-19a-3p decreases cerebral ischemia/reperfusion injury by targeting IGFBP3 in vivo and in vitro. Biol. Res. 2020, 53, 17. [Google Scholar] [CrossRef]
  107. Wei, W.; Li, H.; Deng, Y.; Zheng, X.; Zhou, Y.; Xue, X. The combination of Alisma and Atractylodes ameliorates cerebral ischaemia/reperfusion injury by negatively regulating astrocyte-derived exosomal miR-200a-3p/141-3p by targeting SIRT1. J. Ethnopharmacol. 2023, 313, 116597. [Google Scholar] [CrossRef]
  108. Ruan, Z.F.; Xie, M.; Gui, S.J.; Lan, F.; Wan, J.; Li, Y. MiR-370 accelerated cerebral ischemia reperfusion injury via targeting SIRT6 and regulating Nrf2/ARE signal pathway. Kaohsiung J. Med. Sci. 2020, 36, 741–749. [Google Scholar] [CrossRef]
  109. Gu, C.; Mo, W.; Wang, K.; Gao, M.; Chen, J.; Zhang, F.; Shen, J. Exosomal miR-370-3p increases the permeability of blood-brain barrier in ischemia/reperfusion stroke of brain by targeting MPK1. Aging 2023, 15, 1931–1943. [Google Scholar] [CrossRef]
  110. Liang, J.; Wang, Q.; Li, J.Q.; Guo, T.; Yu, D. Long non-coding RNA MEG3 promotes cerebral ischemia-reperfusion injury through increasing pyroptosis by targeting miR-485/AIM2 axis. Exp. Neurol. 2020, 325, 113139. [Google Scholar] [CrossRef]
  111. Shi, Y.; Li, K.; Xu, K.; Liu, Q.H. MiR-155-5p accelerates cerebral ischemia-reperfusion injury via targeting DUSP14 by regulating NF-κB and MAPKs signaling pathways. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 1408–1419. [Google Scholar] [PubMed]
  112. Shi, Y.; Li, Z.; Li, K.; Xu, K. miR-155-5p accelerates cerebral ischemia-reperfusion inflammation injury and cell pyroptosis via DUSP14/ TXNIP/NLRP3 pathway. Acta Biochim. Pol. 2022, 69, 787–793. [Google Scholar] [CrossRef]
  113. Zhang, L.; Liu, C.; Huang, C.; Xu, X.; Teng, J. miR-155 Knockdown Protects against Cerebral Ischemia and Reperfusion Injury by Targeting MafB. BioMed Res. Int. 2020, 2020, 6458204. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, M.; Wang, J.; Li, J.; Kong, F.; Lin, S. miR-101-3p improves neuronal morphology and attenuates neuronal apoptosis in ischemic stroke in young mice by downregulating HDAC9. Transl. Neurosci. 2023, 14, 20220286. [Google Scholar] [CrossRef] [PubMed]
  115. Fan, W.; Qin, Y.; Tan, J.; Li, B.; Liu, Y.; Rong, J.; Shi, W.; Yu, B. RGD1564534 represses NLRP3 inflammasome activity in cerebral injury following ischemia-reperfusion by impairing miR-101a-3p-mediated Dusp1 inhibition. Exp. Neurol. 2023, 359, 114266. [Google Scholar] [CrossRef]
  116. Ye, X.; Fang, H.; Feng, Y.; Qian, H.; Wu, Z.; Chen, S. MiR-10b-3p Protects Cerebral I/R Injury through Targeting Programmed Cell Death 5 (PDCD5). Crit. Rev. Eukaryot. Gene Expr. 2021, 31, 85–98. [Google Scholar] [CrossRef]
  117. Sun, K.; Zhang, J.; Yang, Q.; Zhu, J.; Zhang, X.; Wu, K.; Li, Z.; Xie, W.; Luo, X. MiR-10b-3p alleviates cerebral ischemia/reperfusion injury by targeting Krüppel-like factor 5 (KLF5). Pflügers Arch. -Eur. J. Physiol. 2022, 474, 343–353. [Google Scholar] [CrossRef]
  118. Yang, L.; Wang, L.; Wang, J.; Liu, P. Long non-coding RNA Gm11974 aggravates oxygen-glucose deprivation-induced injury via miR-122-5p/SEMA3A axis in ischaemic stroke. Metab. Brain Dis. 2021, 36, 2059–2069. [Google Scholar] [CrossRef]
  119. Huang, P.; Wei, S.; Ren, J.; Tang, Z.; Guo, M.; Situ, F.; Zhang, D.; Zhu, J.; Xiao, L.; Xu, J.; et al. MicroRNA-124-3p alleviates cerebral ischaemia-induced neuroaxonal damage by enhancing Nrep expression. J. Stroke Cerebrovasc. Dis. 2023, 32, 106949. [Google Scholar] [CrossRef]
  120. Zhang, K.L.; Li, S.M.; Hou, J.Y.; Hong, Y.H.; Chen, X.X.; Zhou, C.Q.; Wu, H.; Zheng, G.H.; Zeng, C.T.; Wu, H.D.; et al. Elabela, a Novel Peptide, Exerts Neuroprotective Effects Against Ischemic Stroke Through the APJ/miR-124-3p/CTDSP1/AKT Pathway. Cell. Mol. Neurobiol. 2023, 43, 2989–3003. [Google Scholar] [CrossRef]
  121. Zhang, R.; Zhou, W.; Yu, Z.; Yang, L.; Liu, G.; Yu, H.; Zhou, Q.; Min, Z.; Zhang, C.; Wu, Q.; et al. miR-1247-3p mediates apoptosis of cerebral neurons by targeting caspase-2 in stroke. Brain Res. 2019, 1714, 18–26. [Google Scholar] [CrossRef]
  122. Yan, Q.; Sun, S.Y.; Yuan, S.; Wang, X.Q.; Zhang, Z.C. Inhibition of microRNA-9-5p and microRNA-128-3p can inhibit ischemic stroke-related cell death in vitro and in vivo. IUBMB Life 2020, 72, 2382–2390. [Google Scholar] [CrossRef] [PubMed]
  123. Hou, H.; Wang, Y.; Yang, L.; Wang, Y. Exosomal miR-128-3p reversed fibrinogen-mediated inhibition of oligodendrocyte progenitor cell differentiation and remyelination after cerebral ischemia. CNS Neurosci. Ther. 2023, 29, 1405–1422. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, H.; Wang, X.; Chen, W.; Yang, Y.; Wang, Y.; Wan, H.; Zhu, Z. Danhong injection alleviates cerebral ischemia-reperfusion injury by inhibiting autophagy through miRNA-132-3p/ATG12 signal axis. J. Ethnopharmacol. 2023, 300, 115724. [Google Scholar] [CrossRef] [PubMed]
  125. Yang, X.; Xu, J.; Lan, S.; Tong, Z.; Chen, K.; Liu, Z.; Xu, S. Exosomal miR-133a-3p Derived from BMSCs Alleviates Cerebral Ischemia-Reperfusion Injury via Targeting DAPK2. Int. J. Nanomed. 2023, 18, 65–78. [Google Scholar] [CrossRef]
  126. Dai, Q.; Ma, Y.; Xu, Z.; Zhang, L.; Yang, H.; Liu, Q.; Wang, J. Downregulation of circular RNA HECTD1 induces neuroprotection against ischemic stroke through the microRNA-133b/TRAF3 pathway. Life Sci. 2021, 264, 118626. [Google Scholar] [CrossRef]
  127. Chi, W.; Meng, F.; Li, Y.; Li, P.; Wang, G.; Cheng, H.; Han, S.; Li, J. Impact of microRNA-134 on neural cell survival against ischemic injury in primary cultured neuronal cells and mouse brain with ischemic stroke by targeting HSPA12B. Brain Res. 2014, 1592, 22–33. [Google Scholar] [CrossRef]
  128. Wang, C.; Hu, F. Long noncoding RNA SOX2OT silencing alleviates cerebral ischemia-reperfusion injury via miR-135a-5p-mediated NR3C2 inhibition. Brain Res. Bull. 2021, 173, 193–202. [Google Scholar] [CrossRef]
  129. Yi, M.; Li, Y.; Wang, D.; Zhang, Q.; Yang, L.; Yang, C. KCNQ1OT1 Exacerbates Ischemia-Reperfusion Injury Through Targeted Inhibition of miR-140-3P. Inflammation 2020, 43, 1832–1845. [Google Scholar] [CrossRef]
  130. Zhang, Y.; Su, Q.; Xia, W.; Jia, K.; Meng, D.; Wang, X.; Ni, X.; Su, Z. MiR-140–3p directly targets Tyro3 to regulate OGD/R-induced neuronal injury through the PI3K/Akt pathway. Brain Res. Bull. 2023, 192, 93–106. [Google Scholar]
  131. Liu, T.; Feng, J.; Sun, Z.; He, M.; Sun, L.; Dong, S.; Guo, Z.; Zhang, G. Inhibition of miR-141-3p attenuates apoptosis of neural stem cells via targeting PBX1 to regulate PROK2 transcription in MCAO mice. Cell Cycle 2023, 22, 403–418. [Google Scholar] [CrossRef]
  132. Meng, S.; Wang, B.; Li, W. LncRNA MALAT1 improves cerebral ischemia-reperfusion injury and cognitive dysfunction by regulating miR-142-3p/SIRT1 axis. Int. J. Neurosci. 2023, 133, 740–753. [Google Scholar] [CrossRef]
  133. Li, J.; Ma, L. MiR-142-3p Attenuates Oxygen Glucose Deprivation/Reoxygenation-Induced Injury by Targeting FBXO3 in Human Neuroblastoma SH-SY5Y Cells. World Neurosurg. 2020, 136, e149–e157. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, N.; Zhang, L.; Lu, Y.; Zhang, M.; Zhang, Z.; Wang, K.; Lv, J. Down-regulation of microRNA-142-5p attenuates oxygen-glucose deprivation and reoxygenation-induced neuron injury through up-regulating Nrf2/ARE signaling pathway. Biomed. Pharmacother. 2017, 89, 1187–1195. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, S.; Liu, Z. Inhibition of microRNA-143-3p Attenuates Cerebral Ischemia/Reperfusion Injury by Targeting FSTL1. Neuromol. Med. 2021, 23, 500–510. [Google Scholar] [CrossRef] [PubMed]
  136. Li, Y.; Zhao, Y.; Cheng, M.; Qiao, Y.; Wang, Y.; Xiong, W.; Yue, W. Suppression of microRNA-144-3p attenuates oxygen-glucose deprivation/reoxygenation-induced neuronal injury by promoting Brg1/Nrf2/ARE signaling. J. Biochem. Mol. Toxicol. 2018, 32, e22044. [Google Scholar] [CrossRef]
  137. Zhou, H.; Zhou, J.; Teng, H.; Yang, H.; Qiu, J.; Li, X. MiR-145 enriched exosomes derived from bone marrow-derived mesenchymal stem cells protects against cerebral ischemia-reperfusion injury through downregulation of FOXO1. Biochem. Biophys. Res. Commun. 2022, 632, 92–99. [Google Scholar] [CrossRef]
  138. Du, Y.; Ma, X.; Ma, L.; Li, S.; Zheng, J.; Lv, J.; Cui, L.; Lv, J. Inhibition of microRNA-148b-3p alleviates oxygen-glucose deprivation/reoxygenation-induced apoptosis and oxidative stress in HT22 hippocampal neuron via reinforcing Sestrin2/Nrf2 signalling. Clin. Exp. Pharmacol. Physiol. 2020, 47, 561–570. [Google Scholar] [CrossRef]
  139. Wang, X.; Xu, Q.; Wang, S. Overexpression of miR-149-5p Attenuates Cerebral Ischemia/Reperfusion (I/R) Injury by Targeting Notch2. Neuromol. Med. 2022, 24, 279–289. [Google Scholar] [CrossRef] [PubMed]
  140. Zhang, A.; Qian, Y.; Qian, J. MicroRNA-152-3p protects neurons from oxygen-glucose-deprivation/reoxygenation-induced injury through upregulation of Nrf2/ARE antioxidant signaling by targeting PSD-93. Biochem. Biophys. Res. Commun. 2019, 517, 69–76. [Google Scholar] [CrossRef]
  141. Li, Y.; Peng, B.; Li, Y.; Huang, A.; Peng, Y.; Yu, Q.; Li, Y. MiR-203a-3p/153-3p improves cognitive impairments induced by ischemia/reperfusion via blockade of SRC-mediated MAPK signaling pathway in ischemic stroke. Chem. Biol. Interact. 2022, 358, 109900. [Google Scholar] [CrossRef] [PubMed]
  142. Wang, H.J.; Tang, X.L.; Huang, G.; Li, Y.B.; Pan, R.H.; Zhan, J.; Wu, Y.K.; Liang, J.F.; Bai, X.X.; Cai, J. Long Non-Coding KCNQ1OT1 Promotes Oxygen-Glucose-Deprivation/Reoxygenation-Induced Neurons Injury Through Regulating MIR-153-3p/FOXO3 Axis. J. Stroke Cerebrovasc. Dis. 2020, 29, 105126. [Google Scholar] [CrossRef] [PubMed]
  143. Zha, H.; Miao, W.; Rong, W.; Wang, A.; Jiang, W.; Liu, R.; Liu, L.; Wang, Y. Remote ischaemic perconditioning reduces the infarct volume and improves the neurological function of acute ischaemic stroke partially through the miR-153-5p/TLR4/p65/IkBa signalling pathway. Folia Neuropathol. 2021, 59, 335–349. [Google Scholar] [CrossRef]
  144. Li, S.; Zhu, P.; Wang, Y.; Huang, S.; Wu, Z.; He, J.; Hu, X.; Wang, Y.; Yuan, Y.; Zhao, B.; et al. miR-181a targets PTEN to mediate the neuronal injury caused by oxygen-glucose deprivation and reoxygenation. Metab. Brain Dis. 2023, 38, 2077–2091. [Google Scholar] [CrossRef] [PubMed]
  145. Song, X.; Xue, Y.; Cai, H. Down-Regulation of miR-181a-5p Prevents Cerebral Ischemic Injury by Upregulating En2 and Activating Wnt/β-catenin Pathway. J. Stroke Cerebrovasc. Dis. 2021, 30, 105485. [Google Scholar] [CrossRef]
  146. Song, H.; Zhang, X.; Chen, R.; Miao, J.; Wang, L.; Cui, L.; Ji, H.; Liu, Y. Cortical Neuron-Derived Exosomal MicroRNA-181c-3p Inhibits Neuroinflammation by Downregulating CXCL1 in Astrocytes of a Rat Model with Ischemic Brain Injury. Neuroimmunomodulation 2019, 26, 217–233. [Google Scholar] [CrossRef]
  147. Li, S.; Chen, S.; Wang, Y.; Hu, X.; Wang, Y.; Wu, Z.; Huang, S.; He, J.; Deng, F.; Zhao, B.; et al. Direct targeting of DOCK4 by miRNA-181d in oxygen-glucose deprivation/reoxygenation-mediated neuronal injury. Lipids Health Dis. 2023, 22, 34. [Google Scholar] [CrossRef]
  148. Zhang, T.; Tian, C.; Wu, J.; Zhang, Y.; Wang, J.; Kong, Q.; Mu, L.; Sun, B.; Ai, T.; Wang, Y.; et al. MicroRNA-182 exacerbates blood-brain barrier (BBB) disruption by downregulating the mTOR/FOXO1 pathway in cerebral ischemia. FASEB J. 2020, 34, 13762–13775. [Google Scholar] [CrossRef]
  149. Alhadidi, Q.M.; Xu, L.; Sun, X.; Althobaiti, Y.S.; Almalki, A.; Alsaab, H.O.; Stary, C.M. MiR-182 Inhibition Protects Against Experimental Stroke in vivo and Mitigates Astrocyte Injury and Inflammation in vitro via Modulation of Cortactin Activity. Neurochem. Res. 2022, 47, 3682–3696. [Google Scholar] [CrossRef]
  150. Ding, W.; Gu, Q.; Liu, M.; Zou, J.; Sun, J.; Zhu, J. Astrocytes-derived exosomes pre-treated by berberine inhibit neuroinflammation after stroke via miR-182-5p/Rac1 pathway. Int. Immunopharmacol. 2023, 118, 110047. [Google Scholar] [CrossRef]
  151. Wang, W.; Hu, Y.; Zhang, Y. FTX Attenuates Cerebral Ischemia-Reperfusion Injury by Inhibiting Apoptosis and Oxidative Stress via miR-186-5p/MDM4 Pathway. Neurotox. Res. 2022, 40, 542–552. [Google Scholar] [CrossRef]
  152. Ren, Z.; Hu, Y.; Guo, D.; Guan, Z.; Chen, L.; He, J.; Yu, W. Increased miR-187-3p expression after cerebral ischemia/reperfusion induces apoptosis via initiation of endoplasmic reticulum stress. Neurosci. Lett. 2021, 759, 135947. [Google Scholar] [CrossRef]
  153. Fan, W.; Rong, J.; Shi, W.; Liu, W.; Wang, J.; Tan, J.; Yu, B.; Tong, J. GATA6 Inhibits Neuronal Autophagy and Ferroptosis in Cerebral ischemia-reperfusion Injury Through a miR-193b/ATG7 axis-dependent Mechanism. Neurochem. Res. 2023, 48, 2552–2567. [Google Scholar] [CrossRef] [PubMed]
  154. Li, F.; Liang, J.; Tong, H.; Zhu, S.; Tang, D. Inhibition of microRNA-199a-5p ameliorates oxygen-glucose deprivation/reoxygenation-induced apoptosis and oxidative stress in HT22 neurons by targeting Brg1 to activate Nrf2/HO-1 signalling. Clin. Exp. Pharmacol. Physiol. 2020, 47, 1020–1029. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, G.; Li, T.; Chang, X.; Xing, J. Long Noncoding RNA SNHG14 Promotes Ischemic Brain Injury via Regulating miR-199b/AQP4 Axis. Neurochem. Res. 2021, 46, 1280–1290. [Google Scholar] [CrossRef]
  156. Ge, X.L.; Wang, J.L.; Liu, X.; Zhang, J.; Liu, C.; Guo, L. Inhibition of miR-19a protects neurons against ischemic stroke through modulating glucose metabolism and neuronal apoptosis. Cell. Mol. Biol. Lett. 2019, 24, 37. [Google Scholar] [CrossRef] [PubMed]
  157. Zhou, F.; Wang, Y.K.; Zhang, C.G.; Wu, B.Y. miR-19a/b-3p promotes inflammation during cerebral ischemia/reperfusion injury via SIRT1/FoxO3/SPHK1 pathway. J. Neuroinflamm. 2021, 18, 122. [Google Scholar] [CrossRef] [PubMed]
  158. Gao, N.; Tang, H.; Gao, L.; Tu, G.L.; Luo, H.; Xia, Y. LncRNA H19 Aggravates Cerebral Ischemia/Reperfusion Injury by Functioning as a ceRNA for miR-19a-3p to Target PTEN. Neuroscience 2020, 437, 117–129. [Google Scholar] [CrossRef]
  159. Zhang, N.; Yang, L.; Wen, J.; Cui, H. miR-200b-3p antagomir inhibits neuronal apoptosis in oxygen-glucose deprivation (OGD) model through regulating β-TrCP. Brain Res. 2023, 1800, 148192. [Google Scholar] [CrossRef] [PubMed]
  160. Li, B.; Huang, Z.; Meng, J.; Yu, W.; Yang, H. MiR-202-5p attenuates neurological deficits and neuronal injury in MCAO model rats and OGD-induced injury in Neuro-2a cells by targeting eIF4E-mediated induction of autophagy and inhibition of Akt/GSK-3β pathway. Mol. Cell. Probes. 2020, 51, 101497. [Google Scholar] [CrossRef]
  161. Xie, X.; Cao, Y.; Dai, L.; Zhou, D. Bone marrow mesenchymal stem cell-derived exosomal lncRNA KLF3-AS1 stabilizes Sirt1 protein to improve cerebral ischemia/reperfusion injury via miR-206/USP22 axis. Mol. Med. 2023, 29, 3. [Google Scholar] [CrossRef] [PubMed]
  162. Sampath, D.; Branyan, T.E.; Markowsky, K.G.; Gunda, R.; Samiya, N.; Obenaus, A.; Sohrabji, F. Sex differences in cognitive impairment after focal ischemia in middle-aged rats and the effect of iv miR-20a-3p treatment. Neurobiol. Aging 2023, 129, 168–177. [Google Scholar] [CrossRef] [PubMed]
  163. Branyan, T.E.; Selvamani, A.; Park, M.J.; Korula, K.E.; Kosel, K.F.; Srinivasan, R.; Sohrabji, F. Functional Assessment of Stroke-Induced Regulation of miR-20a-3p and Its Role as a Neuroprotectant. Transl. Stroke Res. 2022, 13, 432–448. [Google Scholar] [CrossRef]
  164. Liu, N.; Fan, Y.; Li, Y.; Zhang, Y.; Li, J.; Wang, Y.; Wang, Z.; Liu, Y.; Li, Y.; Kang, Z.; et al. OL-FS13 Alleviates Cerebral Ischemia-reperfusion Injury by Inhibiting miR-21-3p Expression. Curr. Neuropharmacol. 2023. [Google Scholar] [CrossRef]
  165. Peng, Z.; Li, M.; Tan, X.; Xiang, P.; Wang, H.; Luo, Y.; Yang, Y.; Huang, H.; Chen, Z.; Xia, H.; et al. miR-211-5p alleviates focal cerebral ischemia-reperfusion injury in rats by down-regulating the expression of COX2. Biochem. Pharmacol. 2020, 177, 113983. [Google Scholar] [CrossRef] [PubMed]
  166. Tian, Y.S.; Zhong, D.; Liu, Q.Q.; Zhao, X.L.; Sun, H.X.; Jin, J.; Wang, H.N.; Li, G.Z. Upregulation of miR-216a exerts neuroprotective effects against ischemic injury through negatively regulating JAK2/STAT3-involved apoptosis and inflammatory pathways. J. Neurosurg. 2018, 130, 977–988. [Google Scholar] [CrossRef]
  167. Cao, Y.; Liu, H.; Zhang, J.; Dong, Y. Circular RNA cZNF292 silence alleviates OGD/R-induced injury through up-regulation of miR-22 in rat neural stem cells (NSCs). Artif. Cells Nanomed. Biotechnol. 2020, 48, 594–601. [Google Scholar] [CrossRef]
  168. Kui, L.; Li, Z.; Wang, G.; Li, X.; Zhao, F.; Jiao, Y. CircPDS5B Reduction Improves Angiogenesis Following Ischemic Stroke by Regulating MicroRNA-223-3p/NOTCH2 Axis. Neurol. Genet. 2023, 9, e200074. [Google Scholar] [CrossRef]
  169. Zhao, Y.; Gan, Y.; Xu, G.; Hua, K.; Liu, D. Exosomes from MSCs overexpressing microRNA-223-3p attenuate cerebral ischemia through inhibiting microglial M1 polarization mediated inflammation. Life Sci. 2020, 260, 118403. [Google Scholar] [CrossRef]
  170. Wei, X.Y.; Zhang, T.Q.; Suo, R.; Qu, Y.Y.; Chen, Y.; Zhu, Y.L. Long non-coding RNA RPL34-AS1 ameliorates oxygen-glucose deprivation-induced neuronal injury via modulating miR-223-3p/IGF1R axis. Hum. Cell 2022, 35, 1785–1796. [Google Scholar] [CrossRef]
  171. Zhang, Y.; Liu, J.; Su, M.; Wang, X.; Xie, C. Exosomal microRNA-22-3p alleviates cerebral ischemic injury by modulating KDM6B/BMP2/BMF axis. Stem Cell Res. Ther. 2021, 12, 111. [Google Scholar] [CrossRef] [PubMed]
  172. Zhang, H.S.; Ouyang, B.; Ji, X.Y.; Liu, M.F. Gastrodin Alleviates Cerebral Ischaemia/Reperfusion Injury by Inhibiting Pyroptosis by Regulating the lncRNA NEAT1/miR-22-3p Axis. Neurochem. Res. 2021, 46, 1747–1758. [Google Scholar] [CrossRef] [PubMed]
  173. Deng, Y.; Ma, G.; Dong, Q.; Sun, X.; Liu, L.; Miao, Z.; Gao, F. Overexpression of miR-224-3p alleviates apoptosis from cerebral ischemia reperfusion injury by targeting FIP200. J Cell. Biochem. 2019, 120, 17151–17158. [Google Scholar] [CrossRef] [PubMed]
  174. Huo, H.; Hu, C.; Lu, Y.; Zhou, J.; Mai, Z. Silencing of circCDC14A prevents cerebral ischemia-reperfusion injury via miR-23a-3p/CXCL12 axis. J. Biochem. Mol. Toxicol. 2022, 36, e22982. [Google Scholar] [CrossRef]
  175. Kuai, F.; Zhou, L.; Zhou, J.; Sun, X.; Dong, W. Long non-coding RNA THRIL inhibits miRNA-24-3p to upregulate neuropilin-1 to aggravate cerebral ischemia-reperfusion injury through regulating the nuclear factor κB p65 signaling. Aging 2021, 13, 9071–9084. [Google Scholar] [CrossRef]
  176. Di, G.; Yang, X.; Cheng, F.; Liu, H.; Xu, M. CEBPA-AS1 Knockdown Alleviates Oxygen-Glucose Deprivation/Reperfusion-Induced Neuron Cell Damage by the MicroRNA 24-3p/BOK Axis. Mol. Cell. Biol. 2021, 41, e0006521. [Google Scholar] [CrossRef]
  177. Li, Y.; Zhang, J.K.; Yu, Z.T.; Jiang, J.W.; Tang, H.; Tu, G.L.; Xia, Y. LncRNA XIST Exacerbates Oxygen-Glucose Deprivation/Reoxygenation-Induced Cerebral Injury Through the miR-25-3p/TRAF3 Axis. Mol. Neurobiol. 2023. [Google Scholar] [CrossRef]
  178. Wei, R.; Zhang, L.; Hu, W.; Wu, J.; Zhang, W. Long non-coding RNA AK038897 aggravates cerebral ischemia/reperfusion injury via acting as a ceRNA for miR-26a-5p to target DAPK1. Exp. Neurol. 2019, 314, 100–110. [Google Scholar] [CrossRef]
  179. Li, Z.; Li, J.; Tang, N. Long noncoding RNA Malat1 is a potent autophagy inducer protecting brain microvascular endothelial cells against oxygen-glucose deprivation/reoxygenation-induced injury by sponging miR-26b and upregulating ULK2 expression. Neuroscience 2017, 354, 1–10. [Google Scholar] [CrossRef]
  180. Xiao, Y.; Zheng, S.; Duan, N.; Li, X.; Wen, J. MicroRNA-26b-5p alleviates cerebral ischemia-reperfusion injury in rats via inhibiting the N-myc/PTEN axis by downregulating KLF10 expression. Hum. Exp. Toxicol. 2021, 40, 1250–1262. [Google Scholar] [CrossRef]
  181. Li, W.; Zhu, Q.; Xu, X.; Hu, X. MiR-27a-3p suppresses cerebral ischemia-reperfusion injury by targeting FOXO1. Aging 2021, 13, 11727–11737. [Google Scholar] [CrossRef] [PubMed]
  182. Zhang, Z.; He, J.; Wang, B. Circular RNA circ_HECTD1 regulates cell injury after cerebral infarction by miR-27a-3p/FSTL1 axis. Cell Cycle 2021, 20, 914–926. [Google Scholar] [CrossRef]
  183. Li, J.; Peng, L.; Bai, W.; Peng, P.; Chen, W.; Yang, W.; Shao, J. Biliverdin Protects Against Cerebral Ischemia/Reperfusion Injury by Regulating the miR-27a-3p/Rgs1 Axis. Neuropsychiatr. Dis Treat. 2021, 17, 1165–1181. [Google Scholar] [CrossRef]
  184. Gao, X.Z.; Zhang, Z.X.; Han, G.L. MiR-29a-3p Enhances the Viability of Rat Neuronal Cells that Injured by Oxygen-Glucose Deprivation/Reoxygenation Treatment Through Targeting TNFRSF1A and Regulating NF-κB Signaling Pathway. J. Stroke Cerebrovasc. Dis. 2020, 29, 105210. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, Z.; Wang, N.; Zhang, Y.; Zhao, J.; Lv, J. Downregulation of microRNA-302b-3p relieves oxygen-glucose deprivation/re-oxygenation induced injury in murine hippocampal neurons through up-regulating Nrf2 signaling by targeting fibroblast growth factor 15/19. Chem. Biol. Interact. 2019, 309, 108705. [Google Scholar] [CrossRef]
  186. Zhou, X.; Wang, Z.; Xu, B.; Ji, N.; Meng, P.; Gu, L.; Li, Y. Long non-coding RNA NORAD protects against cerebral ischemia/reperfusion injury induced brain damage, cell apoptosis, oxidative stress and inflammation by regulating miR-30a-5p/YWHAG. Bioengineered 2021, 12, 9174–9188. [Google Scholar] [CrossRef] [PubMed]
  187. Tao, H.; Dong, L.; Shan, X.; Li, L.; Chen, H. MicroRNA-32-3p facilitates cerebral ischemia/reperfusion injury through inhibiting Cab39/AMPK. Int. Immunopharmacol. 2023, 121, 110504. [Google Scholar] [CrossRef]
  188. Yi, S.; Zhang, C.; Li, N.; Fu, Y.; Li, H.; Zhang, J. miR-325-3p Protects Neurons from Oxygen-Glucose Deprivation and Reoxygenation Injury via Inhibition of RIP3. Dev. Neurosci. 2020, 42, 83–93. [Google Scholar] [CrossRef]
  189. Wang, Y.; Pan, W.; Wang, Y.; Chen, S. MicroRNA-32-5p attenuates cerebral ischemia/reperfusion injuries by modulating the phosphatase and tensin homologous protein. Metab. Brain Dis. 2021, 36, 2495–2504. [Google Scholar] [CrossRef]
  190. Yan, F.; Wang, P.; Yang, X.; Wang, F. Long non-coding RNA HOXA11-AS regulates ischemic neuronal death by targeting miR-337-3p/YBX1 signaling pathway: Protective effect of dexmedetomidine. Aging 2023, 15, 2797–2811. [Google Scholar] [CrossRef]
  191. Yu, Z.; Zhu, M.; Shu, D.; Zhang, R.; Xiang, Z.; Jiang, A.; Liu, S.; Zhang, C.; Yuan, Q.; Hu, X. LncRNA PEG11as aggravates cerebral ischemia/reperfusion injury after ischemic stroke through miR-342-5p/PFN1 axis. Life Sci. 2023, 313, 121276. [Google Scholar] [CrossRef] [PubMed]
  192. Zhu, H.; Zhang, Y.; Zhu, Y. MiR-342-5p protects neurons from cerebral ischemia induced-apoptosis through regulation of Akt/NF-κB pathways by targeting CCAR2. J. Stroke Cerebrovasc. Dis. 2023, 32, 106901. [Google Scholar] [CrossRef] [PubMed]
  193. Ye, X.; Song, H.; Hu, H.; Zhou, C.; Chen, Q.; Hong, L.; Huang, M.; Zhu, H. MiR-361-3p alleviates cerebral ischemia-reperfusion injury by targeting NACC1 through the PINK1/Parkin pathway. J. Mol. Histol. 2022, 53, 357–367. [Google Scholar] [CrossRef] [PubMed]
  194. Jia, Y.; Liu, J.; Hu, H.; Duan, Q.; Chen, J.; Li, L. MiR-363-3p attenuates neonatal hypoxic-ischemia encephalopathy by targeting DUSP5. Neurosci. Res. 2021, 171, 103–113. [Google Scholar] [CrossRef]
  195. Yang, B.; Zang, L.; Cui, J.; Wei, L. Circular RNA TTC3 regulates cerebral ischemia-reperfusion injury and neural stem cells by miR-372-3p/TLR4 axis in cerebral infarction. Stem Cell Res. Ther. 2021, 12, 125. [Google Scholar] [CrossRef]
  196. Shen, J.; Han, Q.; Li, W.; Chen, X.; Lu, J.; Zheng, J.; Xue, S. miR-383-5p Regulated by the Transcription Factor CTCF Affects Neuronal Impairment in Cerebral Ischemia by Mediating Deacetylase HDAC9 Activity. Mol. Neurobiol. 2022, 59, 6307–6320. [Google Scholar] [CrossRef]
  197. He, Z.; Zhao, Y.; Zhu, Y.; Wang, W.; Liu, X.; Lu, F. Interfering TUG1 Attenuates Cerebrovascular Endothelial Apoptosis and Inflammatory injury After Cerebral Ischemia/Reperfusion via TUG1/miR-410/FOXO3 ceRNA Axis. Neurotox. Res. 2022, 40, 1–13. [Google Scholar] [CrossRef]
  198. Xu, J.; Huang, X.; Liu, S.; Chen, D.; Xie, Y.; Zhao, Z. The protective effects of lncRNA ZFAS1/miR-421-3p/MEF2C axis on cerebral ischemia-reperfusion injury. Cell Cycle 2022, 21, 1915–1931. [Google Scholar] [CrossRef]
  199. Zheng, L.; Tang, X.; Lu, M.; Sun, S.; Xie, S.; Cai, J.; Zan, J. microRNA-421-3p prevents inflammatory response in cerebral ischemia/reperfusion injury through targeting m6A Reader YTHDF1 to inhibit p65 mRNA translation. Int. Immunopharmacol. 2020, 88, 106937. [Google Scholar] [CrossRef]
  200. Gan, C.; Ouyang, F. Exosomes Released from Bone-Marrow Stem Cells Ameliorate Hippocampal Neuronal Injury Through transferring miR-455-3p. J. Stroke Cerebrovasc. Dis. 2022, 31, 106142. [Google Scholar] [CrossRef]
  201. Fan, Y.; Wei, L.; Zhang, S.; Song, X.; Yang, J.; He, X.; Zheng, X. LncRNA SNHG15 Knockdown Protects Against OGD/R-Induced Neuron Injury by Downregulating TP53INP1 Expression via Binding to miR-455-3p. Neurochem. Res. 2021, 46, 1019–1030. [Google Scholar] [CrossRef]
  202. Chen, X.; Zhang, S.; Shi, P.; Su, Y.; Zhang, D.; Li, N. MiR-485-5p Promotes Neuron Survival through Mediating Rac1/Notch2 Signaling Pathway after Cerebral Ischemia/Reperfusion. Curr. Neurovasc. Res. 2020, 17, 259–266. [Google Scholar] [CrossRef] [PubMed]
  203. Zheng, H.; Zhang, G.; Liu, G.; Wang, L. Up-regulation of lncRNA NEAT1 in cerebral ischemic stroke promotes activation of astrocytes by modulation of miR-488-3p/RAC1. Exp. Brain Res. 2023, 241, 395–406. [Google Scholar] [CrossRef] [PubMed]
  204. Zhou, L.; Yang, W.; Yao, E.; Li, H.; Wang, J.; Wang, K.; Zhong, X.; Peng, Z.; Huang, X. MicroRNA-488-3p Regulates Neuronal Cell Death in Cerebral Ischemic Stroke Through Vacuolar Protein Sorting 4B (VPS4B). Neuropsychiatr. Dis. Treat. 2021, 17, 41–55. [Google Scholar] [CrossRef] [PubMed]
  205. Jia, T.; Wang, M.; Yan, W.; Wu, W.; Shen, R. Upregulation of miR-489-3p Attenuates Cerebral Ischemia/Reperfusion Injury by Targeting Histone Deacetylase 2 (HDAC2). Neuroscience 2022, 484, 16–25. [Google Scholar] [CrossRef]
  206. Song, L.; Mu, L.; Wang, H. MicroRNA-489-3p aggravates neuronal apoptosis and oxidative stress after cerebral ischemia-reperfusion injury. Bioengineered 2022, 13, 14047–14056. [Google Scholar] [CrossRef]
  207. Sun, L.; Ji, D.; Zhi, F.; Fang, Y.; Zhu, Z.; Ni, T.; Zhu, Q.; Bao, J. MiR-494-3p Upregulation Exacerbates Cerebral Ischemia Injury by Targeting Bhlhe40. Yonsei Med. J. 2022, 63, 389–398. [Google Scholar] [CrossRef]
  208. Yin, K.J.; Deng, Z.; Huang, H.; Hamblin, M.; Xie, C.; Zhang, J.; Chen, Y.E. miR-497 regulates neuronal death in mouse brain after transient focal cerebral ischemia. Neurobiol. Dis. 2010, 38, 17–26. [Google Scholar] [CrossRef]
  209. Shan, W.; Ge, H.; Chen, B.; Huang, L.; Zhu, S.; Zhou, Y. Upregulation of miR-499a-5p Decreases Cerebral Ischemia/Reperfusion Injury by Targeting PDCD4. Cell. Mol. Neurobiol. 2022, 42, 2157–2170. [Google Scholar] [CrossRef]
  210. Wang, H.; Yang, H.; Chang, M.; Sun, F.; Qi, H.; Li, X. Long non-coding RNA TTTY15 sponges miR-520a-3p to exacerbate neural apoptosis induced by cerebral ischemia/reperfusion via targeting IRF9 both in vivo and in vitro. Biomed. J. 2023, 46, 100530. [Google Scholar] [CrossRef]
  211. Shi, Y.; Yi, Z.; Zhao, P.; Xu, Y.; Pan, P. MicroRNA-532-5p protects against cerebral ischemia-reperfusion injury by directly targeting CXCL1. Aging 2021, 13, 11528–11541. [Google Scholar] [CrossRef] [PubMed]
  212. Wei, L.; Peng, Y.; Yang, X.J.; Zhou, P. Knockdown of long non-coding RNA RMRP protects cerebral ischemia-reperfusion injury via the microRNA-613/ATG3 axis and the JAK2/STAT3 pathway. Kaohsiung J. Med. Sci. 2021, 37, 468–478. [Google Scholar] [CrossRef] [PubMed]
  213. Gao, J.D.; Li, R.J.; Ma, P.L.; Yu, L.L.; Li, J.T.; Tian, H.T. Knockdown of lncRNA HCP5 protects against cerebral ischemia/reperfusion injury by regulating miR-652-3p. J. Biol. Regul. Homeost. Agents 2020, 34, 893–900. [Google Scholar]
  214. Zhang, X.; Feng, Y.; Li, J.; Zheng, L.; Shao, Y.; Zhu, F.; Sun, X. MicroRNA-665-3p attenuates oxygen-glucose deprivation-evoked microglial cell apoptosis and inflammatory response by inhibiting NF-κB signaling via targeting TRIM8. Int. Immunopharmacol. 2020, 85, 106650. [Google Scholar] [CrossRef]
  215. Kong, L.Y.; Li, Y.; Rao, D.Y.; Wu, B.; Sang, C.P.; Lai, P.; Ye, J.S.; Zhang, Z.X.; Du, Z.M.; Yu, J.J.; et al. miR-666-3p Mediates the Protective Effects of Mesenchymal Stem Cell-derived Exosomes Against Oxygen-glucose Deprivation and Reoxygenation- induced Cell Injury in Brain Microvascular Endothelial Cells via Mitogen-activated Protein Kinase Pathway. Curr. Neurovasc. Res. 2021, 18, 20–77. [Google Scholar] [CrossRef]
  216. Xu, H.; Nie, B.; Liu, L.; Zhang, C.; Zhang, Z.; Xu, M.; Mei, Y. Curcumin Prevents Brain Damage and Cognitive Dysfunction During Ischemic-reperfusion Through the Regulation of miR-7-5p. Curr. Neurovasc. Res. 2019, 16, 441–454. [Google Scholar] [CrossRef]
  217. Hu, X.; Ma, F.; Cheng, Z.; Zeng, S.; Shen, R.; Li, X.; Hu, J.; Jin, Z.; Cheng, J. LncRNA PEG11as silencing sponges miR-874-3p to alleviate cerebral ischemia stroke via regulating autophagy in vivo and in vitro. Aging 2022, 14, 5177–5194. [Google Scholar] [CrossRef] [PubMed]
  218. Shen, J.; Li, G.; Zhu, Y.; Xu, Q.; Zhou, H.; Xu, K.; Huang, K.; Zhan, R.; Pan, J. Foxo1-induced miR-92b down-regulation promotes blood-brain barrier damage after ischaemic stroke by targeting NOX4. J. Cell. Mol. Med. 2021, 25, 5269–5282. [Google Scholar] [CrossRef]
  219. Zhang, G.; Guo, J.; Zeng, J.; Zhang, X.; Chen, R.; Wang, G.; Liang, W. LncRNA SNHG14 is beneficial to oxygen glucose deprivation/reoxygenation-induced neuro-2a cell injury via mir-98-5p sequestration-caused BCL2L13 upregulation. Metab. Brain Dis. 2022, 37, 2005–2016. [Google Scholar] [CrossRef]
  220. Qi, D.; Wang, W.; Zhang, Y.; Zhang, T. MiR-99b regulates cerebral ischemia neuronal injury through targeting IGF1R. Panminerva Medica 2023, 65, 30–36. [Google Scholar] [CrossRef]
  221. Huang, Y.; Tang, J.; Li, X.; Long, X.; Huang, Y.; Zhang, X. miR-92b-3p Exerts Neuroprotective Effects on Ischemia/Reperfusion-Induced Cerebral Injury via Targeting NOX4 in a Rat Model. Oxidative Med. Cell. Longev. 2022, 2022, 3494262. [Google Scholar] [CrossRef] [PubMed]
  222. McConnell, H.L.; Mishra, A. Cells of the Blood-Brain Barrier: An Overview of the Neurovascular Unit in Health and Disease. Methods Mol. Biol. 2022, 2492, 3–24. [Google Scholar]
  223. Ago, T. The neurovascular unit in health and ischemic stroke. Nihon Rinsho. 2016, 74, 583–588. (In Japanese) [Google Scholar] [PubMed]
  224. Pivoriūnas, A.; Verkhratsky, A. Astrocyte-Endotheliocyte Axis in the Regulation of the Blood-Brain Barrier. Neurochem. Res. 2021, 46, 2538–2550. [Google Scholar] [CrossRef] [PubMed]
  225. Guo, X.; Liu, R.; Jia, M.; Wang, Q.; Wu, J. Ischemia Reperfusion Injury Induced Blood Brain Barrier Dysfunction and the Involved Molecular Mechanism. Neurochem. Res. 2023, 48, 2320–2334. [Google Scholar] [CrossRef] [PubMed]
  226. Cai, Z.; Qiao, P.F.; Wan, C.Q.; Cai, M.; Zhou, N.K.; Li, Q. Role of Blood-Brain Barrier in Alzheimer’s Disease. J. Alzheimer’s Dis. 2018, 63, 1223–1234. [Google Scholar] [CrossRef]
  227. Sweeney, M.D.; Sagare, A.P.; Zlokovic, B.V. Blood-brain barrier breakdown in Alzheimer disease and other neurodegenerative disorders. Nat. Rev. Neurol. 2018, 14, 133–150. [Google Scholar] [CrossRef]
  228. Sweeney, M.D.; Zhao, Z.; Montagne, A.; Nelson, A.R.; Zlokovic, B.V. Blood-Brain Barrier: From Physiology to Disease and Back. Physiol. Rev. 2019, 99, 21–78. [Google Scholar] [CrossRef]
  229. Andjelkovic, A.V.; Stamatovic, S.M.; Phillips, C.M.; Martinez-Revollar, G.; Keep, R.F. Modeling blood-brain barrier pathology in cerebrovascular disease in vitro: Current and future paradigms. Fluids Barriers CNS 2020, 17, 44. [Google Scholar] [CrossRef]
  230. Chekhonin, V.P.; Lebedev, S.V.; Dmitrieva, T.B.; Blinov, D.V.; Gurina, O.I.; Semenova, A.V.; Volodin, N.N. Enzyme immunoassay of NSE and GFAP as the criterion of dynamic evaluation of the rat blood-brain barrier in perinatal hypoxic ischemic injury of the CNS. Bull. Exp. Biol. Med. 2003, 136, 261–265. [Google Scholar] [CrossRef]
  231. Chalbot, S.; Zetterberg, H.; Blennow, K.; Fladby, T.; Andreasen, N.; Grundke-Iqbal, I.; Iqbal, K. Blood-cerebrospinal fluid barrier permeability in Alzheimer’s disease. J. Alzheimer’s Dis. 2011, 25, 505–515. [Google Scholar] [CrossRef] [PubMed]
  232. Helgudottir, S.S.; Routhe, L.J.; Burkhart, A.; Jonsson, K.; Pedersen, I.S.; Lichota, J.; Moos, T. Epigenetic Regulation of Ferroportin in Primary Cultures of the Rat Blood-Brain Barrier. Mol. Neurobiol. 2020, 57, 3526–3539. [Google Scholar] [CrossRef] [PubMed]
  233. Sun, P.; Hamblin, M.H.; Yin, K.J. Non-coding RNAs in the regulation of blood-brain barrier functions in central nervous system disorders. Fluids Barriers CNS 2022, 19, 27. [Google Scholar] [CrossRef] [PubMed]
  234. Sookoian, S.; Gianotti, T.F.; Burgueño, A.L.; Pirola, C.J. Fetal metabolic programming and epigenetic modifications: A systems biology approach. Pediatr. Res. 2013, 73, 531–542. [Google Scholar] [CrossRef] [PubMed]
  235. Yang, K.; Zeng, L.; Ge, A.; Wang, S.; Zeng, J.; Yuan, X.; Mei, Z.; Wang, G.; Ge, J. A systematic review of the research progress of non-coding RNA in neuroinflammation and immune regulation in cerebral infarction/ischemia-reperfusion injury. Front. Immunol. 2022, 13, 930171. [Google Scholar] [CrossRef] [PubMed]
  236. Meng, Q.; Yang, G.; Yang, Y.; Ding, F.; Hu, F. Protective effects of histone deacetylase inhibition by Scriptaid on brain injury in neonatal rat models of cerebral ischemia and hypoxia. Int. J. Clin. Exp. Pathol. 2020, 13, 179–191. [Google Scholar]
  237. Jambhekar, A.; Dhall, A.; Shi, Y. Roles and regulation of histone methylation in animal development. Nat. Rev. Mol. Cell Biol. 2019, 20, 625–641, Erratum in Nat. Rev. Mol. Cell Biol. 2020, 21, 59. [Google Scholar] [CrossRef]
  238. Churilova, A.V.; Gluschenko, T.S.; Rybnikova, E.A.; Samoilov, M.O. The Effect of Histone Deacetylase Inhibitor on the Expression Level of Glucococrticoid Receptor in Rat Forebrain under Hypoxia. Cell Tissue Biol. 2019, 13, 79–84. [Google Scholar] [CrossRef]
  239. Deniz, B.F.; Confortim, H.D.; Miguel, P.M.; Bronauth, L.; Fernandes, I.R.; Muotri, A.R.; Pereira, L.O. High gestational folic acid supplementation prevents hypoxia-ischemia-induced caspase-3 augmenting without changing synapsin and H3 methylation levels in the rat hippocampus. Int. J. Dev. Neurosci. 2021, 81, 510–519. [Google Scholar] [CrossRef]
  240. Paz, A.A.; González-Candia, A. Potential pharmacological target of tight junctions to improve the BBB permeability in neonatal Hypoxic-Ischemic encephalopathy Diseases. Biochem. Pharmacol. 2023, 207, 115356. [Google Scholar] [CrossRef]
  241. Yang, C.; Hawkins, K.E.; Doré, S.; Candelario-Jalil, E. Neuroinflammatory mechanisms of blood-brain barrier damage in ischemic stroke. Am. J. Physiol. Cell Physiol. 2019, 316, C135–C153. [Google Scholar] [CrossRef] [PubMed]
  242. Shen, G.; Ma, Q. MicroRNAs in the Blood-Brain Barrier in Hypoxic-Ischemic Brain Injury. Curr. Neuropharmacol. 2020, 18, 1180–1186. [Google Scholar] [CrossRef] [PubMed]
  243. Fan, F.; Yang, J.; Xu, Y.; Guan, S. MiR-539 Targets MMP-9 to Regulate the Permeability of Blood-Brain Barrier in Ischemia/Reperfusion Injury of Brain. Neurochem. Res. 2018, 43, 2260–2267. [Google Scholar] [CrossRef]
  244. Yao, X.; Wang, Y.; Zhang, D. microRNA-21 Confers Neuroprotection Against Cerebral Ischemia-Reperfusion Injury and Alleviates Blood-Brain Barrier Disruption in Rats via the MAPK Signaling Pathway. J. Mol. Neurosci. 2018, 65, 43–53. [Google Scholar] [CrossRef]
  245. Zuo, X.; Lu, J.; Manaenko, A.; Qi, X.; Tang, J.; Mei, Q.; Xia, Y.; Hu, Q. MicroRNA-132 attenuates cerebral injury by protecting blood-brain-barrier in MCAO mice. Exp. Neurol. 2019, 316, 12–19. [Google Scholar] [CrossRef] [PubMed]
  246. Deng, L.; Zhang, J.; Chen, S.; Wu, Y.; Fan, X.; Zuo, T.; Hu, Q.; Jiang, L.; Yang, S.; Dong, Z. miR-671-5p Upregulation Attenuates Blood-Brain Barrier Disruption in the Ischemia Stroke Model Via the NF-кB/MMP-9 Signaling Pathway. Mol. Neurobiol. 2023, 60, 3824–3838. [Google Scholar] [CrossRef]
  247. Bai, Y.; Zhang, Y.; Han, B.; Yang, L.; Chen, X.; Huang, R.; Wu, F.; Chao, J.; Liu, P.; Hu, G.; et al. Circular RNA DLGAP4 Ameliorates Ischemic Stroke Outcomes by Targeting miR-143 to Regulate Endothelial-Mesenchymal Transition Associated with Blood-Brain Barrier Integrity. J. Neurosci. 2018, 38, 32–50. [Google Scholar] [CrossRef]
  248. 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]
  249. Ponnusamy, V.; Yip, P.K. The role of microRNAs in newborn brain development and hypoxic ischaemic encephalopathy. Neuropharmacology 2019, 149, 55–65. [Google Scholar] [CrossRef]
  250. Ma, Q.; Dasgupta, C.; Li, Y.; Huang, L.; Zhang, L. MicroRNA-210 Suppresses Junction Proteins and Disrupts Blood-Brain Barrier Integrity in Neonatal Rat Hypoxic-Ischemic Brain Injury. Int. J. Mol. Sci. 2017, 18, E1356. [Google Scholar] [CrossRef]
  251. Wang, Y.; Wang, M.D.; Xia, Y.P.; Gao, Y.; Zhu, Y.Y.; Chen, S.C.; Mao, L.; He, Q.W.; Yue, Z.Y.; Hu, B. MicroRNA-130a regulates cerebral ischemia-induced blood-brain barrier permeability by targeting Homeobox A5. FASEB J. 2018, 32, 935–944. [Google Scholar] [CrossRef] [PubMed]
  252. Pan, J.; Qu, M.; Li, Y.; Wang, L.; Zhang, L.; Wang, Y.; Tang, Y.; Tian, H.L.; Zhang, Z.; Yang, G.Y. MicroRNA-126-3p/-5p Overexpression Attenuates Blood-Brain Barrier Disruption in a Mouse Model of Middle Cerebral Artery Occlusion. Stroke 2020, 51, 619–627. [Google Scholar] [CrossRef] [PubMed]
  253. 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]
  254. Ebrahimi, V.; Rastegar-Moghaddam, S.H.; Mohammadipour, A. Therapeutic Potentials of MicroRNA-126 in Cerebral Ischemia. Mol. Neurobiol. 2023, 60, 2062–2069. [Google Scholar] [CrossRef] [PubMed]
  255. Li, F.; Zhou, F.; Yang, B. MicroRNA152-3p Protects Against Ischemia/Reperfusion-Induced Bbb Destruction Possibly Targeting the MAP3K2/JNK/c-Jun Pathway. Neurochem. Res. 2023, 48, 1293–1304. [Google Scholar] [CrossRef]
  256. Hu, H.; Hone, E.A.; Provencher, E.A.P.; Sprowls, S.A.; Farooqi, I.; Corbin, D.R.; Sarkar, S.N.; Hollander, J.M.; Lockman, P.R.; Simpkins, J.W.; et al. MiR-34a Interacts with Cytochrome c and Shapes Stroke Outcomes. Sci. Rep. 2020, 10, 3233. [Google Scholar] [CrossRef]
  257. Yin, K.J.; Deng, Z.; Hamblin, M.; Xiang, Y.; Huang, H.; Zhang, J.; Jiang, X.; Wang, Y.; Chen, Y.E. Peroxisome proliferator-activated receptor delta regulation of miR-15a in ischemia-induced cerebral vascular endothelial injury. J. Neurosci. 2010, 30, 6398–6408. [Google Scholar] [CrossRef]
  258. Wan, Y.; Jin, H.J.; Zhu, Y.Y.; Fang, Z.; Mao, L.; He, Q.; Xia, Y.P.; Li, M.; Li, Y.; Chen, X.; et al. MicroRNA-149-5p regulates blood-brain barrier permeability after transient middle cerebral artery occlusion in rats by targeting S1PR2 of pericytes. FASEB J. 2018, 32, 3133–3148. [Google Scholar] [CrossRef]
  259. Perales, G.; Westenskow, M.; Gutierrez, R.; Caldwell, K.K.; Allan, A.M.; Gardiner, A.S. MicroRNA-150-5p is upregulated in the brain microvasculature during prenatal alcohol exposure and inhibits the angiogenic factor Vezf1. Alcohol. Clin. Exp. Res. 2022, 46, 1953–1966. [Google Scholar] [CrossRef]
Figure 1. The process of RNA interference involving microRNAs in the example of miR-27a-3p. DNA—deoxyribonucleic acid; RNA Pol II/III—RNA polymerase II or III; pri-miRNA—primary miRNA; Drosha—class 2 ribonuclease III enzyme; DGCR8—DiGeorge syndrome critical region 8; pre-miRNA—precursor miRNA; RAN—RAs-related nuclear protein; GTP—guanosine triphosphate; Dicer—endoribonuclease Dicer or helicase with RNase motif; TRBP—transactivation response RNA binding protein; RISC—RNA-induced silencing complex; Ago2—Argonaute 2; mRNA—messenger RNA; RGS1—regulator of G protein signaling 1; FSTL1—follistatin-related protein 1; FOXO1—forkhead box protein O1, ↓—inhibition.
Figure 1. The process of RNA interference involving microRNAs in the example of miR-27a-3p. DNA—deoxyribonucleic acid; RNA Pol II/III—RNA polymerase II or III; pri-miRNA—primary miRNA; Drosha—class 2 ribonuclease III enzyme; DGCR8—DiGeorge syndrome critical region 8; pre-miRNA—precursor miRNA; RAN—RAs-related nuclear protein; GTP—guanosine triphosphate; Dicer—endoribonuclease Dicer or helicase with RNase motif; TRBP—transactivation response RNA binding protein; RISC—RNA-induced silencing complex; Ago2—Argonaute 2; mRNA—messenger RNA; RGS1—regulator of G protein signaling 1; FSTL1—follistatin-related protein 1; FOXO1—forkhead box protein O1, ↓—inhibition.
Ijms 24 12899 g001
Figure 2. Examples of effects of various miRNAs on signaling pathways in the pathogenesis of ischemia–reperfusion brain injury. EPR = endoplasmic reticulum; BBB = blood–brain barrier; green arrows = positive biological effect of RNA interference; red arrows = negative biological effect of RNA interference. The dark blue color of the molecules indicates microRNA-3p (sense); the olive color indicates microRNA-5p (antisense); the aquamarine color indicates microRNAs with an unknown formation end from pre-microRNA.
Figure 2. Examples of effects of various miRNAs on signaling pathways in the pathogenesis of ischemia–reperfusion brain injury. EPR = endoplasmic reticulum; BBB = blood–brain barrier; green arrows = positive biological effect of RNA interference; red arrows = negative biological effect of RNA interference. The dark blue color of the molecules indicates microRNA-3p (sense); the olive color indicates microRNA-5p (antisense); the aquamarine color indicates microRNAs with an unknown formation end from pre-microRNA.
Ijms 24 12899 g002
Table 1. Effects of various miRNAs on target proteins and signaling pathways in the pathogenesis of ischemia–reperfusion brain injury.
Table 1. Effects of various miRNAs on target proteins and signaling pathways in the pathogenesis of ischemia–reperfusion brain injury.
microRNATarget Protein/Signaling PathwayDirection of Gene ExpressionIschemia/OGD EffectRef.Biological Effect of RNA Interference
miR-101-3p↓ HDAC9↓ apoptosis
↑ morphology of neurons
[114]Positive
miR-101a-3p↑ Dusp1↓ apoptosis[115]Negative
miR-10b-3p↓ PDCD5↓ apoptosis[116]Positive
↓ KLF5↓ apoptosis
↓ swelling
↓ inflammation
[117]Positive
miR-122-5p↓ SEMA3A↓ apoptosis
↓ oxidative stress
[118]Positive
miR-124-3p↓ Nrep/Rnf38↓ apoptosis[119]Positive
↓ CTDSP1/AKT↓ apoptosis
↓ axon damage
[120]Positive
miR-1247-3p↓ Caspase-2↓ apoptosis[121]Positive
miR-128-3p *↑ FOXO/Relaxin↓ apoptosis[122]Negative
↓ ACVR1/BMP↓ demyelination
↑ differentiation
[123]Positive
miR-132-3p↓ ATG12
↑ p-PI3K/p-AKT/mTOR
↓ oxidative stress
↓ autophagy
[124]Positive
miR-133a-3p↓ DAPK1
↑ p-Akt и p-mTOR
↓ apoptosis
↓ autophagy
[125]Positive
miR-133b↓ TRAF3/NF-κB↓ apoptosis[126]Positive
miR-134↓ HSPA12B↓ apoptosis[127]Negative
miR-135a-5p↓ NR3C2↓ apoptosis
↓ oxidative stress
↓ inflammation
[128]Positive
miR-140-3p *↑ HIF-1α↓ apoptosis
↓ oxidative stress
↓ inflammation
[129]Positive
↓ Tyro3/PI3K/Akt
↓ Bax/Caspase-3
↑ Bcl-2
↑ apoptosis
↑ oxidative stress
[130]Negative
miR-141-3p↑ SIRT1/ZO-1/Occludin/Claudin-5/CD31
↓ p-NF-κB/IL-1β/TNF-α/GFAP
↓ inflammation[107]Negative
↑ PBX1/PROK2↓ apoptosis[131]Negative
miR-142-3p *↑ SIRT1/SOD/Catalase
↓ TNF-α/IL-6/IL-1β/ROS/MDA
↓ apoptosis[132]Negative
↓ FBXO3↓ apoptosis
↓ inflammation
[133]Positive
miR-142-5p↑ Nrf2/ARE↓ oxidative stress[134]Negative
miR-143-3p↑ FSTL1
↓ Bax/Caspase-3
↑ Bcl-2
↓ apoptosis[135]Negative
miR-144-3p↑ Brg1/Nrf2/ARE↓ apoptosis
↓ oxidative stress
[136]Negative
miR-145↓ FOXO1↓ apoptosis
↓ oxidative stress
↓ inflammation
[137]Positive
miR-148b-3p ↑ Sestrin2/Nrf2↓ apoptosis
↓ oxidative stress
[138]Negative
miR-149-5p↓ Notch2↓ apoptosis
↓ inflammation
[139]Positive
miR-152-3p ↓ PSD-93
↑ Nrf2/ARE
↓ apoptosis
↓ oxidative stress
[140]Positive
miR-153-3p ↓ SRC/MAPK↓ apoptosis
↓ oxidative stress
↓ inflammation
[141]Positive
↓ FOXO3↓ apoptosis[142]Positive
miR-153-5p↓ TLR4/p65/IkBa↓ apoptosis[143]Positive
miR-155↑ MafB
↓ IL-1β/IL-6/TNF-α
↓ iNOS/COX-2
↓ apoptosis
↓ inflammation
[113]Negative
miR-155-5p↑ DUSP14/TXNIP/NLRP3↓ inflammation
↓ pyroptosis
[112]Negative
↑ DUSP14/NF-kB/MAPKs↓ apoptosis[111]Negative
miR-181a↑ PTEN↓ apoptosis
↓ oxidative stress
[144]Negative
miR-181a-5p ↑ En2/Wnt/β-catenin↓ apoptosis[145]Negative
miR-181c-3p↓ CXCL1↓ inflammation[146]Positive
miR-181d↑ DOCK4↓ apoptosis
↓ inflammation
[147]Negative
miR-182↑ mTOR/FOXO1/Bcl-2/Bax↓ apoptosis[148]Negative
↑ cortactin↓ mitochondrial dysfunction
↓ inflammation
[149]Negative
miR-182-5p↓ TLR4↓ inflammation[101]Positive
↓ Rac1↓ inflammation[150]Positive
miR-186-5p↑ MDM4↓ apoptosis
↓ oxidative stress
[151]Negative
miR-187-3p↓ GRP78/Seipin↑ apoptosis
↑ endoplasmic reticulum stress
[152]Negative
miR-193b↓ ATG7↓ autophagy
↓ ferroptosis
[153]Positive
miR-193b-3p↓ 5-lipoxigenase↓ inflammation[102]Positive
miR-199a-5p↑ Brg1/Nrf2/HO-1↓ apoptosis
↓ oxidative stress
[154]Negative
miR-199b↓ AQP4↓ oxidative stress
↓ inflammation
[155]Positive
miR-19a↑ ADIPOR2↓ apoptosis
↑ glucose metabolism
[156]Negative
miR-19a-3p *↑ FOXO3/SPHK1/NF-kB p65
↓ SIRT1
↑ inflammation[157]Negative
↑ IGFBP3↓ apoptosis
↓ inflammation
[106]Negative
↓ PTEN/PI3K/AKT↓ apoptosis
↓ oxidative stress
[158]Positive
miR-19b-3p↑ FOXO3/SPHK1/NF-kB p65
↓ SIRT1
↑ inflammation[157]Negative
miR-200a-3p↑ SIRT1/ZO-1/Occludin/Claudin-5/CD31
↓ p-NF-κB/IL-1β/TNF-α/GFAP
↓ inflammation[107]Negative
miR-200b-3p↑ β-TrCP↓ apoptosis[159]Negative
miR-202-5p↓ eIF4E/
↑ Akt/GSK-3β
↓ apoptosis
↓ autophagy
[160]Positive
miR-203a-3p↓ SRC/MAPK↓ apoptosis
↓ oxidative stress
↓ inflammation
[141]Positive
miR-206↑ USP22/Sirt1↓ apoptosis
↓ inflammation
[161]Negative
miR-20a-3p-↓ cognitive dysfunction[162]Positive
↓ MMP, IL-17A↓ mitochondrial dysfunction
↓ inflammation
[163]Positive
miR-21-3p↑ CAMKK2/AMPK/Nrf-2↓ oxidative stress
↓ inflammation
[164]Negative
miR-211-5p↓ COX2
↓ PGD2/PGE2/TNF-α/IL-1β
↓ apoptosis
↓ inflammation
[165]Positive
miR-216a↓ JAK2/STAT3
↓ iNOS и MMP-9/TNF-α и IL-1β
↓ apoptosis
↓ inflammation
[166]Positive
miR-22↑ Wnt/β-catenin and PKC/ERK↓ apoptosis[167]Positive
miR-223-3p *↓ NOTCH2↑ angiogenesis[168]Positive
↓ CysLT 2 R↓ inflammation[169]Positive
↑ IGF1R↓ apoptosis
↑ glucose metabolism
[170]Negative
miR-22-3p↓ KDM6B/BMP2/BMF↓ apoptosis[171]Positive
↓ IL-1β, IL-18/Caspase-1↓ pyroptosis
↓ inflammation
[172]Positive
miR-224-3p↓ FIP200↓ apoptosis[173]Positive
miR-23a-3p↓ CXCL12↓ apoptosis[174]Positive
miR-24-3p↑ NRP1/NF-κB p65↑ apoptosis
↑ inflammation
[175]Positive
↓ BOK↓ apoptosis
↓ oxidative stress
[176]Positive
miR-25-3p↓ TRAF3↓ apoptosis
↓ inflammation
[177]Positive
miR-26a-5p↓ DAPK1↓ apoptosis[178]Positive
miR-26b↑ ULK2↑ autophagy[179]Positive
miR-26b-5p↓ KLF10/N-myc/PTEN↓ apoptosis
↓ inflammation
[180]Positive
miR-27a-3p↓ FOXO1/p27 Kip1↓ apoptosis[181]Positive
↓ FSTL1↓ apoptosis[182]Positive
↓ Rgs1↓ apoptosis
↓ inflammation
[183]Positive
miR-29a-3p↓ NF-κB↓ apoptosis[184]Positive
miR-302b-3p↑ FGF15/Nrf2/ARE↓ apoptosis
↓ oxidative stress
[185]Negative
miR-30a-5p↑ YWHAG↓ apoptosis
↓ oxidative stress
↓ inflammation
[186]Negative
miR-30c↓ SOX9/MAPK↓ apoptosis[103]Positive
miR-30c-5p↑ Rock2/MAPK↓ apoptosis[104]Negative
miR-32-3p↓ Cab39/AMPK↑ apoptosis
↑ oxidative stress
[187]Negative
miR-325-3p↓ RIP3↓ apoptosis[188]Positive
miR-32-5p↓ PTEN/PI3K/AKT↓ cell necrosis[189]Positive
miR-337-3p↑ YBX1↓ apoptosis[190]Negative
miR-342-5p↑ PFN1↑ apoptosis[191]Positive
CCAR2/Akt/NF-κB↓ apoptosis[192]Positive
miR-361-3p↑ NACC1/PINK1/Parkin↓ apoptosis
↓ oxidative stress
[193]Positive
miR-363-3p↓ DUSP5↓ cell necrosis[194]Positive
miR-370↓ SIRT6/Nrf2/ARE↓ apoptosis
↓ oxidative stress
↓ inflammation
[108]Positive
miR-370-3p↓ MAPK1↑ BBB permeability[109]Negative
miR-372-3p↓ TLR4↓ apoptosis[195]Positive
miR-383-5p↓ HDAC9↓ apoptosis
↓ endoplasmic reticulum stress
[196]Positive
miR-410↓ FOXO3↓ apoptosis
↓ oxidative stress
[197]Positive
miR-421-3p *↑ MEF2C↓ apoptosis[198]Negative
↓ YTHDF1/NF-κB p65↓ inflammation[199]Positive
miR-449a↓ AREG/EGFR/PI3K/Akt↓ apoptosis[105]Positive
miR-455-3p↓ PDCD7↓ apoptosis[200]Positive
↓ TP53INP1↓ apoptosis
↓ oxidative stress
↓ inflammation
[201]Positive
miR-485↑ AIM2↑ pyroptosis[110]Positive
miR-485-5p↓ Rac1/Notch2↓ apoptosis
↓ inflammation
[202]Positive
miR-488-3p↓ RAC1↓ apoptosis
↓ inflammation
[203]Positive
↓ VPS4B↓ cell necrosis[204]Positive
miR-489-3p *↓ HDAC2↓ apoptosis[205]Positive
↓ SIRT1↑ apoptosis
↑ oxidative stress
[206]Negative
miR-494-3p↓ Bhlhe40↑ apoptosis
↑ oxidative stress
[207]Negative
miR-496↓ BCL2L14↓ apoptosis[98]Positive
miR-497↑ bcl-2/bcl-w↓ apoptosis[208]Negative
miR-499a-5p↓ PDCD4↓ apoptosis[209]Positive
miR-520a-3p↓ IRF9↓ apoptosis[210]Positive
miR-532-5p↓ CXCL1/CXCR2/NF-κB↓ apoptosis
↓ oxidative stress
↓ inflammation
[211]Positive
miR-613↓ ATG3↓ apoptosis[212]Positive
miR-652-3p↑ Bcl-2
↓ Bax
↓ apoptosis[213]Positive
miR-665-3p ↓ TRIM8/NF-κB↓ apoptosis
↓ inflammation
[214]Positive
miR-666-3p↓ MAPK1↓ apoptosis[215]Positive
miR-7-5p↑ RelA p65↓ apoptosis
↓ oxidative stress
↓ inflammation
[216]Negative
miR-874↓ BMF/BCL2L13↓ apoptosis[99]Positive
miR-874-3p↓ ATG16L1↓ apoptosis[217]Positive
miR-92b↑ NOX4↑ BBB damage[218]Positive
miR-92b-3p↓ TRAF3↓ apoptosis
↓ mitochondrial dysfunction
↓ inflammation
[100]Positive
↓ NOX4↓ apoptosis
↓ oxidative stress
↓ inflammation
[100]Positive
miR-9-5p↑ FOXO/Relaxin↓ apoptosis[122]Negative
miR-98-5p↓ BCL2L13↓ apoptosis
↓ oxidative stress
↓ inflammation
↓ endoplasmic reticulum stress
[219]Positive
miR-99b↓ IGF1R↓ apoptosis[220]Positive
Note: *—conflicting data from different studies. ↓—inhibition; ↑—induction; ACVR1—Activin A receptor type I; ADIPOR2—Adiponectin receptor 2; AIM2—absent in melanoma 2; AKT—alpha serine/threonine-protein kinase; AMPK—Adenosine monophosphate-activated protein kinase; AQP4—Aquaporin-4; ARE—antioxidant response element; AREG—amphiregulin; ATG16L1—autophagy-related 16 like 1; ATG3/7/12—autophagic conjugate 3/7/12; Bax—BCL2-associated X protein; BBB—blood–brain barrier; bcl-2—B-cell lymphoma 2; BCL2L13/14—B-cell lymphoma 2-like protein 13/14; bcl-w—anti-apoptotic member of the BCL-2 family of proteins; Bhlhe40—basic helix–loop–helix family member e40; BMF—bone marrow failure; BMP—basic metabolic panel; BOK—Bcl-2-related ovarian killer; Brg1—brahma-related gene-1; Cab39—calcium-binding protein 39; CAMKK2—calcium/calmodulin-dependent protein kinase 2; CCAR2—cell cycle and apoptosis regulator 2; CD—cluster of differentiation; COX-2—Cyclooxygenase-2; CTDSP1—C-terminal domain small phosphatase 1; CXCL1/2/12—chemokine (C–X–C motif) ligand 1/2/12; CysLT 2 R—Cysteinyl leukotriene receptor 2; DAPK1—death-associated protein kinase 1; DOCK4—dedicator for cytokinesis 4; DUSP1/5/14—dual specificity phosphatase 1/5/14; EGFR—epidermal growth factor receptor; eIF4E—eukaryotic translation initiation factor 4E; En2—Engrailed-2; ERK—extracellular signal-regulated kinase; FBXO3—F-Box protein 3; FGF15—fibroblast growth factor 15; FIP200—family-interacting protein of 200 kDa; FOXO1/3—forkhead box transcription factors 1/3; FSTL1—follistatin-like 1; FSTL1—follistatin-related protein 1; GFAP—glial fibrillary acidic protein; GRP78—Glucose-regulated protein 78; GSK-3β—Glycogen synthase kinase-3 beta; HDAC2—Histone Deacetylase 2; HDAC9—Histone Deacetylase 9; HIF-1α—hypoxia-inducible factor 1-alpha; HO-1—Heme oxygenase-1; HSPA12B—heat shock protein A12B; IGF1R—insulin-like growth factor 1 receptor; IGFBP3—insulin-like growth factor binding protein 3; IkBa—nuclear factor of kappa light polypeptide gene enhancer in B-cells inhibitor, alpha; IL-1β/6/17A/18—interleukin-1β/6/17A/18; iNOS—inducible nitric oxide synthase; IRF9—interferon regulatory factor 9; JAK2—Janus Kinase 2; KDM6B—Lysine demethylase 6B; KLF5/10—Krueppel-like factor 5/10; MafB—V-maf musculoaponeurotic fibrosarcoma oncogene homolog B; MAPK—mitogen-activated protein kinase; MDA—maternally-derived antibodies; MDM4—Murine double minute 4; MEF2C—myocyte-specific enhancer factor 2C; MMP-9- Matrix metalloproteinase 9; mTOR—mammalian target of rapamycin; NACC1—nucleus accumbens-associated 1; NF-κB—nuclear factor kappa B; NLRP3—nucleotide-binding domain, leucine-rich family, pyrin domain-containing-3; N-myc—N-myc proto-oncogene; NOTCH2—neurogenic locus notch homolog protein 2; NOX4—NADPH oxidase 4; NR3C2—nuclear receptor subfamily 3, group C, member 2; Nrep—neuronal protein; Nrf2—nuclear factor erythroid 2–related factor 2; NRP1—neuropilin-1; p27 Kip1—Cyclin-dependent kinase inhibitor 1B; PBX1—pre-B-cell leukemia transcription factor 1; PDCD4/5/7—programmed cell death 4/5/7; PFN1—Profilin 1; PGD2/E2—Prostaglandin D2/E2; PINK1—PTEN-induced putative kinase 1; PKC—Protein Kinase C; p-PI3K—Phosphoinositide 3-kinases; PROK2—Prokineticin 2; PSD-93—post-synaptic density 93; PTEN—Phosphatase and Tensin homolog; Rac1—Ras-related C3 botulinum toxin substrate 1; RelA p65—v-rel avian reticuloendotheliosis viral oncogene homolog A; Rgs1—regulator of G protein signaling 1; RIP3—receptor-interacting Serine/Threonine-protein Kinase 3; Rnf38—RING finger protein 38; Rock2—Rho-associated coil-containing protein kinase 2; ROS—reactive oxygen species; SEMA3A—Semaphorin3a; SIRT1/6—Sirtuin 1/6; SOD—Sphincter of Oddi Dysfunction; SOX9—SRY-Box transcription factor 9; SPHK1—sphingosine kinase 1; SRC—non-receptor tyrosine kinases; STAT3—signal transducer and activator of transcription 3; TLR4—Toll-like receptor 4; TNF-α—tumor necrosis factor alpha; TP53INP1—tumor protein 53-induced nuclear protein 1; TRAF3—tumor necrosis factor receptor-associated factor 3; TRIM8—Tripartite motif-containing 8; TXNIP—Thioredoxin-interacting protein; Tyro3—Tyrosine-protein kinase receptor 3-member; ULK2—Unc-51-like autophagy-activating Kinase 2; USP22—Ubiquitin-specific Peptidase 22; VPS4B—Vacuolar protein-sorting 4B; Wnt—Wingless-related integration site; YBX1—Y-Box binding protein 1; YTHDF1—YTH domain family, member 1; YWHAG—Tryptophan 5-Monooxygenase activation protein gamma; ZO-1—zonula occludens 1; β-TrCP—beta-Transducin repeat-containing E3 Ubiquitin protein Ligase.
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

Tregub, P.P.; Ibrahimli, I.; Averchuk, A.S.; Salmina, A.B.; Litvitskiy, P.F.; Manasova, Z.S.; Popova, I.A. The Role of microRNAs in Epigenetic Regulation of Signaling Pathways in Neurological Pathologies. Int. J. Mol. Sci. 2023, 24, 12899. https://doi.org/10.3390/ijms241612899

AMA Style

Tregub PP, Ibrahimli I, Averchuk AS, Salmina AB, Litvitskiy PF, Manasova ZS, Popova IA. The Role of microRNAs in Epigenetic Regulation of Signaling Pathways in Neurological Pathologies. International Journal of Molecular Sciences. 2023; 24(16):12899. https://doi.org/10.3390/ijms241612899

Chicago/Turabian Style

Tregub, Pavel P., Irada Ibrahimli, Anton S. Averchuk, Alla B. Salmina, Peter F. Litvitskiy, Zaripat Sh. Manasova, and Inga A. Popova. 2023. "The Role of microRNAs in Epigenetic Regulation of Signaling Pathways in Neurological Pathologies" International Journal of Molecular Sciences 24, no. 16: 12899. https://doi.org/10.3390/ijms241612899

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